Worldwide Predatory Insects in Agroecosystems 9819909996, 9789819909995

This book brings out the world record of various predatory insects and their role in pestiferous insect management in a

133 28 14MB

English Pages 880 [868] Year 2023

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Introduction
Contents
About the Authors
1: Predatory Insects Collection Methods
1.1 Introduction
1.2 Heteroptera
1.3 Hymenoptera
1.4 Dermaptera
1.5 Coleoptera
1.5.1 Coccinellidae
1.5.2 Carabidae
1.6 Neuroptera
1.7 Thysanoptera
1.8 Collection Device
1.8.1 Direct Observation/Visual Count Method/Whole Plant Samples
1.8.2 Knocked Down/Plant Shaking Methods
1.8.3 Sweep Net Method
1.8.4 Leaf Vacuum/Vacuum Net Method
1.8.5 Square Beating Trays
1.8.6 Traps
1.8.6.1 Light Trap
1.8.6.2 Suction Traps
1.8.6.3 Sticky Cards
1.8.6.4 Pitfall Traps
1.8.6.5 Malaise Traps
1.8.6.5.1 Townes Trap
1.8.6.5.2 Gressitt Trap
1.8.6.5.3 Schacht Trap
1.8.6.6 Combinations of Traps
1.8.6.7 Other Methods
1.9 Combinations of Methods
1.10 Recommendations´
References
2: Distribution and Diversity of Predatory Insects in Agroecosystems
2.1 Introduction
2.2 Distribution of Predatory Insects at Sole/Inter Crop(s)
2.3 Distribution of Predatory Insects in Genetically Modified Crop(s)
2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise
2.4.1 Mantodea
2.4.2 Odonata
2.4.3 Diptera
2.4.4 Hymenoptera
2.4.5 Distribution of Dermaptera
2.4.6 Distribution and Diversity of Various Coleopterans
2.4.7 Neuroptera: Chrysopidae
2.4.8 Hemiptera/Heteroptera: Pentatomidae, Miridae, Geocoridae, Anthocoridae, Nabidae, Reduviidae
2.4.9 Orthoptera
2.5 Factors Responsible for Predators´ Populations
2.5.1 Pests and Their Natural Enemy´s Complex
2.5.2 Conventional Crops
2.5.3 Genetically Modified Crops
2.5.4 Field Conditions
2.5.5 Pesticides Application
2.5.6 Landscape/Urbanisation
2.6 Recommendation
References
3: Identification of Various Insect Predators
3.1 Coleoptera
3.1.1 Coccinellidae
3.2 Dermaptera
3.3 Orthoptera
3.4 Odonata
3.5 Neuroptera
3.6 Thysanoptera
3.7 Heteroptera
3.7.1 Lygaeidae
3.7.2 Anthocoridae
3.7.3 Reduviidae
3.7.4 Nabidae
3.7.5 Pentatomidae
3.7.6 Miridae
3.8 General Features of Vespidae
3.9 General Features of Formicidae
3.10 Predatory Lepidopterans
References
E-Resources
4: General Ecology of Insect Predators
4.1 Introduction
4.2 Association with Tropics
4.2.1 Hymenoptera
4.2.2 Syrphids
4.2.3 Neuroptera
4.2.4 Nabide
4.2.4.1 Agroecosystems
4.2.5 Coleoptera
4.2.5.1 Carabidae
4.2.5.2 Coccinellidae
4.2.6 Reduviidae
4.2.7 Anthecoridae
4.2.8 Dermaptera
4.3 Intra- and Inter-Specific Relations
4.4 Niche
4.5 Future Areas
References
5: Egg Biology of Insect Predators
5.1 General Structure of Insect Eggs
5.2 Incubation
5.3 Description of Eggs
5.3.1 Mantidae
5.3.2 Hymenoptera
5.3.3 Coccinellids
5.3.4 Carabide
5.3.5 Syrphids
5.3.6 Neuropteran
5.3.7 Reduviid
5.3.8 Nabidae
5.3.9 Anthocorids
5.3.10 Pentatomidae
5.3.11 Thysanoptera
5.4 Egg-Laying Pattern and Protecting Mechanism
5.4.1 Plant Chemistry
5.4.2 Volatile and Non-volatile Organic Compounds (VOCs)
5.4.2.1 Coccinellids
5.4.2.2 Syrphids
5.4.2.3 Chrysopidae
5.4.3 Protecting Mechanism
5.4.3.1 Site Selection
5.4.3.1.1 Anthocorid
5.4.3.2 Diptera
5.4.3.3 Reduviids
5.4.3.4 Nabidae
5.4.3.5 Syrphids
5.4.4 Plants and Their Phenology
5.4.5 Other Factors
5.4.5.1 Natural or Laboratory Hosts
5.4.5.1.1 Coleoptera
5.4.6 Inter-, Intra-guild Predation and Cannibalism (IGP)
5.5 Egg Dumping
5.6 Future Focus
References
6: Insect Predators Immature Stages Biology
6.1 Introduction
6.2 Neuropteran Larva
6.3 Lygeidae
6.4 Hemiptera
6.4.1 Nabidae
6.4.2 Pentatomidae
6.4.3 Miridae
6.4.4 Reduviids
6.4.5 Anthocoridae
6.4.6 Geocoridae
6.5 Syrphidae
6.5.1 Larval General Characters
6.5.2 Pupa
6.5.3 Eclosion
6.5.4 Development Periods
6.5.5 Survival
6.5.6 Larval Feeding Behaviour
6.6 Odonata
6.7 Dermaptera
6.8 Coccinelidae
6.9 Thysanoptera (Thrips)
6.10 Mantodea
6.11 Tri-Trophic Interactions of Bt Toxins-Preys-Predator
6.12 Future Area of Research
References
7: Predatory Insects: Adults Biology of Various Orders
7.1 Introduction
7.2 Coleopterans
7.2.1 Plant Varieties/Cultivars on Biology
7.3 Mantodea
7.4 Syrphids
7.5 Dermaptera
7.6 7.6. Hemipteran Predators
7.6.1 Reduviids
7.6.1.1 Reduviids and Storage Pests
7.6.2 Nabidae
7.6.3 Miridae Biology
7.6.4 Anthecoridae
7.6.5 Lygaeidae
7.7 Lacewings (Neuropteran)
7.8 Miscellaneous Predators (Orthopera, Lepidotpera, Others)
7.9 GM Crops or Bt Proteins on the Biology of Predators
7.10 Future Area of Research
References
8: Polymorphisms in Insect Predators
8.1 General Introduction
8.2 Definitions
8.3 Wing Polymorphism
8.3.1 Neuropteran: Lacewings
8.3.2 Coleopteran
8.3.3 Syrphid
8.3.4 Redviidae
8.3.5 Nabidae
8.3.6 Pentatomidae
8.3.7 Odonates
8.3.8 Anthocoridae
8.3.9 Hymenoptera
8.4 Molecular Mechanisms
8.5 Sexual Dimorphism
8.5.1 Mantodean
8.5.2 Orthoptera
8.5.3 Odonate
8.5.4 Anthocoridae
8.5.5 Reduviids
8.6 Mimic Polymorphism
8.7 General Reasons for Polymorphism
8.8 Future Directions
References
Untitled
9: Influence of Ecological/Climatic Change
9.1 Introduction
9.2 Bioefficiency
9.2.1 Dermaptera
9.2.2 Miridae
9.2.3 Pentatomidae
9.2.4 Anthocoridae
9.2.5 Reduviidae
9.2.6 Coleoptera
9.3 Biology
9.3.1 Dermaptera
9.3.2 Damsel Bugs
9.3.3 Aleyrodidae
9.3.4 Anthocoridae
9.3.5 Miridae Biology
9.3.6 Geocoridae
9.3.7 Reduviidae
9.3.8 Coleoptera
9.3.9 Thrips
9.4 Macromolecular and Antioxidant Responses
9.4.1 Macromolecular Profile
9.4.2 Antioxidant Responses
9.5 Temperature Tolerance Factors
9.5.1 Cuticular Permeability
9.6 Cold Storage
9.6.1 Reduviids
9.6.2 Pentatomidae
9.6.3 Nabidae
9.6.4 Geocoridae
9.6.5 Neuroptera
9.6.6 Coleoptera
9.7 Wind Velocity
9.8 Distribution at Field Level
9.9 Elevated Atmospheric CO2
9.10 Conclusions
References
10: Predation Ethology of Various Orders
10.1 Introduction
10.2 Characteristics of Good Predator
10.3 Hymenoptera
10.3.1 Lepidoptera as a Prey
10.3.2 Collembolans as a Prey
10.3.3 Millipedes as Prey
10.3.4 Beetle and Weevil as Prey
10.3.5 Group Feeding
10.4 Coleopterans
10.5 Isopterans
10.6 Dermaptera
10.7 Comparison Between Coleoptera and Dermaptera
10.8 Orthoptera
10.9 Hemiptera
10.9.1 Reduviids
10.9.2 Miridae
10.9.3 Nabidae
10.9.4 Pentatomidae
10.9.4.1 Nymphal Foraging Behaviour
10.9.4.2 Adult Feeding Behaviour
10.9.5 Anthocoridae
10.9.6 Neuroptera
10.9.7 Syrphids
10.9.8 Diptera
10.10 Expression of Bt Toxins Through Their Preys
10.10.1 In Cotton-Coleoptera
10.10.2 In Rice
10.10.3 Geocoridae
10.10.4 Neuroptera
10.10.5 Thysanura
10.11 Intraguild Predation (IGP) or Competition Dominates
10.12 Best Examples
10.12.1 Miridae and Anthocoridae
10.12.2 Coleoptera and Coleoptera
10.12.3 Coleoptera and Dermaptera
10.12.4 Coleoptera and Neuroptera
10.12.5 Coleoptera, Hemiptera, and Diptera
10.12.6 Among Hemipterans
10.13 Role of Molecular and Other Techniques
10.14 Anti-Predatory Acts
10.15 Chemical Ecology
10.15.1 Field Application
10.16 Future Recommendation
References
11: Mating Behaviour and Reproductive Biology of Insect Predators
11.1 Introduction
11.2 Multiple Mating
11.3 Dictyoptera
11.4 Hemiptera
11.4.1 Reduviidae
11.4.2 Pentatomidae
11.4.3 Anthocoridae
11.5 Syrphids
11.6 Hymenoptera
11.7 Dermaptera
11.8 Coleoptera
11.9 Chemical Ecology
11.10 Future Recommendations
References
12: Offense and Defence Mechanism of Insect Predators
12.1 Introduction
12.2 Behavioural Defence
12.3 Structural Defences of Plants-Herbivory-Predators
12.3.1 Mouthparts
12.3.2 Claw
12.3.3 Plant Trichomes
12.3.4 Epicuticular Waxes
12.4 Chemical Defence
12.4.1 Cardiotonic Steroids (CTS) as a Chemical Defence
12.5 Cryptic Colouration
12.5.1 Aposematism
12.5.2 Green-Brown Polymorphism
12.6 Masquerade
12.7 Parental Care
12.7.1 Types
12.7.2 Reduviidae
12.7.3 Dermaptera
12.7.4 Hymenoptera
12.7.5 Coleoptera
12.8 Hibernation
12.9 Camouflaging
12.9.1 Mantids
12.9.2 Reduviids
12.9.3 Pentatomidae
12.10 Recommendations for Future Works
References
13: Venomous and Other Body Fluids in Insect Predators
13.1 Introduction
13.2 Chemistry
13.3 Salivary Gland, Venomous Saliva Collection, and Quantification
13.3.1 Salivary Gland
13.3.2 Venomous Saliva Collection Methods
13.3.3 Saliva Genders on Venom Quantity
13.3.4 Impact of Prey Deprivation on Venom Quantity
13.3.5 Influence of Prey on VS Yield
13.4 Venom Chemical Composition
13.4.1 Proteinaceous Components
13.4.2 Enzymes
13.4.3 Genes
13.5 Non-Proteinaceous Components
13.6 Salivary Venom of Predators
13.6.1 Hymenoptera
13.6.2 Pentatomidae
13.6.3 Reduviidae
13.6.4 Diptera
13.6.5 Dermaptera
13.6.6 Neuroptera
13.6.7 Coleoptera
13.7 Other Fluids
13.8 Biological Activities
13.8.1 Pesticidal Activity
13.8.2 Primary Metabolites Modulation
13.8.3 Immunomodulatory Activity
13.8.4 Inhibition of Haemocyte Aggregation
13.8.5 Spreading Inhibitory Behaviour
13.8.6 Anti-Microbial and Cytotoxic Activities
13.9 Physiological Role
13.10 Future Recommendations
References
14: Prey Record of Various Predators
14.1 Coleoptera
14.2 Orthoptera
14.3 Neuroptera
14.4 Hymenoptera
14.5 Diptera
14.6 Dermaptera
14.7 Hetroptera
14.8 Future Recommendations
References
15: Mass Production of Insect Predators
15.1 Introduction
15.1.1 Levels of Mass Production
15.1.2 Concepts of Artificial Diet
15.2 Hemipteran Predators
15.2.1 Anthocoridae
15.2.2 Nabidae
15.2.3 Lygeidae
15.2.3.1 Artificial Diet for Lygeidae
15.2.4 Reduviids
15.2.4.1 Artificial Diet
15.2.4.2 Consistency of Artificial Diet
15.2.4.3 Preparation
15.2.4.4 Object Preference
15.2.4.5 Feeding Behaviour on Meridic Diet
15.2.4.6 Feeding Behavioral on Meridic Diet in Relation to Starvation
15.2.5 Miridae
15.2.6 Pentatomidae
15.3 Syrphids
15.4 Lacewings
15.5 Coleoptera
15.5.1 Mass of Adults and Eggs and Ovariole Number
15.6 Quality Control
15.7 Future Recommendation
References
16: Bioefficacy of Insect Predators Under Laboratory
16.1 Introduction
16.2 Predators in General
16.3 Hemipteran Predators
16.3.1 Anthecoridae
16.3.1.1 Numerical Response
16.3.2 Pentatomidae
16.3.3 Lygaeidae
16.3.4 Nabidae
16.3.5 Miridae
16.3.6 Reduviidae
16.4 Lacewings
16.5 Coleoptera
16.6 Trichoptera
16.7 Storage Pests
16.8 Factors Influencing Functional Responses
16.8.1 Plant Species and Their Morphology
16.8.2 Pesticides and Biopesticides
16.9 Interaction Multiple Natural Enemies
16.10 Recommendations
References
17: Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen House/Polyphagous
17.1 Introduction
17.2 Hemiptera
17.2.1 Anthocoridae
17.2.2 Reduviids
17.2.3 Pentatomide
17.2.4 Miridae
17.2.5 Entatomidae
17.3 Coleoptera
17.4 Neuroptera
17.5 Multiple Predators
17.5.1 Predation by a Complex of Predators Made Up of Adults of the Lygaeid
17.6 Multiple Controlled Conditions in Field Cage for Life Traits and Bioefficacy
17.7 Recommendation
References
18: Controlled Field Cage and Field Evolution
18.1 Introduction
18.2 Predators in General
18.3 Hemipterans
18.3.1 Miridae
18.3.2 Reduviidae
18.3.3 Anthocoridae
18.3.4 Pentatomidae
18.4 Neuroptera
18.5 Coleoptera
18.6 Dermaptera
18.7 Syrphids
18.7.1 Effectiveness of Syrphids as Field-Based Biocontrol Agents
18.8 Predator Density
18.9 Comparison Between Organic and Conventional Fields
18.10 Genetically Modified (GM) Crops
18.11 Bio-Intensive Pest Management (BIPM)
18.12 Recommendations
References
19: Commercially Available Predators
19.1 Overview
19.2 Commercial Products
19.2.1 University Extension Service
19.2.2 Commercial Producers
19.2.2.1 India
19.2.2.2 Korea
19.2.2.3 Greenhouses
19.3 Quality Control
19.3.1 Species Identification Is Verified
19.3.2 Stock Deterioration
19.3.3 Shipping and Handling
19.4 Future Focus
Annexure 19.1: Contact Address for Selected Commercial Producers
Annexure 19.2: Company Website
Annexure 19.3: Government, University, Commercial and Non-profit Websites on Biological Control (O´Neil et al. 2003)
United States
References
20: Biosafety Assessment of Synthetic Pesticides
20.1 Other Molecules
20.2 Systemic Insecticides
20.3 Hemiptera
20.3.1 Pentatomidae
20.3.2 Lygaeidae
20.3.3 Anthocoridae
20.3.4 Reduviidae
20.4 Lacewings
20.5 Dermapterans
20.6 Coleopterans
20.7 Neuroptera
20.8 Syrphids
20.9 Field Conditions
20.10 Insect Growth Regulators (IGRs)
20.10.1 Botanical IGRs
20.11 Neuropeptides
20.12 Indirect Effects of Pesticides on Natural Enemies
20.12.1 Development, Adult Longevity, and Reproduction
20.12.2 Functional Response
20.12.3 Biological Traits: Bioefficacy
20.13 Fungicides
20.14 Beneficial Impacts: Hormesis
20.15 Future Recommendations
References
21: Biocompatibility of Biopesticides with Predatory Insects
21.1 Role of Biopesticides in Insect Pest Management
21.2 Microbial Pesticides
21.2.1 Coleoptera
21.2.2 Heteroptera
21.2.2.1 Reduviids
21.2.2.2 Pentatomidae
21.2.3 Neuroptera
21.2.4 Other Predators
21.2.5 Microbial Metabolites
21.2.5.1 Spinosad
21.2.5.1.1 Cry Proteins
21.2.5.2 Abamectin
21.2.5.2.1 Hemiptera
21.2.5.2.2 Coleoptera
21.2.5.2.3 Larvae of the Neuroptera
21.2.5.2.4 Dermaptera
21.2.5.2.5 Various Insects
21.2.6 Genetically Modified (GM)/Recombinant Microorganisms
21.2.6.1 Hemiptera
21.2.6.2 Chrysoperla spp.
21.2.6.3 Indian Perspective
21.2.6.4 Spain´s Perspective
21.2.6.5 Turkey´s Perspective
21.2.6.6 Philippines´ Viewpoint
21.2.6.7 Brazil and China´s Viewpoints
21.2.6.8 Korean Perspective
21.3 Botanical Biopesticides
21.3.1 Coleoptera
21.3.2 Neuroptera
21.3.3 Heteroptera
21.3.3.1 Miridae
21.3.4 Different Predators
21.3.5 Combined Effects of Different Botanicals
21.3.6 Botanical Products, Including Oil
21.4 Field Trials
21.4.1 Botanicals
21.4.2 Botanical Products
21.4.3 Post-harvest Storage
21.5 Recommendations
References
Index
Recommend Papers

Worldwide Predatory Insects in Agroecosystems
 9819909996, 9789819909995

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Kitherian Sahayaraj Errol Hassan

Worldwide Predatory Insects in Agroecosystems

Worldwide Predatory Insects in Agroecosystems

Kitherian Sahayaraj • Errol Hassan

Worldwide Predatory Insects in Agroecosystems

Kitherian Sahayaraj Department of Zoology St. Xavier’s College Palayamkottai, Tamil Nadu, India

Errol Hassan School of Agriculture and Food Sciences The University of Queensland Gatton, QLD, Australia

ISBN 978-981-99-1000-7 ISBN 978-981-99-0999-5 https://doi.org/10.1007/978-981-99-1000-7

(eBook)

# The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd. The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721, Singapore

Dedicated to our Parents in Heaven

Introduction

This book covers the predatory insects at a global level, which are generally considered as natural enemies that make them potentially important agents for the eco-friendly pest control of economically important crop pests. They belong to Coleoptera, Neuroptera, Hemiptera, Dermatoptera, Hymenoptera, etc. Many books are available in the literature about each and every insect order. However, no one tried to bring out the information of predators from a global perspective. Chapters included in this book are predatory insects collection methods, distribution and diversity in agro-ecosystems, general characteristics and diagnosis of various insect predators, general ecology, egg biology of various orders, immature stages biology, adults biology of various orders, polymorphisms, influence of ecological/climatic change, predation ethology of various orders, mating behaviour and reproductive biology, offensive and defensive mechanism, venomous and other body fluids, prey records, their stage and selection and chemical ecology, mass production, bioefficacy under laboratory, pot condition and controlled field cage and field evolution, commercially available predators, biosafety to synthetic pesticides, biocompatibility of biopesticides to predatory insects, concluding remarks and future recommendations, detailed contents and citations. Many novel concepts such as global warming and predatory insects; egg, nymph, and adult biology; chemical ecology in prey stage and prey preference of predators; mass production using natural, factitious and artificial diets; venom role in predator; marketable natural enemies and list of organisations engaged in this aspect; biosafety of various biopesticides; natural enemies enhancement by botanicals; compatibility of botanicals/microbial and natural enemies, also trophic level interactions; the impact of various genetically engineered crops on the distribution of predators and their synergistic interaction with cry proteins; bio-nano pesticides and their utility values in pest control are included in the book. The uniqueness of this book lies in its distinctive features. The book covers all aspects (A-Z) of predatory insects, providing a global perspective. Each chapter delves into relevant content related to the title, supported by the latest evidence. It offers a comprehensive review of the distribution, biology, mass production, release,

vii

viii

Introduction

and evaluation of predatory insects. Its content is easy, catering to both graduate students and researchers. All chapters include well-supported evidence through citations, tables, illustrations, and photographs. Detailed contents with Species Index and Content Index are provided, and the book features updated citations that are valuable for those looking to write a research manuscript.

Contents

1

2

Predatory Insects Collection Methods . . . . . . . . . . . . . . . . . . . . . . . 1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Heteroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5.1 Coccinellidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5.2 Carabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.6 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.7 Thysanoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8 Collection Device . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.1 Direct Observation/Visual Count Method/Whole Plant Samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.2 Knocked Down/Plant Shaking Methods . . . . . . . . . . 1.8.3 Sweep Net Method . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.4 Leaf Vacuum/Vacuum Net Method . . . . . . . . . . . . . 1.8.5 Square Beating Trays . . . . . . . . . . . . . . . . . . . . . . . 1.8.6 Traps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.9 Combinations of Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.10 Recommendations’ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Distribution and Diversity of Predatory Insects in Agroecosystems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Distribution of Predatory Insects at Sole/Inter Crop(s) . . . . . . . 2.3 Distribution of Predatory Insects in Genetically Modified Crop(s) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.1 Mantodea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.2 Odonata . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.3 Diptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1 1 2 4 4 5 5 5 5 6 6 7 9 9 10 10 11 18 20 20 25 25 26 33 34 34 34 37 ix

x

Contents

2.4.4 2.4.5 2.4.6

Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Distribution of Dermaptera . . . . . . . . . . . . . . . . . . . Distribution and Diversity of Various Coleopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.7 Neuroptera: Chrysopidae . . . . . . . . . . . . . . . . . . . . . 2.4.8 Hemiptera/Heteroptera: Pentatomidae, Miridae, Geocoridae, Anthocoridae, Nabidae, Reduviidae . . . . 2.4.9 Orthoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Factors Responsible for Predators’ Populations . . . . . . . . . . . . 2.5.1 Pests and Their Natural Enemy’s Complex . . . . . . . . 2.5.2 Conventional Crops . . . . . . . . . . . . . . . . . . . . . . . . 2.5.3 Genetically Modified Crops . . . . . . . . . . . . . . . . . . . 2.5.4 Field Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.5 Pesticides Application . . . . . . . . . . . . . . . . . . . . . . . 2.5.6 Landscape/Urbanisation . . . . . . . . . . . . . . . . . . . . . 2.6 Recommendation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39 39

46 54 54 54 54 55 57 58 58 59 60

3

Identification of Various Insect Predators . . . . . . . . . . . . . . . . . . . . 3.1 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.1 Coccinellidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Orthoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 Odonata . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6 Thysanoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7 Heteroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.1 Lygaeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.2 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.3 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.4 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.5 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.6 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.8 General Features of Vespidae . . . . . . . . . . . . . . . . . . . . . . . . . 3.9 General Features of Formicidae . . . . . . . . . . . . . . . . . . . . . . . 3.10 Predatory Lepidopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . E-Resources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

71 72 72 74 76 78 79 81 83 84 85 86 88 90 91 92 93 93 93 94

4

General Ecology of Insect Predators . . . . . . . . . . . . . . . . . . . . . . . . 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Association with Tropics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.1 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.2 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.3 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

95 95 96 96 96 97

41 44

Contents

5

6

xi

4.2.4 Nabide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.6 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.7 Anthecoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.8 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Intra- and Inter-Specific Relations . . . . . . . . . . . . . . . . . . . . . 4.4 Niche . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5 Future Areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

98 99 101 104 105 106 108 110 110

Egg Biology of Insect Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1 General Structure of Insect Eggs . . . . . . . . . . . . . . . . . . . . . . 5.2 Incubation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3 Description of Eggs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.1 Mantidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.2 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.3 Coccinellids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.4 Carabide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.5 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.6 Neuropteran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.7 Reduviid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.8 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.9 Anthocorids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.10 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.11 Thysanoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4 Egg-Laying Pattern and Protecting Mechanism . . . . . . . . . . . . 5.4.1 Plant Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.2 Volatile and Non-volatile Organic Compounds (VOCs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.3 Protecting Mechanism . . . . . . . . . . . . . . . . . . . . . . . 5.4.4 Plants and Their Phenology . . . . . . . . . . . . . . . . . . . 5.4.5 Other Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.6 Inter-, Intra-guild Predation and Cannibalism (IGP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5 Egg Dumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6 Future Focus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

113 113 114 115 115 116 117 117 119 120 120 123 124 124 124 125 125

Insect Predators Immature Stages Biology . . . . . . . . . . . . . . . . . . . 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2 Neuropteran Larva . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3 Lygeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4 Hemiptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.1 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.2 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

149 150 150 153 154 154 156

125 128 131 134 139 141 141 142

xii

Contents

6.4.3 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.4 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.5 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.6 Geocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5 Syrphidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.1 Larval General Characters . . . . . . . . . . . . . . . . . . . . 6.5.2 Pupa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.3 Eclosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.4 Development Periods . . . . . . . . . . . . . . . . . . . . . . . 6.5.5 Survival . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.6 Larval Feeding Behaviour . . . . . . . . . . . . . . . . . . . . 6.6 Odonata . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.7 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.8 Coccinelidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.9 Thysanoptera (Thrips) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.10 Mantodea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.11 Tri-Trophic Interactions of Bt Toxins–Preys–Predator . . . . . . . 6.12 Future Area of Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

158 158 164 166 166 167 168 168 169 169 170 170 171 173 174 175 175 177 177

7

Predatory Insects: Adults Biology of Various Orders . . . . . . . . . . . 7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2 Coleopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2.1 Plant Varieties/Cultivars on Biology . . . . . . . . . . . . 7.3 Mantodea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.5 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6 7.6. Hemipteran Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6.1 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6.2 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6.3 Miridae Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6.4 Anthecoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6.5 Lygaeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.7 Lacewings (Neuropteran) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.8 Miscellaneous Predators (Orthopera, Lepidotpera, Others) . . . . 7.9 GM Crops or Bt Proteins on the Biology of Predators . . . . . . . 7.10 Future Area of Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

183 183 184 190 190 192 195 196 196 203 205 206 207 209 211 213 214 215

8

Polymorphisms in Insect Predators . . . . . . . . . . . . . . . . . . . . . . . . 8.1 General Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3 Wing Polymorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.1 Neuropteran: Lacewings . . . . . . . . . . . . . . . . . . . . 8.3.2 Coleopteran . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

221 221 222 222 224 225

. . . . . .

Contents

9

xiii

8.3.3 Syrphid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.4 Redviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.5 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.6 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.7 Odonates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.8 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.9 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.4 Molecular Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5 Sexual Dimorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5.1 Mantodean . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5.2 Orthoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5.3 Odonate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5.4 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5.5 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.6 Mimic Polymorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.7 General Reasons for Polymorphism . . . . . . . . . . . . . . . . . . . . 8.8 Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

227 228 233 234 235 235 236 237 239 239 240 240 241 242 243 243 243 244

Influence of Ecological/Climatic Change . . . . . . . . . . . . . . . . . . . . . 9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2 Bioefficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.1 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.2 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.3 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.4 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.5 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.6 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3 Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.1 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.2 Damsel Bugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.3 Aleyrodidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.4 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.5 Miridae Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.6 Geocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.7 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.8 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.9 Thrips . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.4 Macromolecular and Antioxidant Responses . . . . . . . . . . . . . . 9.4.1 Macromolecular Profile . . . . . . . . . . . . . . . . . . . . . . 9.4.2 Antioxidant Responses . . . . . . . . . . . . . . . . . . . . . . 9.5 Temperature Tolerance Factors . . . . . . . . . . . . . . . . . . . . . . . 9.5.1 Cuticular Permeability . . . . . . . . . . . . . . . . . . . . . .

249 250 250 251 252 254 256 257 259 261 262 264 265 266 267 268 271 273 274 275 275 276 277 279

xiv

10

Contents

9.6

Cold Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.1 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.2 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.3 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.4 Geocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.5 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.6 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.7 Wind Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.8 Distribution at Field Level . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.9 Elevated Atmospheric CO2 . . . . . . . . . . . . . . . . . . . . . . . . . . 9.10 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

279 280 280 282 282 283 283 285 287 288 291 291

Predation Ethology of Various Orders . . . . . . . . . . . . . . . . . . . . . . 10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.2 Characteristics of Good Predator . . . . . . . . . . . . . . . . . . . . . . 10.3 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.3.1 Lepidoptera as a Prey . . . . . . . . . . . . . . . . . . . . . . . 10.3.2 Collembolans as a Prey . . . . . . . . . . . . . . . . . . . . . . 10.3.3 Millipedes as Prey . . . . . . . . . . . . . . . . . . . . . . . . . 10.3.4 Beetle and Weevil as Prey . . . . . . . . . . . . . . . . . . . . 10.3.5 Group Feeding . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.4 Coleopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.5 Isopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.6 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.7 Comparison Between Coleoptera and Dermaptera . . . . . . . . . . 10.8 Orthoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9 Hemiptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.1 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.2 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.3 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.4 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.5 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.6 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.7 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.8 Diptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.10 Expression of Bt Toxins Through Their Preys . . . . . . . . . . . . . 10.10.1 In Cotton-Coleoptera . . . . . . . . . . . . . . . . . . . . . . . 10.10.2 In Rice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.10.3 Geocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.10.4 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.10.5 Thysanura . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.11 Intraguild Predation (IGP) or Competition Dominates . . . . . . .

299 300 301 302 303 303 304 304 304 305 306 307 308 309 309 309 315 316 318 321 321 324 324 325 326 327 328 328 329 332

Contents

10.12

11

12

xv

Best Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.12.1 Miridae and Anthocoridae . . . . . . . . . . . . . . . . . . . . 10.12.2 Coleoptera and Coleoptera . . . . . . . . . . . . . . . . . . . 10.12.3 Coleoptera and Dermaptera . . . . . . . . . . . . . . . . . . . 10.12.4 Coleoptera and Neuroptera . . . . . . . . . . . . . . . . . . . 10.12.5 Coleoptera, Hemiptera, and Diptera . . . . . . . . . . . . . 10.12.6 Among Hemipterans . . . . . . . . . . . . . . . . . . . . . . . . 10.13 Role of Molecular and Other Techniques . . . . . . . . . . . . . . . . 10.14 Anti-Predatory Acts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.15 Chemical Ecology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.15.1 Field Application . . . . . . . . . . . . . . . . . . . . . . . . . . 10.16 Future Recommendation . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

333 333 334 334 335 335 336 337 339 341 345 347 347

Mating Behaviour and Reproductive Biology of Insect Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Multiple Mating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.3 Dictyoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.4 Hemiptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.4.1 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.4.2 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.4.3 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.5 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.6 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.7 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.8 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.9 Chemical Ecology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.10 Future Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

355 355 356 356 358 358 362 364 365 366 368 370 372 373 373

Offense and Defence Mechanism of Insect Predators . . . . . . . . . . . . 12.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.2 Behavioural Defence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.3 Structural Defences of Plants–Herbivory–Predators . . . . . . . . . 12.3.1 Mouthparts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.3.2 Claw . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.3.3 Plant Trichomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.3.4 Epicuticular Waxes . . . . . . . . . . . . . . . . . . . . . . . . . 12.4 Chemical Defence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.4.1 Cardiotonic Steroids (CTS) as a Chemical Defence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.5 Cryptic Colouration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.5.1 Aposematism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.5.2 Green-Brown Polymorphism . . . . . . . . . . . . . . . . . .

377 378 378 379 380 382 382 383 384 388 390 390 391

xvi

13

Contents

12.6 12.7

Masquerade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Parental Care . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7.1 Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7.2 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7.3 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7.4 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.8 Hibernation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.9 Camouflaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.9.1 Mantids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.9.2 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.9.3 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.10 Recommendations for Future Works . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

392 392 392 393 395 395 396 396 398 399 400 402 402 402

Venomous and Other Body Fluids in Insect Predators . . . . . . . . . . 13.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.2 Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3 Salivary Gland, Venomous Saliva Collection, and Quantification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3.1 Salivary Gland . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3.2 Venomous Saliva Collection Methods . . . . . . . . . . . 13.3.3 Saliva Genders on Venom Quantity . . . . . . . . . . . . . 13.3.4 Impact of Prey Deprivation on Venom Quantity . . . . 13.3.5 Influence of Prey on VS Yield . . . . . . . . . . . . . . . . . 13.4 Venom Chemical Composition . . . . . . . . . . . . . . . . . . . . . . . . 13.4.1 Proteinaceous Components . . . . . . . . . . . . . . . . . . . 13.4.2 Enzymes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.4.3 Genes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.5 Non-Proteinaceous Components . . . . . . . . . . . . . . . . . . . . . . . 13.6 Salivary Venom of Predators . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.1 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.2 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.3 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.4 Diptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.5 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.6 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.7 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.7 Other Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.8 Biological Activities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.8.1 Pesticidal Activity . . . . . . . . . . . . . . . . . . . . . . . . . 13.8.2 Primary Metabolites Modulation . . . . . . . . . . . . . . . 13.8.3 Immunomodulatory Activity . . . . . . . . . . . . . . . . . . 13.8.4 Inhibition of Haemocyte Aggregation . . . . . . . . . . . 13.8.5 Spreading Inhibitory Behaviour . . . . . . . . . . . . . . . . 13.8.6 Anti-Microbial and Cytotoxic Activities . . . . . . . . . .

409 410 410 411 411 415 416 417 418 419 419 421 424 425 426 426 428 429 431 432 432 433 434 434 436 437 438 438 439 439

Contents

xvii

13.9 Physiological Role . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 440 13.10 Future Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . 440 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 440 14

Prey Record of Various Predators . . . . . . . . . . . . . . . . . . . . . . . . . . 14.1 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.2 Orthoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.3 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.4 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.5 Diptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.6 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.7 Hetroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.8 Future Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

447 448 460 462 466 470 473 476 478 501

15

Mass Production of Insect Predators . . . . . . . . . . . . . . . . . . . . . . . . 15.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.1.1 Levels of Mass Production . . . . . . . . . . . . . . . . . . . 15.1.2 Concepts of Artificial Diet . . . . . . . . . . . . . . . . . . . . 15.2 Hemipteran Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.1 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.2 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.3 Lygeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.4 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.5 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.6 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.3 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.4 Lacewings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.5.1 Mass of Adults and Eggs and Ovariole Number . . . . 15.6 Quality Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.7 Future Recommendation . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

525 525 526 527 530 530 534 534 536 552 552 554 556 559 564 567 568 572

16

Bioefficacy of Insect Predators Under Laboratory . . . . . . . . . . . . . . 16.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.2 Predators in General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3 Hemipteran Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.1 Anthecoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.2 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Lygaeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.3 16.3.4 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.5 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.6 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.4 Lacewings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

579 579 581 582 582 587 590 591 592 593 600 604

xviii

Contents

16.6 16.7 16.8

Trichoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Storage Pests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Factors Influencing Functional Responses . . . . . . . . . . . . . . . . 16.8.1 Plant Species and Their Morphology . . . . . . . . . . . . 16.8.2 Pesticides and Biopesticides . . . . . . . . . . . . . . . . . . 16.9 Interaction Multiple Natural Enemies . . . . . . . . . . . . . . . . . . . 16.10 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

18

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen House/Polyphagous . . . . . . . . . . . . . . . . . . . . . . . 17.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2 Hemiptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2.1 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2.2 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2.3 Pentatomide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2.4 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2.5 Entatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.3 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.4 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.5 Multiple Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.5.1 Predation by a Complex of Predators Made Up of Adults of the Lygaeid . . . . . . . . . . . . . . . . . . . . . . . 17.6 Multiple Controlled Conditions in Field Cage for Life Traits and Bioefficacy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.7 Recommendation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Controlled Field Cage and Field Evolution . . . . . . . . . . . . . . . . . . . 18.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.2 Predators in General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.3 Hemipterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.3.1 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.3.2 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.3.3 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.3.4 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.4 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.6 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.7 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.7.1 Effectiveness of Syrphids as Field-Based Biocontrol Agents . . . . . . . . . . . . . . . . . . . . . . . . . . Predator Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.8 18.9 Comparison Between Organic and Conventional Fields . . . . . . 18.10 Genetically Modified (GM) Crops . . . . . . . . . . . . . . . . . . . . .

608 608 611 612 614 618 619 619 627 627 629 630 634 636 637 640 641 645 648 650 657 659 659 665 665 666 671 671 672 676 679 681 684 687 687 687 688 688 688

Contents

xix

18.11 Bio-Intensive Pest Management (BIPM) . . . . . . . . . . . . . . . . . 693 18.12 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695 19

20

Commercially Available Predators . . . . . . . . . . . . . . . . . . . . . . . . . 19.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19.2 Commercial Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19.2.1 University Extension Service . . . . . . . . . . . . . . . . . . 19.2.2 Commercial Producers . . . . . . . . . . . . . . . . . . . . . . 19.3 Quality Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19.3.1 Species Identification Is Verified . . . . . . . . . . . . . . . 19.3.2 Stock Deterioration . . . . . . . . . . . . . . . . . . . . . . . . . 19.3.3 Shipping and Handling . . . . . . . . . . . . . . . . . . . . . . 19.4 Future Focus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Annexure 19.1: Contact Address for Selected Commercial Producers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Annexure 19.2: Company Website . . . . . . . . . . . . . . . . . . . . . . . . . . . Annexure 19.3: Government, University, Commercial and Non-profit Websites on Biological Control (O’Neil et al. 2003) . . . . . . . . . . . . . . United States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Biosafety Assessment of Synthetic Pesticides . . . . . . . . . . . . . . . . . . 20.1 Other Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.2 Systemic Insecticides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.3 Hemiptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.3.1 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.3.2 Lygaeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.3.3 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.3.4 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.4 Lacewings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.5 Dermapterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.6 Coleopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.7 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.8 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.9 Field Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.10 Insect Growth Regulators (IGRs) . . . . . . . . . . . . . . . . . . . . . . 20.10.1 Botanical IGRs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.11 Neuropeptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.12 Indirect Effects of Pesticides on Natural Enemies . . . . . . . . . . 20.12.1 Development, Adult Longevity, and Reproduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.12.2 Functional Response . . . . . . . . . . . . . . . . . . . . . . . . 20.12.3 Biological Traits: Bioefficacy . . . . . . . . . . . . . . . . .

703 703 704 705 705 707 712 715 715 716 724 730 731 731 732 735 735 737 738 738 740 741 746 751 752 754 756 757 758 759 764 764 765 767 768 770

xx

21

Contents

20.13 Fungicides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.14 Beneficial Impacts: Hormesis . . . . . . . . . . . . . . . . . . . . . . . . . 20.15 Future Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

771 773 775 775

Biocompatibility of Biopesticides with Predatory Insects . . . . . . . . . 21.1 Role of Biopesticides in Insect Pest Management . . . . . . . . . . 21.2 Microbial Pesticides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.1 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.2 Heteroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.3 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.4 Other Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.5 Microbial Metabolites . . . . . . . . . . . . . . . . . . . . . . . 21.2.6 Genetically Modified (GM)/Recombinant Microorganisms . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3 Botanical Biopesticides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.1 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.2 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.3 Heteroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.4 Different Predators . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.5 Combined Effects of Different Botanicals . . . . . . . . . 21.3.6 Botanical Products, Including Oil . . . . . . . . . . . . . . 21.4 Field Trials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.4.1 Botanicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.4.2 Botanical Products . . . . . . . . . . . . . . . . . . . . . . . . . 21.4.3 Post-harvest Storage . . . . . . . . . . . . . . . . . . . . . . . . 21.5 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

785 785 786 786 789 792 793 793 798 802 806 807 809 810 811 811 812 812 813 814 815 815

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 821

About the Authors

Kitherian Sahayaraj Ph.D., D.Sc., received his undergraduate, postgraduate, and doctoral education from Madurai Kamaraj University, India. Prior to his current position, he worked at St. Joseph’s College and Arulanadar College, India. Over the past 33 years, Dr. Sahayaraj’s research efforts have been dedicated to multidisciplinary, integrated approaches applied to pest management, especially through bio-intensive pest management. Dr. Sahayaraj has over 217 scientific papers published (180 peer-reviewed journals, 15 in proceedings, and 22 book chapters). He has published eight books with national and international publishers. He is an internationally recognised expert in many areas of advanced entomology including insect chemical ecology, artificial diet formulation, biopesticide formulation, bio-intensive integrated pest management, bionanomaterial, and insect molecular biology. He has organised five international conferences (BIOCICON) and seven national conferences. He is the Editor-in-Chief of a journal and also the Editor of more than 10 reputed peer-reviewed journals including Science Reports. He is the reviewer of more than 25 journals. He has been honoured with several awards from regional (Best Researcher in Science, St. Xavier’s College), national (Scientists of the year 2008, NESA, New Delhi; Young Achievers Award 2010 by SADHNA, Solan; Bharat Seva Ratan Gold Medal Award 2014 by GEPRA, New Delhi; Bharat Seven Rethan Gold Medal by GEPBRA, New Delhi), and international agencies (Hyoshi Environmentalist Award, Japan; YOUNG IOBC travel grant award; Excellent Scholar Award by XIX International Botanical Congress, China). Dr. Sahayaraj has operated 14 research projects funded by national (DST, DBT, CSIR, MOEs, MEFs) and international (IFS) funding agencies. He has guided and supervised 22 Ph.D. scholars and supervising five researchers. Errol Hassan Ph.D., Dipl. Eng, received his undergraduate degree from the University of Ankara, Faculty of Agriculture as Dipl. Eng (Ziraat Yuksek Muhendisi). After graduation, he worked at American Aid Office in Ankara, Turkey, as an Agricultural Liaison officer for 2 years. He returned to Cyprus in 1960 and was employed as Agricultural Teacher at a Turkish High School for 2 years. In 1962, he applied for and received a scholarship from the Federal Republic of Germany to study postgraduate in Plant Protection in Hanover at Plant Protection Institute for 2 years. xxi

xxii

About the Authors

After that he received another scholarship from the University of Gottingen from Forest Zoological Institute in 1964 to undertake Doctoral Research under Professor Dr. E. Schimitschek, working on the Ecology and feeding behaviour of parasitic insects utilising flowering plants. He graduated in 1966 with “Magna Cum Laude”. After graduation, he was employed at the same Institute as an assistant Lecturer. Towards the end of 1967, he was offered a position by the Australian External Territories Department as an entomologist at the Research and Surveys Department of Agriculture and Fisheries. His other duty was to provide lectures on Applied Entomology at the University of Papua New Guinea (PNG). The research area covered insect pests of cacao, vegetable, and spices. He worked in PNG until 1973 and then received an offer from the Queensland Agricultural College (QAC) to become a lecturer and teach Plant Protection and Entomology. He was subsequently promoted to Senior Lecturer. In 1990, the Australian Federal Government amalgamated the colleges with Universities and QAC joined the University of Queensland (UQ) and all staff was absorbed by the University.

1

Predatory Insects Collection Methods

Contents 1.1 1.2 1.3 1.4 1.5

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Heteroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5.1 Coccinellidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5.2 Carabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.6 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.7 Thysanoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8 Collection Device . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.1 Direct Observation/Visual Count Method/Whole Plant Samples . . . . . . . . . . . . . . . . 1.8.2 Knocked Down/Plant Shaking Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.3 Sweep Net Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.4 Leaf Vacuum/Vacuum Net Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.5 Square Beating Trays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.8.6 Traps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.9 Combinations of Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.10 Recommendations’ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.1

1 2 4 4 5 5 5 5 6 6 7 9 9 10 10 11 18 20 20

Introduction

Both generalised and specific predators are considered as an important component of the bio-intensive integrated pest management (BIPM) in most of the cropping systems. Long-running discussions have focused on the optimal methods to gather the diverse spectrum of terrestrial arthropods. All entomologists have their own fascinating methods to collect insects with their rich experience and expertise. However, predatory insects could be gathered using either active (sweep netting, foliage pounding, looking under rocks, sifting leaves, peeling tree bark), or passive

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_1

1

2

1

Predatory Insects Collection Methods

(looking under rocks, sifting leaves, etc.) methods (pitfall traps or baited pitfall traps, Malaise and flight intercept traps, light traps and collecting at a light sheet, sticky traps, pheromone traps and yellow pans). Before entering into the actual content of the chapter, we would like to introduce the predatory insects to the readers. Predatory insects (general and specific) or Entomophagbous insects were recorded from more than 11 insect orders; however, the most important and frequently available in various agroecosystems are listed in Table 1.1. Insects may need to be collected and preserved for many reasons: (a) (b) (c) (d)

Collection for later examination and identification Preservation for the master collection Sampling to check population numbers Collection, preservation, and mailing for positive identification.

In general, net, beating tray, aspirator, light traps, pitfall traps, food lures or sex lures, berlese funnel (extraction of insects from soil samples) have been utilised for inset collection. How an insect of each order has been collected is described below.

1.2

Heteroptera

Anthocorid predators are present in every zoogeographical area on the planet and are considered as potential biocontrol agents. Typically referred to as tiny flower bugs or tiny pirate bugs, they consume small lepidopteran larvae, small grubs, psocids, mites, thrips, aphids, other storage pests (Ballal et al. 2016). Generalised numerous greenhouses and horticultural crops have seen a significant increase of predatory bugs of the genus Orius in iron. Typically, samples were taken by beating flowers or terminal buds onto a plate made of white plastic (Hassanzadeh et al. 2013). In another study, ten sweet pepper flowers were randomly collected from plots. Thrips and other natural enemies were then removed from flowers while they were immersed in 70% ethyl alcohol under a stereo microscope at a magnification of 40 (Funderburk et al. 2000). Similar method was also used to collect the anthocorids in the flowering plants in gardens and along roadsides in Hawaii Island. They were Orius persequens, Orius tristicolor, Paratriphlepslaeviusculus, Montandonio laconfusa, and Blaptostethus pallescens (Calvert et al. 2019). Stall bugs, or Berytidae, are a small family of true bugs in the Heteroptera order. They are mostly phytophagous. Some species are facultatively carnivorous, saprophagous, and occasionally omnivorous. The family of insects known as the Miridae belongs to the order Heteroptera. They are significant global crop pests that attack crops like alfalfa, apple, cocoa, cotton, sorghum, and tea. Mirid bugs that prey on invertebrate pests of horticultural crops are endemic, generalist zoophytophagous predators known as predatory mirid bugs (Hemiptera: Miridae). A variety of tiny invertebrates, primarily arthropods, are preyed upon by the Nabidae (order Heteroptera). Entomologists have taken notice of some species’ predaceous

1.2 Heteroptera Table 1.1 Predatory insects including their order, families, and common names

3

Order Coleoptera

Hymenoptera

Diptera

Thysanoptera

Heteroptera

Neuroptera

Odonata

Orthoptera Dermaptera

Mantoidea

Family Coccinellidae Carabidae Melridae Staphylinidae Anthicidae Vespidae Sphecidae Hybotidae Formicidae Cecidomyiidae Syrphidae Dolichopodidae Hybotidae Chloropidae Phlaeothripidae Aeolothripidae Thripidae Anthocoridae Berytidae Lygaeidae Miridae Nabidae Pentatomidae Reduviidae Chrysopidae Ascalaphidae Hemerobiidae Coniopterygidae Coenagrionidae Aeschnidae Libellulidae Tettigoniidae Labiduridae Anisolabididae Chelisochidae Forficulidae Mantoideae

Common name Lady beetles Ground beetle Soft-winged flower beetle Rove beetle Flower beetles Wasp Digger wasps Dance flies Ants Gall midges Hoverflies Long-legged flies Dance flies Chloropid flies Thunderflies Vespiform thrips Pear thrips Minute piratebugs Stilt bugs Big-eyedbug Plant bug Damselbug Stingbug Hunterbugs Lacewing Owlflies Brown lacewings Dustywings Damselfly Dragon-fly Groundling dragonfly Long-horned grasshopper Striped earwigs Ringlegged earwig Black earwigs European earwig Preying Mantids

behaviour as well as their widespread prevalence in various environments, particularly agroecosystems. True bugs are referred to as Lygaeidae. The family of insects includes some of the seed bugs and insects popularly referred to as milkweed bugs. About 60 genera in six subfamilies make up the family. The insect family Pentatomidae, which belongs to

4

1

Predatory Insects Collection Methods

the order Heteroptera, includes several shield bugs and stink bugs. Most of them eat plants, while some are predators of other insects. Many species are omnivores, whether they are primarily herbivorous or carnivorous. The rostrum is firmly increscent, the labium is inserted extremely close to the base of the labrum, and the posterior borders of the buccula are fused. These changes to the head are related with the feeding habits. A sizable, global family of predatory insects is known as the Reduviidae. It comprises three-legged bugs, wheel bugs, assassin bugs, and ambush bugs. It is one of the larger families in the Heteroptera with over 7000 species total. A vital part of ecosystems, assassin bugs are also crucial to agriculture and medicine.

1.3

Hymenoptera

Nearly 5000 species make up the huge family Vespidae in the Hymenoptera order. It is regarded as a diverse and international family of wasps, containing both numerous solitary wasps and nearly all of the social wasps. Many species act as pollen carriers, potentially or even successfully pollinating a variety of plants, while others are wellknown predators of insect pest species. The most ferocious predators, Vespidae, can consume enormous quantities of invertebrate food. Additionally, social wasps are key agents for biological control of phytophagous insects in natural habitats and agroecosystems and are generalist predators. In the order Hymenoptera, the family of ants known as Formicidae are social insects. Out of an estimated 22,000 species, more than 12,500 have been classified.

1.4

Dermaptera

Family Labiduridae is a sizable family of earwigs in the suborder Forficulinae, and its members are frequently referred to as striped earwigs. The majority of the family is global. Scale insects, aphids, spider mites, and psyllids have been shown to be able to consume earwig pests in a number of laboratory investigations that targeted specific orchard pests. The capacity to assault and feed on a variety of prey, particularly eggs and immature stages of insects of the orders Lepidoptera, Hemiptera, Coleoptera, and Diptera, has led to the designation of earwigs as voracious predators. On the other hand, some labidurids are thought to be plant pests. The family of earwigs known as Forficulidae belongs to the Forficulinae suborder. With 250 species globally, it is the earwig family with the most diversity. The description of the earwig, Doru lineare, as a predator of Spodoptera frugiperda, the autumn army worm larvae (Lepidoptera: Noctuiidae).

1.6 Neuroptera

1.5

Coleoptera

1.5.1

Coccinellidae

5

Six subfamilies make up the well-known beetle family Coccinellidae, which is found all over the world. Coccinellidae is popularly known as ladybird beetles or ladybugs. As the term “lady” refers to the biblical Mother Mary, ladybugs, ladybirds, or, preferably, lady beetles have been used to depict them throughout history. All remaining coccinellids are predators of hemipteran insects from the suborder Sternorrhyncha (such as aphids, scales, psyllids, and whiteflies), mites, and finally other insect larvae, with the exception of the mycophagous Coccinellinae and the phytophagous Epilachninae. The Coccinellidae, the family to which these insects belong, has a very varied range of behaviours. All terrestrial ecosystems, including tundra, forests, and grassland agroecosystems, as well as those in the plains and mountains, are home to them. To make the most of the employment of coccinellids in biological control, more attention must be placed on evaluation, predator specificity, comprehension of colonisation of novel settings, and assessment of community-level interactions (BC).

1.5.2

Carabidae

The Carabidae are a sizable, diverse family of beetles. There are more than 40,000 species of them worldwide, and they are generally referred to as ground beetles. In both natural and agricultural settings, carabids are generalist predators. They are mentioned as being predators of slugs, herbaceous plant seeds, lepidopteran larvae, and aphids. Additionally, a variety of these beetle species play a part in the natural BC for numerous lepidopteran pests in various crops.

1.6

Neuroptera

A vast and widely distributed family of insects known as Chrysopidae has between 1300 and 2000 species in 85 genera. Green lacewing is the name given to them. They eat mites, aphids, and other small arthropods in addition to pollen, nectar, and honeydew, and some, like Chrysopa, are primarily predatory. As a result, their predatory larvae are often known as “aphid lions” or “aphid wolves” in common parlance. Millions of ravenous Chrysopidae are raised for sale in numerous nations as BC agents of insect and mite pests in agriculture and gardens (Stelzl and Devetak 1999).

6

1.7

1

Predatory Insects Collection Methods

Thysanoptera

The tiny, elongated insects known as thrips have fringed wings. The many different plants and animals that thrips species feed on are punctured, and the contents are then sucked up. The species of manythrips are regarded as pests because they consume plants with a marketable value. While certain thrips species are seen to be beneficial because they eat on other insects or mites, others feed on pollen or fungus spores. Thrips can be general feeders, predacious, mycetophagous, phytophagous, or pollinophagous. By using female parthenogenesis, thrips reproduce. The four pre-imaginal instars of these thrips are the larva I, larva II, pupa I, and pupa II. A frequent and significant predator in Japanese fields of bean, orchard fruit, citrus, pear, and tea is the scolothrip stakahashii. Scolothrips longicornis is a natural helpful thrips found in regions of bean (Phaseolis spp.), cucumber (Cucumis sativus), and eggplant (Solanum melongena) in the Mediterranean and Middle East. With over 3550 species in 460 genera, the Phlaeothripidae is the biggest family in the order Thysanoptera (for systematic details). Some species feed on other insects or mites and are seen as beneficial; however, many species are considered pests because they graze on plants of commercial value. Many animals that live in the tropics are fungivores that consume pollen or spores from fungi. The majority of Haplothrips species are found in Europe, practically all of which reproduce solely in flowers and many of which are host-specific. The thrips Haplothrips subtilissimus feeds on other tiny arthropods as a predator, in contrast to the majority of its relatives in the genus, which consume plants. It is known that some Haplothrips species feed on thrips. For instance, it has been noted that the spider mite T. urticae has a natural opponent in Haplothrips victoriensis. Haplothrips brevitubus adults and larvae were seen feeding on the mulberry thrips, Pseudodendro thrips mori, in a mulberry field in Kagoshima, Japan (Thysanoptera: Thripidae). However, certain Phlaeothripidae thrips species have been seen feeding on Tuta absoluta. The holarctic region is where the Aeolothripidae family of thrips is most prevalent, while certain species can also be found in the drier regions of the subtropics. Many of them also consume flowers in addition to their usual meal of other arthropods. A pantropical genus of obligate predators that mimic ants is called Franklinothrips. Different species in the genus are thought to be effective BC agents against the thrips pest, and they are occasionally marketed as such. Franklinothrips vespiformis is a pest management product that has been marketed in Europe to combat pests like the greenhouse thrips, Heliothrips haemorrhoidalis. Additionally, it consumes mites; nymphs of a certain species of whitefly, and agromyzid fly larvae.

1.8

Collection Device

Predatory insects can be sampled using a variety of techniques, and the effectiveness and efficiency of various techniques can differ. This chapter discusses various active and passive sampling techniques for flying insects, predatory insects, and other natural enemies (predators) of agricultural pests. There is also discussion of sample

1.8 Collection Device

7

sizes, lethal sampling, and the potential utility of by catch. Furthermore, trustworthy sampling techniques are required to assess the situation of these natural enemies and the success of conservation actions. The eight most popular methods for capturing predatory insects for biodiversity studies include traps (Yellow sticky traps, bowl traps, pitfall traps, Malaise traps, and light traps), aspirators, direct observation/ visual count method, knocked down method, Sweep net method, suction trap, Square beating trays, etc., or combinations of these methods. Yi et al. (2021) divided sampling techniques into three categories: 1. Active sampling techniques with an inherent “activity density” 2. Passive sampling techniques with no “activity density” bias 3. Passive sampling techniques with an “activity density” bias In certain articles, procedures with or without human power for collecting specimens are referred to as “active” and “passive” sampling, respectively. Sampling techniques can be broadly categorised into three groups: passive sampling techniques without a “activity density” bias, such as collecting soil or leaf litter samples, sweep netting, and knocking down using chemical fogging; passive sampling techniques with a “activity density” bias, such as pitfall traps, sticky traps, suction traps, Malaise traps, and window traps; and active sampling techniques with an inherent “activity density” bias (Yi et al. 2021). Table 1.2 illustrates numerous approaches that can be utilised to gather predators from various insect orders. The majority of insect orders contain predatory species, including some families of flies (Diptera), such as hoverflies (Syrphidae) and robber flies, as well as dragonflies (Odonata), mantids (Mantodea), true bugs (Heteroptera, Hemiptera), thrips (Thysanoptera), lacewings (Neuroptera), and beetles (Col (Asilidae). Coccinellini (Halyziini) Casey [Halyziats chitscherini] include Coccinella septempunctata, Hippodamia (Adonia) variegate, Calvia punctata, Adalia bipunctata, Adalia tetraspilota, Aiolocariah exaspilota, Macroilleis (=Halyzia) hauseri, and Oenopia conglobate. Trifolium, maize, and wheat all included Coccinella septempunctata and Ippodamia (Adonia) variegate; walnut contained Aiolocaria hexaspilota; and wheat contained Oenopia conglobata.

1.8.1

Direct Observation/Visual Count Method/Whole Plant Samples

Predatory insects can typically be sampled by direct population counts on vegetation in studies. Predatory insects that are just moderately active or hardly moving respond well to this technique. Because precise counts of active, swiftly moving, or quickly disturbed arthropods are challenging to obtain. Natural enemies connected to the crop may be visually counted in fields with uniformly spaced plants, and densities per plant or per unit area may be computed. However, in agroecosystems that are more structurally complex, a measure of density can be produced by utilising a square area or a particular plant structure as the sample unit.

8

1

Predatory Insects Collection Methods

Table 1.2 Comparisons among different terrestrial predators sampling methods at various crops Devices Malaise traps Malaise traps, traditional beat sampling of branches, suction sampling Square beating trays, aspirator, knocking to branches, holding and pulling of twigs sweep net Sweep net sampling Hand picking, sweep netting, light trap, direct observation Sweep net, D-Vac®, Ground-cloth Whole plant, A-vac beating tray, D-vac UC-vac methods Leaf litter sampling + pitfall trapping Sweep net, D-Vac®, ground-cloth Sweep net sampling Sweep net sampling Leaf litter Pitfall traps Vacuum insect net (D-Vac), pitfall and soil cores Pitfall traps Whole plant, A-vac beating tray, D-vac UC-vac methods Sweep net Sweep net Beating flowers or terminal buds

Insect to be collected Flying insects (e.g.: wasps) Flying insects, predators Young once and adult earwigs Adult Nabis spp. Reduviids

Insect group Hymenoptera Diptera Dermaptera

Heteroptera

Geocoris spp. nymphs and adults Geocoris spp. Predators with small body size Nabis spp. nymphs and adults Adult Chrysopa spp. Adult coccinellidae Ground-dwelling predatory beetles Mobile, grounddwelling predators Ground-dwelling predatory beetles Mobile, grounddwelling predators Lacewings Life stages except eggs Life stages except eggs Thrips

Coleoptera

Coleoptera (Carabidae)

Coleoptera (Staphylinidae) Neuroptera

Mantodea Thysanoptera

Both the D-VAC vacuum method and entire plant samples have been employed for sampling. Big-eyed bugs, Geocoris spp., and tiny pirate bugs, Orius tristicolor (White), were used in the first technique. Geocoris punctipes and Geocoris pallens, damsel bugs (Nabi spp.), green lacewings (Chrysopa spp.), and the notoxus beetle (Notoxus calcaratus) in California cotton for 3 years (Gonzalez et al. 1977). Chambers and Adams (1986) counted the eggs and larvae of syrphid and coccinellid aphid predators on winter wheat shoots and in quadrats. Using direct

1.8 Collection Device

9

observation techniques in soybean, Woltz and Landis (2014) found that Anthocorid nymphs and adults were the most prevalent predators (44 individuals), followed by spider adults and immatures (8 individuals), and coccinellid adults (6 individuals). At the National Corn and Sorghum Research Center (NCSRC), Pakchong, Nakhon Ratchasima, Thailand, earwig sampling was done using the visual count method on sweet corn plants over two growing seasons: from June to August 2008 (rainy season) and from August to October 2008 (late-rainy season) (Nawanich et al. 2010). The NCSRC corn fields were home to Proreus simulans all year long, according to the findings. Depending on how challenging it is to see and count the target arthropods and how complicated their environment is, direct visual counting of predatory insects is relatively labour-intensive, which can be a significant drawback.

1.8.2

Knocked Down/Plant Shaking Methods

Proreus simulans were found on each plant, which had an average height of 39.9 cm, and the stalks were knocked down with a rotary chain chopper (Patanakamjorn et al. 1978). For the majority of predators in Iowa soybeans, plant shaking was shown to yield the most accurate and economical estimates (Bechinski and Pedigo 1982).

1.8.3

Sweep Net Method

Sweep netting is a popular technique for collecting a range of beneficial hemipterans, beetles, and neuropterans as well as other predatory arthropods on vegetation. There is, however, a dearth of knowledge regarding the best time to do sweep net sampling in order to gauge overall predator abundance. Sweep nets were used in South Africa’s corn fields between 2014 and 2015 to capture arthropod samples. The most prevalent species found during this investigation were lacewings (Chrysoperla congrua), coccinellids (Hippodamia variegate, Cheilomenes lunata), and praying mantids (Episcopomantis sp., Galepsus sp.) (Greyvenstein et al. 2020). Coccinellid densities in wheat fields were calculated using sweep netting in a study. Using a sweep net, aspirator, and hand picking, adult ladybird beetles (Coccinellinae, Chilocorinae) were collected in 2001. In certain places, more than one method was utilised to gather insects, but in Pakistan, juvenile stages were taken directly from their natural habitats (Khan et al. 2007). Coccinellidae (Coleoptera) (53.7%), Nabidae (Rhynchota) (21.6%), and Anthocoridae (9.18%) were also collected using the same technique in the Italian province of Bologna (Burgio et al. 2006). Sweep net sampling was more advantageous for adult Nabis spp., Chrysopa spp., and Coccinellidae in terms of cost and variability analysis (Bechinski and Pedigo 1982). Important benefits of this technology include inexpensive equipment costs and a possibility for high specimen yields per unit of labour. Repeatable and comparative findings can be obtained by standardising efforts depending on the quantity of sweeps, the area sampled, or the length of the sampling. Sweep net

10

1

Predatory Insects Collection Methods

collections, however, might also be influenced by the taxon being sampled as well as the vegetation structure. A typical sweep net was used to collect adult Nabidae samples. However, they are also helpful in agroecosystems, such as alfalfa, cereals, and winter wheat. These tools have been utilised extensively in the sampling of grassland invertebrates. Similarly, sweep net sampling was preferable for adult Nabis spp., Chrysopa spp., and Coccinellidae in Iowa soybeans in terms of cost and variability analysis (Bechinski and Pedigo 1982). Additionally, Bechinski and Pedigo (1982) discovered that the predator species Nabis spp. and Chrysopa spp. Sweep net sampling outperformed plant shake, absolute, and vacuum sampling in terms of cost and variability for Coccinellidae. The method with the lowest efficiency was vacuum sampling. Sycanus dichotomus and Cosmolestes picticeps (Hemiptera: Reduviidae) were captured using sweep nets from oil palm fields in Malaysia (Jamian et al. 2017). When evaluating the relative species richness and abundance of tiny, vegetationdwelling predators between various places with similar vegetation types, netting is highly helpful. But the collector’s abilities are the only factor that affects the nets capture rate. It takes a lot of time and is best suited for open habitat types like grassland and bushes, but it is difficult to standardise in forest situations with dense vegetation. Additionally, because it requires strong vision, netting is typically done during the day, which limits its wider applicability, such as for catching nocturnal taxa. Netting is a very economical and unobtrusive technique.

1.8.4

Leaf Vacuum/Vacuum Net Method

Using a leaf vacuum (BG 56 C-E; Stihl, Waiblingen, Germany) to suck arthropods off of plants into fine mesh collecting bags, predators were sampled on the foliage and soil beneath soybean plants. The study’s findings indicate that ten families, four orders of predatory insects, and two orders of predatory arachnids were sampled. Included in the first group are the families Carabidae, Coccinellidae, Elateridae, Lampyridae, Staphylinidae, Anthocoridae, Nabidae, and Chrysopidae (Woltz and Landis 2014). The writers, Bechinski and Pedigo (1982), do not propose the vacuum net since it was the least satisfactory. In Australia, the spined predatory shield insect (Oechalia schellembergii), pollen beetle (Dicranolaius bellulus), and transverse ladybird beetle (Coccinella transversalis) were all captured using the vacuum approach in lucerne (Hossain et al. 1999).

1.8.5

Square Beating Trays

Albouy and Caussanel (1990) advise using the square beating trays method to sample earwigs. Using 0.50 m2 beating trays, earwigs and aphids in each canopy were sampled once each month (three vigorous hits of the tree canopy in opposite

1.8 Collection Device

11

directions). In 2006, 2007, and 2008, eight trees were sampled, and nine trees in 2009 and 2010. From April to November, Forficula auricularia was typically active in canopies; it was also present in January, March, and December. In canopies, first instars have never been photographed. Second instars could infrequently be discovered in canopies, and as the next stages developed, they became increasingly common until adults started to emerge. In canopies, there were no variations in the frequencies of male and female species. In canopies, Forficula pubescens did not appear until May but persisted through December. In May and June, Forficula pubescens was only ever discovered. First instars have never been discovered in canopies. In canopies, there were equal numbers of males and females (RomeuDalmau et al. 2012).

1.8.6

Traps

As they look for hosts, prey, or plant-based food supplies, many predatory insects, parasitic predators, and flies are energetic flyers and frequently numerous on vegetation. Some of the procedures discussed above, such as netting and suction sampling, can be helpful for these insects as well. These insects can be caught using pan traps, Malaise traps, and sticky traps. Pan traps are also known as Moericke traps or bowl traps in bee literature. These traps have been used to study parasitoids in a range of settings, including cultivated habitats and agroecosystems. In Madhya Pradesh, India, during the Kharif season of 2004, a study was carried out to examine the potential of light traps as integrated pest management tools in the paddy ecosystem. The study included 17 species of predatory insects from the orders Hemiptera, Coleoptera, Odonata, Hymenoptera, and Dictyoptera (Sharma et al. 2010).

1.8.6.1 Light Trap Numerous insects, the majority of which are nocturnal, and a few diurnal species are positively phototropic and drawn to light. One of the oldest, most traditional, most indigenous methods for collecting insects is the use of light traps (Singh et al. 2018). This approach was widely used in the first decade of the twentieth century, primarily for the management of insect pests. In the Jabalpur region of Madhya Pradesh, a study was carried out during the Kharif season (July–October) of 2004 to document the scope of light traps in the paddy ecosystem and collect 17 species of predatory insects (Sharma et al. 2010). Later, during the Kharif season of 2006 at Balaghat, India, Crospedophorus elegans (Coleoptera), Ectomocoris cordiger, Canthecona furcellata (Hemiptera), and Statilia maculata (Dictyoptera) were recorded from the vegetable environment using light trap (Sharma and Bisen 2013). Between 2003 and 2005, predatory insects were observed at Changping (soybean–corn), Haidian (wheat–corn), and Zijiao (wheat–cotton) in China. By using black light trap, predatory insects including Chrysopidae spp., Harmonia axyridis, and Propylea japonica (1503 individuals from 6 species) were collected (Ma and Ma 2012).

12

1

Predatory Insects Collection Methods

From June to December 2015, Jabalpur (MP), India, employed the new Jawahar light trap model created at JNKVV, Jabalpur with a mercury vapour lamp (80 W) as the light source to record the predatory insects at paddy fields (Mishra et al. 2017). Predators from the Coleoptera (462), Hemiptera (290), Odonata (267), Neuroptera (767), Dermaptera (135), Dictyoptera (62), Diptera (694), and Orthoptera orders were collected (159). The information is given below: Coleoptera were the most numerous with 11 species and 4 families, including Prothyma sp. (272), Chlaenius pictus (61), Brachinus sexmaculeatus (52), Chlaenius nigricans (18), and Cicindela flexuosa making up the majority of the family Carabidae’s 8 predatory species. Onitis falcutus, a species of dung beetle, was the only member of the family Scarabaeidae to be documented (165 beetles) (Mishra et al. 2017). Sirthenea carinata (135), Antilochus conqueberti (69), Ectomocoris ululans (61), and Eocanthecona furcellata (25) are among the hemipterous predatory species (Mishra et al. 2017). According to Mishra et al. (2017), the order Odonata contained two species, Pantala flavescens (165) and Coenagrion sp. (102), which are members of the Libellulidae and Coenagrionidae families, respectively. The order Neuroptera was represented by two species, Ascalaphus sp. (694) and Chrysoperla sille (Mishra et al. 2017). Only one species of each of the orders Dermaptera, Dictyoptera, and Orthoptera, namely the earwig Elaunonbipartitus (135), mantis Archimantis latistyla (62), and long-horned predatory grasshopper Conocephalus sp. (159), were found in the study area (Mishra et al. 2017). Ambrose and Livingstone (1989) used light traps to capture up to 16 different species of assassin bugs from Peninsular India. Light draws a lot of predatory reduviids (Ambrose et al. 2007; Hribar and Henry 2007; Lucas et al. 2016). The following species were identified: Coranus spiniscutis, Oncocephalus sannulipes, Polididus armatissmus, Pygolam pisfoeda, Ectomocoris sp., and Emesaya sp. From a total of nine species of reduviid predators, there were five subfamilies: Harpactorinae, Peiratinae, Emesinae, Reduviinae, and Stenopodinae. With two species belonging to two genera (Coranus and Polididus), Harpactorinae was the most dominant subfamily, followed by Peiratinae, Stenopodinae, Emesinae, and Red (Farheen 2017). In 2016, as many as predatory insects were collected from Adhartal, Jabalpur (MP) using the Jawahar light trap model (SM-96) and an 80-W mercury vapour lamp. Carabidae predators, Prothyma sp., Cicindela flexuosa, Chlaenius pictus, Chlaenius nigricans, Diplocheila polita, Brachinus sexmaculeatus, Brachinus longipalpis, Crosopedophorus elegans, Scarabaeidae, Onitis falcutus (Family), Scirtidae, Cyphon padi, Reduviidae, Sirthenea carinata, Ectomocoris ululans; Pentatomidae, Eocanthecona furcellata, and Pyrrhocoridae, Antilochus conqueberti (Singh et al. 2018) were collected.

1.8.6.2 Suction Traps Devices that use an air flow directed across a net to capture arthropods are called suction traps. The Johnson–Taylor suction trap and the Dietrick vacuum bug net are the two primary varieties of suction traps (D-vac). Suction traps made by Johnson– Taylor are mostly employed to capture aerial arthropods. A motor fan is employed in

1.8 Collection Device

13

the D-vac suction trap, which is mostly used for sampling ground-dwelling arthropods. Typically, it comprises of a nylon gathering net and an electric motorised exhaust fan. To collect samples of arthropods in their terrestrial habitat, various types of suction samplers have been created (or) adapted from devices created for other uses, such as leaf blowers). Arthropods are routinely collected from plants using suction samplers. These samplers draw arthropods into a collection net or bag using the suction created by a motorised fan. The D-vac was the first suction machine for sampling terrestrial arthropods that was commercially accessible, and it has been widely used to sample arthropods in a variety of terrestrial settings, including agricultural crops. The Dietrick vacuum insect net, sometimes known as “D-vac”, is the most frequently used suction sampler (Sunderland et al. 1995; Elliott et al. 2006). Hemiptera, Neuroptera, Carabidae, Coccinellidae, Syrphidae, Staphylinidae, and different parasitoids were among the arthropod taxa that were gathered. D-vac sample for adult Coccinellidae and Carabidae was insufficient. However, this technique was very effective at detecting coccinellid larvae, adult and juvenile Nabidae and Chrysopidae, Araneae, and adult Staphylinidae in wheat fields (Elliott et al. 2006). Gonzalez et al. (1977) were the first to note that D-vac was the most sexual method and was unfit for cotton predator sampling.

1.8.6.3 Sticky Cards Sticky traps are frequently used to sample insects and monitor pest populations. Traditional sticky traps are frequently coloured intentionally to draw particular arthropod taxa, despite the fact that they are generally thought of as passive sampling techniques. With the sticky side facing away from the plant, sticky cards were placed on plants in the inner rows of a plant and fastened with wire ties to the plant halfway between the ground and the top of the plant. Each card was arranged in a unique row with a unique height. After 48 h, the cards were retrieved and wrapped in transparent plastic. A dissecting microscope was afterwards used to identify and count the insects. The counts from the three cards at one location were then combined. The day after filed counts were conducted, sticky cards were collected. Sticky cards were used to gather insects including Coleoptera (CoccinellidaeCoccinella maculate, Harmonia axyridis, Coccinella septempunctata, Propyleaquatuor decimpunctata, and Hippodamia spp.), hemiptera (Anthocoridae: Orius insidiosus, Nabidae, Reduviidae), diptera (Syrphidaea, Cecidomyiidae), neuroptera (Chrysopidaea), and hymenoptera (Formicidae) from sweet corn plants. The coccinellids Coccinella maculata and Harmonia axyridis as well as the anthocorid Orius insidiosus were the most prevalent predators (Musser et al. 2004). 1.8.6.4 Pitfall Traps Pitfall traps have only ever been used to gather ground beetles and offer a simple, affordable method to sample ground-dwelling arthropods (Coleoptera: Carabidae). The most significant families of helpful predatory insects are earwigs (Dermaptera), rove beetles (Coleoptera: Staphylinidae), and carabid beetles (Coleoptera: Carabidae) (Winder et al. 2001). Carabids are the group that has been the subject

14

1

Predatory Insects Collection Methods

of the most research and understanding. Pitfall traps are the primary method used to sample carabids and other beneficial predatory insects. Since at least the early 1900s, pitfall traps have been employed to gather and count helpful predatory insects (Fichter 1941). Preservatives include water, brine, formalin (which is extremely toxic and is rarely used), ethylene glycol, propylene glycol, acetic acid, alcohol, kerosene, and chloral hydrate. In Iowa, where Agropyron repens, Amaranthus retroflexus, Physalis heterophylla, Phalaris arundinacea, Bromus japonicus, Polygonum convolvulus, Solanum nigrum, Lactucas cariola, Cannabis sativa, and Bromus inermis were common, a research using pitfall traps was done to record the predators. Pasimachus elongates (3), Scarites quadriceps (1), Dyschirius globulosus (2), Carabus meander (2), Calosoma calidu (4), Galosoma obsoletum (5), Scaphinotus elevates (1), Notiophilus semistriatus (1), and Pasimachus elongates (3) were also recorded (Esau 1968). In Alabama’s conventionally tilled, irrigated, and unirrigated peanut (Arachis hypogaea) fields in 1987 and 1988, this technique was employed to track the seasonal abundance of predators. The most prevalent species was Labidur arlparia (Dermaptera: Labiduridae). Carabids, Staphylinids, Iycosids, Geocorls spp., and Solenopsis invicta (Hymenoptera: Formicidae) were also prevalent (Kharboutli and Mack 1991). Carabids have been the main subject of most investigations using pitfall traps in agroecosystems (Halsall and Wratten 1988). To catch earwigs, 30 homemade bamboo traps and pitfall traps (Barber 1934) were used (Huth et al. 2011). For instance, in the Vineyards, approximately 86,000 earwigs were captured in 2007 and 2008. Additionally, it was stated that throughout the entire summer, nocturnal earwigs spend the day in dark, congested areas of the grapevine, such as in the grapes, under the leaves, between the canes, and inside woody poles (Huth et al. 2011). Pitfall traps offer a lot of drawbacks despite the fact that this trapping method is often used. For instance, they frequently fail to capture small and “trap-shy” species, eventually deplete the local carabid population, necessitate the capture of ground-dwelling species, and yield various results depending on the size and type of the trap, the type of preservative used, and the placement of the trap (Ulyshen et al. 2005). It is suitable to collect Carabidae predators using four pitfall traps (Cockfield and Potter 2017). When using pitfall traps to study a particular arthropod taxon, a good combination of trap designs should be taken into account. Pitfall capture results are affected by the structure of the ground vegetation, trap size and shape, meteorological conditions, material of the trap and their concentration, cover used, etc.

1.8.6.5 Malaise Traps The first Malaise trap was created by Swedish hymenopterist Dr. Rene Edmond Malaise in Burma in 1934. Subsequently, three other varieties of Malaise traps were proposed: the original unilateral, a bilateral type with a lateral collector and another with a collector in the middle. There have been numerous Malaise trap designs created during the past few decades. The basic design of the Malaise trap is a tent with two large openings—one in the front and the other in the back, just across from

1.8 Collection Device

15

the first. A cloth barrier to stop the flying insects is in the centre. In their upward movement in an attempt to escape, insects become caught in the collecting jar that is installed directly above at the summit and is loaded with a lethal chemical. The cloth arrangement is mounted on four logs or poles and held up by ropes (Sheikh et al. 2016). 1.8.6.5.1 Townes Trap A straightforward Malaise trap design created by Dr. Henry Keith Townes is the most widely used because of its practical layout and lightweight construction. The trap has a central diaphragm and a lateral collector at the peak, and it is open at both ends. Typically, the colour of the trap is black and either all-white or all-black. Due to their positive phototropism, the majority of insects enter the trap, strike the diaphragm, move upward to a light opening, and then become trapped in the collection. The mesh size, microhabitat, and trap design all have a significant impact on how effective a Malaise trap is. The trap design, together with the trap’s placement in the ideal location, is the most crucial of the three variables mentioned above. The target insect’s size will determine the mesh size, which should be extremely fine. According to Matthews and Matthews (1983), the majority of commercially available traps feature an opening measuring 3 m2 (total sampling surface of both sides), or around 1.92 m2 per length of diaphragm (Sheikh et al. 2016) (Fig. 1.1a). 1.8.6.5.2 Gressitt Trap Malaise trap has been expanded into the Gressitt trap. It actually consists of two Malaise traps linked at the back, giving it two peaks and two collectors, which combine to form a big trap. The trap’s aperture is around 6 m wide, making it 2.3 times bigger than Townes trap (Sheikh et al. 2016) (Fig. 1.1b). 1.8.6.5.3 Schacht Trap Without a diaphragm, the Schacht trap operates on the tenet that insects striking an angled surface will travel toward the collecting bottle. Because it would serve as an

Fig. 1.1 Different types of Malaise traps. Malaise traps (a), Gressitt trap (b), and Schacht trap (c)

16

1

Predatory Insects Collection Methods

insect deterrent, the diaphragm is missing. According to Sheikh et al. (2016), the trap has so far produced good results, notably for sampling Diptera (Fig. 1.1a–c). Smaller Townes design and larger Gressitt design stand out among them. Flying insects are collected with Malaise traps, especially wasps (Hymenoptera) and flies (Diptera). A collection head is located at the highest point of the trap’s tent-like framework. Flying insects hit the screen and move upward, where they are caught in a plastic bottle and prepared for analysis. The traps can also be used to catch other insects, such as beetles, which fall to the ground when they contact the trap when used alongside trays. A study was carried out in Brazil between February 2010 and January 2015 utilising Malaise traps to assess the seasonality and diversity of predatory insects (Diptera: Syrphidae and Asilidae) in Brachiaria decumbens monoculture and silvopastoral systems. A total of 11 hoverfly species (Diptera: Syrphidae) were gathered, of which five and three were unique to silvopasture and monoculture, respectively. In comparison to the silvopasture, the monoculture had a much higher number of specimens.

1.8.6.6 Combinations of Traps The majority of studies show that combining approaches yields more insects than doing so separately. By collecting weekly samples of harmful and helpful arthropods associated with soybeans, such as insect pests, in 1972, sweep net, D-Vac®, and ground-cloth sampling techniques were contrasted. During the monitoring period, nymphs of Geocoris and Nabis species were observed. Adult geocorids and nabids could be sampled using any of the three techniques as well (Shepard et al. 1974). In the Chitral District of Pakistan, 12 species of predatory beetles were identified, including members of the Coccinellinae and Chilocorinae subfamilies of the Coccinellidae family and three distinct tribes (Chilocorini, Coccinellini, and Psylloborini). Adult specimens were gathered using hand picking, aspirator, and sweep net techniques (Khan et al. 2007). The predatory insects have been collected using the procedures listed below. Neuropterans are one of the few predatory insects that are very migratory and unable to locate their population in a crop. The ecological arthropod trophic pyramid is topped by Neuropters. Under the protection of natural habitats, they were captured using yellow sticky traps and an aspirator. Results from Sorribas et al. (2016) reveal that yellow sticky traps collected a substantially larger proportion of Coniopterygidae (dustywings) (97% of the total lacewings captured with this method) than the suction device did for lacewings of the Chrysopidae family (green lacewings). Utilising several traps, including light traps, yellow-pan traps, and pitfall traps, predatory insects were recorded from oil palm fields in Sumatra, Indonesia, from February to August 2019. There were documented insects from the following Coleoptera families: Anthicidae, Cantharidae, Carabidae, Cerambycidae, Cicindelidae, Coccinellidae, Hydrophilidae, Pyralidae, Salpingidae, Scaritidae, Staphylinidae and Dermaptera species include Anisolabididae and Chelisochidae (Husni and Jauharlina 2021). At the Agro-Techno Center (ATC), Sriwijaya University, Indralaya, South Sumatra, Indonesia, soybeans were studied using pitfall traps, net traps, and ocular

1.8 Collection Device

17

observations. Odontoponera denticulata, Solenopsis sp., Coccinella transversalis, and Micraspis discolour are the four entomophagous insect species that have been identified. Common soybean pathogens included Odontoponera denticulate, Solenopsis sp., Coccinella transversalis, Micraspis discolour, and Lophyraintermedia (Anggraini et al. 2021). However, other arthropodesd like Lycosa pseudomonas, Odontoponera denticulate, Lophyra intermedia, Drapestica socialis, Oxyopes javanus, Pardosa distincta were also collectred from the agroecosystem (Anggraini et al. 2021). In another study, sweep nets, pitfall traps, visual records, and rice hill dissection were used at Sakha Agricultural Research Station throughout the course of two successive rice seasons (2017 and 2018). In the nursery, the predatory insect population was noted (Hegazy et al. 2021). According to the findings, Coleoptera (44.12%) predominated, followed by Hymenoptera (20.59%), Hemiptera (2.94%), Odonata (14.71%), Neuroptera (2.94%), Diptera (11.76%), and Orthoptera (2.94%) (Hegazy et al. 2021). Hambäck et al. (2021) employed standard branch beat sampling and suction sampling to describe the ecology of predatory arthropods in Swedish apple orchards in the years 2015 and 2016. According to the results, suction sampling significantly increases the number of predatory dipterans (Diptera: 32% vs. 20%, Hymenoptera: 25% vs. 7%). Contrary to beat sampling, where spiders were the most prevalent group, suction sampling revealed that predatory dipterans were the most numerous predatory group. For instance, out of 2083 predatory arthropods that we caught, 532 were caught in 2015 (beat sampling), and 1551 were caught in 2016 (suction sampling). Dance flies in the family Hybotidae were one group of predatory flies that was notably abundant in both species and individuals in the investigated apple orchards. Even though the sample method’s bias was obvious, it was gratifying to see that it had no impact on the management of predatory arthropod ecosystems. Using both techniques, organic apple orchards had higher densities of dipteran and coleopteran predators, whereas IPM-managed orchards had higher densities of Opilionids. However, the response differed in sign between predatory groups. The inclusion of landscape factors further demonstrated effects of landscape variety and deciduous forest cover. Spiders, Opilionids, and dipterans were significantly less prevalent in orchards that were flanked by more complex landscapes (high landscape variety and/or high deciduous forest cover), although both Coleoptera and Heteroptera were more prevalent. They came to the conclusion that predatory dipterans may be important in apple orchards and strongly advised that apple and other crops aggressively incorporate predatory Diptera. Priscibrumus uropygialis, an apple and pear orchard, Chilocorus circumdatus, and apricot trees all yielded Chilocorus rubidus (Khan et al. 2007). In Cherry orchards (Cerasus avium), Tezcan and Kocarek (2009) used pitfall traps, fermenting bait traps, and beating (50 trees were beaten in each orchard) to capture four Dermapteran predators, including Forficula auricularian, Forficula lurida, Forficula smyrnensis, and Guanchia hincksi in Muradiye (Man 140 specimens (66.99%) from three species were obtained using fermenting bait traps, ten specimens (4.78%) from three species were obtained using pitfall traps, and 59 specimens (28.23%) from four species were obtained by beating vegetation.

18

1

Predatory Insects Collection Methods

The best strategy for monitoring, specifically for Forficula smyrnensis, in cherry orchards was collection using fermenting bait traps.

1.8.6.7 Other Methods For collecting samples of spiders and predatory insects in cotton, the pounding sheet technique was said to be the most efficient. However, beating sheets would not be a useful sample technique in maize fields due to the structure of the plants there. A unique apparatus was created in 2004 by Teshler and co-workers to capture and preserve a coleopteran predator. The biological control agent of common ragweed, Ambrosia artemisiifolia, Ophraella communa (Coleoptera: Chrysomelidae), has been collected, transported, and delivered using a multipurpose device. The instrument is a 125-ml plastic specimen jar that can accommodate either adults or pupae of Ophraella communa. Insect collecting and counting times are shortened when utilised as an aspirator. The median survival times for Ophraella communa adults and pupae held inside the container at 3 °C are 41 and 21 days, respectively. Insects are fed while being transported using a cotton wick soaked in water or a solution of 5% sugar, and the container’s design reduces insect mortality by creating the ideal microclimate for insect storage and transportation. The containers can be used for field releases of Ophraella communa adults and pupae. They are made to shield insects from rainfall and reduce contacts with predators and parasites. The container is quite adaptable and could potentially be used with a number of predatory insect species, even though it has been built and tested for Ophraella communa.

1.9

Combinations of Methods

Using the Whole Plant A-Vac UC-vac and tray D-vac techniques, Geocoris spp., Hemerobius spp., Chrysopa camea, and Orius tristicolor were collected from commercial strawberry farms in California in 1988 and 1989. Although the numbers of these insects acquired in D-vac, A-vac, and beating-tray samples, in that order, were consistently higher than those found in the UC-vac sample (Zalom et al. 1993). It reported no significant difference across sampling methods for captures of either Geocoris spp. or the lacewings (Zalom et al. 1993). From September 1996 to January 1997, predatory insects in a commercial wheat crop (cv. Otto) in Chile were sampled using vacuum insect nets (D-Vac), pitfalls, and soil cores (Carrillo et al. 2007). The huge beetle Calosoma vagans was the one with the most carabids caught, according to the results. Staphylinid beetles were the most numerous. In Indonesian zinnia (Zinnia sp., hybrids) and soybean (Glycine max) planting sites, visual observation, pitfall traps, and net-traps were used (Anggraini et al. 2021). Additionally, the findings show that zinnia had a higher total number of individual predatory arthropods than soybean did in nets and pitfall traps. From May 2018 to November 2018, reduviid predators were surveyed in Indonesia using three methods: sweep net, pitfall trap, and Malaise trap at ungrazed oil palm plantations or non-agropastoral sites (T) and grazed oil palm plantations or agropastoral sites (Nazilah et al. 2020). Formicidae and Reduviidae were in general

1.9 Combinations of Methods

19

dominant among predatory insects. The Reduviidae family has 405 members and 8 morphospecies, while the Formicidae family has 6617 members and 56 morphospecies. The most prevalent predatory insect species in an oil palm plantation is Cosmolestes sp. (Reduviidae), according to the authors, who do not mention any collection procedure data. Bechinski and Pedigo (1982) employed the sweep net, plant shaker, and vacuum net as three collection tools to capture adult Orius insidiosus (Anthocoridae), Nabis spp. (Nabidae), Araneida, Chrysopa spp. (Neuroptera: Chrysopidae), adult Coccinellidae (Coleoptera), and adult Anthicus cer. Plant shaking is advised for sampling the majority of predators since it typically produces the most accurate and economical estimates. Furthermore, there were yearly variations in the connection of relative estimates with absolute population changes, which appeared to be impacted by soybean plant growth parameters. Turkey between 2008 and 2010: Aspirator, knocking to branches, handing, and removing of twigs from olive trees (Olea europae). Guanchia brignolii, Forficula lurida, Guanchia hincksi, Forficula sp., Forficula aetolica, Forficula auricularia, Forficul adecipiens, Forficula lurida, Forficula sp., and Forficula sp. were all found in olive groves. In this study, Forficula lurida (41.8%) was shown to be the most prevalent species (Kaçar and Nishikawa 2018). With a sweep net, drop cloth, beat bucket, shake bucket, and a visual search of cotton plants, big-eyed bugs, Geocoris punctipes, green lacewing larvae, Orius spp., flower bugs, lady beetles (Coccinellidae-Hippodamia convergens), adult lady beetles, and green lacewing larvae were all recorded. The beat bucket and shake bucket methods, according to the authors, took a lot less time to complete than the other three sample techniques (Allen and Wilson 1999). Predatory insects were collected using a sweep net and a hand picking technique on Bt (SRCH-639 BG II) and non-Bt cotton plants in farmer’s fields in Nalgonda District, Telangana State, from July 2018 to January 2019. In both Bt and non-Bt cotton fields, predatory insects such as Brumoides suturalis (three-stripped ladybird beetle), Cheilomenes sexmaculata (six-spotted zigzag ladybird beetle), Coccinella transverslis (transverse ladybird beetle), Paederus sp. (rove beetle), Cochliomyia sp. (blow predatory insects Propylea dessecta (spotless ladybird beetle), Vespula sp. (wasp), and Calopteryx splendens (banded demoiselle damselfly) were only found in non-Bt cotton fields, while Chrysopa sp. (green lacewing neuropteran) was only found in Bt cotton (Mallesh and Sravanthy 2021). In order to gather predators from commercial Bt (variant DPL 458, DPL 555) and non-Bt cotton (DPL 491, 493) in Tift County, Georgia, from 2002 to 2004, whole plants were bagged, drop cloth samples, and pitfall traps were used. This technique concentrated on the stationary plant-eating stages of predator life (Torres and Ruberson 2005). Data on 21 and 18 predatory taxa were gathered in Bt and non-Bt cotton fields using drop cloth and whole plant sampling, respectively, and the results were generated from those data. The common predators in the cotton fields included Chrysoperla rufilabris larvae, Coccinella septempunctata, Coleomegilla maculate, Dorutae niatum, Geocoris floridanus, Geocoris punctipes, Geocoris uliginosus, Harmonia axyridis, Hippodamia convergens, Notoxus monodon, and

20

1

Predatory Insects Collection Methods

Orius insidiosus. Predatory heteropterans, including all damsel bug species combined (Nabis roseipennis, Nabis americoferus, Tropiconabis capsiformis, and Nabis alternatus), were more common in Geocoris punctipes and Orius insidiosus in all years (Torres and Ruberson 2005). Arthropod predator activity patterns may also be influenced by external factors like ground cover and vegetation. The response of the overall catch to temperature changes has been theorised to be different from the responses seen at the species level. Furthermore, the population of predators is also determined by wind speed, field water levels, and humidity. Daily variations in light intensity, temperature, and humidity might affect coccidioid behaviour. Lacewings are more active at night and spend the day resting in their natural surroundings. Along with prey numbers, plant phonology and its combinations also influenced the population of predators.

1.10

Recommendations’

• When sampling predators or insects, the timing of the sampling should be taken into and the predators’ daily activity patterns. • The monitoring of beneficial insects in a crop can be done using appropriate methods. • The recording of natural enemies can be done using a combination of two or more approaches. The key factors to take into account when choosing a sampling method are cost and time.

References Albouy V, Caussanel C (1990) Dermapteresou perce oreilles. Federation Francaise des Societes de Sciences Naturelles, Paris Allen EK, Wilson WT (1999) The beat bucket: a rapid, reliable method for sampling predatory insects and spiders in cotton. In: Proceedings of the Beltwide cotton conference. National Cotton Council, Memphis Ambrose DP, Livingstone D (1989) Population dynamics of assassin bugs from Peninsular India (Insecta-Heteroptera-Reduviidae). J Bombay Nat Hist Soc 86(3):388–395 Ambrose DP, Sudha PG, Nagaraja K, Kumar AG (2007) Biodiversity of light-attracted entomofauna in a rice agroecosystem, Tirunelveli, Tamil Nadu, South India. Insect Environ 13(3): 118–120 Anggraini E, Irsan C, Gunawan B (2021) Phytophagous insects and predatory arthropods in soybean and zinnia. Biodivers J Biol Divers 22(3):1405–1414 Ballal A, Chandish R, Yamada K (2016) Anthocorid predators. In: Ecofriendly pest management for food security. Academic, Cambridge, pp 183–216 Barber H (1934) Traps for cave-inhabiting insects. J Elisha Mitchell Sci Soc 46:259–266 Bechinski EH, Pedigo LP (1982) Evaluation of methods for sampling predatory arthropods in soybeans. Environ Entomol 11:756–76l Burgio G, Ferrari R, Boriani L, Pozzati M, van Lenteren J (2006) The role of ecological infrastructures on Coccinellidae (Coleoptera) and other predators in weedy field margins within northern Italy agroecosystems. Bull Insectol 59(1):59

References

21

Calvert F, Hollingsworth RG, Wall M, Follett PA (2019) Survey of flowering plants in Hawaii as potential banker plants of anthocorid predators for thrips control. J Asia-Pac Entomol 22(3): 638–644 Carrillo R, Alarcón R, Neira M (2007) The effects of carabid beetles (Coleoptera: Carabidae) on the arthropod fauna of wheat fields in Chile. Rev Biol Trop 55(1):101–111 Chambers RJ, Adams TH (1986) Quantification of the impact of hoverflies (Diptera: Syrphidae) on cereal aphids in winter wheat: an analysis of field populations. J Appl Ecol 23:895–904 Cockfield SD, Potter DA (2017) Predatory insects and spiders from suburban lawns in Lexington, Kentucky. Great Lakes Entomol 17(3):179–184 Elliott NC, Tao FL, Fuentes-Granados R, Giles KL, Elliott DT, Greenstone MH, Shufran KA, Royer TA (2006) D-vac sampling for predatory arthropods in winter wheat. Biol Control 38(3): 325–330 Esau KL (1968) Carabidae (coleoptera) and other arthropods collected in pitfall traps iniowa cornfields, fencerows, and prairies. Iowa State University Ph.D. thesis Farheen F (2017) Biological control potential evaluation of Coranus spiniscutis (Insecta: Hemiptera: Reduviidae) against three selected insect pests. Ph.D. thesis, Manonmaniam Sundaranar University, Tamil Nadu, India, p 121 Fichter E (1941) Apparatus for the comparison of soil surface arthropod populations. Ecology 22: 338–339 Funderburk J, Stavisky J, Olson S (2000) Predation of Frankliniella occidentalis (Thysanoptera: Thripidae) in field peppers by Oriusinsidiosus (Hemiptera: Anthocoridae). Environ Entomol 29(2):376–382 Gonzalez D, Ramsey DA, Leigh TF, Ekbom BS, Van Den Bosch R (1977) A comparison of vacuum and whole-plant methods for sampling predaceous arthropods on cotton. Environ Entomol 6(5):750–760 Greyvenstein B, Siebert SJ, Van Den Berg J (2020) Effect of time of day on efficacy of sweep net sampling of arthropod predators in maize agro-ecosystems in the Northwest Province, South Africa. Afr Entomol 28:150–163 Halsall NB, Wratten SD (1988) The efficiency of pitfall trapping for polyphagous predatory Carabidae. Ecol Entomol 13:293–299 Hambäck PA, Porcel M, Tasin M, Samnegård U (2021) Predatory arthropod community composition in apple orchards: orchard management, landscape structure and sampling method. J Appl Entomol 145(1-2):46–54 Hassanzadeh H, Shishehbor P, Esfandiari M, Rajabpour A (2013) Collection of Orius species (Hemiptera: Anthocoridae) in some counties of Yazd province, Iran. Munis Entomol Zool 8(1): 507–508 Hegazy FE, Hendawy EA, Mesbah II, Salem FA (2021) The insect pests, the associated predatory insects and prevailing spiders in rice fields. J Plant Prot Pathol 12(5):365–371 Hossain Z, Gurr GM, Wratten SD (1999) Capture efficiency of insect natural enemies from tall and short vegetation using vacuum sampling. Ann Appl Biol 135(2):463–467 Hribar LJ, Henry TJ (2007) Empicoris subparallelus (Hemiptera: Heteroptera: Reduviidae), a predatory bug new to the fauna of Florida. Fla Entomol 90(4):738–741 Husni H, Jauharlina J (2021) Diversity and abundance of predatory arthropods on immature and mature oil palm (Elaeisguineensis) plantations. Pak J Biol Sci 24(1):25–34 Huth C, Schirra KJ, Seitz A, Louis F (2011) A beneficial species becomes a pest–the common earwig Forficula auricularia (Linnaeus 1758). IOBC WPRS Bull 67:249–256 Jamian S, Norhisham A, Ghazali A, Zakaria A, Azhar B (2017) Impacts of 2 species of predatory Reduviidae on bagworms in oil palm plantations. Insect Sci 24(2):285–294 Kaçar G, Nishikawa M (2018) Forficulidae fauna of olive orchards in the Southeastern Anatolia and Eastern Mediterranean Regions of Turkey (Dermaptera). J Entomol Res Soc 20(1):27–35 Khan I, Din S, Khalil SK, Rafi MA (2007) Survey of predatory coccinellids (Coleoptera: Coccinellidae) in the Chitral district, Pakistan. J Insect Sci 7(1):7. https://doi.org/10.1673/031. 007.0701

22

1

Predatory Insects Collection Methods

Kharboutli MS, Mack TP (1991) Relative and seasonal abundance of predaceous arthropods in Alabama peanut fields as indexed by pitfall traps. J Econ Entomol 84(3):1015–1023 Lucas M, Forero D, Basset Y (2016) Diversity and recent population trends of assassin bugs (Hemiptera: Reduviidae) on Barro Colorado Island, Panama. Insect Conserv Divers 9(6): 546–558 Ma G, Ma C-S (2012) Differences in the nocturnal flight activity of insect pests and beneficial predatory insects recorded by light traps: possible use of a beneficial-friendly trapping strategy for controlling insect pests. Eur J Entomol 109(3):395 Mallesh M, Sravanthy C (2021) Diversity of predatory arthropods in Bt and non-Bt cotton fields of Nalgonda district, Telangana state, India. J Entomol Zool Stud 9(2):1199–1203 Matthews RW, Matthews RJ (1983) Malaise trap: the townes trap catches more insects. Contribution of the American Entomological Institute 20:428–432 Mishra Y, Sharma AK, Pachori R, Kurmi A (2017) Taxonomic documentation of insect pest fauna of rice collected in light trap at Jabalpur district of Madhya Pradesh. J Entomol Zool Stud 5(6): 1212–1218 Musser FR, Nyrop JP, Shelton AM (2004) Survey of predators and sampling method comparison in sweet corn. J Econ Entomol 97(1):136–144 Nawanich S, Chongrattanameteekul W, Supamongkol P (2010) Diversity of arthropods in sweet corn. Agric Nat Resour 44(6):1094–1099 Nazilah H, Sahari B, Prabowo R, Buchori D (2020) Effect of cattle integration on the diversity of predatory Reduviidae in oil palm plantation, Central Kalimantan, Indonesia. IOP Conf Ser Earth Environ Sci 468(1):012004 Patanakamjorn S, Guthrie WD, Young WR (1978) Iowa State J Res 52(4):277 Romeu-Dalmau C, Espadaler X, Piñol J (2012) Abundance, interannual variation and potential pest predator role of two co-occurring earwig species in citrus canopies. J Appl Entomol 136(7): 501–509 Sharma AK, Bisen UK (2013) Taxonomic documentation of insect pest fauna of vegetable ecosystem collected in light trap. Int J Environ Sci 4(3):4–10 Sharma AK, Barche S, Mishra PK (2010) Pest and predatory insect species inhabiting paddy ecosystem in Jabalpur, Madhya Pradesh collected with the help of light traps. Pest Manage Econ Zool 18(1/2):125–133 Sheikh AH, Thomas M, Bhandari R, Meshram H (2016) Malaise trap and insect sampling: mini review. Bio Bull 2(2):35–40 Shepard M, Carner GR, Turnipseed SG (1974) A comparison of three sampling methods for arthropods in soybeans. Environ Entomol 3(2):227–232 Singh S, Sharma AK, Saxena AK, Panday AK, Kakade SH (2018) Taxonomic analysis of phototactic beneficial insects as biocontrol agents (predators and parasites) collected in light trap in rice ecosystem at Jabalpur. J Entomol Zool Stud 6(3):850–853 Sorribas J, González S, Domínguez-Gento A, Vercher R (2016) Abundance, movements, and biodiversity of flying predatory insects in crop and non-crop agroecosystems. Agron Sustain Dev 36(2):34 Stelzl M, Devetak D (1999) Neuroptera in agricultural ecosystems. Agric Ecosyst Environ 74(1-3): 305–321 Sunderland KD, De Snoo GR, Dinter A, Hance T, Helenius J, Jepson P, Kromp B, Lys J-A, Samu F, Sotherton NW et al (1995) Density estimation for invertebrate predators in agroecosystems. In: Arthropod natural enemies in arable land I—density, spatial heterogeneity and dispersal. Aarhus University Press, Aarhus Tezcan S, Kocarek P (2009) Dermaptera fauna of the ecologically managed cherry orchards in western Turkey. Munis Entomol Zool 4(2):572–576 Torres JB, Ruberson JR (2005) Canopy-and ground-dwelling predatory arthropods in commercial Bt and non-Bt cotton fields: patterns and mechanisms. Environ Entomol 34(5):1242–1256 Ulyshen MD, Hanula JL, Horn S (2005) Using Malaise traps to sample ground beetles (Coleoptera: Carabidae). Can Entomol 137(2):251–256

References

23

Winder L, Holland JM, Perry JN, Woolley C, Alexander CJ (2001) The use of barrier-connected pitfall trapping for sampling predatory beetles and spiders. Entomol Exp Appl 98:249–258 Woltz JM, Landis DA (2014) Comparison of sampling methods of A phis glycines predators across the diel cycle. J Appl Entomol 138(7):475–484 Yi Z, Jinchao F, Dayuan X, Weiguo S, Axmacher JC (2021) A comparison of terrestrial arthropod sampling methods. J Resour Ecol 3(2):174–182 Zalom FG, Pickel C, Walsh DB, Welch NC (1993) Sampling for Lygus hesperus (Hemiptera: Miridae) in strawberries. J Econ Entomol 86(4):1191–1195

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Contents 2.1 2.2 2.3 2.4

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Distribution of Predatory Insects at Sole/Inter Crop(s) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Distribution of Predatory Insects in Genetically Modified Crop(s) . . . . . . . . . . . . . . . . . . . . . . . Distribution of Various Predators at Different Agroecosystems Order-Wise . . . . . . . . . . . . . 2.4.1 Mantodea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.2 Odonata . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.3 Diptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.4 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.5 Distribution of Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.6 Distribution and Diversity of Various Coleopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.7 Neuroptera: Chrysopidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.8 Hemiptera/Heteroptera: Pentatomidae, Miridae, Geocoridae, Anthocoridae, Nabidae, Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.9 Orthoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Factors Responsible for Predators’ Populations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.1 Pests and Their Natural Enemy’s Complex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.2 Conventional Crops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.3 Genetically Modified Crops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.4 Field Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.5 Pesticides Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.6 Landscape/Urbanisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.6 Recommendation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.1

25 26 33 34 34 34 37 39 39 41 44 46 54 54 54 54 55 57 58 58 59 60

Introduction

Terrestrial arthropods particularly predatory insects (natural enemies) are extremely important ecosystem components particularly in cropping systems. By significantly influencing the control of crop-damaging insects, these natural enemies from the insect orders Coleoptera, Hymenoptera, Diptera, Thysanoptera, Heteroptera,

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_2

25

26

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Neuroptera, Odonata, Orthoptera, and Dermaptera provide ecosystem services in agroecosystems. Recent decades have seen a plethora of studies on the ecology and distribution of insects, and the emphasis on tree cropping systems has emphasised the need of understanding the distribution, diversity, and ecology of natural enemies. The biological effectiveness of natural enemies at the ecosystem level is influenced by the intricacy of interactions between natural enemies and/or pest insects. The natural enemies that are present in diverse agroecosystems are listed in tables in this chapter. However, initially, reports with list of families or genus of various insect orders Coleoptera, Hymenoptera, Diptera, Thysanoptera, Heteroptera, Neuroptera, Odonata, and Orthoptera were provided.

2.2

Distribution of Predatory Insects at Sole/Inter Crop(s)

Following formulae are very much useful for the population dynamics studies. Mean density =

Xi=N × 100

where Xi = No of insects N = Total area sampled or number of plants sampled. Relative density (RD-%) The species abundance was estimated expressing the data as relative density. RD% = Number of individual species=Total number of individuals of all species × 100 Orius insidiosus, Nabis spp., and other predators that live in foliage and litter were common in a complex of more than 80 species (Hemiptera, Carabidae, Staphylinidae, and Formicidae) that was recorded (Bechinski and Pedigo 1981). Oecophylla longinoda and Oecophylla smaragdina are found in Asia and Australia and are beneficial predators in the timpercocoa, citrus, coffee, oil palm, mango, coconut, and timber production. From 1973 to 1985 in alfalfa fields from the eastern South Dakota, eight species of aphidophagous predators commonly occurred in alfalfa in the following order of decreasing relative abundance: Nabis americojerus Carayon, Hippodomia convergens, Chrysoperla porabundo, Hippodomia tredecimpunctata tibialis, Hippodomia parenthesis, Coleomegilla maculata, Coccinella transversoguttata, and Cycloneda munda (Elliott and Kieckhefer 1990). At the Appalachian Fruit Research Station in Kearneysville, West Virginia, four orchards were tested four times in 1991. In total, 1176 predators from 22 families and 7 orders were gathered. The Coleoptera family Coccinellidae had the most species (283) and individuals (283) overall (16). However, the species with the highest population density was Coniopteryx sp. (Coniopterygidae: Neuroptera), which had almost as many individuals (271) as coccinellids. In the list of predators, three omnivorous families [Gryllidae (Orthoptera), Forculidae (Dermaptera), and

2.2 Distribution of Predatory Insects at Sole/Inter Crop(s)

27

Miridae (Hemiptera)] accounted for 18.94% of all predators (Brown and Schmitt 2001). During 1996, from the UK, 21 species of the Carabidae family were found in wheat fields, including Agonum dorsalea, Amara aenea, Amara eurynota, Amara familiarisa, Amara ovate, Bembidion guttulaa, Bembidion lamprosa, Bembidion obtusum, Bradycellus harpalinusa, Carabus violaceous, Clivina fossora, Demetrias atricapillusa, Harpalus rufipesa, Loricera pilicornisa, Nebria brevicollisa, Notiophilus biguttatus, Pterostichus cupreusa, Pterostichus melanarius, Pterostichus nigera, Pterostichus strennusa, and Trechus quadristriatus and 12 Staphylinidae species (Aleocharinae, Consoma spp., Lathrobium spp., Omaliinae spp., Oxytelinae spp., Quedius spp., Staphyliniae spp., Tachinus spp., Tachyporus chrysomelinusa, Tachyporus hypnoruma, Tachyporus spp., and Xantholinus spp.) (Collins et al. 2002). Episyrphus balteatus, Eupeodes corolla, Eupeodes lundbeckii, Melanostoma mellinum, Scaeva pyrastri and Sphaerophoria scripta were observed in wheat plantation from 1988 to 1990 at Germany (Tenhumberg and Hans-Michael 1995). In Peru, from potato field, predators of Carabidae, Coccinelidae, Nabidae, Lygaeidae, Chrysopidae, and Syrphidae are some of the families (Cisneros 1995) have been reported. Reduviidae, Anthocoridae, Miridae, Geocorinae, and Nabidae were recorded from soybean cultivars like Cubasoy-23, Incasoy-24, Incasoy-27, and Doko in Cuban (Marrero and Martínez 2003). The three most prevalent natural enemies of Dysaphis plantaginea (Homoptera: Aphididae) in Asturian of NW Spain apple orchards are Episyrphus balteatus (Diptera: Syrphidae), Adalia bipunctata (Coleoptera: Coccinellidae), and Aphidoletes aphidimyza (Diptera: Cecidomyiidae) (Miñarro et al. 2005). Almost all predators from Heteroptera (Miridae, Nabidae, Reduviidae, Lygaedae, Veliidae, Mesoveliidae, Hydrometridae), Coleoptera (Coccinellidae, Staphylinidae, Carabidae, Tenebrionidae), Orthoptera (Tettigoniidae, Gryllidae, Tridactylidae), Hymenoptera (Formicidae, Vespidae, Pompilidae, Sphecidae, Eumenidae), Diptera (Ephydridae, Platystomatidae), Odonata (Libellulidae, Gomphidae, Coenagrionidae, Protoneuridae), Dermaptera (Carcinophoridae), Mantodea (Mantidae), Phasmatoidea (Phasmatidae), and Neuroptera (Ascalaphidae) were recorded from rice ecosystems at Sri Lanka when insects recorded from 1995 to 1998 (Bambaradeniya et al. 2004). Later, Bambaradeniya and Edirisinghe (2009) also recorded in rice ecosystems; there are 25 species of Coleoptera, 25 species of Hymenoptera, and 19 species of Odonata. Eggs, larvae, and pupae of various syrphids species were collected from organically grown romaine lettuce at 14 sites between 16 March and 20 September 2005 (Smith and Chaney 2007). According to Pavuk (2007) during 2007 in northwest Ohio in soybean agroecosystems field several species of ladybird beetles (Coccinellidae: Harmonia axyridis, Coccinella septempunctata, and Coleomegilla maculate), hemipterans like Anthocoridae (Orius insidiosus), Nabidae and Reduviidae, green lacewings (Neuroptera: Chrysopidae), and a number of these predatory arthropod species may have potential as biological control agents of Aphis glycines were observed.

28

2

Distribution and Diversity of Predatory Insects in Agroecosystems

From India, paddy fields of five districts of Uttar Pradesh yielded dragonflies (Crocothemis servilia, Orthetrum sabina (Libullulidae), damselflies (Ischnura senegalensis-Odonata Agrionidae), Agriocnemis pygmaea-Agrionidae), ladybird beetles (Harmonia octomaculata, Micraspis discolor, Micraspis inopsCoccinellidae), ground beetles (Ophionea nigrofasciata, Casnoidea ishii ishii, Casnoidea indica-Carabidae), long-horned grasshopper (Conocephalus longipennis-Tettigonidae); kets (Anaxipha longipennis-Trigoniidae) July end to mid-October Moderate Metioche vittaticollis Orthoptera Trigoniidae July end to mid-October Moderate Ripple bug Microvelia douglasi (Hemiptera: Veliidae) from August to September. Moderate Mesovelia vittigera (Hemiptera: Veliidae) from August to September. Moderate plant and leaf bugs Cyrtorhinus lividipennis (Hemiptera: Miridae) from August to September. Low earwig Euborellia stali (Dermaptera: Carcinophoridae) recorded from July end to mid-October, low rove beetles Paederus fuscipes (Coleoptera: Staphylinidae) (Pathak et al. 2011). Borkakati et al. (2018) have reported the following predators in various cropping systems. The common natural enemies viz., were Conocephalus longipennis, Agriocnemis femina Brauer, Micraspis crocea, Cicindela undulate and Cicindela melancholia, predated upon different rice pest. Dipha aphidovora, Micromus igorotus, and Chrysoperla spp. were the most common predators on Ceratovacuna lanigera. Spalgius epius and chrysopids were observed on papaya mealybug. In cabbage, Bhut jalakia and okra ecosystem, Coccinella septempunctata, and Coccinella transversalis were the predominant predators. From Uzbekistan, a total of 1472 beetles were collected from vegetable crops like tomatoes (varieties Volgograd 5/95 and Vostok-36), two agrocenoses of potatoes (varieties Sante, Pikasso), two agrocenoses of cabbage (varieties Slava 1305 and Toshkent 10), and more belongs to 27 species of rove beetles and 22 species of ground beetles (Carabidae) (Staphylinidae). The most common species are: Calathus melanocephalus (5.39%), Poecilus cupreus, 1758 (5.3%), Bembidion femoratum (5.10%), Aleochara bilineata (17.6%), Aloconota gregagia (10.21%), Amischa analis (6.01%), and Amischa bifoveotata (5.41%). The dominant species are: Harpalus rufipes (Halimov 2020). During the same time, Coccinella septempunctata (Coccinellidae) (19.7% of the total species density), Chrysoperla carnea (Chrysopidae) (17.8%), Aphelinus sp. (Aphelinidae) (16.2%), and Coccinella undecimpunctata (Coccinellidae) (12.0%) were collected from alfalfa (20 species), followed by cotton (13 species) and faba bean (9 species) (EL-Sheikh et al. 2020). In 2011, Coccinellidae, Syrphidae, Anthocoridae, and Neuroptera are commonly present in chilli pepper ecosystems at Brazil (Amaral et al. 2013). Phenology and quantity of the first brood of the common earwig, Forficula auricularia, were observed over a 13-year period in a Dutch apple orchard (Helsen et al. 1998). From Melon crop predators of Dermaptera (Labiduridae), Hymenoptera (Formicidae, Vespidae, Mutillidae, Sphecidae), Diptera (Syrphidae, Dolichopodidae), Hemiptera (Reduviidae, Pyrrhocoridae), Coleoptera (Staphylinidae, Coccinellidae), and Neuroptera (Chrysopidae, Myrmeleontidae) were observed in 2012. Pirate bugs (Anthocoridae), praying mantis (Mantodea:

2.2 Distribution of Predatory Insects at Sole/Inter Crop(s)

29

Mantidae), and earwigs (Dermaptera: Forficulidae) were observed in okra plots in 2009 at Nigeria (Ratnadass et al. 2014). A total of 2986 individuals of natural enemies grouped into seven orders, 12 genera, and 37 species were gathered from an organic rice field in the Philippines by Labe et al. in 2016. It was possible to identify insects belonging to the Odonata, Strepsiptera, Orthoptera, Hemiptera, Coleoptera, Diptera, and Hymenoptera. There was additional evidence of sub-class Araneae (Class Arachnida). Agriocnemis species made up 312 of the individual Odonata gathered, followed by Strepsiptera (Halictophagus) with 59 and Coleoptera (Micraspis spp.) with 328. Condylustylus (Dolichopodidae), with 113 genera (counts), Focipomyia sp. (67), and Octhera sauteri (179), were the most numerous genera in the Order Diptera (Ephydridae). Trichogramma (Trichogramatidae) ranked first among the Hymenoptera, with a total of 64 individuals, Amauromorpha accepta (85), and Trichomalopsis oryzae (178). There are 989 Tetragnatha species in the subclass Araneae. The most popular crop farmed in Oklahoma between 2016 and 2027 is winter wheat (Triticum aestivum), which is typically sown in the fall and harvested in June. It is mostly attacked by the greenbug (Schizaphis graminum) and bird cherry-oat aphid (Rhopalosiphum padi), two cereal aphid species. Larval coccinellids, which are arthropod predators, were discovered during each growth season and ranged in number from 2.23 to 15.38 per sample (Tenhumberg and Hans-Michael 1995). In 2000/2001, 16 study locations in Switzerland looked at the species richness and abundance of staphylinid and carabid beetles overwintering in winter wheat fields and regions with 1- to 3-year-old wildflowers. During the survey, 46 staphylinid and 20 carabid species of beetles were gathered. Anotylus rugosus, Platystethus nitens, Gabrius pennatus, Carpelimus corticinus, and Lathrobium longulum were the staphylinid beetles with the greatest abundance. Agonum mülleri, Pterostichus anthracinus, Acupalpus meridianus, and Tachys bistriatus were the species most commonly found among carabid beetles (Frank and Reichhart 2004). In India, survey was conducted in cashew field and results revealed the presence of 18 species of predatory assassin bugs viz., Cydnocoris gilvus, Endochus albomaculatus, Epidaus bicolor, Euagoras plagiatus, Irantha armipes, Panthous bimaculatus, Rhynocoris fuscipes, Rihirbus trochantericus var. sanguineous, Rihirbus trochantericus var.luteous, Sphedanolestes signatu, Sycanus galbanus, Alcmena sp., Biasticus sp., Endochus sp., Epidaus sp., Lanca sp., and Scadra sp. Of these 14 species viz., Cydnocoris gilvus, Epidaus bicolor, Endochus albomaculatus, Euagoras plagiatus, Panthous bimaculatus, Rihirbus trochantericus var. sanguineous, Rhynocoris fuscipes, Sycanus galbanus, Alcmena sp., Biasticus sp., Endochus sp., Epidaus sp., Lanca sp., and Scadra sp. are newly recorded from cashew as predator on tea mosquito bug (TMB). The reduviids, Endochus sp., Epidaus bicolor, Panthous bimaculatus, and Sycanus galbanus were categorised as very common (VC) in occurrence of which Panthous bimaculatus and Sycanus galbanus were observed as very efficient predator of TMB in field. Alcmena sp., Biasticus sp., C. gilvus, E. plagiatus, I. armipes, Lanca sp., Rihirbus trochantericus var. sanguineous, and Sphedanolestes signatu were recorded as common (C) of which Rihirbus trochantericus var. sanguineous and Irantha armipes were observed

30

2

Distribution and Diversity of Predatory Insects in Agroecosystems

as efficient predator of TMB in field. Endochus albomaculatus, Epidaus sp., Rhynocoris fuscipes, and Scadra sp. were categorised as very rare (VR) (Saroj et al. 2016). In agricultural soils, rove beetles (Staphylinidae) and ground beetles (Carabidae) are frequent generalist predators. The staphylinids and the carabids both consume small nematodes, mites, and collembola, as well as ants, aphids, caterpillars, insect eggs, and springtails. They so can limit some prey populations in agroecosystems. Most carabids and staphylinids species were substantially more numerous near the boundaries than in the fields, demonstrating the critical role that boundaries play for these predator species in the agricultural landscape. Only the carabid Clivina fossor and the genus Lathrobium staphylinids were uniformly spread throughout the entire region. It was determined that the grass fields were of minimal value as a reservoir for the predatory beetles because none of the species with higher densities in the grass fields and their boundaries were among the significant predators of insect pests in agricultural fields (Andersen 1997). In Brazil, fieldwork was done in the seasons 2004–2005, 2005–2006, and 2006–2007. While Tetracha brasiliensis, Selenophorus seriatoporus, and Pentacomia cupricollis were prominent solely in the region under the no-tillage regime, and Abaris basistriata, Odontocheila nodicornis, and Calosoma granulatum were dominant in both areas. More dominating Carabidae were present than dominant Staphylinidae. Eulissus chalybaeus, a dominating member of the Staphylinidae, was found in all no-tillage soybean and maize patches and forest fragments (Martins et al. 2012). Rove beetles (Coleoptera: Staphylinidae) are acknowledged as significant agroecosystem contributors and are best known for their role in biological management as arthropod pest predators. Unfortunately, little is known about their bionomics in agroecosystems in North America. As a result, surveys of soybean–hedgerow agroecosystems in Ontario, Canada were conducted in 2009 and 2010 to characterise the assemblage’s common, widespread members and their seasonal activity patterns. Outside of the growing season, the possibility for refuge habitat in nearby hedgerows was evaluated. Particularly in terms of native species, it was discovered that the rove beetle assemblage on soybean plants during the growth season was a less diversified subset than that in the nearby hedgerow regions. There were numerous native and non-native species (>1% activity density). Predaceous, univoltine, and adult overwintering species were the most prevalent, according to records in the literature. Midway through the growth season, the most common species showed the highest activity density, and all were found in hedgerow environments outside of the growing season. The staphylinid fauna of various North American agroecosystems was compared, and it was found that several common species dominated assemblages under a variety of circumstances (Brunke et al. 2014). From Okra field Coccinella transversalis, Harmonia octomaculata, Hyperaspis maindroni, Illeis cincta, Micraspis discolor, Propylea dissecta (Coccinellidae), Ophionea nigrofasciata (Carabidae), Paederus fuscipes (Staphylinidae), Camponotus sericeus, Camponotus compressus, Paratrechinica longicornis, Solenopsis geminate, Tetraponera rufonigra (Formicidae), Ropalidia fasciata, Vespa cincta (Vespidae), Sceliphron madraspatanum conspicillatum, Ropalidia

2.2 Distribution of Predatory Insects at Sole/Inter Crop(s)

31

fasciata (Sphecidae), Chrysoperla carnea (Chrysopidae), Ischnura aurora, Brachythemis contaminate, Crocothemis servillia, Diplocodes trivialis (Coenagrionidae), Neurothemis fulvia, Orthetrum Sabina, Pantala flavescens, Rhythemis variegate, Tholymis tillarga (Libullelidae-Odonta), Conocephalus longipenni, and Conocephalus maculatus (Tettigoniidae-Orthoptera) were observed during Kharif, 2012 and Rabi, 2012–2013 at Karaikal of Tamil Nadu, India (Chakraborty et al. 2014). In Malaysia, the natural enemies of most rice pest insects, such as damselflies, dragonflies, spiders, and mirid carnivores, keep their populations at low levels. Besides hoppers, damselfly, Agriocnemis sp. (Odonata: Coenagrionidae) feeds on moths in the rice fields. The dragonfly, Diplacodes sp. (Odonata: Libellulidae), is a common insect in rice fields (Ooi 2015). From July to November 2018, at Myanmar rice fields, Bledius filipes, Paederus riparius (Staphylinidae), Micraspis crocea, Harmonia octomaculata, Menochilus sexmaculatus (Coccinellidae), Casnoidea indica, Ophionea nigrofasciata, Eucolliuris fuscipennis (Carabidae) (Coleoptera), Euborellia annulipes (Dermaptera), Tornosvaryella oryzaetora (Diptera), Cyrtorhinus lividipennis (Miridae), Microvelia douglasi (Veliidae), Polytoxus fuscoviftatus (Reduviidae), Agriocnemis pygmaea, Agriocnemis femina (Odonata), and Anaxipha longipennis, Metioche vittaticollis (Orthoptera) were recorded (Phyu et al. 2020). The predominant predatory insects in potato fields in the USA were Nabis and Geocoris, and organic farms had more of them than conventional farms did (Krey et al. 2021). A total of 5220 predators from the families Aeolothripidae, Phlaeothripidae, Formicidae, Anthocoridae, Chrysopidae, Araneidae, Tetragnathidae, and Uloboridae were gathered from the mango agroecosystems of Mexico between 2008 and 2009 (Rocha et al. 2015). Three species, Orius albidipennis, Orius maxidentex, and Orius minutus, as well as Orius (Dimorphella) albidipennis, Orius maxidentex, Orius minutus, Orius laevigatus, and Orius niger, were previously collected from different host plants of Tazian and Minab (Sadegh Nejad et al. 2019). Ceraeochrysa cincta, Ceraeochrysa claveri, and Ceraeochrysa cubana are among the 28 species of Chrysopidae that can be found on different crops in Brazil (Santos et al. 2020). Predators in organic soybean fields from 2017 to 2019 included species from Ukraine’s Orius sp. (Anthocoridae), Adonia variegate (Coccinellidae), and Sphaerophoria scripta (Syrphidae), while those in conventional fields included Orius sp. (Anthocoridae), Propyleaquatuord ecimpunctata (Coccinellidae) (Grabovska et al. 2021). A study was conducted from 2018 to 2019 on 34 ornamental trees and shrubs located in 16 parks near the center of Burdur, Turkey. This study recorded that Coccinellidae (24), Cantharidae (1) (Coleoptera), Nabidae (1), Miridae (1) (Hemiptera), Sryphidae (1) (Diptera), and Forficulidae (1) (Dermaptera) were shown to be insect predators of aphids (Patlar et al. 2021). Insects from Hemiptera (Nabidae), Nabis pseudoferus (Miridae), Deraeocoris lutescens; Coleoptera (Coccinellidae) Adalia bipunctata, Adalia decempunctata, Adalia fasciatopunctatarevelierei, Chilocorus bipustulatus, Clitostethus arcuatus,

32

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Coccinella septempunctata, Exochomusqua dripustulatus, Harmonia axyridis, Harmonia quadripunctata, Hippodamia variegate, Hippodamia undecimnotata, Myrrha octodecimguttata, Oenopia conglobata, Oenopia lyncea, Propylaea quatuordecimpunctata, Scymnus (Scymnus) apetzi, Scymnus (=Scymnus) bivulnerus, Scymnus (=Mimopullus) flagellisiphonatus, Scymnus (Scymnus) frontalis, Scymnus (Scymnus) interruptus, Scymnus (Scymnus) rubromaculatus, Scymnus (=Pullus) subvillosus, Scymnus pallipediformis, Stethorus gilvifrons, Diptera (Syrphidae), Scaevadi gnota, and Dermaptera (Forficulidae), Forficula auricularia were observed from the study area (Patlar et al. 2021). In Iran, from a cotton field, 43 predatory insects belong to order Coleoptera (eight species of Coccinellidae and five species of Carabidae), order Diptera (seven species of the family Syrphidae), order Heteroptera (seven species of the family Anthocoridae, three species of the family Geocoridae, order Mantodea (one species of the family Empusidae, three species of the family Mantidae), family Nabidae (three species), family Reduviidae (four species), order Neuroptera (one species each of the families Chrysopidae and Hemerobiidae), and family Reduviidae (Sakenin et al. 2021). In Spain, a field survey was conducted in alfalfa field from 2010 to 2021. Total of 9124 coccinellids in all were gathered for the investigation. There were 16 species/genera identified, including Coccinella septempunctata, Hippodamia variegata, Propyleaquatuordecempunctata, Scymnus spp., Coccinella quinquepunctata, Coccinulaquatuordecimpustulata, Exochomusnigromaculatus, Tytthaspissedecimpunctata, Hyperaspis sp., Adalia bipunctata, Oenopialyncealyncea, Chilochorus bipustulatus, Psyllobora vigintiduopunctata, Stethorus punctillum, Subcoccinella vigintiquattuorpunctata, and Coccinella undecimpunctata (Meseguer et al. 2021). In Tamil Nadu of India, different predators such as Cryptolaemus montrouzieri, Menocheilussexmaculatus, Anegleiscardoni, Hyperaspis maindroni, Brumoides suturalis, Scymnus spp. (Coleoptera: Coccinellidae), Chrysoperla species, Mallada species, Spalgisepeus Lycaenidae (Lepidoptera), Geocoris species, Anthocoridae (Hemiptera), Cardiastethus species, and Diadiplosis species (Diptera: Cecidomyiidae) were recorded from the cassava field (Thennarasi et al. 2021). The diversity of predatory fauna at wheat during Rabi, 2019–2020 revealed the presence of Coccinellids (Coccinella transversalis, Cheilomenes sexmaculata, Illeis cincta), green lacewings (Chrysoperla sp.), and syrphids (Jambagi et al. 2022). Previously, Amala and Shivalingaswamy (2018) conducted the survey in the guava, sapota, and mulberry fields of the Kanakapura district of Karnataka, India, on a biweekly basis between 2016 and 2017. The families Anthocoridae (Cardiastethus sp., Geocoris sp., and Orius sp.), Carabidae (Carabus sp.), Coccinellidae (Brumus suturalis, Coccinella transversalis, Coccinella septempunctata, Pharascym nushorni, Scymnus coccivora), Chrysopida (Chrysoperla zastrowisillemi, Mallada boniensis, Hemerobia sp.), Pentatomidae (Eocanthecona furcellata), and Syrphidae (Ischiodon scutellaris) were recorded.

2.3 Distribution of Predatory Insects in Genetically Modified Crop(s)

2.3

33

Distribution of Predatory Insects in Genetically Modified Crop(s)

The use of GM insect-resistant plants in agriculture has evolved into a potent tool for eradicating major pests. In order to suppress the European corn borer, Ostrinia nubilalis, transgenic maize MON810 containing the Cry1Ab gene from Bacillus thuringiensis Berliner var. kurstaki expressing Cry1Ab insecticidal protein was initially licenced for cultivation in Europe in 1998. On 32.1 million hectares worldwide in 2009, transgenic cotton and maize cultivars that express Bt proteins were planted. It is also anticipated that several crops that express novel insecticidal proteins will soon be commercialised. Even though there is a wide variety of scientific data on the effects of Bt maize on non-target arthropods, several pertinent taxa have not undergone in-depth research in Europe. One of the largest beetle families, Staphylinidae, contains more than 47,000 species and is found in practically all types of environments worldwide. However, owing of taxonomic restrictions and a lack of knowledge on species ecology and prey preferences, rove beetles have only occasionally been used in integrated pest management. According to research on maize, ants, spiders, rove beetles, predaceous mites, and ground beetles make up the majority of the plant’s natural enemies. West of Budapest, Hungary, a 3-year field experiment (2001, 2002, and 2003) was conducted in an experimental maize stand surrounded by sizable peach and apricot orchards. In six Bt and six isogenic maize plots during the course of the 3-year survey, 1538 individuals and 21 species were identified. About 681 individuals and 18 species in total were caught in Bt stands, while 857 individuals and 18 species were caught in isogenic stands. Expression of the Cry1Ab protein or genetic alteration as a whole had no discernible impact on the community as a whole. Predators that consume aphids had significantly and barely significantly larger abundances in isogenic maize stands in 2002 and 2003, respectively (Balog et al. 2010). Temporal phenology and community structure of the predatory insects dwelled above the ground level in commercial Bt maize fields in Central Spain was investigated over a period of 3 years by Farinós et al. (2008). This study reveals that Bt maize was frequently infested by rove beetles, carrion beetles, click beetles, earwigs, and damsel bugs. Additionally, recent research in Mexico indicates that the Bt corn hybrids Agrisure™ 3000 GT, Agrisure® VipteraTM 3110, and Agrisure® VipteraTM 3111 were planted between 2009 and 2013. From 2009 to 2011, a survey of transgenic rice was undertaken in China, where it discovered numerous Coleopteran predators, including Harmonia axyridis, Propylaea japonica, and Thinopinus pictus (Zhang et al. 2015). As essential components of Integrated Pest management (IPM) systems, transgenic cotton, maize, and brinjal are of utmost importance. On the question of whether transgenic crops will harm the local anthocorids, there was much conjecture. In southern Bohemia, there have been no reports of Bt maize having any unintended impact on these local natural adversaries, particularly anthocorid carnivores. Additionally, it has been shown that O. majusculus ingesting Bt protein through plant leaves, pollen, or the food web has no detrimental effects on

34

2

Distribution and Diversity of Predatory Insects in Agroecosystems

its biological parameters, but it does have a favourable impact on fecundity and developing time (Ballal and Yamada 2016). When exposed to some lepidopteran rice pests, a transgenic rice line (T2A-1) with a synthetic Cry2A (Cry2Aa) gene demonstrated effective resistance. The rove beetle (Paederus fuscipes), a generalist predator in rice ecosystems, may eat a variety of rice insect pests, including planthoppers. It is important to evaluate the potential dangers of T2A-1 rice to this predator given the possibility that Cry2Aa could come into contact with Paederus fuscipes through the tritrophic food chain. This study used a tritrophic experiment to examine the effects of Cry2Aa on Paederus fuscipes through the food chain of T2A-1 rice, Nilaparvata lugens, and Paederus fuscipes. No accumulating Cry2Aa could be found in Paederus fuscipes adults after feeding on N. lugens nymphs raised on T2A-1, despite Cry2Aa being found in Nilaparvata lugens. Additionally, P. fuscipes’ life table parameters in this tritrophic chain showed no negative consequences. Additionally, Paederus fuscipes did not significantly suffer any negative consequences after direct exposure to a high dose of purified Cry2Aa protein, which was the worst-case scenario. These findings demonstrated that Paederus fuscipes were not harmed by transgenic Cry2Aa rice (Guo et al. 2020). About 17,626 predators in all, representing 09 taxonomic groups and 30 families, were gathered across all treatments in various locations. Compared to non-GM lines, the common predator population was marginally greater on the GM hybrids, but the changes were not statistically significant (Aguirre et al. 2021).

2.4

Distribution of Various Predators at Different Agroecosystems Order-Wise

2.4.1

Mantodea

Mostly large predatory insects make up the Mantodea group, which is found throughout the world’s tropical and subtropical climates. Insects belong to Empusidae, Hymenopodidae, Liturgusidae, Mantidae, and Toxoderidae are usually distributed in various agroecosystems. From the literature survey, it is very clear that a greater number of mantides were observed from rice field followed by fruits (paddy, mango, and banana). Mantis field surveys at various countries reveals that they were distributed in orchids, cereals and pulses, fruits, vegetables and ornamental plants, etc. (Table 2.1).

2.4.2

Odonata

From India, Lestesthor acicus, Lestesum brinus, Lestes viridulus (Lestidae), Coperam arginipes, Elattoneura nigerrima (Platycnemididae); Agriocnemis pygmaea, Ceriagrioncorom andelianum, Enallagma cyathigerum, Ischnura aurora, Ischnura elegans, Ischnura nursei, Ischnura senegalensis, Paracercion malayanum, Pseudagrion decorum, Pseudagrion hypermelas, Pseudagrion microcephalum,

2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise

35

Table 2.1 Mantodea predators recorded from various agroecosystems: a worldwide report with reference Crops Rice

Location Sri Lanka

Year 1995–1998

Insects Mantides

India

2011–2013

India

2016–2017

Shorgum

India

2011–2013

Creobroter kolhapurensis Aethalochroa ashmoliana Archimantis latistyla Creobroter apicalis Gongylus gongylodes Hierodula coarctata Hierodula grandis Hierodula keraleness Hierodula membranacea Hierodula venosa Hierodula viridis Humbertiella ceylonica Mantis religiosa Schizocephala bicornis Tenodera sinensis Creobroter maharashtri

Rose

India

2011–2013

Eremoplana elongata

Cotton

India

2011–2013

Hierodula orientalis s

Tanzania Iran

2018–2019 2020

Cotton- (Bt and non-Bt) Cajanus

India

2018–2019

Sphodromantis viridis Bolivaria brachyptera Mantis religiosa Mantis religiosa

India

2011–2013

Hierodula shivajiensis

Castor

India

2011–2013

Humbertiella mulberae

Guava

India

2011–2013

Medicinal plants Sugarcane Paddy, mango, banana

Pakistan

2016

Schizocephala gramminae Mantis religiosa

Pakistan India

2016 2016–2017

Thesprotia graminis Aethalochroa ashmoliana Ameles fasciipennis Creobroter apicalis Empusa guttula

References Bambaradeniya et al. (2004) Sathe and Vaishali (2014) Patel et al. (2018)

Sathe and Vaishali (2014) Sathe and Vaishali (2014) Sathe and Vaishali (2014) Lusana et al. (2019) Sakenin et al. (2021) Mallesh and Sravanthy (2021) Sathe and Vaishali (2014) Sathe and Vaishali (2014) Sathe and Vaishali (2014) Gul et al. (2017) Kausar et al. (2017) Hiral et al. (2018)

(continued)

36

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Table 2.1 (continued) Crops

Location

Year

Mango

India

2016–2017

Maize

South Africa

2014–2015

Banana

India

2016–2017

Yellow, pepper

Nigeria

2015

Insects Gongylus gongylodes Gongylus trachelophyllus Archima ntislatistyla Hierodula coarctata Hierodula grandis Hierodula keraleness Hierodula membranacea Schizocepha labicornis Aethalochroa insignis Hierodula venosa Hierodula viridis Humbertiella ceylonica Mantis religiosa Statilia maculata Tenodora sinensis Toxoderopsis spinigera Tropiodo guttatipennis Mantis religiosa Hierodula viridis Statilia maculate Hierodula grandis Tropiodo guttatipennis Hierodula membranacea Aethalochroa insignis Hierodula venosa Compsothespis sp. Entella sp. Episcopomantis sp. Galepsus sp. Harpogomantis sp. Hemiempusa sp. Pyrgomantis sp. Tenodera sp. Aethalochroa ashmoliana Aethalochroa insignis Hierodula venosa Hierodula viridis Mantis religiosa Tenodera sinensis Stagmomantis carolina

References

Patel et al. (2018)

Greyvenstein et al. (2020)

Patel et al. (2018)

Agwu et al. (2018) (continued)

2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise

37

Table 2.1 (continued) Crops Ornamental plants Graphs

Location Pakistan

Year 2018–2019

Insects Sphodromantis viridis

References Khan et al. (2019)

USA

2007

Leptothrips mali

Papaya tree Eupatorium lindleyanum Longan and litchi orchards Soybean

Brazil Japan

2015–2016 2018

Carica papaya Tendera sinensis

China

2019

Hierodula patellifera

Costello et al. (2021) Lanna et al. (2021) Sakagami et al. (2021) Wang et al. (2020)

Indonesia

2019

Mantis religiosa

Kharif crops

India

2010–2011

Mantis religiosa

Anggraini et al. (2020) Naikwadi and Javalage (2019)

Pseudagrio nrubriceps (Coenagrionidae); Anax guttatus, Anaximma culifrons (Aeshnidae); Ictinogomphu srapax (Gomphidae); (Libellulidae) Acisomapa norpoides, Brachydiplax sobrina, Brachythemis contaminate, Crocothemis servilia, Diplacodes lefebvrii, Diaplacodes trivialis, Neurothemis tullia, Orthetrumlu zonicum, Orthetrumpruin osumneglectum, Orthetrum sabina, Orthetrumtaeni olatum, Taeniolate marshhawk, Pantalafla vescens, Rhyothemis variegate, Tholymis tillarga, Trameabasilaris burmeisteri, Trithemis aurora, Trithemis pallidinervis, Urothemis signata, Zyxomma petiolatum were collected from the rice field (Rohmare et al. 2017). During the same period, another study was conducted in West Bengal of India. In West Bengal of India, a study was conducted from April 2011 to February 2014 in various agroecosystems and recorded 17 species of odonates belong to the sub-orders Anisoptera (12 species), and Zygoptera (5 species). The recorded species are Tholymis tillarga, Orthetrum sabina, Diplocodes trivialis, Trithemis pallidinervis, Rhyothemis variegate, Neurothemistullia, Crocothemisservilia, Aethriamanta brevipennis, Acisomapanorpoides, Brachydiplax sobrina, Pantala flavescens, Brachythemis contaminate, Ceriagrioncoro mandelianum, Agriocnemis lacteola, Ischnura aurora, Ceriagrioncerino rubellum, Agriocnemis pygmaea, and Ceriagrioncoro mandelianum (Dwari and Mondal 2017). Anggraini et al. (2020) recorded Ischnura heterosticta and Arigomphus villosipes from Soybean ecosystems. Dragonflies were recorded from the rice field at China in (Qian et al. 2021).

2.4.3

Diptera

All known Chamaemyiidae members have larvae that feed on homopteran species with soft bodies, mainly homopterous pests of economic importance, including adelgids and aphids, as well as mealy bugs and scales. However, Syripdes are

38

2

Distribution and Diversity of Predatory Insects in Agroecosystems

more commonly present in the agroecosystems rather than Cecidomyiidae, Dolichopodidae and Hybotidae (Table 2.2). More than 400 species of Syrphids act as predators. Predatory larvae, primarily aphidophagous ones, are present in more than one-third of the Syrphidae and play a significant role in managing the pest Table 2.2 Dipteran predators recorded from various agroecosystems: a worldwide report with reference Crops Soybean

Location Indonesia

Year 2019

Fruit orchards

Turkey

2011

Vegetables

India

2013–2014

Cotton

Egypt

1996, 1997

Orchards

Tanzania Italy

2018–2019 1995 and 1997

Insects Promachus vertebrates Diptera Leucopis annulipes Leucopis formosana Leucopis glyphinivora Leucopis revisenda Leucopis rufithorax Leucopis spyrothecae Episyrphus balteatus Eupeodes (Macrosyrphus) confrater Sphaerophoria scripta Melanostoma univitatum Ischiodon scutellaris Metasyrphus corolae Metasyrphus confrator Paragus tibialis Paraguss eratus Betasyrphus serarius Paragusa egyptius Sphaerophoria flavicauda Eupeodes confrate Episyrphus balteatus Malanosto mamellinum Sphaerophoria scripta Eristalisar bustorum Eumerus sogdianus Episyrphus belteotus Sphaerophoria sulphuripes

Cabbage Oregon broccoli Citrus

India Corvallis

2017–2019 2000

Spain

2015–2016

Coconut

Brazil

2017–2018

Jetropa

Mexico

2016

Cecido myiidae Aphidoletes aphidimyza Condylostylus depressus Condylostylus electus Chironomus plumosus

Brinjal

Bangladesh

2011–2012

Syrphus confractor

References Anggraini et al. (2020) Satar et al. (2015)

Bhat and Bhagat (2017)

Lusana et al. (2019) Burgio et al. (2006)

Sarma et al. (2021) Ambrosino et al. (2006) Bouvet et al. (2020) Penner et al. (2021) Ramos-Robles et al. (2021) Akter et al. (2018)

2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise

39

population. Totally, 15 Syrphinae of Diptera were collected from agroecosystems in 2003–2004 from Iran. They were Chrysotoxum bactrianum, Episyrphus balteatus, Eupeodes (=Eupeodes) nuba, Eupeodes corolla, Ishidona egyptius, Melanostoma melenium, Paragus (=Paragus) Compeditus, Paragus (=Paragus) quadrifasciatus, Paragus albifrons, Scaeva (=Scaeva) pyrastari, Scaevaalbo maculata, Sphaerophoria rueppelli, Spherophoria scripta, and Spherophoria turkmenica (Jalilian 2019). Because Dolichopodidae (Diptera) are one of the most prevalent generalist highly mobile predators in Brazilian agroecosystems, Harterreiten-Souza et al. (2021) studied the spatiotemporal dynamics of these species in organic vegetable farms. Condylostylus and Chrysotus adults, both Dolichopodidae, were observed during a 2-year period in five organic farms’ fields, fallow areas, agroforests, and forest remnants. In contrast to agroforests and forest remnants, predators favoured openfield settings. Probably because there are more options for prey hunting in open-field settings, there are higher population densities. As abiotic circumstances tighten up during the dry season, agroforests and woodlands serve as nesting grounds and refuges, preserving the predator populations throughout the entire farmland ecosystem. Within the farm, both predatory taxa coexist in habitats with distinctive timerelated characteristics, constituting a dynamic and continuous population unit. The significance of agroforests and forest fragments in species conservation should therefore be taken into account in organic crop conservation biological management measures (Harterreiten-Souza et al. 2021).

2.4.4

Hymenoptera

Most of the predators of Hymenoptera belong to Vespidae, Sphecidae, and Formicidae. This category has drawn increased attention since it includes important species for ecosystem structure, such as ants and wasps that manage phytophagous insects or bees that pollinate native plants. Predators of Formicidae dominated in various agroecosystems than Vespidae and Sphecidae (Table 2.3).

2.4.5

Distribution of Dermaptera

In general, earwigs (Dermaptera), which are extensively scattered around the world and pestiferous predators, are polyphagous. However, they are reported to be primarily aphidophagous, occurring in many agricultural ecosystems. These insects are part of a small order of insects with over 1800 species spread around the globe (apart from the polar regions), with the tropics having the highest variety. For example, of more than 60 species that have been described from Australia, there are only ten predators. Additionally, there are 22 species in the USA, some of which are crop pest predators. Dermapteran predators are mainly reported from orchids like apple, cherry, etc. (Table 2.4).

2019–2020

2019–2020 2007

Burkina Faso

Burkina Faso Mexico

India

Mexico Bangladesh India

Sri Lanka

Maize

Coffee

Agroecosystems

Jetropa Brinjal Agroecosystems

Rice

2007

2016 2011–2012 2014–2016

Insects Ondotomachus bauri Oecophyllas maragdina Oecophylla longinoda Crematogaster nr. sumichrasti Ceranisu smenes Odontoponera denticulate Aphelinus mali Encarsia perniciosi Aphytis proclia Diaperasticus erythrocephalus Forficula senegalensis Pheidole megacephala Azteca instabilis Camponotu stextor Crematogaster spp. Dorylus molestus Myrmicaria brunnea Oecophylla suaragdina Tetraponera rufonigra Lasius alienus Camponotus compressus Icaria ferruginea Phimenes flavopictum Polistes fuscatus Polistes hebraeus Vespa orientalis Brachystegu sdecoratus Sphecidae

Formicidae Formicidae Vespidae

Formicidae

Formicidae Formicidae

Forficulidae

Aphelinidae

Order Hymenoptera Hymenoptera Hymenoptera Formicidae Eulophidae

Bambaradeniya and Edirisinghe (2009)

Ramos-Robles et al. (2021) Akter et al. (2018) Taye et al. (2017)

Taye et al. (2017)

Ahissou et al. (2021)

Ahissou et al. (2021)

References Duyck et al. (2011) Vayssières et al. (2013) Ativor et al. (2012) Rocha et al. (2015) Nyasani et al. (2013) Anggraini et al. (2020) Khursheed et al. (2021)

2

2014–2016

2008, 2009 2009 2019 2018–2019

Mango French bean Soybean Apple

Year 2006

Location West Indies Southeast Asia Ghana México Kenya Indonesia India

Crops Banana Citrus

Table 2.3 Hymenopteran predators recorded from various agroecosystems: a worldwide report with reference

40 Distribution and Diversity of Predatory Insects in Agroecosystems

2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise

41

Table 2.4 Dermapteran predators recorded from various agroecosystems: a worldwide report with references Crops Faba beans Cotton

Location Victoria

Year 1993

Insect Labidirra truncata

Genus Labiduridae

Egypt

1995

Staphylinidae

Egypt

1996, 1997

Phenobremia aphidivora Paederus alfierii

Egypt

1996

Labidura riparia

Dermapteran

Forficula auricularia Forficula auricularia Forticula auricularia Euborellia caraibea

Forficulidae

Apple Apple Kharif crops Banana Cherry

Germany and Spain India West Indies Turkey

2017 2010–2011 2006 1998–1999

Forficula smyrnensis Guanchia hincksi Forficulas myrnensis Forficula lurida Forficula auricularia

Staphylinidae

Forficulidae Forticulidae Dermaptera Dermaptera

References Curtis and Horne (1995) El-Heneidy et al. (1996, 1997) El-Heneidy et al. (1996, 1997) El-Heneidy et al. (1996, 1997) Helsen et al. (1998) Happe et al. (2018) Naikwadi and Javalage (2019) Duyck et al. (2011) Serdar and Petr (2009)

Four species, Forficula auricularia, Forficula lurida, Forficula smyrnensis, and Guanchia hincksi, members of the family Forficulidae, were discovered in environmentally friendly cherry orchards (Cerasus avium) in the western Turkish towns of Muradiye (Manisa) and Oren (Izmir) (Tezcan and Kocarek 2009)

2.4.6

Distribution and Diversity of Various Coleopterans

Ladybirds (Coccinellidae), Carabidae, Melridae, Staphylinidae, and Anthicidae have been widely employed to control economically significant herbivorous insects (Table 2.5). Cryptolaemus montrouzieri, Rodolia cardinalis, Cryptogna thanodiceps, Rhyzobius lophanthae, Diomus hennesseyi, Pseudoazya trinitatis, Hippodamia convergens, Chilocorusbipustulatus, Chilocorusnigritus, Chilocorus cacti, Coccinella septempunctata, Curinus coeruleus, Clitostethus oculatus, Halmus chalybeus, Hyperaspis notate, Chilocorus kuwanae, Harmonia axyridis, and Rodolia pumila are very important, worldwide commercially available Coccinellidae (Rondoni et al. 2021).

42

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Table 2.5 Coleopteran predators recorded from various agroecosystems: a worldwide report with reference Crops Location Coccinellidae Alfalfa USA

Year

Insect

References

1988–1992

Elliott et al. (2002)

Wheat

South Dakota

1988–1990

Tomato

India

2016–2018

Cabbage

India

2017–2019

Wheat

Spain

2007

Hippodamia convergens Hippodamia parenthesis Coccinella septempunctata Hippodamia convergens Hippodamia parenthesis Coleomegilla maculate Coccinella septempunctata Brumoides suturalis Cheilomenes sexmaculata Coccinella septempunctata Coccinella transversalis Maenochilus sexmaculatus Coccinellaseptem punctata Micraspis discolor Harmonia dimidiata Adonia variegata

Watermelon

Nigeria

2016

Orchards

Italy

1995–1997

Cotton

Egypt

1996, 1997

Maize

Burkina Faso

2019–2020

Cheilomenes sulphurea Epilachna chrysomelina Exochomus flavipes Hippodamia variegate Coccinella septempunctata Propyleaquatuo rdecimpunctata Scymnu srubromaculatus Scymnus apetzi Chilomenes vicina Coccinella undecimpunctata Coccinella septempunctata Scynmus interruptus Scymnus syriacus Cheilomenes sulphurea

Elliott et al. (1999)

Khokhar and Rolania (2021)

Pérez-Fuertes et al. (2015) Emmanuel et al. (2019) Burgio et al. (2006)

El-Heneidy et al. (1996, 1997)

Ahissou et al. (2021) (continued)

2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise

43

Table 2.5 (continued) Crops Maize

Location South Africa

Year 2014–2015

Insect Buleae anceps Cheilomenes lunata Declivitata bohemani Declivitataha matapygmaea Exochomus flavipes Harmonia axyridis Hippodamia variegate Lioadalia flaromaculata Harmonia xyridis

References Greyvenstein et al. (2020)

Yellow pepper Alfalfa

Nigeria

2015

USA

2014–2015

Shrestha et al. (2021)

2019

Coccinella septempunctata Adalia bipunctata Cocinella repanda

Soybean

Indonesia

Maize

Burkina Faso India

2019–2020

Cheilomenes sulphurea

2010–2011 2010–2011

Cheilomenes sexmaculata Illeis cincta

India Mexico

2016

Cycloneda sanguinea

Bangladesh

2011–2012

Micraspis crocera

Nigeria

2016

Faba beans

Victoria

1993

Barley

Sweden

2011

Orchards Banana Cotton

Italy West Indies Egypt

1995–1997 2006 1996, 1997

Transgenic corn

USA

2004

Cicindela melancholica Megacephala denticollis Platymetopus vestitus Gonocepha lutnadelaidae Pterostichus melanarius Poecilus cupreus Harpalu srufipes Bembidion lampros Trechus secalis Demetrias atricapillus Galerita tristis Calosoma chlorostictum De Geer Coleomegilla maculata Harmonia axyridis Cycloneda munda

Kharif crops Kharif crops Jetropa Brinjal Carabidae Water melon

Agwu et al. (2018)

Anggraini et al. (2020) Ahissou et al. (2021) Naikwadi and Javalage (2019) Naikwadi and Javalage (2019) Ramos-Robles et al. (2021) Akter et al. (2018) Emmanuel et al. (2019) Curtis and Horne (1995) Roubinet et al. (2018)

Burgio et al. (2006) Duyck et al. (2011) El-Heneidy et al. (1996, 1997) Harwood et al. (2005) (continued)

44

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Table 2.5 (continued) Crops

Location

Year

Brinjal Bangladesh Staphylinidae Soybean Indonesia

2011–2012

Insect Diabroticaun decimpunctata howardi Popillia japonica Hypera postica Photinus pyralis Chaetocnema pulicaria Ophioneanigro fasciata

2019

Paederus fuscipes

Egg plant Cleridae Jetropa

Bangladesh

2011–2012

Paederus indica

Mexico

2016

Hippodamia convergens

References

Anggraini et al. (2020) Akter et al. (2018)

Akter et al. (2018)

Ramos-Robles et al. (2021)

In Uzbekistan, in the period 2015–2017, from tomatoes, potatoes, and cabbage (variety Slava 1305, variety Toshkent 10), a total of 1472 beetles were gathered, including 27 species of rove beetles and 22 species of ground beetles (Carabidae) (Staphylinidae). In terms of dominance, the following species were identified: Harpalus rufipes (17.6%), Aleochara bilineata (17.6%), Amara fulva (13.28%), Bembidion properans (10.39%), Trechus quadristriatus (6.20%), Calathus melanocephalus (5.39%), Poecilus cupreus (5.3%), and Bembidion femoratum (5.10%) (Halimov 2020, b)

2.4.7

Neuroptera: Chrysopidae

The Neuropterans are the most important predators, which are numerous and aggressive, hold a prominent position among the many different natural enemies of pests in agricultural settings. The terrestrial Neuroptera, mainly those of the Chrysopidae, Hemerobiidae, and Coniopterygidae families are considered of economic importance, because their larvae are predators of agricultural pests (aphids, whiteflies, and scale insects). The adult predator consumes nectar, pollen, and even insect honeydew as part of their palyno-glycophagous diet in order to receive the protein and carbohydrates they need to survive. The larvae’s diet primarily consists of small arthropods including aphids, thrips, mites, scales, springtails, and moths, yet when prey is scarcer, non-prey food sources like pollen and nectar may be added. A predator of Chrysopidae is commonly present in most of the agroecosystems. Alfalfa, cotton, apple, citrus, corn, maize, mango, olive orchards, orchards, potato, wheat, and jetropa are common crops where we find these predators (Table 2.6). In Mexico, particularly in citrus, orchards, Chrysopidae like Ceraeochrys acincta, Ceraeochrys aclaveri, Ceraeochrys acubana, Ceraeochrysa elegans, Ceraeochrys aeveres, Ceraeochrys asmithi, Ceraeochrys avalida, Chrysopa nigricornis, Chrysoperla carnea, Chrysoperla comanche, Chrysoperla exotera, Chrysoperla externa, Chrysoperla rufilabris, Chrysopodes (Neosuarius) collaris, Eremochrysa (Eremochrysa) punctinervis, Leucochrysa (Nodita) americana, and

2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise

45

Table 2.6 Neuropteran predators recorded from various agroecosystems: a worldwide report with reference Crop Alfalfa

Country USA

Period 1988–1992

Cotton

Egypt

1996, 1997

Alfalfa + cotton Wheat

California

2006–2008 1988–1990

Corn

South Dakota USA

Orchards

Italy

1995–1997

Olive orchards Mango

Spain

2000

México

2008, 2009

Maize

South Africa

2014–2015

Potato

2015, 2016

Apple

Central Minnesota India

2018–2019

Citrus

Spain

2015–2016

Jetropa

Mexico

2016

Jetropa

Mexico

2016

Coffee

Brazil

2013

2004

Predator Chrysoperla plorabunda Chrysoperla carnea Chrysopa carnea

Genus Neuroptera Neuroptera Chrysopidae

Chrysoperla plorabunda Chrysoperla carnea Chrysoperla carnea Chrysopa perla Chrysoperla carnea Ceraeochrysa claveri Ceraeochrysa cubana Ceraeochrys aeveres Chrysoperla congrua Italochrysa similis Chrysopa spp.

Neuroptera

Chrysoperla zastrowi Chrysopidae larvae

Neuroptera

Chrysoperla carnea Mantis pectinicornis Ceraeochrysa cubana Ceraeochrysa claveri Ceraeochrysa everes Leucochrysa (Nodita) cruentata Chrysoperla externa

Neuroptera

Neuroptera Chrysopidae

Chrysopidae

References Elliott et al. (2002) El-Heneidy et al. (1996, 1997) Sivakoff et al. (2012) Elliott et al. (1999) Harwood et al. (2005) Burgio et al. (2006) Corrales and Campos (2004) Rocha et al. (2015)

Chrysopidae

Greyvenstein et al. (2020)

Neuroptera

Middleton et al. (2021) Khursheed et al. (2021) Bouvet et al. (2020) Ramos-Robles et al. (2021) Ramos-Robles et al. (2021) Martins et al. (2019)

Neuroptera

Neuroptera Neuroptera

46

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Leucochrysa (Leucochrysa) arizonica are recorded. Similarly, predators from Coniopterygidae [Coniopteryx (Scotoconio pteryx) josephus, Neoconis szirakii and Semidalis boliviensis] and Hemerobiidae [Symphero biussubcostalis] are recorded from the citrus orchids (Sarmiento-Cordero et al. 2021).

2.4.8

Hemiptera/Heteroptera: Pentatomidae, Miridae, Geocoridae, Anthocoridae, Nabidae, Reduviidae

Predators from Pentatomidae, Miridae, Geocoridae, Anthocoridae, Nabidae, and Reduviidae are included under this sub-order. All these predators are considered as important biological control agent worldwide. Most of the groups have been produced commercially, and farmers have been utilising these predators to protect their crops (Table 2.7). Anthocoridae: With 11 genera and over 150 known species worldwide, the Anthocorini tribe of the Anthocoridae family has the largest species diversity. Acompocoris, Anthocoris (Northern Hemisphere, with a few exceptions), Arnulphus, Coccivora, Compsobiella, Elatophilus, Galchana, Macrotrachelia, Melanocoris, and Temnostethus are the 11 genera now included in the Anthocorini (Table 2.8).

Table 2.7 Pentatomidae, Miridae, and Geocoridae predators recorded from various agroecosystems: a worldwide report with reference Crop Country Pentatomidae Potato Central Minnesota Cotton India

Period

Predator

References

2015 and 2016 2019

Podisus maculiventri Antilochus coquebertii

Apple Miridae Brinjal Geocoridae Potato

USA

2018–2019

Halyomorpha halys

Middleton et al. (2021) Sahayaraj and Fernandez (2021) Ogburn et al. (2021)

Bangladesh

2011–2012

Cyrtorhinus lividipennis

Akter et al. (2018)

Central Minnesota

2015 and 2016 2003–2004

Geocoris spp. Geocoris punctipes

Middleton et al. (2021) Torres and Ruberson (2006) Sivakoff et al. (2012)

Cotton Alfalfa + cotton Alfalfa

California

2006–2008

Geocoris spp.

Spain

1994–2003

Cotton Cotton

Maricopa Iran

2018 2020

Lygus sp., Adelphocoris lineolatus Geocoris punctipes Geocoris megacephalus

Pons et al. (2005) Hagler et al. (2021) Sakenin et al. (2021)

2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise

47

Table 2.8 Anthocoridae predators recorded from various agroecosystems: a worldwide report with reference Orchards

Italy

Alfalfa

Spain

1995 and 1997 1994–2003

Wheat

Spain

2007

Graphs Vegetable crops

Tunisia Spain

2004 2009, 2010

Vegetables, orchards crop fields

Turkey

2015–2016

Crops

East Africa



Mango

México

2018

México

2008, 2009

Kenya Spain

2009 2009, 2010

French bean Lettuce

Anthocoris nemoralis Orius majusculus Orius niger Orius niger Orius albidipennis Orius niger Orius laevigatus Orius pallidicornis Orius vicinus Orius horvati Orius laevigatus Orius majusculus Orius laevigatus Orius minutus Orius laticollis Orius horvathi Orius albidipennis Orius niger Orius albidipennis Orius horvathi Orius laevigatus Orius minutes Orius niger Orius vicinus O. albidipennis O. tantillus O. thripoborus Orius insidiosus Orius insidiosus Orius tristicolor Orius perpunctatus Orius spp. Orius majusculus Orius minutus Orius laticollis Orius horvathi Orius laevigatus Orius albidipennis Orius niger

Burgio et al. (2006) Pons et al. (2005) Pérez-Fuertes et al. (2015)

Ben Moussa (2004) Gomez-Polo et al. (2013)

Pehlivan and Atakan (2020)

Hernández (1999)

Carrillo-Arámbula and Infante (2021) Rocha et al. (2015)

Nyasani et al. (2013) Gomez-Polo et al. (2016)

(continued)

48

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Table 2.8 (continued) Cucumber Rice Sweet pepper Soybeans

Beijing China Italy Iowa

Corn Transgenic corn Cotton

China USA Maricopa Egypt

2011 2015 2002–2003 1977 and 1978 2015 2004 2018 1996, 1997

Iran USA Mexico

2020 2007 2016

Graphs Jetropa

Orius sauteri Orius tantilus Orius laevigatus Orius insidiosus Orius sauteri Orius insidiosus Orius tristicolor Orius albidipennis O. laevigatus Anthocoris minki Orius spp. Orius tristicolor

Raen et al. (2016) Bosco et al. (2008) Bechinski and Pedigo (1981) Zhang et al. (2021) Harwood et al. (2005) Hagler et al. (2021) El-Heneidy et al. (1996, 1997) Sakenin et al. (2021) Costello et al. (2021) Ramos-Robles et al. (2021)

Ballal and Yamada (2016) in their review paper highlighted the following account related to Indian anthecorids. The most prevalent anthocorids found in various crop ecosystems in India are Orius spp., with Orius maxidentex and Orius tantillus being the most widespread. In the sunflower environment, Orius maxidentex was noted to be a potential predator of Helicoverpa armigera. It has also been reported that Orius species have a seasonal abundance and distribution pattern that closely resembles that of Helicoverpa armigera eggs. On the reproductive sections of pigeon pea and sorghum plants, Orius tantillus has been reported to be an active predator of Helicoverpa armigera eggs and first instar larvae. Orius albidipennis records from India need to be verified because they are speculative. It is understood that this species is only found in southern Europe, Africa, and the Middle East. Orius maxidentex is more than likely the species that has previously been reported from India as Orius albidipennis. Orius spp., including Orius niger, Orius dravidiensis Muraleedharan, Orius shyamavarna, Orius niger aegypitiacus, Orius minutus, and Orius amnesius from various host plants, were further documented as a result of the authors’ surveys; the last species set a new distribution record for India. Anthocorids C. exiguus, Cardiastethus affinis, and B. sodalist have been identified as possible predators of the serious pest Orius arenosella that damages coconut plantations in south India. Despite the fact that C. exiguus is listed as a potential predator of Orius arenosella, the authors were able to document its association with a wide range of pests, including thrips, mites, and mealybugs on cashew, papaya, rose, mango, jamun, Tecoma stans, Butea monosperma, Thespesia, Cassia javanica, Caesalpinia pulcherrima, Aegle marmelos, Areca, Ligustrum, and the dry fruits of Adenanthera pavonina and Delonix regia. The authors found Orthesia (on Lantana) and Hemiberlesia lataniae (on Agave) to be linked with Cardiastethus affinis, a predator of Orius arenosella, and they also found fallen leaves and flowers of Spathodea campanulata. In Kerala state, Orthaga exvinacea, which infests mango, is preyed upon by Cardiastethus affinis. India has

2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise

49

very little information about the genus Blaptostethus. The only records that exist are those of Blaptostethus kumbi, which was discovered in Mysore sugarcane fields, and Blaptostethus pallescens, which was discovered in Bombay and Tamil Nadu. B. pallescens was discovered in Karnataka as a potential predator of the Tetranychus urticae, Chilo partellus, and maize stem borer, respectively. The mealybug Ferrisia virgata, which infests the purple orchid tree Bauhinia purpurea, has been preyed upon by Anthocoris muraleedharani, whereas M. indica has been observed in Karnataka as a predator of G. uzeli, which infests Ficus retusa. Xylocoris species have been found under leaf litter, under plants, behind tree bark, and even within grains that have been preserved. Two species of the genus Xylocoris, Xylocoris (Proxylocoris) clarus and Xylocoris (Arrostelus) flavipes, were found in India. Xylocoris (Proxylocoris) afer, Xylocoris (Proxylocoris) confusus, and Xylocoris (Arrostelus) ampoli were identified in India from the dried fruits of Ficus and Lagerstroemia. In India, Xylocoris flavipes is known to be a frequent predator of pests found in stored grains. Other Xylocoris species, including Xylocoris afer, Xylocoris ampoli, and Xylocoris confusus, may also serve a use in managing pests in various crop ecosystems and storage systems. In our country’s mountainous, steep terrain, anthocorids have been found. Tetraphleps raoi was discovered on Benguet pine trees in Shillong that were growing at a height of roughly 5000 ft. and were afflicted by Pineus sp. The Ladakh region produced Anthocoris flavipes and Temnostethus (Ectemnus) paradoxus, whereas the Kashmir region produced Anthocoris sp., Orius minutus, and Orius lindbergi. In Himachal Pradesh, Anthocoris minki and Anthocoris confusus were discovered. Nabidae: There are only a few species, and the most of them are from the genus Nabis. There are a lot of data to suggest that nabids react to habitat alteration, like what happens in agroecosystems. More nabids were sustained by maize and bean inter-cropped fields than by monocultures. Nabids were more prevalent in soybean fields than other fields. Damselflies can be found on a wide variety of plants, including agricultural crops like alfalfa, cotton, soybeans, orchards, potato, organic strawberry fields, etc. in the USA from North Carolina to Texas and Florida, and southward into Central America. They also have a range in South America, Russia, Africa, and Europe. These predators are recorded both in normal and genetically modified crop worldwide (Table 2.9). Reduviids are distributed mainly in various agroecosystems such as ailanthus, alfalfa, bendhi, cabbage, cardamom ecosystem, chillies, citrus, coco, coconut, oil palm, cotton, cowpea, fruit crops and pecan, jetropa, lucerne, maize, mustard, organic vineyard, pigeon pea, pumpkin, rice, sugarcane, sunflower, sweet potato, tobacco, watermelon, etc. (Table 2.10). In Italy, Zelus renardii has been collected from various agroecosystems listed in Table 2.11. It also lived in storage facilities including warehouses, godowns, grain mills, and shipment sites in addition to agroecosystems. For instance, Amphibolus venator preys on a variety of insects found in stored goods. It preys on Alphitobius diaperinus, Corcyra cephalonica, Latheticus oryzae, Trogoderma granarium, Tribolium castaneum, and Tribolium castaneum (Pingale 1954; Haines 1991).

50

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Table 2.9 Anthocoridae predators recorded from various agroecosystems: a worldwide report with reference Crops Alfalfa, soybean

Country USA

Period 1983 and 1984

Wheat

South Dakota Germany

1988–1990

Spain USA

1994–2003 1988–1992

USA

2014–2015

Orchards

Italy

1995 and 1997

Alfalfa + cotton Soybeans

California’s

2006–2008

Iowa

Nabis spp.

Cotton

Iran

1977 and 1978 2020

Transgenic corn

USA

2004

Nabis roseipennis

Alfalfa

1993–1995

Predator Nabis americoferus, Nabis roseipennis, and Nabis rufusculus Nabis americoferus Aptus mirmicoides Nabis brevis Nabicula flavomarginata Nabis major Nabis ferus Nabis pseudoferus Nabis provencali Nabis americoferus Nabis americoferus Nabis ferus Aptus mirmicoides Nabis punctatus Nabis rugosus Nabis spp.

Nabis palifer

References Braman and Yeargan (1990) Elliott et al. (1999) Roth (2003)

Pons et al. (2005) Elliott et al. (2002) Shrestha et al. (2021) Burgio et al. (2006) Sivakoff et al. (2012) Bechinski and Pedigo (1981) Sakenin et al. (2021) Harwood et al. (2005)

According to Pingale (1954), Amphibolus venator is successful in controlling Ephestia cautella and Alphitobius diaperinus in a warehouse trial. Alphitobius venator has frequently been discovered in warehouses in Thailand as well as groundnut shipments from Africa to England (Hill 1990). Amphibolus venator seems to have adapted to live in rice milling operations. Peregrinator biannulipes, another assassin insect, is a well-known natural adversary of pests found in stored goods. It feeds on moths, beetles, Tribolium castaneum, Tribolium confusum, Stegobium paniceum, and Lasioderma serricorne in addition to Anagasta kuehniella, Plodia interpunctella, Corcyra cephalonica, Pyraris farinalis, and moths (Tawfik and Awadallah 1983; Tawfik et al. 1983a, b). Amphibolus venator and P. biannulipes were both discovered in the same rice milling facility.

2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise

51

Table 2.10 Distribution of reduviids (Heteroptera: Reduviidae) in agricultural ecosystems with references Crop name Sugarcane

Pigeon pea

Cabbage Alfalfa Cardamom ecosystem

Cotton

Insect name Acanthaspis quinquespinosa Pristhesancus plagipennis Rhynocoris marginatus Rhynocoris marginatus, Rhynocoris fuscipes, Paralenaeus pyrrhomelas, Ectomocoris Rhynocoris marginatus Irantha armipes Sycanus pyrrhomelas Rhynocoris longifrons Polybia sp. Rhynocoris segmentarius Nagusta goedelii Sycanus indagator Rhynocoris longifrons Endochus migratorius Endochus atricapillus Rihirbus trochantericus Epidaus bicolor Acanthaspis siva Ectomocoris tibialis Acanthespis pedestris Catamiarus brevipennis Ectomocoris tibialis Lophocephalus guerini Rhynocoris marginatus Rhynocoris fuscipes Rhynocoris kumarii Onocephalus pilicornis Coranus aegyptiacus Coranus nodulosus Phonoctonus nigrofasciatus Phonoctonus fasciatus Phonoctonus subimpictus Pisilus tipuliformis Pristhesancus papuensis

Location India Queensland India India

Location references India Butani (1958) Illingworth (1921) Sahayaraj (1999) Claver (2011)

India

Ambrose and Claver (2001)

Sweden South Africa Iran India

Miranda Ortiz (2011) Niba (2011) Rakhshani et al. (2010) Nagarajan and Varadarasan (2013)

India

Kalidas and Sahayaraj (2012)

Egypt 1996, 1997 India India

El-Heneidy et al. (1996, 1997) Singh et al. (1987) Sahayaraj (1991) Evans (1962) Parker (1972)

Nigeria West Africa Australia

Zelus renardii

California

Zelus renardii

America

Parker (1965) Martin and Brown (1984) Cisenros and Rosenheim (1997) van den Bosch and Hagen (1966) (continued)

52

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Table 2.10 (continued) Crop name

Insect name Zelus exsanguis Zelus cervicalis Zelus socius Pristhesancus plagipennis

Location North America

Location references Ables (1978)

Australia

Rhynocoris fuscipes Acanthaspis pedestris Acanthaspis quinquespinosa Acanthaspis subrufa Oncocephalus fuscirostris

India India

Grundy and Maelzer (2000) Singh and Sing (1987) Rajagopal (1984)

Onocephalus pilicornis

Tobacco

Coconut

Fruit crops and pecan Oil palm Coco Sweet potato Cowpea Mustard Lucerne Maize

Rhynocoris punctiventris Agriocleptes bahianus Apiomerus lanipes Coranus atricapillus Coranus spiniscutis Rhynocoris squaliua Catamiarus brevipennis Ectrychotes dispar Rhynocor ismarginatus Coranus atricapillus Coranus spiniscutis Cosmocolopiusnigro annulatus Apiomerus floridensis Arilus cristatus Sineas pinipes Zelus exsanguis Cosmelestes picticeps

Australia Egypt— 1996, 1997 Iran, 2020 Brazil Brazil India India India India India

Miles and Bull (2000), Murray (1982) El-Heneidy et al. (1996, 1997) Sakenin et al. (2021) Marques et al. (2006) Marques et al. (2006) Singh (1985) Bose (1949) Singh (1985) Pawar et al. (1986) Singh (1985)

Brazil

Jahnke et al. (2002)

USA

Mizell and Tedders (1995)

Malaysia

Cheong et al. (2010)

Carcinoma astrologus Oncocephalus subspinosus Coranus spiniscutis Coranus spiniscutis

Ghana Rep.

Babin (2009)

India India

Bose (1949) Bose (1949)

Coranus spiniscutis Pirates ephippiger

India Australia

Coranus spiniscutis Ectomocoris cordiger Cydnocoris gilvus Oncocephalus impudicus Coranus spiniscutis

India India India

Bose (1949) Miles and Bull (2000), Murray (1982) Jalali and Singh (2002) Misra (1975) Bose (1949)

(continued)

2.4 Distribution of Various Predators at Different Agroecosystems Order-Wise

53

Table 2.10 (continued) Crop name

Panthous bimaculatus Polytoxus fuscovittatus Coranus spiniscutis Rhynocoris fuscipes Pristhesancus plagipennis

Location Burkina Faso 2019–2020 India India India India Australia

Sunflower

Pristhesancus plagipennis

Australia

Citrus

Rhynocoris albopunctatus Pristhesancus plagipennis

Uganda Australia

Rhynocoris fuscipes Rhynocoris fuscipes Rhynocoris fuscipes

India India India

Pumpkin

Rhynocoris lapidicola Rhynocoris nitidulus

Water melon

Rhynocoris rubricus Chorosomas chillingi Corizushyoscya mihyoscyami Liorhyssus hyalinus Sinea diadema Barce uhleri

India Nigeria, 2016 Cortia 2010–2012

Ailanthus Rice

Bhindi Chillies

Organic vineyard

Insect name Rhynocoris sp.

Mexico, 2016

Location references Ahissou et al. (2021)

Varma (1989) Satpathy et al. (1975) Bose (1949) Singh (1985) Grundy and Maelzer (2000) Grundy and Maelzer (2000) Nyirra (1970) Grundy and Maelzer (2000) Singh and Sing (1987) Singh and Sing (1987) Cherian and Brahmachari (1941) Joseph (1959) Emmanuel et al. (2019) Franin et al. (2021)

Ramos-Robles et al. (2021)

Jetropa

Table 2.11 Zelus renardii field survey in southern Italy and Albania (Lahbib et al. 2022) Location Bari, Italy Copertino, Lecce, Italy Moro, Italy

Date of collection September 2014, 2015 May 2016 September 2014

Bari, Italy

August 2015

Bari-Bitritto, Italy Bari, Italy Lecce, Italy Tirana, Albania Laknas, Albania

April 2015 September 2017 March 2018 March 2019 March 2019

Plant name Quercus ilex Quercus ilex Citrus spp., Pyracantha coccinea, Pyrus communis, Olea europaea, Ficus carica Ailanthus altissima, Capsicum annuum, Aloysia citriodora Olea europaea Laurus nobilis Laurus nobilis Ocimum basilicum Oleae europaea

54

2.4.9

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Orthoptera

Gryllotalpa gryllotalpa (42) (Gryllotalpidae), Gryllus desertus (07), and Gryllus campestris (05) (Gryllidae) were recorded from the vegetable crops from the south of Oltenia in 2020 (Bîrzanu and Mitrea (2021).

2.5

Factors Responsible for Predators’ Populations

2.5.1

Pests and Their Natural Enemy’s Complex

Predators’ population evolution, geographic distribution, trophic dynamics, and ecosystem function are key factors of insects, as well as inter-specific and intraspecific interactions (such as competition and predation) (Song et al. 2021). Predatory nabid bugs gathered in high aphid density areas, indicating that the rise in aphid abundance might indirectly affect predation (Ostman and Ives 2003). In another study, similarly effects were recorded for ants. Solenopsis invicta is an important biological control agent in sugarcane, cotton in the USA. Solenopsis invicta (ants) were a prevalent predator in Alabama cotton and had a negative correlation with 16 herbivorous species, including lepidopterous larvae and a number of hemipterous pests (Eubanks 2001). Whenever the aphid populations are high in cotton field, its predation population declines (Lewis et al. 1997). Foods that predators need are frequently unavailable in crops, thus they eat another predator. Presences of predator or predators in a crop also play a vital role on the distribution of natural enemies in various crops. According to Cardinale et al. (2003), the pea aphid predation levels in alfalfa were higher when all three species of natural enemies were present than when each predator acted alone. In South Carolina, higher S. invicta concentrations in conservation tillage cotton led to lower levels of numerous other significant predators. Other major natural adversaries tend to have lower numbers in environments that favour ants (McCutcheon 1999).

2.5.2

Conventional Crops

Crop types and their phenology, presence and absence of trichome, structural modified structures, space within and between rows, and inter and trap cropping systems are few factors which govern the distribution and abundance of prey’s and subsequently to the natural enemies. Rotation had a greater impact on predator populations in maize agroecosystems, with multi-year rotation systems having a lot more predators than non-rotation systems (Brust and King 1994). Adult predatory hoverflies need access to floral resources since many of their species depend on nectar for energy and pollen for gametogenesis. Coccinella septempunctata, a species of seven-spotted lady beetle, responded favourably to manipulations of the fragmentation scale but not to treatments that altered vegetation composition. The changes of the vegetation’s mix and

2.5 Factors Responsible for Predators’ Populations

55

fragmentation size had no effect on the beetle Pterostichus melanarius (Banks 1999). Carabids showed stronger system and crop effects than spiders or staphylinids. The benefits of farm management on species richness were rather marginal. Once more, the crop itself appeared to be the primary structural element. It is concluded that crop structure and factors connected to crops play a major role in determining the quantity and quality of predator populations. The agroecological architecture of the landscape is explored in regard to the impact of favourable crops and field size in predator enhancement (Booij and Noorlander 1992). Corn in monocultures had a higher prevalence of the predatory coccinellid beetle Coleomegilla maculata (Coccinellidae) than corn in two different polycultures (Andow and Risch 1985). Additionally, it was discovered that the total number of predators was 20% greater in cotton grown in wheat straw (Waters et al. 1999). An evaluation of the impact of six groundcover management strategies (pinebark, plastic and straw mulches, tillage, herbicide, and natural soil) on the presence of ground beetles was done in the field in a cider-apple orchard (Coleoptera: Carabidae). Pitfall traps were used to gather eight different species of carabids. More than 98% of the total catches were made up of the three most prevalent beetles: Steropus gallega (65.8%), Pseudophonus rufipes (18.2%), and Poecilus cupreus (14.6%) (Miñarro and Dapena 2003). On Langqi Island in Fujian, PR China, field tests were conducted in 2004 to examine the community composition and diversity of predatory arthropods in vegetable fields when Chinese cabbage (Brassica chinensis) was inter-cropped with green cabbage (Brassica oleracea), garlic (Allium sativum), and lettuce (Lactuca sativa). Plot 1 of the inter-cropping plots had two rows of Chinese cabbage and one ridge each of green cabbage, lettuce, or garlic (CG1) (CB1). Plot 2’s Chinese cabbage was planted in the middle of the ridge and had the same ridge’s margins under-sown with either green cabbage (CB2), lettuce (CL2), or garlic (CG2). For comparison with plots 1 and 2, a Chinese cabbage (CK) monoculture plot was set up. CG1 had the highest species richness while CK had the lowest. CL1 had the largest abundance, while CB1 had the lowest. Inter-cropping treatments had much greater diversity indices than CK, with the exception of CL1 (Cai et al. 2010).

2.5.3

Genetically Modified Crops

Five crops make up the majority of the GM crops, and two of them—cotton and corn—are resistant to either just insects or to both insects and herbicides. The other three (canola, sugar beet, and soybean) are herbicide-resistant (Abbas 2018). Different concepts as well as observations have been recorded worldwide about the distribution of natural enemies in genetically engineered crops at field level. Additionally, field experiments showed that Bt crops had significantly higher levels of beneficial arthropods than crops treated with conventional pesticides (Abbas 2018). For instance, Souza et al. (2019) and Liu et al. (2021a, b) proclaimed that predators’ populations can be directly and indirectly affected; therefore, monitoring and studying how these organisms interact in GM plants are necessary.

56

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Yu and colleagues’ (2014) field study also supported this claim. They reports that the dominant distribution of predators and parasitoids in China was unaffected by Bt soybean cultivation. Similar findings were previously noted by Naranjo (2005) in a cotton field in the USA. The author claims that a long-term field plot research in Arizona, USA, evaluating the impact of Bt cotton releasing Cry1Ac toxins on 22 taxa of plant-dwelling arthropod natural enemies over several generations discovered no long-term consequences of Bt cotton. In a different field investigation, it was discovered that populations of common predators like Hyppodamia convergens, Orius insidiosus, and Scymnus spp. in a Bt corn field were similar to those on a normal corn field (Al-Deeb and Wilde 2003). Similar findings in Bt maize were also demonstrated. Bt maize with Cry1Ab toxin was proven to be less harmful to predatory arthropod populations in Spain (De la Poza et al. 2004). In 2019, the same was observed. Based on species richness, variety, and evenness indices, it was discovered that Bt maize expressing Cry1Ab toxins had no effects on arthropod communities in Brazil (Frizzas et al. (2017). According to Hernandez-Juárez et al. (2019), the latter was also observed in Mexico in the same crop that expressed the Cry1Ab, Vip3Aa20, and mCry3A poisons on three non-target predators. A finding demonstrates that the number and frequency of the predators were not negatively impacted by the Bt maize. According to Han et al. (2016), the genetically crops may have an impact on natural enemies in one of three ways: 1. Directly through exposure to insecticidal proteins through feeding on IRGM crop tissues (e.g. omnivorous predators). 2. Indirectly because IRGM crops may cause changes in the crop environment, such as the quantity or nutritional quality of non-prey foods, as well as plant cues that natural enemies rely on when searching for food or shelter. 3. Indirectly because changes in “plant-herbivore-paras”. Direct or indirect effect of GM crops in natural enemy is proposed by O’Callaghan et al. (2005). According to them, the natural enemies could be affected directly through the consumption of GM pollen, plant tissue, or live recombinant proteins in their hosts’ bodies. The indirect consequences could be caused by prey that has consumed GM plants becoming smaller, sicker, or less appetising. Furthermore, in actual situations, many biotic or abiotic restrictions might affect behavioural features like as mobility, foraging (host/prey location, selection, and appropriateness), mating/oviposition, orientation/associative learning, and other behaviours relevant to individual species. A field study demonstrating that three predator species did not differ in abundance before, during, or after pollen discharge in both Bt maize and ordinary maize is one of the few cases (Pilcher et al. 1997). In India, the single (Cry1Ac) and dual (Cry1Ac and Cry2Ab) toxins produced by the transgenic Bt cotton had no discernible impact on the fitness of the predator Chrysoperla zastrowisillemi (Neuroptera: Chrysopidae) (Shera et al. 2018). In Mexico, the impact of the genetically modified (GM) corn hybrid (Agrisure® VipteraTM 3111) on the abundance of non-target predators Orius insidiosus

2.5 Factors Responsible for Predators’ Populations

57

(Hemiptera: Anthocoridae), Coleomegilla maculata (Coleoptera: Coccinellidae), and Chrysoperla carnea (Neuroptera: Chrysopidae). According to the findings, Agrisure® VipteraTM 3111 was more abundant than the isolines. The genus Orius was the most abundant group, while Coccinellidae was least. Less than 10% of the predators observed on GM maize plants were represented by the other predatory groups that were observed (Albajes et al. 2012). Due to the danger they may present to predators as non-target organisms, genetically engineered corn (maize), Zea mays (Poaceae), expressing Bacillus thuringiensis (Bt) Berliner (Bacilaceae), is a contentious topic in Mexico. Therefore, it is critical to assess that risk prior to the introduction of Bt corn for commercial planting in Mexico. In a randomised complete block design with three treatments, the effect of the genetically modified corn hybrid Agrisure® VipteraTM 3111 on the abundance of non-target predators Orius insidiosus (Hemiptera: Anthocoridae), Coleomegilla maculata (Coleoptera: Coccinellidae), and Chrysoperla carnea (Neuroptera: Chrysopidae) was assessed. About 5228 predators in total, with two peaks before and after pollination, were tallied in all hybrids from both localities: 2431 at Oso Viejo and 2797 at El Camalote. Each predator population fluctuated in both places in a manner that was comparable for all hybrids. Agrisure® VipteraTM 3111 was more abundant than the isolines treated with and without insecticide in every instance, despite the lack of a statistically significant difference between treatments. According to the findings, Orius insidiosus, Coleomegilla maculata, and Chrysoperla carnea predator abundance is unaffected by Agrisure® VipteraTM 3111 (Hernández-Juárez et al. 2019).

2.5.4

Field Conditions

Field edge, refuges/residues, and field margins are crucial in agriculture because they offer safe haven for beneficial predators. Remains give soil-dwelling detritivores more substrate, increasing their population density, and, in turn, the amount of prey available to soil-dwelling predators. Additionally, residues provide a safe haven for predatory insects like carabids and rove beetles (House and Brust 1989). The abundance of foliar predators and overall predators in field margins was dramatically raised by floral plantings. Predators in general or foliar predators were not significantly affected by the amount of floral cover in field margins. Additionally, the presence of floral plantings greatly increased the abundance of epigeal predators, but increased floral cover had no appreciable effects on epigeal abundance (Middleton et al. 2021). Impact of tillage practice such as organic matter, soil moisture, and nutrient regime on the predators abundance as also recorded in wheat at India. Coccinella septempunctata, wasps (Ropalidia spp.), and rove beetle (Paederus fuscipes) were commonly present in the field (Jasrotia et al. 2021), which shows no influence of various tillage systems in wheat. Field boundary is another factor which regulates govers fly movement from field to field (Wratten et al. 2003). Richer plant species are another major aspect that influences the predatory environment. It is anticipated

58

2

Distribution and Diversity of Predatory Insects in Agroecosystems

to make plants more structurally complex, which could limit or eliminate the need for predators to migrate in search of food, partners, or shelter. Reduced movement may lower the trapping and result in an under estimating of the abundance of predators. This can be caused by the eco-accessibility systems to substitute herbivore prey. The establishment and growth of chrysopid populations is also aided by ground cover plants. Numerous elements will affect and decide the variety and quantity of Carabidae in agricultural settings. Some of these factors are understood, but because predators respond to a wide variety of environmental factors, it is often difficult to quantify the impact of any individual factor. Furthermore, a wide diversity of species with different environmental requirements has been found within agricultural areas, although many are relatively rare or are migrants from adjacent non-crop habitats. Even within agricultural areas, there are distinctions between the fauna of arable areas with frequent soil cultivation and less disturbed grassland habitats (Holland and Luff 2000). Presence of too many natural enemies in an agroecosystem during sampling might have both negative effects and positive impacts, but a comparison between these two effects farmer is less prominent whereas the latter is more useful to the crops, environments and human beings. Negative impact includes both intraguild and inter-guild competitions amoung the natural enemies’ communities into account for considering competition for space or habitat occupation, which results either be positive or negative to crops.

2.5.5

Pesticides Application

Common pesticides, in general, are not targeted and frequently kill natural enemy populations, which can disturb ecosystems and lead to the recovery of other pest populations. Additionally, it is not unusual for insect biotypes to evolve resistance to the popular insecticides. There are countless studies that have documented how pesticides killed other non-target arthropods, created major ecological imbalances, and impacted the biodiversity of arthropods.

2.5.6

Landscape/Urbanisation

Urbanisation is spreading around the globe, changing environmental and habitat conditions and having a negative impact on creatures that live in urban settings. Arthropod abundance can be impacted by urbanisation in a variety of ways. Many habitat generalist species with good dispersing capacity attain large populations in urban areas, in contrast to species with restricted habitat range and low dispersal ability, which frequently react negatively to urban conditions. Because of the filtering effect that urban settings have on predator–prey and mutualist interactions, both direct and indirect effects on the number of predatory and phytophagous species may result. In this study, we evaluated how urbanisation affected aphids, which are predatory arthropods (Korányi et al. 2021).

2.6 Recommendation

59

Over the past 10 years, biological diversity has significantly decreased as a result of habitat loss and degradation. One of the main causes of habitat loss has been determined to be the transformation of natural habitats into agricultural landscapes. The population of natural enemies benefits greatly from localised increases in plant diversity. However, these beneficial outcomes frequently depend on the landscape context, which affects the variety of natural enemies present and their capacity to colonise newly developed habitats. Additionally, they discovered that diversified cover crops (2 versus 20 plant species) in nine pairs of vineyards increased natural enemy numbers by 140% overall (Beaumelle et al. 2021). The composition of local species able to inhabit the last remaining patches of fragmented habitat can change as a result of changes in landscape heterogeneity brought on by urbanisation, which typically results in a decline in native species abundance and richness when combined with human interventions. Urban insect biodiversity is also known to be influenced by local habitat variables that depend on human design and management, such as vegetation structure (i.e. vegetation height, bloom abundance). The direction and amount of these effects on insect populations and crop damages are determined by landscape structure, which has an impact on natural enemies and trophic interactions (Karp et al. 2018). The following idea was recently put up by Beaumelle et al. (2021), i.e. semi-natural environments like hedgerows, grasslands, or woods support essential resources and habitats for natural enemies including alternative prey, nectar, pollen, or overwintering locations. To filter fertilisers and solid materials from agricultural runoff, Finland has built new wetlands. Numerous studies show that the environment (site occupancy, extinction, or colonisation) favours odonates. Odonata may gain from these wetlands as new breeding grounds. But it is still unclear whether aspects of artificial wetlands’ environments are crucial for Odonata (Huikkonen et al. 2020). The profusion of odonates in rice paddies in Tochigi Prefecture, Japan, was influenced by eco-friendly agricultural practices in addition to the landscape (area of forest at the edge of the paddy field) (Baba et al. 2019). Three odonates including Tramea onusta, Epitheca princeps, Pantala flavescens, and Calopteryx maculate are best preserved in Iowa due to the state’s landscape (Harms et al. 2014). A 14-year study in Mexico found a quadratic link between human footprint and species richness, and a positive relationship between species richness and forest cover (Cuevas-Yáñez et al. 2017). The consequences of urbanisation on ground beetles at all scales, from the population to the (sub) community to the ecosystem, have not yet been summarised in a review study. We attempt to close this gap in the current article by examining 139 urban research on ground beetles that have been published (Fig. 2.1).

2.6

Recommendation

• Exact factor responsible for the distribution and abundance of each predatory group should be studied with worldwide network. • Within the ecosystems how a predator or combination of predators affected by biotic and abiotic factors should be studied in detail.

60

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Fig. 2.1 Urban studies on ground beetles that were published each year between 1975 and 2020. The following search phrases were used on 4 April 2020, in Web of Science to conduct the literature search: Topics include “urbanisation OR urbanisation” and “carabid*”. A substantial link can be seen in the fitted second-order polynomial curve (F2, 30 = 52.33, p < 0.001, R2 = 0.76). (After Magura and Lövei 2021)

• Influence of various plant protections on the population abundance should be highlighted. • Various urbanisation on the populations of different predators should be studied.

References Abbas MST (2018) Genetically engineered (modified) crops (Bacillus thuringiensis crops) and the world controversy on their safety. Egypt J Biol Pest Control 28(1):1–12 Ables JR (1978) Feeding behaviour of an assassin bug, Zelus renardii. Ann Entomol Soc Am 71:476–478 Ambrose DP, Claver MA (2001) Survey of reduviid predators in seven pigeonpea agroecosystems in Tirunelveli, Tamil Nadu, India. Int Chickpea and Pigeonpea News 8:44–45 Aguirre LA, Hernández-Juárez A, Cerna E, Flores M, Frías GA, Ochoa YM (2021) Diversity, abundance, and effect of genetically modified maize on nontarget predators in Sinaloa, Mexico. J Entomol Sci 56(4):541–555 Agwu EJ, Odo GE, Ekeh F, Attamah GN, Uwagbae M, Eze C (2018) A survey of the insect pests and farmers practices in the cropping of yellow pepper Capsicum annuum Linnaeus in Enugu State of Eastern Nigeria. Afr J Agric Res 13(15):742–752 Ahissou BR, Sawadogo WM, Bonzi S, Baimey H, Somda I, Bokonon-Ganta AH, Verheggen FJ (2021) Natural enemies of the fall armyworm Spodoptera frugiperda (Smith) (Lepidoptera: Noctuidae) in Burkina Faso. Tropicultura 1881 39(2021):2295–8010

References

61

Akter S, Alam MZ, Rana MS, Saif HB (2018) Impacts of three pesticides on arthropod biodiversity in brinjal agro-ecosystem. Int J Agric Res Innov Technol 8(1):55–61 Albajes R, Farinós GP, Pérez-Hedo M, Poza MDL, Lumbierres B, Ortego F, Castañera P (2012) Post-market environmental monitoring of Bt maize in Spain: non-target effects of varieties derived from the event MON810 on predatory fauna. Span J Agric Res 10(4):977–985 Al-Deeb MA, Wilde GE (2003) Effect of Bt corn expressing Cry3Bb1 toxin for corn rootworm on above ground non-target arthropods. Environ Entomol 32(5):1164–1170 Amala U, Shivalingaswamy TM (2018) Effect of intercrops and border crops on the diversity of parasitoids and predators in agroecosystem. Egypt J Biol Pest Control 28(1):1–4 Amaral DSSL, Venzon M, Marcus VA, Duarte FFS, Pallini A, James DH (2013) Non-crop vegetation associated with chili pepper agroecosystems promote the abundance and survival of aphid predators. Biol Control 64(3):338–346 Ambrosino MD, Luna JM, Jepson PC, Wratten SD (2006) Relative frequencies of visits to selected insectary plants by predatory hoverflies (Diptera: Syrphidae), other beneficial insects, and herbivores. Environ Entomol 35(2):394–400 Andersen A (1997) Densities of overwintering carabids and staphylinids (Col, Carabidae and Staphylinidae) in cereal and grass fields and their boundaries. J Appl Entomol 121(1–5):77–80 Andow DA, Risch SJ (1985) Predation in diversified agroecosystems: relations between a coccinellid predator Coleomegilla maculata and its food. J Appl Ecol 22:357–372 Anggraini E, Pardingotan R, Herlinda S, Irsan C, Harun MU (2020) Diversity of predatory arthropods in soybean (Glycine max L.) Refugia. J Agric Sci Technol 4(2):101–117 Ativor IN, Afreh-Nuamah K, Billah MK, Obeng-Ofori D (2012) Weaver ant, Oecophyllalonginida (Latreille) (Hymenoptera: Formicidae) activity reduces fruit fly damage in citrus orchards. J Agric Sci Technol 2:449–458 Baba YG, Kusumoto Y, Tanaka K (2019) Positive effect of environmentally friendly farming on paddy field odonate assemblages at a small landscape scale. J Insect Conserv 23(3):467–474 Babin R (2009) Contribution à l'amélioration de la lutte contre le miride du cacaoyer Sahlbergella singularis Hagl (Hemiptera, Miridae) Influence des facteurs agroécologiques sur la dynamique des populations du ravageur. Doctoral dissertation, UPV Ballal CR, Yamada K (2016) Anthocorid predators. In: Omkar (ed) Ecofriendly pest management for food security. Academic Press, London, pp 183–216 Balog A, Kiss J, Szekeres D, Szenasi A, Marko V (2010) Rove beetle (Coleoptera: Staphylinidae) communities in transgenic Bt (MON810) and near isogenic maize. Crop Prot 29(6):567–571 Bambaradeniya CNB, Edirisinghe JP (2009) Composition, structure and dynamics of arthropod communities in a rice agro-ecosystem. Ceylon J Sci (Biol Sci) 37(1):23–48 Bambaradeniya CNB, Edirisinghe JP, De Silva DN, Gunatilleke CVS, Ranawana KB, Wijekoon S (2004) Biodiversity associated with an irrigated rice agro-ecosystem in Sri Lanka. Biodivers Conserv 13(9):1715–1753 Banks JE (1999) Differential response of two agroecosystem predators, Pterostichusmelanarius (Coleoptera: Carabidae) and Coccinellaseptempunctata (Coleoptera: Coccinellidae), to habitatcomposition and fragmentation-scale manipulations. Can Entomol 131(5):645–657 Beaumelle L, Auriol A, Grasset M, Pavy A, Thiéry D, Rusch A (2021) Benefits of increased cover crop diversity for predators and biological pest control depend on the landscape context. Ecol Solut Evid 2(3):e12086 Bechinski EJ, Pedigo LP (1981) Ecology of predaceous arthropods in Iowa soybean agroecosystems. Environ Entomol 10(5):771–778 Ben Moussa H (2004) Importance of the genus Orius (Heteroptera: Anthocoridae) in the biological control of Thripidae in organic grapevine particularly in Apulia region (Italy) and Tunisia. Théses et Masters (CIHEAM) 364:97 Bhat DM, Bhagat RC (2017) Host range and diversity of syrphid predators (Insecta: Diptera) of aphids on vegetable crops of Kashmir, with new host aphid/plant records. Trends Biosci 10(6): 1446–1448

62

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Bîrzanu FO, Mitrea I (2021) Research on the entomofauna of orthoptera in some vegetable crops in southern Oltenia. Curr Trends Nat Sci 10(19):187–193 Booij CJH, Noorlander J (1992) Farming systems and insect predators. Agric Ecosyst Environ 40(1–4):125–135 Borkakati RN, Saikia DK, Buragohain P (2018) Natural enemy fauna of paddy and horticultural ecosystems in upper Assam. Indian J Entomol 80(3):658–661 Bosco L, Giacometto E, Tavella L (2008) Colonization and predation of thrips (Thysanoptera: Thripidae) by Orius spp. (Heteroptera: Anthocoridae) in sweet pepper greenhouses in Northwest Italy. Biol Control 44(3):331–340 Bose M (1949) On the biology of Coranus spiniscutis Reuter, an assassin bug (Heteroptera: Reduviidae). Indian J Entomol 11:203–208 Bouvet JPR, Urbaneja A, Monzó C (2020) Aphid predators in citrus crops: the least voracious predators are the most effective. J Pestic Sci. https://doi.org/10.1007/s10340-020-01265-z Bouvet JPR, Urbaneja A, Pérez-Hedo M, Monzó C (2019) Contribution of predation to the biological control of a key herbivorous pest in citrus agroecosystems. J Anim Ecol 88(6): 915–926 Braman SK, Yeargan KV (1990) Phenology and abundance of Nabis americoferus, N. roseipennis, and N. rufusculus (Hemiptera: Nabidae) and their parasitoids in alfalfa and soybean. J Econ Entomol 83(3):823–830 Brown MW, Schmitt JJ (2001) Seasonal and diurnal dynamics of beneficial insect populations in apple orchards under different management intensity. Environ Entomol 30(2):415–424 Brunke AJ, Bahlai CA, Klimaszewski J, Hallett RH (2014) Rove beetles (Coleoptera: Staphylinidae) in Ontario, Canada soybean agroecosystems: assemblage diversity, composition, seasonality, and habitat use. Can Entomol 146(6):652–670 Brust GE, King LR (1994) Effects of crop rotation and reduced chemical inputs on pests and predators in maize agroecosystems. Agric Ecosyst Environ 48(1):77–89 Burgio G, Ferrari R, Boriani L, Pozzati M, van Lenteren J (2006) The role of ecological infrastructures on Coccinellidae (Coleoptera) and other predators in weedy field margins within northern Italy agroecosystems. Bull Insectol 59(1):59 Butani DK (1958) Parasites and predators recorded on sugarcane pests in India. Indian J Entomol 20:270–282 Cai H, You M, Lin C (2010) Effects of intercropping systems on community composition and diversity of predatory arthropods in vegetable fields. Acta Ecol Sin 30(4):190–195 Cardinale BJ, Harvey CT, Gross K, Ives AR (2003) Biodiversity, and biocontrol: emergent impacts of a multi-enemy assemblage on pest suppression and crop yield in an agroecosystem. Ecol Lett 6:857–865 Carrillo-Arámbula L, Infante F (2021) Development, survival, and reproduction of Oriusinsidiosus (Hemiptera: Anthocoridae) reared on Frankliniellainvasor (Thysanoptera: Thripidae). J Insect Sci 21(3):4 de Carvalho AS, Molina-Rugama AJ, Silva FEL, Costa EM, Araujo EL (2018) Predator insect families associated with melon crop in a semiarid region in Brazil. J Agric Sci 10(10):267–273 Chakraborty A, Kumar K, Rajadurai G (2014) Biodiversity of insect fauna in Okra (Abelmoschus esculentus (L.) Moench) ecosystem. Trends Biosci 7(16):2206–2211 Cheong YL, Sajap AS, Hafidzi MN, Omar D, Abood F (2010) Outbreak and natural enemy of bagworm. J Entomol 7(3):141–151 Cherian MC, Brahmachari K (1941) Notes on three predatory hemipteran from South India. Indian J Entomol 3:115–118 Cisneros F (1995) Control de plagasagrícolas. Editorial Impresión, Lima Cisenros JJ, Rosenheim AJ (1997) Ontogenetic change of prey preference in the generalist predator Zelus renardii and its influence on predator-predator interactions. Ecol Entomol 22(4):339–407 Claver MA (2011) Biodiversity of pegionpea insect pests and their predatory insects in five districts of North Eastern Uttar Pradesh, India. In: Proceeding of National conference on biotechnology for sustainable development. Gorakpur University, Uttar Pradesh, pp 127–132

References

63

Collins KL, Boatman ND, Wilcox A, Holland JM, Chaney K (2002) Influence of beetle banks on cereal aphid predation in winter wheat. Agr Ecosyst Environ 93(1–3):337–350 Corrales N, Campos M (2004) Populations, longevity, mortality, and fecundity of Chrysoperla carnea (Neuroptera, Chrysopidae) from olive-orchards with different agricultural management systems. Chemosphere 57(11):1613–1619 Costello MJ, Thrupp A, McGourty G (2021) Influence of vineyard vegetational borders on western grape leafhopper (Erythroneuraelegantula Osborn), its egg parasitoids (Anagrus spp.) and generalist insect predators. S Afr J Enol Vitic 42(1):25–35 Cuevas-Yáñez K, Benítez M, Rocha M, Córdoba-Aguilar A (2017) Large-scale human environmental intervention is related to a richness reduction in Mexican odonates. Rev Mex Biodivers 88(3):664–673 Curtis JE, Horne PA (1995) Effect of Chlorpyrifos and cypermethrin applications on non-target invertebrates in a conservation-tillage crop. Aust J Entomol 34(3):229–231 Duyck P-F, Lavigne A, Vinatier F, Achard R, Okolle JN, Tixier P (2011) Addition of a new resource in agroecosystems: do cover crops alter the trophic positions of generalist predators? Basic Appl Ecol 12(1):47–55 Dwari S, Mondal AK (2017) Diversity of Odonates in agricultural fields of Howrah district, West Bengal, India. J Entomol Zool Stud 5(3):1588–1595 El-Heneidy AH, Fayad YH, Gonzalez D, Abdel-Salam NM, Ellington J, Moawad GM (1996) Pestpredator-interactions in untreated cotton fields at three plant growth stages. 1- location impact. Egypt J Biol Pest Control 6(1):75–81 El-Heneidy AH, Ibrahin AA, Gonzalez D, Abdel-Salam NM, Ellington J, Moawad GM (1997) Pest-predator -interactions in untreated cotton fields at three plant growth stages. 2-Planting date impact. Egypt J Agric Res 75(1):137–155 Elliott NC, Kieckhefer RW (1990) A thirteen-year survey of the aphidophagous insects of alfalfa. Prairie Nat 22(2):87–96 Elliott NC, Kieckhefer RW, Lee JH, French BW (1999) Influence of within-field and landscape factors on aphid predator populations in wheat. Landsc Ecol 14(3):239–252 Elliott NC, Kieckhefer RW, Michels GJ, Giles KL (2002) Predator abundance in alfalfa fields in relation to aphids, within-field vegetation, and landscape matrix. Environ Entomol 31(2): 253–260 EL-Sheikh W, El-Kenway A, Soliman M (2020) Biodiversity and population dynamics of natural enemies in the western desert agro-ecosystem, Egypt. Egypt Acad J Biol Sci A Entomol 13(4): 1–14 Emmanuel O, Emmanuel Oludele O, MondayUnwabunne U (2019) Diversity, spatial and temporal distribution of above-ground arthropods associated with watermelon in the Nigerian southern guinea savanna. J Insect Biodivers Syst 5(1):11–32 Eubanks MD (2001) Estimates of the direct and indirect effects of red imported fire ants on biological control in field crops. Biol Control 21:35–43 Evans DE (1962) The food requirements of phonoctonus nigrofasciatus stål.(hemiptera, reduviidae). Entomologia Experimentalis et Applicata 5(1):33–39 Farinós GP, de la Poza M, Hernández-Crespo P, Ortego F, Castañera P (2008) Diversity and seasonal phenology of aboveground arthropods in conventional and transgenic maize crops in Central Spain. Biol Control 44(3):362–371 Franin K, Gabrijela KF, BrankaMaricic SM, Pavlovic M, Kos T, Baric B, Laznik Z (2021) True bugs (Heteroptera) assemblage and diversity in the ecological infrastructures around the Mediterranean vineyards. Bull Insectol 74(1):65–78 Frank T, Reichhart B (2004) Staphylinidae and Carabidae overwintering in wheat and sown wildflower areas of different age. Bull Entomol Res 94(3):209–217 Frizzas MR, Oliveira CMD, Omoto C (2017) Diversity of insects under the effect of Bt maize and insecticides. Arq Inst Biol 84:e0062015

64

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Gomez-Polo P, Alomar O, Castané C, Riudavets J, Agustí N (2013) Identification of Oriusspp. (Hemiptera: Anthocoridae) in vegetable crops using molecular techniques. Biol Control 67(3): 440–445 Gomez-Polo P, Alomar O, Castañé C, Aznar-Fernández T, Lundgren JG, Piñol J, Agustí N (2016) Understanding trophic interactions of Oriusspp. (Hemiptera: Anthocoridae) in lettuce crops by molecular methods. Pest Manag Sci 72(2):272–279 Grabovska T, Lavrov V, Putchkov O (2021) Diversity of entomofauna in organic versus conventionally managed soybean fields protected by forest shelter belts in Ukraine. Org Agric 11(4):625–638 Greyvenstein B, Siebert SJ, Van Den Berg J (2020) Effect of time of day on efficacy of sweep net sampling of arthropod predators in maize agro-ecosystems in the Northwest Province, South Africa. Afr Entomol 28(1):150–163 Grundy PR, Maelzer DA (2000) Predation by the assassin bug Pristhesancus plagipennis (Walker) (Hemiptera, Reduviidae) of Helicoverpa armigera (Hübner) (Lepidoptera, Noctuidae) and Nezara viridula (L) (Hemiptera, Pentatomidae) in the laboratory. Australian J Entomol 39:280–282 Gul SU, Rehman HA, Ahmad N, Din A, Jawad MS, Noureen S, Khatoon N, Ali M, Ashfaq MA (2017) Medicinal plants feeding insect fauna of district Karak, KP, Pakistan. J Entomol Zool Stud 5(6):2563–2566 Guo M, Wang Z, Cai W, Hua H, Zhao J (2020) Safety assessment of transgenic Cry2Aa rice to a generalist predator, Paederus fuscipes Curtis (Coleoptera: Staphylinidae). Ecotoxicol Environ Saf 200:110719 Hagler JR, Thompson AL, Machtley SA, Casey MT (2021) Arthropod demography, distribution, and dispersion in a novel trap-cropped cotton agroecosystem. J Insect Sci 21(1):20 Haines CP (1991) Insects and arachnids of tropical stored products: their biology and identification (a training manual), 2nd edn. Natural Resources Institute, Chatham, pp 75–76 Halimov F (2020) Seasonal dynamics of dominant species of soil predators (Coleoptera: Carabidae, Staphylinidae) in agrolandscapes and their potential gluttony. Trav Mus Nat Hist Nat Grigore Antipa 63:175 Hammond RB, Stinner BR (1987) Soybean foliage insects in conservation tillage systems: effects of tillage. previous cropping history. and soil insecticide application. Environ Entomol 16:524– 531 Han P, Velasco-Hernández MC, Ramirez-Romero R, Desneux N (2016) Behavioral effects of insect-resistant genetically modified crops on phytophagous and beneficial arthropods: a review. J Pest Sci 89(4):859–883 Happe AK, Roquer-Beni L, Bosch J, Alins G, Mody K (2018) Earwigs and woolly apple aphids in integrated and organic apple orchards: responses of a generalist predator and a pest prey to local and landscape factors. Agric Ecosyst Environ 268:44–51 Harms TM, Kinkead KE, Dinsmore SJ (2014) Evaluating the effects of landscape configuration on site occupancy and movement dynamics of odonates in Iowa. J Insect Conserv 18(3):307–315 Harterreiten-Souza ÉS, Togni PH, Capellari RS, Bickel D, Pujol-Luz JR, Sujii ER (2021) Spatiotemporal dynamics of active flying Diptera predators among different farmland habitats. Agric For Entomol 23(3):334–341 Harwood JD, Wallin WG, Obrycki JJ (2005) Uptake of Bt endotoxins by nontarget herbivores and higher order arthropod predators: molecular evidence from a transgenic corn agroecosystem. Mol Ecol 14(9):2815–2823 Helsen H, Vaal F, Blommers L (1998) Phenology of the common earwig Forficula auricularia L. (Dermaptera: Forficulidae) in an apple orchard. Int J Pest Manage 44(2):75–79 Hernández LM (1999) A review of the economically important species of the genus Orius (Heteroptera: Anthocoridae) in East Africa. J Nat Hist 33(4):543–568 Hernández-Juárez A, Aguirre LA, Cerna E, Flores M, Frías GA, Landeros J, Ochoa YM (2019) Abundance of non-target predators in genetically modified corn. Fla Entomol 102(1):96–100 Hill DS (1990) Pests of stored products and their control. Belhaven Press, London, 69p

References

65

Hiral P, Shukla A, Pratik S (2018) Biodiversity of praying mantids (Insecta: Mantodea) in Gujarat, India. J Entomol Zool Stud 6(4):455–459 Holland JM, Luff ML (2000) The effects of agricultural practices on Carabidae in temperate agroecosystems. Integr Pest Manage Rev 5(2):109–129 House GJ, Brust GE (1989) Ecology of low-input, no-tillage agroecosystems. Agric Ecosyst Environ 27:331–345 Huikkonen IM, Helle I, Elo M (2020) Heterogenic aquatic vegetation promotes abundance and species richness of Odonata (Insecta) in constructed agricultural wetlands. Insect Conserv Divers 13(4):374–383 Illingworth JF (1921) The linear bug Phadnacantha australia Kirkaldy, a new pest of sugar cane in Queensland. In: Queensland Bureau of Sugar Experimental Station, Division of Entomology. Brisbane Bulletin, p 14 Jahnke SM, Redaelli LR, Diefenbach LMG (2002) Population dynamics of Cosmoclopius nigroannulatus Stal (Hem; Reduviidae) in tobacco culture. Rev Bras Biol 62(4B):819–826 Jalali SK, Singh SP (2002) Seasonal activity of stem borers and their natural. Entomon 27(2):137– 146 Jalilian F (2019) Diversity indeces of predaceuse Syrphid flies (Diptera: Syrphidae) in some agroecosystems of Kermanshah Province. Biocontrol Plant Prot 7(1):1–13 Jambagi SR, Kambrekar DN, Mallapur CP, Naik VR (2022) Insecticide compatibility with abundance and diversity of predatory fauna in wheat (Triticum aestivum) ecosystem Jasrotia P, Bhardwaj AK, Katare S, Yadav J, Kashyap PL, Kumar S, Singh GP (2021) Tillage intensity influences insect-pest and predator dynamics of wheat crop grown under different conservation agriculture practices in rice-wheat cropping system of Indo-Gangetic Plain. Agronomy 11(6):1087 Joseph MT (1959) Biology, binomics and economic importance of some reduviids collected from Delhi. Indian J Entomol 21:46–58 Kalidas S, Sahayaraj K (2012) Survey of reduviids in cotton agro-ecosystem of Tamil-Nadu, India. Middle East J Sci Res 12(9):1216–1223 Karp DS, Chaplin-Kramer R, Meehan TD, Martin EA, DeClerck F, Grab H, Gratton C, Hunt L, Larsen AE, Martínez-Salinas A, O’Rourke ME, Rusch A, Poveda K, Jonsson M, Rosenheim JA, Schellhorn NA, Tscharntke T, Wratten SD, Zhang W et al (2018) Crop pests and predators exhibit inconsistent responses to surrounding landscape composition. Proc Natl Acad Sci U S A 115:E7863–E7870. https://doi.org/10.1073/pnas.1800042115 Kausar S, Abbas MN, Mahmood-ul-Hassan M, Rana SA, Rana N, Maqsood I, Gul I (2017) Diversity and dynamics of macro-invertebrates in sugarcane agro-ecosystem of Punjab (Pakistan). Am Eurasian J Agric Environ Sci 7:128–142 Khan M, Lanjar AG, Chang BH, Bukero A, Rajput A, Magsi FH, Shah R, Solangi AW, Chang AH (2019) Insect pests associated with ornamental plants. Biol Sci 62(3):188–194 Khokhar S, Rolania K (2021) Population dynamics of aphid and coexisting predators in tomato agroecosystem. J Agrometeorol 23(2):200–206 Khursheed S, Bhat ZA, Rather GH, Itoo H, Malik AR, Pandit BA (2021) Occurrence of insect and mite pests and their natural enemies under high density apple agro-ecosystems in Kashmir. J Entomol Zoo Stud 9(1):993–998 Korányi D, Szigeti V, Mezőfi L, Kondorosy E, Markó V (2021) Urbanization alters the abundance and composition of predator communities and leads to aphid outbreaks on urban trees. Urban Ecosyst 24(3):571–586 Krey KL, Smith OM, Chapman EG, Crossley MS, Crowder DW, Fu Z, Snyder WE (2021) Prey and predator biodiversity mediate aphid consumption by generalists. Biol Control 160:104650 Labe MS, Fernado NC, Fiegalan ER (2016) Insects associated with organic and inorganic rice farming. Int J Agric Technol 12(7.1):1345–1366 Lahbib N, Picciotti U, Sefa V, Boukhris-Bouhachem S, Porcelli F, Garganese F (2022) Zelus renardii roaming in Southern Italy. Insects 13(2):158

66

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Lanna LM, Rocha JFH, Cavalcante S, Godoy D, Teixeira MLF (2021) First record of non-carnivore feeding behavior in a wild praying mantis (Mantodea: Mantidae). Entomol Commun 3:ec03003 Lewis WJ, Haney PB, Reed R, Walker A (1997) A total systems approach for sustainable cotton production in Georgia and the Southeast: first-year results. In: Drugger P, Richter D (eds) Proceedings of the beltwide cotton conferences, cotton insect research and control conference, New Orleans, 6–10Jan 1997. National Cotton Council, Memphis, pp 1129–1134 Liu F, Luo J, Zhu X, Zhao C, Niu L, Cui J (2021a) Transgenic Cry1Ac/CpTI cotton assessment finds no detrimental effects on the insect predator Chrysoperla sinica. Ecotoxicol Environ Saf 208:111680 Liu Z, Zifang Q, Pingping L, Yue Y, Gary WF, Wangpeng S (2021b) Influence of plant physical and anatomical characteristics on the ovipositional preference of Orius sauteri (Hemiptera: Anthocoridae). Insects 12(4):326 Lusana MS, Tarimo AJP, Mulungu LS (2019) Abundance and composition of cotton insect species in the western cotton growing area (WCGA) in Tanzania. J Entomol Zool Stud 7(6):1087–1092 Magura T, Lövei GL (2021) Consequences of urban living: urbanization and ground beetles. Curr Landsc Ecol Rep 6(1):9–21 Mallesh M, Sravanthy C (2021) Diversity of predatory arthropods in Bt and non-Bt cotton fields of Nalgonda district, Telangana state, India. J Entomol Zool Stud 9(2):1199–1203 Marques OM, Gil-Santana HR, Coutinho ML, da Silva Júnior DD (2006) Percevejos predadores (Hemiptera, Reduviidae, Harpactorinae) em fumo (Nicotiana tabacum L) no município de Cruz das Almas. Bahia Revista Brasileira de Zoociências 8(1):55–59 Marrero Artabe L, Martínez MD (2003) Occurrence of Heteroptera in Cuban soybean agroecosystems (Glycine max (L) Merril). Revista de Protección Vegetal 18(2):98–103 Martins ICF, Cividanes FJ, Ide S, Haddad GQ (2012) Diversity and habitat preferences of Carabidae and Staphylinidae (Coleoptera) in two agroecosystems. Bragantia 71:471–480 Martin WRJR, Brown JM (1984) The action of acephate in Pseudoplusia includes (Lepidoptera, Noctuidae) and Pristhesancus papuensis (Hemiptera, Reduviidae). Entomol Experimentalis et Applicata 35:3–9 Martins CC, Santos RS, Sutil WP, Oliveira JFAD (2019) Diversity and abundance of green lacewings (Neuroptera: Chrysopidae) in a Conilon coffee plantation in Acre, Brazil. Acta Amazon 49:173–178 McCutcheon GS (1999) Effect of surface tillage on population dynamics of beneficial arthropods in cotton. In: Dugger P, Richter D (eds) Proceedings of the beltwide cotton conferences, Orlando, 3-7 Jan 1999. National Cotton Council, Memphis, pp 1117–1120 Meseguer R, Levi-Mourao A, Pons X (2021) Species complex and temporal associations between Coccinellids and Aphids in Alfalfa Stands in Spain. Insects 12(11):971 Middleton EG, MacRae IV, Philips CR (2021) Floral plantings in large-scale commercial agroecosystems support both pollinators and arthropod predators. Insects 12(2):91 Milanez JM (1984) Ocorrência de artrOpondosem urn sistema de consOrciofeijdo-milho, comparadoaosrespectivosmonocultivos. Resumos IX Congr. Bras. Entomol. Londrina. Brasil. Londrina: Soc. Entomol. Bras, p 38 Miles M, Bull T (2000) Update from the darling downs area-wide management project. Heliothis Stateline 3:8–10 Miñarro M, Dapena E (2003) Effects of groundcover management on ground beetles (Coleoptera: Carabidae) in an apple orchard. Appl Soil Ecol 23(2):111–117 Miñarro M, Hemptinne JL, Dapena E (2005) Colonization of apple orchards by predators of Dysaphis plantaginea: sequential arrival, response to prey abundance and consequences for biological control. Biocontrol 50(3):403–414 Miranda Ortiz F (2011) Biological control of diamondback moth. Acta Universitatis Agriculturae Sueciae. Doctoral Thesis,. Acta University of Agriculture Sciences, p 72 Misra RM (1975) Notes on Anthia sexguttata Fabr (Carabidae, Coleoptera) a new predator of Pyrausta machaeralis Walker and Hyblea puera Crammer. Indian Forester 101:604

References

67

Mizell RF, Tedders WL (1995) A new monitoring method for detection of the stinkbug complex in pecan orchards. Proc Southeast Pecan Growers Assoc 88:36–40 Murray DAH (1982) Heliothis egg parasitism study on sorghum 1982. Unpublished QDPI report. QDPI, Toowoomba Nagarajan K, Varadarasan S (2013) Reduviid fauna associated with cardamom forest ecosystem. Insect Environ 19(1):50–53 Naikwadi B, Javalage S (2019) Study of morphological characters of important predatory fauna found in kharif Agro-ecosystem from the region, Akola, Maharashtra, India. J Entomol Zool Stud 7(5):976–983 Naranjo SE (2005) Long-term assessment of the effects of transgenic Bt cotton on the abundance of nontarget arthropod natural enemies. Environ Entomol 34:1193–1210 Niba AS (2011) Arthropod assemblage dynamics on cowpea (Vigna unguiculata L Walp) in a subtropical agro-ecosystem. South Africa. Afr J Agric Res 6(4):1009–1015 Nyasani JO, Meyhöfer R, Subramanian S, Poehling H-M (2013) Seasonal abundance of western flower thrips and its natural enemies in different French bean agroecosystems in Kenya. J Pest Sci 86(3):515–523 Nyirra ZM (1970) The biology and behaviour of Rhynocoris albopunctatus (Hemiptera, Reduviidae). Ann Entomol Soc Am 63:1224–1227 O’Callaghan M, Glare TR, Burgess EP, Malone LA (2005) Effects of plants genetically modified for insect resistance on nontarget organisms. Annu Rev Entomol 50:271–292 Ogburn EC, Heintz-Botz AS, Talamas EJ, Walgenbach JF (2021) Biological control of Halymorpha halys (Stål)(Hemiptera: Pentatomidae) in apple orchards versus corn fields and their adjacent woody habitats: high versus low pesticide-input agroecosystems. Biol Control 152:104457 Ooi AC (2015) Common insect pests of rice and their natural biological control. Utar Agric Sci J 1(1):49–59 Ostman I, Ives AR (2003) Scale-dependent indirect interactions between two prey species through a shared predator. Oikos 102:505–514 Parker AH (1965) The predatory behaviour and life history of Pisilus tipuliformis Fabricius (Hemiptera: Reduviidae). Entomol Expe et Appl 8(1):1–12 Parker AH (1972) The predatory and sexual behaviour of Phonoctonus fasciatus (P. de B.) and P. subimpictus Stål (Hem., Reduviidae). Bull Entomol Res 62(1):139–150 Patel HN, Abhishek S, Prajapati JN (2018) Diversity of praying Mantids (Insecta: Mantodea) from various ecosystems of south Gujarat, India. J Entomol Zool Stud 6(4):167–169 Pathak M, Rizvi PQ, Kumar P, Garg DK, Singh NK (2011) Natural enemies of rice insect pests in rice-wheat cropping system of northern Indo-Gangetic plains. Oryza 48(4):383–386 Patlar G, Oğuzoğlu Ş, Avci M, Şenol Ö (2021) Aphid (Hemiptera: Aphididae) species in Burdur urban parks with three records for the fauna of Turkey, their host plants and predators. Turk J Entomol 45(3):381–397 Pavuk DM (2007) Predatory arthropod communities in Ohio soybean agroecosystems infested with the soybean aphid, Aphis glycines (Hemiptera: Aphididae). In: The 2007 ESA annual meeting, 9–12 Dec 2007 Pawar CS, Bhatnagar VS, Jadhav DR (1986) Heliothis species and their natural enemies, with their potential in biological control. Proc Indian Acad Sci (Anim Sci) 95:695–703 Pehlivan S, Atakan E (2020) Distribution and seasonal abundance of predatory bugs, Orius spp. (Hemiptera: Anthocoridae) in Adana Province, Turkey. Turkish. J Entomol 44(1):57–69 Penner FV, da Silva YK, Soares MM, Bastos LF, Batista TF (2021) First records of Condylostylusdepressus (Aldrich, 1901) and Condylostyluselectus (Walker, 1852) (Diptera, Dolichopodidae) in coconut plantations in state of Pará, Brazil. Entomol Commun 3:ec03014 Pérez-Fuertes O, García-Tejero S, Pérez Hidalgo N, Mateo-Tomás P, Olea PP (2015) Irrigation effects on arthropod communities in Mediterranean cereal agro-ecosystems. Ann Appl Biol 167(2):236–249

68

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Philpott SM, Perfecto I, Vandermeer J (2014) Behavioral diversity of predatory arboreal ants in coffee agroecosystems. Environ Entomol 37(1):181–191 Phyu W, Phyu MH, Oo T (2020) Diversity of arthropod natural enemy species in intensive monsoon rice growing area, Nay Pyi Taw, Myanmar. Int J Environ Rural Dev 11(2):71–77 Pilcher CD, Obrycki JJ, Rice ME, Lewis LC (1997) Preimaginal development, survival, and field abundance of insect predators on GM Bacillus thuringiensis corn. Environ Entomol 26:446–454 Pingale SV (1954) Biological control of some stored grain pests by the use of a bug predator, Amphibolus venator Klug. Indian J Entomol 16:300–302 Pons X, Núñez E, Lumbierres B, Albajes R (2005) Epigeal aphidophagous predators and the role of alfalfa as a reservoir of aphid predators for arable crops. Eur J Entomol 102(3):519 de la Poza M, Pons X, Farinós GP, López C, Ortego F, Eizaguirre M, Castañera P, Albajes R (2004) Impact of farm-scale Bt maize on abundance of predatory arthropods in Spain. Crop Prot 24: 677–684 Qian P, Bai Y, Zhou W, Yu H, Zhu Z, Wang G, Zhu ZR (2021) Diversified bund vegetation coupled with flowering plants enhances predator population and early-season pest control. Environ Entomol 20(20):1–10 Raen AZ, Dang C, Fang W, Peng YF, Ye GY (2016) Thrips-mediated impacts from transgenic rice expressing Cry1Ab on ecological fitness of non-target predator Orius tantilus (Hemiptera: Anthocoridae). J Integr Agric 15(9):2059–2069 Rajagopal D (1984) Observation on the natural enemies of Odontotermes wallonensis (Wassmann) (Isoptera, Termitidae) in South India. J Soil Biol Ecol 4:102–107 Rakhshani H, Ebadi R, Hatami B, Rakhshani E, Gharali B (2010) A survey of alfalfa aphids and their natural enemies in Isfahan, Iran, and the effect of alfalfa strip-harvesting on their populations. J Entomol Soc Iran 30(1):13–28 Ramos-Robles M, Sánchez-Vega M, Aguirre-Uribe LA, Méndez-López A, Martínez-Amador SY, Leal-Robles AI, González-Méndez LM (2021) Diversity and species turnover of insects in a Jatropha curcas agroecosystem. Southwestern Entomol 46(1):115–128 Ratnadass A, Zakari-Moussa O, Kadi-Kadi HA, Kumar S, Grechi I, Ryckewaert P, Salha H (2014) Potential of pigeon pea as a trap crop for control of fruit worm infestation and damage to okra. Agric For Entomol 16(4):426–433 Rocha FH, Infante F, Castillo A, Ibarra-Nunez G, Goldarazena A, Funderburk JE (2015) Natural enemies of the Frankliniella complex species (Thysanoptera: Thripidae) in Ataulfo mango agroecosystems. J Insect Sci 15(1):114 Rohmare VB, Rathod DM, Parasharya BM (2017) Diversity and population dynamics of Odonata (Insecta: Odonata) in rice growing area of Central Gujarat. J Biol Control 30(3):149–157 Rondoni G, Isabel B, Jana C, Eric C, Alejandro CC, François D, Edward WE (2021) Exotic ladybirds for biological control of herbivorous insects–a review. Entomol Exp Appl 169(1): 6–27 Roth S (2003) Spatial and temporal pattern of colonization of Nabidae (Heteroptera) in alfalfa (Medicago sativa). J Appl Entomol 127(4):221–227 Roubinet E, Jonsson T, Malsher G, Staudacher K, Traugott M, Ekbom B, Jonsson M (2018) High redundancy as well as complementary prey choice characterize generalist predator food webs in agroecosystems. Sci Rep 8(1):8054 Sadegh Nejad A, Rajabpour A, Cultler C (2019) Effect of climate and ecosystem manipulation on biodiversity of Anthocoridae in Hormozgan province (Iran). Plant Prot 42(2):37–49 Sadeghi H, Gilbert F (2000) Aphid suitability and its relationship to oviposition preference in predatory hoverflies. J Anim Ecol 69:771–784 Sahayaraj K (1991) Bioecology, ecophysiology and ethology of chosen predatory hemipterans and their potential in biological control (Insecta, Heteroptera, Reduviidae). PhD thesis, Madurai Kamaraj University, Madurai Sahayaraj K (1999) Field evaluation of the predator, Rhynocoris marginatus (Fab) on two groundnut defoliators. Int Arachis Newsletter 19:41–42

References

69

Sahayaraj K, Fernandez SM (2021) The predation behavior and the prey size preferences of Antilochus coquebertii (Pyrrhocoridae) against Dysdercus koenigii (Pyrrhocoridae). Int J Trop Insect Sci 41(2):1763–1769 Sakagami K, Funamoto D, Sugiura S (2021) The scientific naturalist. Ecology 102(11):e03482 Sakenin H, Hassan G, Majid N (2021) A study on the predators and parasitoids in some cotton fields of Iran. J Anim Environ 13(1):397–406 Santos ADF, Barbosa NCCP, Thomazini TC, Morales AC (2020) Genetic variability of Ceraeochrysa cincta, Ceraeochrysa claveri, and Ceraeochrysa cubana (Neuroptera: Chrysopidae) populations in agroecosystems of Southeast Brazil. Environ Entomol 49 (3):765–775 Sarma D, Saikia DK, Devee A, Borkakati RN (2021) Diversity of insect pests and predators of cabbage ecosystem in different phonological stages of cabbage. Int J Curr Microbiol App Sci 10(07):427–433 Sarmiento-Cordero MA, Rodríguez-Vélez B, Huerta-Martínez FM, Uribe-Mú CA, ContrerasRamos A (2021) Community structure of Neuroptera (Insecta) in a Mexican lime orchard in Colima, Mexico. Rev Mex Biodivers 92:923399 Saroj PL, Bhat PS, Srikumar KK (2016) Tea mosquito bug (Helopeltis spp)–A devastating pest of cashew plantations in India, A review. Indian J Agric Sci 86(2):151–162 Satar S, Raspi A, Özdemir I, Tusun A, Karacaoğlu M, Benelli G (2015) Seasonal habits of predation and prey range in aphidophagous silver flies (Diptera Chamaemyiidae), an overlooked family of biological control agents. Bull Insectol 68(2):173–180 Sathe TV, Vaishali PJ (2014) Report on nine new species of mantids (Insecta: Mantodea) and their insect pest predatory potential from agroecosystems of Kolhapur region. J Entomol Zool Stud 2(5):304–307 Satpathy JM, Patnaik NC, Samalo AP (1975) Observations on the biology and habits of Sycanus affinis Reuter (Hemiptera, Reduviidae) and its status as a predator. J Bombay Nat Hist Soc 72:589–595 Serdar T, Petr K (2009) Dermaptera fauna of the ecologically managed cherry orchards in western turkey. Munis Entomol Zool 4(2):572 Shelton MD, Edwards CR (1983) Effects of weeds on the diversity and abundance of insects in soybeans. Environ Entomol 12:296–298 Shera PS, Karmakar P, Sudhendu S, Sangha KS (2018) Impact of Bt cotton expressing single (Cry1Ac) and dual toxins (Cry1Ac and Cry2Ab) on the fitness of the predator Chrysoperla zastrowisillemi (Esben-Petersen): prey-mediated tri-trophic analysis. Egypt J Biol Pest Control 28(1):1–7 Shrestha G, Rijal JP, Reddy GV (2021) Characterization of the spatial distribution of alfalfa weevil, Hypera postica, and its natural enemies, using geospatial models. Pest Manag Sci 77(2): 906–918 Singh OP (1985) New record of Rhynocoris fuscipes Fabr. as a predator of Dicladispa armigera (Oliver). Agric Sci Dig 5:179–180 Singh OP, Singh KJ (1987) Record of Rhynocoris fuscipes Fabricius as a predator of Green stink bug, Nezara viridula Linn, infesting soybean in India. J Biol Control 1:143–146 Singh J, Arora R, Singh AS (1987) First record of predators of cotton pests in the Punjab. J Bombay Nat Hist Soc 84:456 Sivakoff FS, Rosenheim JA, Hagler JR (2012) Relative dispersal ability of a key agricultural pest and its predators in an annual agroecosystem. Biol Control 63(3):296–303 Smith HA, Chaney WE (2007) A survey of syrphid predators of Nasonovia ribisnigri in organic lettuce on the central coast of California. J Econ Entomol 100(1):39–48 Sommaggio D (1999) Syrphidae, can they be used as environmental bioindicators? Agric Ecosyst Environ 74:343–356 Sommaggio D, Burgio G (2003) Role of Diptera Syrphidae as landscape indicators: analysis of some case studies in Northern Italy. Landscape management for functional biodiversity. IOBC WPRS Bull 26:145–150

70

2

Distribution and Diversity of Predatory Insects in Agroecosystems

Song Y, Yang X, Zhang H, Zhang D, He W, Wyckhuys KA, Wu K (2021) Interference competition and predation between invasive and native herbivores in maize. J Pest Sci 27:1–11 Souza CS, Marucci RC, Chaves DR, Mendes SM (2019) Effects of genetically modified plants with Bt toxins on natural enemies. In: Natural enemies of insect pests in neotropical agroecosystems. Springer, Cham, pp 489–496 Speight MCD, Castella E, Obrdlik P, Ball S (eds) (1997–2003) Syrph the Net, the Database of European Syrphidae. Syrph the Net Publications, Dublin, pp 1–38 Tawfik MFS, Awadallah T (1983) Effect of temperature and relative humidity on various stages of the predated, Allaceocramm biannulipes (Montr.et. Sign) (Hemiptera : reduviidae). Bull Soc Entomol 64:239–250 Tawfik MFS, Awadallah KT, Abdellah MMH (1983a) The biology of the reduviid Allaeocranum biannulipes (Montrouzier and Signoret), a predator of stored product insects. Bull Soc Entomol 64:231–237 Tawfik MFS, Awadallah KT, Abdellah MMH (1983b) Effect of prey on various stages of the predator, Alloeocranum biannulipes (Montr. et Sig.) (Hemiptera: Reduvidae). Bull Soc Entomol Egypt 64:251–258 Taye RR, Bathari M, Borkataki S, Rahman A (2017) Diversity of Hymenopteran predators and parasitoids in Assam Agricultural University Campus, Jorhat. J Entomol Zool Stud 5(6): 2420–2423 Tenhumberg B, Hans-Michael P (1995) Syrphids as natural enemies of cereal aphids in Germany: aspects of their biology and efficacy in different years and regions. Agr Ecosyst Environ 52(1): 39–43 Tezcan S, Kocarek P (2009) Dermaptera fauna of the ecologically managed cherry orchards in western Turkey. Munis Entomol Zool 4(2):572–576 Thennarasi A, Jeyarani S, Sathiah N (2021) Diversity of predators associated with the Mealybug Complex in Cassava growing districts of Tamil Nadu, India. Int J Plant Soil Sci 33(22):62–79 Torres JB, Ruberson JR (2006) Interactions of Bt-cotton and the omnivorous big-eyed bug Geocorispunctipes (Say), a key predator in cotton fields. Biol Control 39(1):47–57 van den Bosch R, Hagen KS (1966) Predaceous and parasitic arthropods in California cotton fields. Calif Agric Exp Stn Bull 820:32 Varma RV (1989) New record of Panthous bimaculatus (Hemiptera, Reduviidae) as predator of pests of Alantha triphysa. Entomon 14:357–358 Vayssières J-F, Sinzogan A, Adandonon A, Van Mele P, Korie S (2013) Ovipositional behaviour of two mango fruit fly species (Diptera: Tephritidae) in relation to Oecophylla cues (Hymenoptera: Formicidae) as compared to natural conditions without ant cues. Int J Biol Chem Sci 7:447–456 Wang S, Wu J, Zhao Y, Li R, Zhao D (2020) Functional response of adult Hierodula patellifera (Serville, 1839) (Mantodea: Mantidae) to Tessaratomapapillosa (Drury) (Hemiptera: Tessaratomidae). Int J Trop Insect Sci 40(4):1053–1058 Waters D, Drysdale R, Kimber K (1999) Planting cotton into wheat stubble. Aust Grain 9:4–6 Wratten SD, Bowie MH, Hickman JM, Evans AM, Sedcole JR, Tylianakis JM (2003) Field boundaries as barriers to movement of hover flies (Diptera: Syrphidae) in cultivated land. Oecologia 134(4):605–611 Yu H, Li Y, Li X, Wu K (2014) Arthropod abundance and diversity in transgenic Bt soybean. Environ Entomol 43:1124–1134 Zhang DJ, Lu ZY, Liu JX, Li CL, Yang MS (2015) Diversity of arthropod community in transgenic poplar-cotton ecosystems. Genet Mol Res 14(4):15713–15729 Zhang L, Qin Z, Liu P, Yin Y, Felton GW, Shi W (2021) Influence of plant physical and anatomical characteristics on the ovipositional preference of Orius sauteri (Hemiptera, Anthocoridae). Insects 12(4):326

3

Identification of Various Insect Predators

Contents 3.1

Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.1 Coccinellidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Orthoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4 Odonata . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6 Thysanoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7 Heteroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.1 Lygaeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.2 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.3 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.4 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.5 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7.6 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.8 General Features of Vespidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.9 General Features of Formicidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.10 Predatory Lepidopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

72 72 74 76 78 79 81 83 84 85 86 88 90 91 92 93 93 93

In areas like systematics, conservation biology, and ecology, the identification of species is crucial. While it may be simple to recognise huge, charismatic creatures, most organisms require specialist knowledge for effective identification, and the inability to distinguish between species poses a significant obstacle known as the taxonomic impediment. The availability of guides for identifying species is the most fundamental prerequisite for anyone researching and working on biodiversityrelated issues. However, there are only a limited number of taxonomic groupings for which the general public and non-taxonomists can easily employ identification aids. Protura, Diplura, Thysanura, and Collembola are examples of ametabolous insects. Other hemimetabolous insects include Ephemeroptera, Odonata, Mantodea,

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_3

71

72

3 Identification of Various Insect Predators

Phasmida, Isoptera, Blattaria, Plecoptera, Dermaptera, Embioptera, Psocoptera, Mallophaga, Anoplura, and Hemiptera (Neoptera, Megaloptera, Mecoptera, Coleoptera, Trichoptera, Lepidoptera, Diptera, Siphonaptera, and Hymenoptera).

3.1

Coleoptera

3.1.1

Coccinellidae

The Coccinellidae family of beetles is a subfamily of the superfamily Cucujoidea, which itself is a suborder of the suborder Polyphaga of the beetles (Coleoptera). The Endomychidae (“handsome fungus beetles”) and Corylophidae are their close relatives within the Cucujoidea (“minute fungus beetles”). There are close to 6000 species of ladybirds in the world. These 105 individuals are divided into native and adventive groups. Some adventive species are immigrants, while others were brought in. Three subfamilies, Epilachninae, Coccinellinae, and Tetrabrachinae, make up the family Coccinellidae (Lithophilinae). Many agroecosystems are home to generalist predatory beetles from the families Carabidae (ground beetles) and Staphylinidae (rove beetles). All insects belonging to the Coleoptera’s Carabidae family of beetles are collectively referred to as “ground beetles”. With almost 40,000 species worldwide, ground beetles, often known as carabids, are one of the largest insect families. They have a wide size range, measuring between 0.7 and 66 mm (https://bugguide.net/node/view/186). The large legs and powerful mandibles of ground beetles make them ravenous predators and vital for the biological control of insect pests on farms (Snyder 2019). Although they mainly forage on the soil’s surface, adult beetles will occasionally venture into the vegetation in search of food. In addition to the fact that the adults of these beetles are helpful predators, the burrowing larvae of these insects hunt down and eat soil pests. Numerous kinds of ground beetles have varied feeding habits, consuming both plant seeds and other insects (including weeds). Size/Colour: Depending on the species, ladybird adults are oval, have wings, and can be anywhere between 1 and 10 mm long. On average, females are bigger than males. Some species’ adults have vivid colours. However, some people contend that some adult ladybirds’ vivid colours—such as red on black or black on red—are aposematic, signalling to potential predators that the beetles are poisonous or repulsive. Head: For chewing, they use their mandibles. Adult ladybirds have a tendency to bleed reflexively from their tibia-femoral joints (leg joints). Ground and tiger beetles belong to the insect family called carabids, and they are crucial biological control agents in agroecosystems. Carabid beetles are dangerous predators in the insect world thanks to their huge eyes, spiny muscular legs, and large jaws. Egg, larva, pupa, and adult are the four separate life stages of ground beetles. The larvae have enormous, pincer-like mandibles, and no wings, allowing them to eat other

3.1 Coleoptera

73

Fig. 3.1 Coleopteran general features dorsal (a) and ventral view (b)

soil-dwelling creatures. Use this guide to distinguish larval carabid beetles from other common insect larvae types. While the adult life stage is mostly spent aboveground, the egg, larval, and pupal phases are mostly spent underground. Although adult carabid beetles have wings, they typically do not fly, and many ground beetles lack the ability to fly altogether. Figure 3.1 shows a diagrammatic illustration of a carabid beetle. Mouthparts feature two bristly tail appendages in addition to sharp, protruding mouthparts. They can remove the snail from its shell thanks to their long, hook-shaped mouthparts.

Thorax: Long legs are a common feature of carabid beetles, which enable them to move quickly to capture prey and evade other predators. The characteristic look of a huge, jellybean-shaped trochanter, or upper leg segment, at the base of a ground beetle’s hind legs, can be used to identify it. When viewed from the ventral surface, this lobed trochanter conceals the whole first abdominal segment. Outside of this group, no other beetles possess this trait. The “false ground beetles” are the only exception to this rule. They have a body form like ground beetles and huge trochanters, but their coxa, which is found next to the trochanter at the base of their hind legs, is much larger. The elytra (wing covers) of ground beetles are lustrous black or brown, ornamented with ridges, and may be fused together along the midline. Many species lack or have shortened hind wings. When agitated, ground beetles typically run rather than fly because they prefer moist, chilly environments. The superfamily Vespoidea, which includes the hornets and yellow jackets, is called Vespidae (yellow jackets). Their wings fold longitudinally when in rest, setting them apart from other wasp groups. They claimed that while they were all globally diverse, they were mostly tropical. Additionally, they mentioned that

74

3 Identification of Various Insect Predators

although adults are typically black or brown, they are frequently spotted with yellow or white. The majority of creatures live alone, although many are social. In solitary species, the larva lives in a cell that the adult female builds and supplies, where it typically preys on other insects, particularly caterpillars. In some cases, pollen and nectar are substituted for the larva while providing sustenance. Adult females of gregarious species consume masticated insects or, less frequently, glandular secretions. In the nests of social insects, a few of them are kleptoparasites. There is pupation inside the cell. These striking-looking vespidae measure around 16 mm (0.63 in.) and have orange antennae, wings, and tarsi. The body could be completely black or brown with segmented patches and thin yellow stripes. Compared to the more well-known species of wasps and hornets, the sting of this insect is painful but less poisonous to people (Vespa, Vespula). The females chewed wood from dead trees, fence posts, or unpainted building materials and combined it with saliva to create a paste, which is then used to construct the nest. Long, slender antennae that are bent but not curled are significant physical characteristics (as in Pompilidae). The pronotum extends laterally to the tegulae; the wings’ discoidal (M-4) cells are typically very lengthy. They are dark, medium in size (9–25 mm), and marked in red, white, or yellow. Insects typically fold their wings longitudinally while they are at rest. Coleopteran predators are classified into six subfamilies: Eumeninae, Euparagiinae, Masarinae, Polistinae, Stenogastrinae, and Vespinae. Prominent Identification • Forewings fold in half longitudinally. • Posterior margin of pronotum distinctly “U”-shaped. • Forewings fold in half longitudinally. • Two pairs of membranous wings, although some may be wingless such as some species of female wasps and the worker caste of ants. • The tiny hooks that hold the forewings together are larger than the ones on the hind wings. • The ovipositor, which is typically hardened in females and may be modified for sawing, piercing, or stinging. • The term “wasp waist” refers to the constriction between the first two segments of the abdomen found in the majority of hymenopterans. • Chewing (mandibulate) mouthparts, though some species, like bees, have transformed their bottom lip into a tongue. • Compounded eyes, often big.

3.2

Dermaptera

The word “dermaptera”, which alludes to the skin-like appearance of the forewings, is derived from the Greek words “derma” and “pteron”. The common name comes from the false myth that earwigs enter into people’s ears while they are asleep. The

3.2 Dermaptera

75

hemimetabolous order of insects known as Dermaptera is small and relatively old. Its external morphology is distinguished by paired cerci (forceps) at the posterior end and short tegmina that cover the hind wings only partially. Up to 50 mm long, dermaptera are elongate insects. With nine families and 181 genera, Forficulina contains the majority of species. Forficulina species are free-living, possess usable wings, and are not parasitic. They change into pincers and have unsegmented cerci. Two suborders, Arixeniina and Hemimerina, have been added as more cases have been discovered. A Malaysian hairless bulldog bat’s body contained an epizoic species of earwig. With a total of five species, Arixeniina represents the two genera Arixenia and Xeniaria. They share Hemimerina’s traits of being blind, without wings, and with filiform segmented cerci. African rodents’ fur is home to viviparous ectoparasites called hemimerina. With a combined 11 species, Hemimerina also has two genera: Hemimerus and Araeomerus. In order to protect and care for their eggs and newly hatched nymphs, female Dermaptera keep them close together until at least their first moult. Size: Their body is elongated, measuring about 6–55 mm. The Saint Helena earwig, which can grow to be 80 mm long, is an exception. Head: The mouthparts are of the chewing type, and the head is prognathous. With mandibulate, forward-projecting mouthparts, compound eyes that range in size from big to missing, no ocelli, and short annulate antennae, adult earwigs are elongate and dorsoventrally flattened. Ten or more segments may be present on the antennae. Thorax: The mesothorax is still separate from the prothorax. Cursorial legs have a tarsus made up of three segments. The tarsi are three-segmented, and the second tarsomere is brief. The term “dermaptera” refers to front wings that are thicker or skin-like. Typically, adults have two sets of wings. The forewings are elytriform or transformed into tegmina, short, leathery, and leatherous. When the insect is at rest, its big, membranous, fan-shaped, or circular hind wings are folded beneath its forewings. The hind wings of this particular group of insects are semi-circular, membranous, and have veins that are organised radially. When the insects are at rest, they fan out beneath the front pair. While the majority of earwig species have short front wings (tegmina) that do not cover the abdomen, certain earwig species are wingless as adults. When at rest, the hind wings are folded fan-like and then lengthwise, projecting slightly from beneath the tegmina. Many species are apterous or, if winged, the fore wings are broad, membranous, semi-circular, and dominated by an anal fan of radiating vein branches connected by cross-veins. Abdomen: The cerci are transformed into forceps that can squeeze if the earwig is touched, and the abdomen is frequently telescopic. Because of this partial metamorphosis, nymphs resemble adults. The cerci, a pair of pincers on the back of their abdomen, help identify these insects. Pincers on men are curled, while those on females are straight. Both protection and prey capture are accomplished with these pincers.

76

3.3

3 Identification of Various Insect Predators

Orthoptera

The order of insects known as Orthoptera contains katydids, crickets, grasshoppers, locusts, and other members of their family. The majority of orthopteran species are phytophagous, meaning they eat fungus, upper plant foliage, and roots. While some animals are omnivorous, many are predatory. Orthoptera have been described in over 25,000 different species, although only around 80 of them are aquatic and another 110 or so are water-dependent. However, numerous species of mole crickets (Gryllotalpidae), pygmy grasshoppers (Tetrigidae), katydids (Tettigoniidae), pygmy mole grasshoppers (Tridactylidae), and grasshoppers (Acrididae) are connected to damp habitats. Size: Few species have bodies that are less than 10 mm long, while the majority is longer than 50 mm, with some species having bodies longer than 100 mm and wingspanes of 200 mm or more. The bodies of orthopterans are typically cylindrical, with enlarged hind legs and muscles designed for jumping. Head: The antenna are filiform and typically long and threadlike, with less than ten to several hundred segments. The head is hypognathous, hardly ever prognathous. The pronotum, or region of the body immediately behind the head, is frequently broad, shield-like, and, in certain circumstances (as with many katydids), completely covers the insect’s body (pygmy grasshoppers). Orthopterans have mandibulate, or chewing/biting, mouthparts. Thorax: The prothorax is big. The pronotum is bent and covers the pleural area ventrally. The front and middle legs are cursorial, or suited for walking; but, in some species, the front pair of legs may be changed for digging (e.g. mole crickets, pygmy mole crickets, false mole crickets), or both the front and middle pairs may be modified for grabbing (predatory katydids). Short and segmentless cerci. Tibial auditory organs are located in the front legs of some orthopterans (most katydids and crickets) (the ear). Most orthopterans have saltatorial hind legs, which are designed for leaping and have robust, muscular femora and long, slender tibiae. About 2.6 m leaps can be made repeatedly by some grasshoppers without showing any telltale indications of exhaustion. This is primarily made feasible by the presence of the protein resilin in their back legs. Resilin has exceptional elastic characteristics, restoring stored energy 97% of the time. This enables the insect to launch itself into the air with an explosive release of energy that is not achievable with just muscle force. Some orthopteran species, particularly those that live underground, lack the capacity to jump, and their hind legs resemble those of normal cursorial species. The inner surface of the hind femur is adapted for sound production in some grasshoppers and some Ensiform (stridulation). Even though the basic shape of the orthopteran thorax shows their kinship to winged insects, the grylloblattids lack wings, and all large orthopteran groups contain a few wingless species. Orthopterans have wings that are either fully

3.3 Orthoptera

77

grown or to varying degrees underdeveloped. Within a species, wing polymorphism, or the existence of individuals with both fully grown and shortened wings, is not unusual. The forewings have become a little thicker and have developed leathery tegmina. The majority of crickets and katydids have modified tegmina for stridulating. When present, the hind wings are fan-shaped and tucked under the first pair when the bird is at rest. The hindwings frequently extend past the tegmina and can be seen behind them. While adult members of micro- and brachypterous species always have the first pair of wings overlapping the second pair, despite their nymphal look, the wing buds of nymphal stages are always positioned so that the second pair of wings overlaps the first one. Abdomen: In grasshoppers, the abdominal tympana, or lateral auditory organs, are located at the base of the abdomen. The first abdominal segment of the Acridoidea (locusts, grasshoppers) has two ears on either side. When seen with the naked eye, the tympanal membranes are approximately spherical, opaque, and easily discernible. The locust can discern the direction of a sound source thanks to the tracheal sacs that connect its two ears. The ears are found on the tibia of the forelegs of the crickets (Gryllidae) and katydids (Tettigoniidae) in the suborder Ensifera. There are two eardrums on each leg, one on each side. The ovipositor, which is formed by the eighth and ninth abdominal segments, is prominent in the females of most orthopterans. Females of grasshoppers and their cousins typically lack a long, external ovipositor, but katydids and crickets typically have a well-developed ovipositor that is shaped like a sword, sickle, or needle. Salient Characters • Two sets of wings, the base of the forewings is leathery or stiffened, and they are narrower than the hind wings. At repose, they are held overlapping the abdomen like a roof. When at rest, the membranous hind wing is folded fan-style under the forewings. • A mandibulate set of teeth. • Substantial compound eyes. • Depending on the species, antennae can range in length from little to very long. Grasshoppers typically have long antennae compared to crickets and katydids, which typically have small antennae. • For jumping, the hind legs are lengthened and changed. • This order is challenging to mistake for other insects due to its distinctive appearance. Orthoptera juveniles resemble tiny, wingless adults. Many orthopterans have sound-producing organs in their legs, wings, or abdomen. These sounds are primarily made by male to female.

78

3.4

3 Identification of Various Insect Predators

Odonata

Odonates consume insects and are hence carnivorous, or more specifically, insectivorous. The population of many other insect species is thought to be controlled by dragonflies, one of the top predators in the insect kingdom. They primarily consume tiny insects like caddisflies, mayflies, gnats, termites, flies (Diptera), and ants. Large, rounded heads with mostly well-developed, faceted eyes, legs that make it easier to grasp prey in flight, two pairs of long, transparent wings that can move independently, and elongated, ten-segmented abdomens are all characteristics of Odonata members. Size: Medium to large in size, dragonflies and damselflies have body lengths between 15 and 120 mm. Head: It has three ocelli and larger compound eyes. The majority of the animal’s head is covered in dragonfly eyes that are in close proximity to one another. Usually, damselflies have a space between their eyes. While damselfly eggs are cylindrical and lengthier, measuring around 1 mm, dragonfly eggs are spherical and roughly 0.5 mm long. The almost 28,000 separate units (ommatidia) that make up each compound eye collectively cover the majority of the head. Odonata have exposed mouthparts and three pairs of jointed appendages. Their mandibulate-type jaws have been modified for biting, making them effective hunters. The prehensile labium that all Odonata possess may be extended forward from beneath the head faster than the majority of prey can respond, making their bite lethal to animals. Antenna is really tiny. Thorax: The three pairs of legs are all close to the head and are more useful for grabbing prey and perching on vegetation to rest or lay eggs than for walking. The broad, numerously veined, and slender wings of the majority of families have a pterostigma, an opaque structure on the leading edge close to the tip of the wing. This is a cell, a thickened, blood-filled, and frequently colourful region. In this context, the term “cell” refers to a closed, by veins bound portion of an insect wing. The pterostigma’s functions are not entirely understood, but it most likely serves both a visual and an aerodynamic purpose. The energy required to raise and lower the wings may be lessened by adding mass at the end of the wing. Thus, the energy required for flight could be decreased with the appropriate wing mass and stiffness combination. Other insects, including bees, also have pterostigmas. A distinctive nodus, or notch, may also be seen in both groups on the front edge of each wing. The dragonfly Petalurain gesntissima has a 162-mm wing span, whereas the damselfly Agriocnemis femina has a 20-mm wing span. The enormous damselfly Megaloprepus coerulatus and the Hawaiian endemic dragonfly Anax strenuus both have wingspans of up to 190 cm. Abdomen: Significantly longer than wings is the abdomen. In contrast to the real genital opening, which is situated close to the tip of the abdomen, males have

3.5 Neuroptera

79

distinctive secondary genitalia on the underside of the second and third abdominal segments (Grzimek et al. 2004). On the second and third abdominal segments of males, there appears to be a pouch containing the secondary genitalia. On the final abdominal segments are the real male genitalia as well as a grabbing mechanism for holding the female during mating. At the tip of the abdomen, the male generates sperm and transports it to the secondary genitalia, where the female can access it. Females have a single genital entrance and a little ovipositor at the end of the abdomen that will be utilised to oviposite her eggs instead of supplementary genitalia or gripping features. Typically, the female dragonfly is dull brown or grey and the male is more colourful. Suborders: True dragonflies belong to the infra order Anisoptera, while damselflies and dragonflies belong to suborders Zygoptera and Epiprocta, respectively. Dragonflies and damselflies are different from one another while sharing many characteristics. When at rest, dragonflies hold their wings either out to the side or outward and downward. They are strong flyers with sturdy bodies (or even somewhat forward). The majority of damselfly species maintain their wings folded back over the abdomen when at rest because they tend to be less robust and even appear somewhat weak in flight. Salient Features: Large, multifaceted compound eyes, two pairs of sturdy, transparent wings, occasionally with coloured patches, and an elongated body are the distinguishing features of adult dragonflies. Many dragonflies have vivid structurally created iridescent or metallic colours that make them stand out in flight.

3.5

Neuroptera

The term Neuropteran, which alludes to the reticulate nature of the wing venation, is derived from the Greek words neuro, which means nerve, and pteron, which means wing. Common names are typically given to particular families or groups of Neuropteran based on their physical characteristics or dietary preferences. The Neuropterans are a tiny, diverse order of neopterous endopterygote insects that are primarily found in tropical climates. There are 17 families in the order Neuroptera, which are currently divided into the suborders Nevrorthiformia, Myrmeleontiformia, and Hemerobiiformia. Because of their uncommon life cycles and distant environments, the members of the Ithonidae family of moth lacewings are a rare breed among lacewings. In contrast to most other lacewings, which normally keep their wings at a steeper, more “roof-like”, angle over the body, adults are robustly built and hold their brownish or pale wings rather flat over the back. The group’s popular name refers to the adults’ apparent resemblance to particular moth species. Ithonids are commonly thought of as “primitive” Neuropterans that are connected to rapismatids and polystoechotids. One particular little family of Neuropterans is the Dilaridae, also known as the pleasant lacewings. They are one of the few families of Neuroptera whose females

80

3 Identification of Various Insect Predators

Fig. 3.2 Adults of Plega signata (Mntispidae) (a) and Chaetoleon pusillus (Myrmeleontidae) (b) (John D. Oswald)

bear ovipositors, have numerous long hairs covering their bodies, have rather broad wings with transverse dark bands, and are quite modest in size. These insects are frequently confused with little moths because of how much they resemble them, especially the males. Large adults of the Polystoechotidae family have hyaline wings with varying dark patterns and dark, very robust, bodies. A medium-sized family of Neuropterans known as the Mantispidae, or mantidflies, are easily identified by their swollen, raptorial forelegs, which give them the appearance of small preying mantids. In both tropical and temperate regions of the planet, the family is multicultural, but the tropics are significantly more diverse and abundant. The adult Plega signata (Mntispidae) and Chaetoleon pusillus (Myrmeleontidae) line drawings from Fig. 3.2 are shown below. Size: Lacewings range in size from 1.5 to 35 mm (0.059–1.377 in.) in length, and from 2 to more than 100 mm in the length of the anterior wing. Soft-bodied insects are known as neuropterans. Medium-sized snakeflies have anterior wing lengths of at least 10 mm and lacewing-like wings. Head: In Megaloptera and Neuropteran, the head is transverse or squarish, while in Raphidiodea, it is elongated. In addition to having special piercing-sucking larval jaws made up of the mandibles (directed downward) and maxillae, Neuropteran adults also have hypognathous (directed downward) mouthparts. Although the ocelli are typically absent, the compound eyes are well-developed. The antennae are lengthy, multisegmented, and can have an enlarged, club-like apex. Thorax: The entire thorax is loosely structured, and the prothorax is mobile. Typically, neuropterans have four enormous, subequally sized membranous wings with extensive, netlike venation. The four wings can be big or extended, but they are all roughly the same size. All species’ wings are translucent, with numerous veins and cross veins, and occasionally have hints of brown or green hue. The wing coupling mechanism is straightforward; the wings are kept roof-like over the back

3.6 Thysanoptera

81

of the insect when it is at rest. When compared to dobsonflies and alderflies, which are similar in general appearance but with wings that appear heavier than those of lacewings, the latter have more many veins and appear more delicate. Abdomen: Except in Chrysopidae, the abdomen is ten-segmented; cerci are lacking. Appendages (cerci) do not exist on the ten segments of the abdomen, while terminal claspers may be present. The last abdominal segment in both sexes may be shrunk, altered, or supported by two trichobothria on its dorsa-lateral sides (short, stiff sensory bristles). Abdominal terminalia in females might vary as well. A long, slender ovipositor is unique to female Raphidiodean species. Pupation often takes place inside of a silken cocoon, and the immature stages differ significantly from the adults. The pupa is exaggerated and decticous, and its mandibles are fully formed.

3.6

Thysanoptera

The word Thysanoptera, which means “fringe-wing” in Greek, refers to the lengthy peripheral fringe that may be seen in specimens with wings. We do not use the term “thrip”, which is both singular and plural for the insect. There are over 4000 identified species in the global order of thrips. Hemiptera are linked to them, but they have the following distinctive traits instead: • • • • • •

Short antennas with 6–10 segments. Anteriorly, the head narrows, creating a conical mouth hole. Cylindrical or spindle-shaped in form. Slender, rod-like front and back wings with a thick fringe of long hairs. Many animals have no wings at all. With ever-movable sticky bladders apically, tarsi are 1–2 segmented.

Size: Though most thrips are tiny and elongated, certain species can grow to a length of 12 mm. The majority of thrips measure 1.5–3 mm (0.06–0.12 in.), with the smallest measuring approximately 0.6 mm (0.02 in.) and the largest being approximately 15 mm (0.6 in.). Head: A thrips has a slightly cocked head, with the mouth cone facing backward. Large facets are typically seen in compound eyes; ocelli are only present in winged adults. The right mandible is lacking in thrips, which is another peculiar characteristic. The labrum and labium form a cone, the left mandible is specialised for rasping, and the maxillae are adapted for piercing. Typically, the nine-segmented antennae of most species are reduced by fusion and protrude forward in front of the eyes. Only the left mandible, which is the only developed mouthpart, and the maxillae, which have been transformed into piercing stylets, remain. The labium and the maxilla both have segmented sensory projections (palps).

82

3 Identification of Various Insect Predators

Thorax: The forelegs of the thrips are carried by the first part of the thorax (prothorax), which may have ridges on the inner surfaces (coxae) that, when rubbed with spurs from the lower surfaces (femurs), produce sounds that, however, cannot be heard by the human ear without amplification. The wings, as well as the mid and hind legs, are supported by the prothorax portion of the thoracic. Typically, the wings are plain and straplike. Never folded, they are placed over the abdomen when at rest. The term Thysanoptera, which means “fringed wings”, comes from the fact that the wings typically have lengthy posterior fringes that resemble hair and shorter fringes that appear on the leading edge. When present, wings are long, slender, and have reduced venation. They are carried over the abdomen in repose but are not folded. The pretarsal creates an eversible bladder that is used for adhesion, and the tarsi are either one or two segmented. Abdomen: The thrips has a ten-segmented abdomen; an ovipositor may or may not be present, but cerci are never present. Particularly in the Tubulifera, the abdomen is elongated and typically flattened dorsa-ventrally. With the exception of a basic 11th segment, there are ten different segments. Spiracles are present in the first and eighth abdominal segments. Several S-shaped setae on the back of the Tubulifera frequently hook into the wing fringes to hold the wings at rest. In the Terebrantia, the terminal segment is split ventrally up to the eighth segment in the females but is rounded in the males. The terminal section of the Tubulifera is tube-like. The paired appendages and the unpaired copulatory organ (aedeagus) that make up the male genital organs all retreat into the belly. The posterior portion of the abdomen contains the pair of testes as well as the accessory glands and their ducts. The Terebrantia female typically possesses an ovipositor made up of two pairs of blades that resemble saws. The Tubulifera does not have any of these. There are eight egg sacs (ovarioles) and a seminal receptacle, which is spherical and frequently coloured and used to store sperm obtained during mating. Thrips have a sophisticated kind of metamorphosis that alternates between simple and full metamorphosis. As a result, nymphal instars 1–2 are active, feeding, and have internal or wingless development; instars 3 (prepupa) and 4 (pupa), on the other hand, are dormant, non-feeding, and have external wing buds. Small, elongated, and usually wingless insects are classified as thysanoptera. Major traits of the order include protruding tarsal bladders and asymmetrically piercing mouthparts. There are about 4000 species of thrips known, and they are either general feeders, phytophagous, polyphagous, mycetophagids, or predaceous. A modest number of animals that prey on spider mites are found in only three families. The key that follows is greatly simplified and can be used to classify them: Females lack an ovipositor, and both sexes’ abdomens are tubular at the top: suborder Aeolothripidae: ovipositor pointing downward; thin, acuminate wings; family Phlaeothripidae: saw-like ovipositor present in females; bluntly rounded apex of abdomen; ovipositor pointing upwards; tubulifera: ovipositor present; both families have broad wings.

3.7 Heteroptera

3.7

83

Heteroptera

A piercing and sucking type of mouth is present, and the head is opisthognathous. There are two sets of bristle-like stylets, which are actually modified mandibles and maxillae. Stylets rest in the rostrum’s or labium’s grooves. The maxillary and labial palps have both atrophied. Scutellum acts as the dorsal representation of the mesothorax. Forewings either have a consistent overall thickness or are basally coriaceous and distally membranous. Cerci are never present. Metamorphosis is frequently gradual but infrequently total. Alimentary canals have been appropriately modified to accommodate liquid food (filtering device). Salivary glands are always present, and extra-oral digesting seems to be common. Fusion of the thoracic and abdominal ganglia. Figure 3.3 shows the basic structural characteristics of a heteropteran inset.

Fig. 3.3 Typical structure of a Heteropteran insect

84

3 Identification of Various Insect Predators

Heteropteran predator species include Anthocoridae, Berytidae, Lygaeidae, Miridae, Nabidae, and Reduviidae species.

3.7.1

Lygaeidae

A number of big-eyed bug species are important economic polyphagous predators that eat prey from at least 3 classes, 10 orders, and 30 different families of arthropods. It has insects that are oblong-oval in shape, robust in build, and somewhat flattened, with a broad head that is wider than it is long. Diagnostic Characters: There are fragrance glands; the hemelytra’s membrane has only four to five veins. This is a sizable and significant family of primarily seedeating insects. The family has now been split up into at least ten different families (see textbook). While some species are predators, many are significant pests of crops, particularly grains. Size and Colour: These insects are tiny to medium in size, oval or elongated in shape, and range in length from 4 to 20 mm. They are typically brown, red, or black, however occasionally they are vividly coloured. Head: As their name suggests, they have distinctively huge, prominent eyes for their size. Their antennae are small, stout, and expanded at the tip. These insects have long proboscises that are used to drill and inject enzymes into their food. The first joint of the rostrum is a little longer than the second. They are distinguished by having prominent, big eyes that are widely spaced apart and curl backward on the sides of their skulls. They have a wide field of vision thanks to their huge eyes and good visual perception, which improves their capacity to locate and pounce on prey. Because of their stylet’s flexibility, they can stretch it in front of their heads when eating. The stylet permits Geocoris to consume food that frequently approaches or exceeds their body weight, which is a key factor in their efficiency as predators. Thorax: The femur, tibial tip, and tarsi of the brown, long, slender legs are marked with black. Hemelytra lack the cuneus. A transverse basal vein gives rise to 4–5 irregular veins in the membrane. Adults have two pairs of completely functional wings and are 3/16 in. or less long, measuring 3–5 mm (forewings and hindwings). The forewings are membranous (thin and malleable) at the tip and rigid at the base. At repose, the forewings form a triangle-shaped pattern behind the pronotum (also known as the shoulders), pointing backward. True bugs have these wing characteristics. Their forewings have a membranous back part and a somewhat sclerotised front section. They always have membranous hind wings. The wings are folded over the back and abdomen like a flattened shield while the bird is at rest. Abdomen: The female genital cleft ends roughly in the middle of the abdomen. The fourth and fifth abdominal segments are typically obscured by the midsection or

3.7 Heteroptera

85

barely noticeable. Since the size and colour of the species tend to vary, it can be quite difficult to distinguish many species of the genus Nysius without using a specimen or the male and female genitalia.

3.7.2

Anthocoridae

Cimicoidea contains a sizable and diversified family by the name of Anthocoridae. Additional taxonomic and evolutionary research on this family is desperately needed. Tropical and subtropical regions require thorough faunal research. If we want to use anthocorids for biocontrol, accurate identification is crucial. In various regions of the world, anthocorid predators are acknowledged as possible biocontrol agents. The low levels of pest populations have been successfully maintained by natural populations of numerous possible anthocorids. They consume tiny lepidopteran larvae, tiny grubs, psocids, mites, thrips, aphids, and a few pests that live in storage facilities. In general, the Anthocoridae play a significant role in the control of phytophagous mites, mite eggs, insect eggs, and a variety of soft-bodied insects. The majority of fruit trees, corn, cotton, soybeans, alfalfa, and grape vine are just a few of the significant crops that these beneficial insects may be found on. Natural selection and the unfavourable characteristics of this group of organisms are the two main causes of sexual dimorphism, which is clearly visible in the majority of anthocorids. Diagnostic Characters: Small, 3–5 mm, rather oval, flattened, black with white patterns; hemelytra with a cuneus; beak divided into three segments. Size and Colour: Medium-sized, delicate bugs called anthocorids have body lengths that range from 1.5 to 5 mm (0.06–0.19 in.). The adult can frequently be found on flowers, under loose bark, or in leaf litter, and it is primarily black with white patterns on its wings. Head: Prenatal disc contains shallow callosities separated by a few punctures; male anterior tibiae have a row of peg-like teeth; long setae on prenatal angles are absent; lateral margins of pronotum are wide. The body is brown to dark brown, with the following features: head, pronotum, last two segments of antennae, scutellum, base of clavus, sternum, and abdomen. The body of an adult minute pirate insect is oblong to oval and seems to be somewhat flattened on top. It also has bulging eyes. The rostrum had three segments, whereas the antenna had four. The antennae are likewise sexually dimorphic; segment II in the female is just slightly thinner and gradually extended towards the apex, whereas segment II in the male is clearly thick and sausage-shaped. In contrast to the female, segments III and IV in the male are also thick and spindle-shaped. Thorax: Front wings with a basal half that is leathery and an apical half that is membranous. Pronotum that is often broad, trapezoidal, or rounded (hemelytra). At

86

3 Identification of Various Insect Predators

repose, the wings form a “X” on the back and are two- or three-segmented on the Tarsi. Abdomen: In males, the anterior margin of the last sternite is laterally excised from the penultimate sternite, and the posterior margin of the seventh sternite has a V-shaped median projection bisecting the eighth sternite. The posterior margins of the last four abdominal sternites are skewed towards the lateral. On the left side of the abdomen were the male genitalia. Most anthocorid species, including Orius, Cardiastethus, Blaptostethus, Anthocoris, etc., clearly display sexual dimorphism. For instance, in Cardiastethus exiguus Poppius, the anterior margin of the last sternite is laterally excised from the penultimate sternite in the male, the posterior margin of the seventh sternite has a V-shaped median projection bisecting the eighth sternite in the female, and the posterior margin of the last four abdominal sternites are skewed towards lateral. The antennae are likewise sexually dimorphic; segment II in the female is only slightly thinner and progressively enlarges towards the apex, whereas segment II in the male is clearly thick and sausage-shaped. Additionally, thick and spindle-shaped in the male, segments III and IV are more slender in the female (Ballal and Yamada 2016).

3.7.3

Reduviidae

Assassin bugs (Hemiptera, Heteroptera, Reduviidae) have a variety of intricate behavioural and morphological adaptations for capturing prey. Assassin bugs refer to members of the heteropteran family Reduviidae because the majority of species attack and consume other insects. Long legs, an extended head with a noticeable constricted “neck”, and conspicuous, segmented, tubular mouthparts—most generally referred to as the proboscis, though some authors use the term “rostrum”—are their most distinguishing features. The majority of species have vivid colours that range from brown to black to red to orange. It contains thread-legged bugs or assassin bugs (genera include Melanolestes, Platymeris, Pselliopus, Rasahus, Reduvius, Rhiginia, Sinea, and Triatoma) (the subfamily Emesinae, including the genus Emesaya). This is one of the larger families in the Hemiptera, with around 7009 species known to exist. Diagnostic Characters: The beak is short, three-segmented, and its tip fits into a stridulatory groove in the prosternum. The head is normally elongate, with a necklike portion behind the eyes, and a transverse groove between the eyes. Size and Colour: Although many Reduviidae have huge bodies, overall body length can vary from very small species, such Tribelocodia, which have a body length of about 2–3 mm, to species with an astonishing 4 cm in length. Typically, adult insects are between 4.1 and 40.4 mm long. The majority of species are dullcoloured, frequently mirroring colour patterns in the microhabitats they occupy, such as bark, leaf litter, or rock crevices. However, species in at least seven subfamilies

3.7 Heteroptera

87

Fig. 3.4 Scanning electron micro photograph of Rhynocoris fuscipes head

display contrasting warning colouring. Assassin bugs have a variety of morphological adaptations and defensive and aggressive actions. With the help of these adapted defence mechanisms, animals may fend off predators, evade larger prey, and avoid cannibalism. Head: Labium is short and stout or elongate, anteriorly declivous, and the head is short and transverse or elongate. Head with or without a buccula that obscures the labium’s with 4 (1 decreased) or 3 segments of the labium are visible (Fig. 3.4). Short antenna with a scape that is either significantly longer than the pedicel or about the same length as the scape. The third labial segment (L3) is thin, bent, and straight. Whether or not the eyes pedunculate. Ocelli is either present or not. Thorax: The majority of assassin bugs feed on other arthropods, and a surprising variety of morphological changes have been made to help them capture their prey. A clade of Phymatinae, the ambush bugs, have developed subchelate (foretibia clamps against distal process on fore femur) and chelate (foretibia folds back against incrassate forefemur) raptorial grasping legs, while Emesinae can steal from spider webs thanks to their long appendages. Finally, species in a clade of Harpactorinae have without or with a clear fossula spongiosa at the ventral foretibial apophysis. There is no noticeable spur on the foretibial comb. The proleg exhibits several of these morphological modifications. Emesinae prolegs are primarily raptorial and are employed for prey hunting, prey grooming, and prey gripping. The proleg possesses a number of physical traits that distinguish Emesinae as a group, including the anterior opening of the acetabulum, the lengthening of the procoxa, and the lateral (campaniform) sensilla on the protibia. The Metapterini family, which consists of 28 genera and about 280 known species, is distinguished by a prominent basal process on the anteroventral series of the profemur and significantly altered pretarsal features (Castro-Huertas et al. 2019). Tarsi is having two or three segments. On the pronotum, there is a transverse sulcus either in the middle or at the front. Winged, apterous, and micropterous Flagellomeres are often divided into four to six pseudo segments, with the scutellum typically having two posteriorly projecting lateral

88

3 Identification of Various Insect Predators

Fig. 3.5 Adult Rhynocoris species such as (a) Rhynocoris kumarii, (b) Rhynocoris marginatus, (c) Rhynooris fuscipes, and (d) Rhynocoris longefroms

prongs and either no median tip or one median point. Bronzy-fuscous forewing membrane that just touches the abdominal apex.

Abdomen: The abdomen frequently overhangs the wings and is broader than the rest of the body. The abdominal borders are more widely exposed in females due to their bigger size compared to males. The tip of the rostrum is rubbed back and forth along transverse ridges on a longitudinal groove on the prosternum by the wheel bug and the majority of other reduviids to create “chirping” noises. Four reduviids belong to Rhynocoris spp. Present in India is representative of reduviid identification (Fig. 3.5).

3.7.4

Nabidae

Nabis currently includes 12 subgenres worldwide: Nabis (Australonabis) for the Australian region, Nabis (Aspilaspis) for the Afrotropical, Oriental, and Palearctic regions, Nabis (Dolichonabis) for the Neoarctic and Palearctic regions, Nabis (Halonabis) for the Oriental and Palearctic regions, Nabis (Limnonabis) for the Neoarctic and Palearctic regions, Na Nabis (Philobatus) (Palearctic region), (Omanonabis) (Nearctic region), and Nabis (Tropiconabis) and (Reduviolus) (Nearctic and Palearctic areas) (Australian, Afrotropical, Nearctic, Neotropical, Oriental, and Palearctic regions). Although there is some controversy regarding the fundamental eating habits of heteropterans, it is recognised that damsel bugs belong to the insect family Nabidae. In 20 genera, there are more than 500 species. Many damsel bugs use their forelegs to grasp and hold prey, much like mantids do. Diagnostic Characters: Oblong and long; ocelli present; prothoracic legs with thinned-out femora; hemelytra membrane with numerous marginal cells. Special Characters: Mostly feeds on aphids, moth eggs, and small caterpillars, such as imported cabbageworm, corn earworm, and armyworms. Leafhoppers, tiny sawfly larvae, mites, nymphs of the tarnished plant bug, eggs and nymphs of the

3.7 Heteroptera

89

asparagus beetle, and Colorado potato beetle are examples of additional prey. While the immature stages are known to prefer staying lower down on the plant, outside the canopy, Nabis adults are known to favour the highest regions of the plant (within the plant canopy). Size and Colour: They are elongated, wingless, soft-bodied terrestrial carnivores. The majority of species are of average size, occasionally growing longer than 10 mm. Some are more stout-bodied, while others are more elongated and dully coloured. Occasionally, nabids have characteristic red and black colour patterns or tan-coloured bugs that resemble small, smooth-looking assassin bugs that range in colour from yellow to reddish brown or fuscous to black. Head: Behind the eyes, the head constricted noticeably and obliquely. When a Nabidae animal prey is penetrated, the mandibles protrude in front of the maxillae. The tips of both mouthpieces are serrated. The rostrum of Nabidae members is flexible and very mobile; it is divided into four segments and, while in the resting position, never rises above the mesocoxae. Approximately or not twice as long as the head is the scape. The third article is typically the longest, whereas the first is rectilinear and thick. The most adaptable joints are the first two. The maxillary stylets, which have sharp denticles that are pointed forward in their anterior region, are longer than the mandibular stylets, which are delicately denticulate. Beyond the base of the head, the second rostral segment. Thorax: Pro- and meso-femoral setae resemble spines and are equipped ventrally with tiny, blunt, piceous teeth. Prothoracic anterior lobe with deep, central impression on basal margin; median lobe broad, trapezoidal, strongly convex, and anteriorly declivous; prominent scent gland; enlarged front legs for grasping prey; hemelytra narrow at middle; membrane round posteriorly; no distinct venation; macropterous. Nabidae have wings as adults, and some species may travel quite a distance. Numerous Nabidae adults have fully formed (macropterous) wings, and numerous species exhibit significant wing variation (microptery and brachyptery). Wing decrease might be little or significant. Though this is an abnormal condition, there are a few cases when the wing reduction is unilateral rather than bilateral. Abdomen: The lateral borders of the abdomen, especially in females, are parallel, and the body is significantly elongated or short. Abdomen sterna with little shining black bare spots on the mesa of each spiracle; abdominal connexiva not separated from venter by noticeable longitudinal depression. Male paramerenarrower, with apex that is longer and narrower and seems to be more sharply twisted or incurved; brachypterous hemelytra that only extends to anterior boundary of third abdominal tergite.

90

3.7.5

3 Identification of Various Insect Predators

Pentatomidae

Pentatomidae species are opportunistic predators that target mainly coleopteran and lepidopteran larvae. Although stink bugs come in a variety of sizes and shapes, they are typically recognised by their five-segmented antennae, well-developed and frequently triangle-shaped scutellum, and round or oval (occasionally shield-shaped) bodies. Diagnostic Characters: Large, triangular scutellum that does not extend past the corium or the top of the abdomen; spines on the tibiae are weak or non-existent. There are several well-known stink bugs. They release pungent gases from their stink glands. The marmorated stinkbug, the southern green stink bug, and the harlequin bug on cabbage are among the pestiferous species; the majority is herbivorous. Some are predatory. Size and Colour: About 7–25 mm in length, moderate in size. Body violaceous blue with dark bronzy punctures on a greyish-yellow background. Head: Each side of the central lobe of the head has a band of blackish punctures, and between the pronotal angles there is a pale, smooth, shining line. The pronotal lateral angles have two spines, the posterior spine of which is extremely little. Body length may range between 12 and 14 mm; membrane and antennae are black; antennae, eyes, two to three spots on the pronotum, two spots on each basal angle of the scutellum, and membrane are black; the abdomen, rostrum, and coxae of the femur are pale reddish. Predatory stink bugs have piercing-sucking mouthparts that are made up of a four-segmented rostrum or beak (labium) that forms a sheath around two mandibular and two maxillary stylets. Asopines are distinguished by having a thicker rostrum as opposed to phytophagous pentatomids, which have a short rostrum. Mandibular stylets on pentatomid predators have sharp hooks and blunt tips for anchoring to prey tissue. Pentatomids typically have harpoon-like stylets, whereas reduviids typically have knife-like stylets. The mandibular teeth of certain predaceous Heteroptera differ from one another. Thorax: Broken transverse fascia on the anterior portion of the pronotum, some more or less prominent longitudinal fasciae on the anterior half, and bronzy black lateral angles. The body above is extremely delicately and sparingly punctate, with the scutellum’s base slightly raised. The front wing completely or almost completely encircles the border of the abdomen. Abdomen: Eight-segmented abdominal wall, with the first segment being significantly shorter than the others. Segment IV typically has the longest lateral margin, with the segments’ lateral lengths getting shorter at the front and back. Segments III– VII have a single spiracle on either side ventrally. White to green in colour, punctate laterally, lacking in the middle. Brown to red tergum with tiny punctures. Green connexiva with larger tergum-sized punctures. Ventral and dorsal sutures are pale in

3.7 Heteroptera

91

hue. Located ventrally on the second abdominal segment and continuing to the middle of the metasternum is the basal spine. Setae are particularly plentiful medially along the male pygophoresetose’s ventral posterior border. Similar in shape to men, but typically larger than them. The apex of the first gonocoxae is sparsely setose, whereas the apices of the eighth and ninth paratergites are fairly setose. Other traits, such as adult males.

3.7.6

Miridae

Although many species of the Miridae are facultatively predatory on all life stages of different insects, primarily Homoptera and Heteroptera, the majority of them are phytophagous. Some researchers, however, believed that the family as a whole mostly relied on other insects for sustenance. Some creatures primarily feed on human blood and are zoophagous. There are species of mirids that frequent insectivorous plants and feed on recently caught prey as well as myrmecophilous mirids. Diagnostic Characters: Small, oval, or elongate; 10 mm or less; hemelytra with cuneus; membrane with one or two closed cells. Size and Colour: The adult specimens range in length from 2.75 to 3.50 mm. Some are gloomy, while others are vividly colourful. Head: Small, convex, sloping, black, and with setae on the vertex, the opisthorhynchous head. Short, extended into the neck, and constrained by the collar, rostrum, apex of first joint of antennae, anterior callosities, broad central area to posterior area of pronotum, base of head, huge spot on inner side of each eye. Filiform, yellow, four-segmented antennae with bristles on every segment are inserted on the outer half of compound eyes. Oblong-reniform, red, and separated from the anterior margin of the pronotum are the compound eyes, which are situated in the middle of the head. Labrum is brief and undetectable. Yellow sucking teeth extending between the middle and posterior coxae. Mandibles, maxillae, and rostrum have four segments each. The labium is cylindrical. The mandibles and maxillae are stiff and closely spaced. Thorax: Strongly punctate thorax that gradually gets narrower towards the abdomen. The pronotum, which was black with a central green patch and extended to the middle of the thorax, was slightly apart from the mesoscutum. The scutum-scutellar suture’s margin was covered with setae that reached the collar. Setae on a black collar pointing towards the vertex. The prosternum is expanded ventrally and parallel to the collar beneath the episternum. Propleural suture separates the epimeron from the episternum. Orange hairs on the mesoscutum facing the hemelytra. Mesoscutum between the scutellum and the scutum-scutellar suture with three black dots. Short, spotty, and with a flattened scutellum, the postnotum. Mesopleural suture divides the meso-epimeron and meso-episternum. As the

92

3 Identification of Various Insect Predators

mesothoracic coxae extended ventrally, so did the mesosternum. Unique corium, clavus, and cuneus on the hemelytra (a triangular apical piece of the basal part of forewing). The forewings have a noticeable posterior to abdominal tilt. The membrane contains loop veins. The wings have a downward inclination. A big black lateral patch that extends to the sternum from each side of the scutellum, with a body beneath it and iridescent legs having conical coxae. Trochanter with visible front and back and coxo-trochanteral articulation. Large, sturdy, proximally expanded femora with six spots and setae. Dark spines expand the tibiae. Two tarsomeres on Tarsus, the first one on the left Clavus is triangular and inverted, and it is enclosed by the wing’s edge and the corium in a straight line. Cuneus fractured and just half the size of the clavus. White and translucent membranous wing with two cells and pale veins. Abdomen: First tergite relates to the first ventricle of the hologastric abdomen, which has ten transversely and laterally compressed abdominal segments. On the visceral abdomen, tergites and vent rites have similar sizes. Reduced terminal ventrites that do not reach the wing’s distal edge.

3.8

General Features of Vespidae

The Hymenoptera (Insecta) family of wasps, or Vespidae, is widely distributed. More than 5000 species of members of this family can be found all over the world, with most of them being found in tropical areas. Adults are typically black or brown with extensive yellow or white markings. In solitary species, the larva lives in a cell built and abundantly supplied by the adult female and feeds primarily on the caterpillars of other insects. Rarely is a mixture of pollen and nectar given to the larva. The larva of social wasps gradually consumes masticated insects or, less frequently, is primarily fed on glandular secretions by adult females. Vespid wasps have significant ecological functions, making them useful bio-indicators of changes in the environment and in their habitat. They have the potential to be efficient biological control agents due to their predatory behaviour. Adults range in size from tiny to large (6–28 mm). Solitary wasps can build their own free mud nests or they can use pre-existing cavities, such as beetle tunnels in wood or abandoned hymenopteran mud nests, to build their nests. Social wasps frequently construct impressive paper nests and fiercely defend their colony by stinging. The following traits are indicative of the family Vespidae: Antennae 13 segmented in male and 12 segmented in female; inner margin of eyes deeply emarginate; dorsal rim of torulus simple; Pronotum extending back to tegula; dorsal rim of torulus simple; inner margin of the eyes deeply emarginate; mesopleuron lacking any oblique sutures; wings longitudinally folded in repose (except in Masarinae and Stenogastrinae); fore wing with first discoidal cell elongate and longer than submedian cell (except in Masarinae); hind wing with an toothed, bifid, or simple tarsal claws; spiracles on 1–7 segments on the gaster; and a sting without sheaths; Both sexes macropterous, with distinctive spiniform parameres on the male genitalia.

References

3.9

93

General Features of Formicidae

Their exoskeleton, hard armour covering their body, is called this. The majority of ants have either a red or a black tint, and their length can range from 1/3″ to 1/2″. They have six legs, each with three joints, just as other insects. Ants have huge heads, elbowed antennae, compound eyes, and strong jaws. Due to its substantial impact on soil structure, the ant family, or Formicidae, is most likely the most important group of soil insects. Ants can be found in polar and tropical settings and are abundant, diversified, and widely distributed. Ants significantly alter their environments. They are important tiny invertebrate predators. Due to their actions, there are less other predators like spiders and carabid beetles.

3.10

Predatory Lepidopterans

Given the vast food diversity displayed by other holometabolous organisms, the scarcity of carnivorous Lepidoptera is particularly surprising. Phytophagecontaining orders include the Coleoptera, Hymenoptera, and Diptera. There are only roughly 200 species in eight superfamilies, known to be parasites or obligate predators. In addition, as predators, Lepidopterans are surprisingly timid creatures; they typically eat slow, soft-bodied scale insects, other insects’ eggs, or ant brood. Balduf (1938) identified four major categories of entomophagous caterpillars: cannibals, which primarily diverge from otherwise phytophagous lifestyles; occasional predators, such as species that occasionally attack non-conspecific caterpillars and scavengers that occasionally take prey living in the same habitat; habitual predators, such as species that regularly feed on homopterans or insects like ants; and parasites/parasitoids, such as thrip. The majority of the animals discussed in this article belong to groups 3 and 4, which together make up the group of obligate carnivores, whereas those in categories 1 and 2 are facultatively entomophagous.

References Balduf WV (1938) The rise of entomophagy among Lepidoptera. Am Nat 72:358–379 Ballal CR, Yamada K (2016) Anthocorid predators. In: Omkar (ed) Ecofriendly pest management for food security. Academic Press, London, pp 183–216 Castro-Huertas V, Forero D, Grazia J (2019) Comparative morphology of the raptorial leg in threadlegged bugs of the tribe MetapteriniStål, 1859 (Hemiptera, Heteroptera, Reduviidae, Emesinae). Zoomorphology 138(1):97–116 Grzimek B (2004) Grzimek’s animal life encyclopedia. Volume 3: Insects, Gale Snyder WE (2019) Give predators a complement: Conserving natural enemy biodiversity to improve biocontrol. Biol Control 135:73–82

94

3 Identification of Various Insect Predators

E-Resources https://www.britannica.com/animal/insect 49-04-412-453-Pierce.pdf (yale.edu) https://www.royensoc.co.uk/entomology/orders/ https://texasinsects.tamu.edu/insect-orders/ https://en.wikipedia.org/wiki/Insect

4

General Ecology of Insect Predators

Contents 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Association with Tropics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.1 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.2 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.3 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.4 Nabide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.6 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.7 Anthecoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.8 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 Intra- and Inter-Specific Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4 Niche . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5 Future Areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.1

95 96 96 96 97 98 99 101 104 105 106 108 110 110

Introduction

The common ecology of generalist insects, particularly predators, has not been studied extensively, with a focus on the habits and associated organism that generalists are expected to exhibit. A survey is more important to know the ecology of any predators. To identify resident natural enemies, field camera aided in the identification of arthropod predators and also to other animals in the field. Many predators have greater limitations on the settings in which they can survive than do their prey. Such predatory insects could have different environmental preferences depending on where they are. However, understanding an organism’s habitat preferences is crucial for understanding its evolution as well as for facilitating conservation efforts for threatened species and for understanding how they may react to manmade challenges like climate change (Sloggett and Zeilstra 2020).

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_4

95

96

4.2

4

General Ecology of Insect Predators

Association with Tropics

During the years 2016–2018, sorghum (Sorghum bicolor) plants underwent a triangular relationship analysis with predatory insects (Coccinellidae, Syrphidae, and Chrysopidae), their prey aphids, and environmental factors. Schizaphis graminum, Rhopalosiphum maidis, and Melanaphis sacchari were the three aphid species that were gathered. The most prevalent species, Melanaphis sacchari, had a sharp decline in population over time. Twelve Coccinellidae predator species, two Chrysopidae, and 13 Syrphidae species were noted. Melanaphis sacchari and predators have a greater association than weather, according to principal component analysis. According to Rodríguez-Vélez et al. (2021), the relationships between Hippodamia convergens, Scymnus dozieri, and Melanaphis sacchari were outstanding.

4.2.1

Hymenoptera

The fifth or sixth pair of leaves on the principal plagiotropic branches of the median third of the canopy served as the optimum sample unit for Vespidae predators when ants were in the vegetative phase. The optimal unit for sampling Vespidae in coffee plants that were already in the reproductive phase was a leaf on the fourth or sixth pair of leaves on principal plagiotropic branches, or on the median third of the fifth pair of leaves on the plant face exposed to the sun in the afternoon (Fernandes 2012). Although ant assemblages and forest type were only moderately correlated, early successional woods had significantly more ants and more species of ants than mid-late successional forests. Micropterous wasp abundance and richness in morphospecies, on the other hand, were positively correlated with aspen basal area but unrelated to successional stage. The findings show that changes in the boreal vegetation brought on by climate warming will also affect the population of predaceous insects, with ants responding favourably to disturbance and wasps favourably to an increase in the amount of aspen in the landscape (Alexandria et al. 2019). The major ant species found throughout the Eurasian continent is the red wood ant (Formica rufa). Due to their status as top predators, these animals have a significant ecological impact on the ecosystems they inhabit (Frizzi et al. 2022).

4.2.2

Syrphids

About one-third of hoverfly species belong to the Syrphinae subfamily, which can live in a wide variety of vegetated terrestrial settings. With the exception of Antarctica, the group is widespread and is particularly well-represented in the Palearctic and Nearctic regions. Syrphid larvae have a wide variety of environments. Syrphids can be divided into two groups: aquatic and terrestrial, depending on their habitat. Syrphids can be scavengers, predators, and phytophagous on land as well as in the water. A few commensal species, such as bees, wasps, and termites, are

4.2 Association with Tropics

97

connected to social insects. The subfamilies Eumerinae and Macrodontinae are primarily home to phytophagous species. On plants from the Liliaceae family (onion and hyacinth), Umbelliferae (carrot), and Solanaceae (potato), phytophagous species have been observed (potato). In extremely contaminated aquatic habitat, scavenger or saprophagous species of Xylota, Syritta, and some species of Eristalinae exist. It is known that some Eristalis species can accidentally myiasis in people and domestic animals. While a few species are aquatic or survive in extremely damp environments, the majority of species are terrestrial. At least 25% of the terrestrial forms are predatory, primarily aphidophagous. Three subfamilies, the Syrphinae, Milesiinae, and Microdontinae, make up the family Syrphidae. The majority of aphidophagous syrphids are members of the Syrphinae subfamily. There are more than 4700 species in the world, and the Indian subcontinent alone is home to 312 species in 71 genera. The majority of aphidophagous syrphids found in Europe belong to the Syrphini and Melastomini tribes of the Syrphinae subfamily. In contrast, the Syrphini and Paragini tribes in India are significant. Different Paragus species are widely spread in the Syrphini, but Ischiodon, Eupeodes, Dideopsis, and Episyrphus are significant genera in the Paragini. The three genera with the most in-depth investigation are Eupeodes corollae, Ischiodon scutellaris, and Episyrphus balteatus (Sunil and Chandish 2013). A long-term study was carried out in Brazil between February 2010 and January 2015 to determine the Syrphidae population level in monoculture and silvopastoral systems. A total of 11 hoverfly species (Diptera: Syrphidae) were gathered, of which 5 and 3 were unique to silvopasture and monoculture, respectively. In comparison to the silvopasture, the monoculture had a much higher number of specimens. A total of 24 species of robber flies (Diptera: Asilidae) were gathered, of which eight were unique to silvopasture and two to monoculture. Robber fly populations were not noticeably different between the monoculture and the silvopasture. The diversity of the Syrphidae was higher in the monoculture and was comparable to the diversity of the Robber fly family. Circular study showed seasonality, with hoverfly distribution clustering during dry seasons and dispersing during robber fly seasons. In each of the systems, the majority of hoverflies and robber flies were unintentional (Veríssimo et al. 2021). The authors recommended more research be done on how these predator species can work as biological control agents for pasture pests.

4.2.3

Neuroptera

It is regarded as a crucial biological control agent globally. The fact that Neuroptera are resistant to a variety of pesticides is another factor contributing to their popularity. This makes them—particularly Ch. carnea—useful control organisms for toxicological testing of agrochemicals. The International Organization of Biological Control uses Ch. carnea as one of its test organisms (IOBC). Larvae of Neuroptera can be semi-aquatic or aquatic. Although some larvae can be found on the water’s surface, most are found in the benthic zone, beneath rocks, on the surface of sponges, or inside their cavities. It depends on the species; larvae

98

4

General Ecology of Insect Predators

are either general predators that eat a variety of small benthic invertebrates or specialised predators of freshwater sponges and maybe Bryozoa. Typically, invertebrate predators use their probosci to sift through the muck in search of prey. Predators of sponges use their proboscis to puncture the sponge cells and get the fluids out. The early, in-depth studies of the biology and ecology of sisyrids by Withycombe (1923, 1925) and Killington (1936) are the best examples. They are the perfect insect models for ecosystem health as well as environmental indicators due to the ecological separation of adults and larvae, which inhabit a variety of habitats (such as arboreal, psammophilous, semi-aquatic, aquatic, and inquilines) and exhibit many different life history strategies (Aspöck 1992). Many families are relatively understudied, and our understanding of their taxonomy, biology, phylogeny, biogeography, and conservation status is inadequate despite the fact that they are of considerable ecological importance (Ohl 2004, 2011). In Saudi Arabia, the Al Figrah Mountains typically had more specimens of the nemopterid Dielocroce chobauti (Nemopteridae) males and females than the Al Hazim Valley (Al-Ahmadi 2016).

4.2.4

Nabide

The Nabidae are part of a guild of terrestrial ecosystems’ arthropod predators. They frequently hang out with members of a number of other predaceous Heteropteran families, including Anthocoridae, Lygaeidae (Geocoris etc.), and Miridae (Deraeocorinae and others). There is an apparent ordering of habitat preferences even though the Nabidae is a small family with only about 380 species (Kerzhner 1986) and very limited ecological diversity. Southwood (1977a) presented a useful explanation of the r-K selection spectrum and described how the habitat affects ecological strategies. K strategists seem to be the most well-known nabids. Particularly in terms of distribution, a significant portion exhibits some degree of lung loss. Additionally, some species have narrow niche breadth. For instance, the Pro-stemmatinae are present in the litter layer and on the ground. Ground-dwelling Lygaeidae are frequently closely allied with their prey (Pericart 1987). The first few nymphal instars of Himacerus apterus reside in the grass-forested zone before moving up into trees (Koschel 1971). Since it can only survive in spider webs, Arachnocoris albomaculatus is thought to eat trapped insects there (Myers 1925). Several species of Chrysomelidae are preyed upon by the typically brachypterous Nabicula subcoleoptrata within its habitat (Messina 1982). Although nabids appear to be able to attack a variety of other arthropods, the majority of the prey species—as was previously mentioned—belong to specific insect groups. In fact, certain species can be regarded as r-strategists since they are generalist feeders. Alloeorhynchus reinhardi has recently been discovered in moist areas near streams, particularly on piles or clusters of dead Cyperaceae spp. leaves under plants in Kyushu, Japan, and Korea. Adults were gathered between March and November (Souma et al. 2022).

4.2 Association with Tropics

99

4.2.4.1 Agroecosystems Numerous research on the functional role of nabids as predators in various agroecosystems have contributed significantly to our understanding of their habits (Stoner 1973). There are only a few species, and the most of them are from the genus Nabis. According to Southwood’s definition, the most, if not all, would be regarded as r-strategists (Southwood 1977b). He suggests that those multi-predator species are likely to play a significant role in the transient habitats frequently produced by agriculture, where K-strategists may not be able to establish themselves. According to Ehler and Miller (1978), r-selected natural enemies can manage r-selected pests. They used Nabis americoferus as an example of one such adversary. There are a lot of data to suggest that nabids react to habitat alteration, like what happens in agroecosystems. More nabids were sustained by inter-cropped bean and maize fields than by monocultures (Milanez 1984). Despite having no effect on predator populations, different soybean maturation dates, locations, and row spacing had more nabids in grassy soybean fields than in broadleaf weed and weed-free areas. Nabids were most prevalent in no-till fields that had previously been planted with soybeans and had not received an insecticide application in conservation tillage schemes for soybean fields (Hammond and Stinner 1987). Although predatory Heteropteran species, such as the Nabidae, could play a significant part in the natural control of pest species, Carayon (1961) noted that it would be challenging to make use of them due to their polyphagy, cannibalism, and change in numbers over time and geography. Heteropteran predators of two different sizes were recognised by him. Since nabids typically belong to a guild of predators, the combined effort of this guild is crucial. A continuum capable of coping with a wide range of prey sizes is provided by the size classes of the Anthocoridae, Lygaeidae, Miridae, Nabidae, Reduviidae, and Pentatomidae (Irwin and Shepard 1980).

4.2.5

Coleoptera

4.2.5.1 Carabidae In arable terrain, carabid beetles develop diverse, abounding groups. Because of their lengthy legs and strong mandibles, ground beetles are recognised for being ferocious predators and crucial for the biological control of insect pests on farms. Although they mainly forage on the soil’s surface, adult beetles will occasionally venture into the surrounding vegetation in search of food. Carabid activity, abundance, and species richness in agricultural settings can all be influenced by environmental conditions. The most crucial variables are crop qualities, food availability, soil type, soil moisture, temperature, humidity, and light. The shade offered by cultivated or weedy plant cover, which maintains soil moisture and lessens temperature change below the canopy, can also have an impact on carabids (Cividanes 2021). Carabids are epigean beetles that live in the top strata of the soil where they occupy a variety of ecological niches and play a significant role in the diversity of predators in both agricultural and natural settings. Both hydrophilic and eurythermic

100

4

General Ecology of Insect Predators

describe Pterostichus melanarius. This species inhabits a range of soil types; however, it stays away from pure sand and gravel. Pterostichus melanarius can live in a wide range of settings, including cities, roadsides, woods, forest edges, meadows, grasslands, lake and river banks, orchards, and arable land because it is able to adapt to a wide range of temperatures and soils and because it is a generalist feeder. Pterostichus melanarius may also spread its range rather quickly, with mean daily dispersal distances of 2.5–5 m. Individuals can, however, move up to 44 m in a single night. When looking for mates in the late summer, the daily dispersal distance may increase (Busch et al. 2021). A variety of insect pest populations, such as beetle larvae, caterpillars, aphids, and slugs, are controlled by Pterostichus melanarius, a critical predator, in many agricultural systems, including vegetable, small and large fruit, and field and fodder crops (Busch et al. 2021).

4.2.5.2 Coccinellidae Some species of aphid-eating ladybird beetles exhibit habitat specialisation that is directly related to the types of aphids they consume; however, some taxa appear to consume a wide variety of aphid species but are still restricted to specific environments that are not the sole source of their prey (Sloggett and Zeilstra 2020). Coccinellidae have a variety of roles in complicated food webs, including that of predators, consumers of non-prey items, and hosts or prey for natural enemies. The claimed variety of coccinellids varies widely between nations as well. There are two types of coccinellid species: stenotopic and eurytopic. A coccinellid habitat is thought to have a particularly significant microclimate. Many ladybird species show a preference for particular flora types or habitat strata. Along with this, there must be an adequate supply of appropriate food. The preference for a certain habitat changes with the seasons as a habitat’s microclimate changes, which in turn affects the distribution of prey populations and the behaviour of coccinellids. A well-researched group of consumers, predatory lady beetles (Coccinellidae), can shed light on the connections between inter-specific niche diversity, species richness, and prey consumption. According to several research, sympatric lady beetles employ different habitats when observing individual plants, individual agricultural fields, and larger landscapes. Additionally, seasonal and daily activity patterns vary among species. By encouraging various predator species to target various subgroups of the prey population, these spatiotemporal disparities in habitat utilisation should have complementary effects on prey. This should then result in stronger biological control at levels of greater predator variety. Indeed, stronger prey suppression is often associated with higher predator biodiversity, according to experimental manipulations of predator species richness in communities that include coccinellids. In these experimental experiments, lady beetles occasionally occupied distinct niches as very ferocious predators, and frequently they also complemented or made it easier for other species to catch their prey. It was discovered that intergroup predation was rarely a powerfully disruptive effect, possibly because differences in spatiotemporal niches prevented inter-species contacts. In conclusion, coccinellid species have unique responses to and effects on their communities, usually to the advantage of herbivore control. Therefore, coccinellids may fit best

4.2 Association with Tropics

101

in a habitat where they complement other species, helping to increase the diversity of predators and improve biological control (Snyder 2009). Coccinellids’ ecology has not been adequately studied. Coccidophagous coccinellids tend to stay in a limited region during their life cycle, whereas aphidophagous coccinellids disperse widely. The majority of the species are most common in recently cultivated agricultural, horticultural, and household settings, gardens and parks, and greenhouses (Roy and Migeon 2010). In 1995 and 1997, there were many trees (including orchard and cultivated crops), shrubs, and weeds in Northern Italy. The authors conducted a survey and found that Coccinellidae, along with other predators, were widespread. These predators included Hemerobiidae, Chrysopidae, Syrphidae, Anthocoridae, Nabidae, Miridae, Carabidae, Cantharidae, Forficulidae, and Staphylinidae (Burgio et al. 2004). In 2006, pear and apple orchards and arable crops were the main features of the province of Bologna. Coccinellidae, Anthocoridae, Nabidae, Carabidae, Staphylinidae, Chrysopidae, and Syrphidae are all present in the 1995–1997 survey (Burgio et al. 2006). The five-spot ladybird, Coccinella quinquepunctata, was studied for its habitat preferences (Sloggett and Zeilstra 2020). The ladybird Coccinella quinquepunctata is found in pioneer habitats near water, according to the results, but it also ranges into non-riverine pioneer habitats in north-western continental Europe, where it is most likely able to do so due to a damper (micro)climate. Inferring from this, it is likely that other predators whose habitat is unrestricted by the presence of prey also depend on microclimatic conditions.

4.2.6

Reduviidae

In tropical evergreen forests, Ectomocirs xavierei can be found among the stones and in the litter, frequently with other assassin bugs like Pirates affinis and Holoptilus melanospilus (Vennison and Ambrose 1991). The assassin bug Acanthaspis quinquespinosa is alate, warningly coloured, crepuscular, entomosuccivorous, polyphagous, and multivoltine. It is found in peninsular India’s semiarid regions, scrub jungles, tropical evergreen forests, and agroecosystems. Neohaematorrhophus therasii, Acanthaspis pedestris, and Acanthaspis quinquespinosa were gathered. Only at lower elevations, where the predators Camponotus compressus, Odentotermes obesus, carabid beetles, and Chrotogonus sp. are dominant, was the bug discovered. Here, as opposed to the other two habitats, there was a greater chance of these bugs catching a lot of prey. Acanthaspis quinquespinosa and Neohaematorrhophus therasii, two reduviids, were discovered behind some stones (Sahayaraj 2007). Sivanthipatti is a scrub jungle located distant from tropical rainforest areas in Palayamkottai. From this environment, researchers gathered Acanthaspis quinquespinosa, Edocla slateri, Neohaematorrhophus therasii, Coranus nodulosus, Rhinocoris marginatus, Ectomocoris tibialis, and Oncocephalus annulipus. There are farmed areas nearby where vegetables, cotton, paddy, and groundnuts are

102

4

General Ecology of Insect Predators

planted. Under the stones, Acanthaspis quinquespinosa, Odentotermes obesus, Chrotogonus sp., and carabid beetles could all be found. Predation on insect pests from nearby agroecosystems was a fairly likely potential (Sahayaraj 2007). Usarathukudieruppu, a semiarid region in Tamil Nadu, is located close to Sathankulam but farther from a tropical rainforest than Sivanthipatti. Odentotermes obesus was the only food source in this environment, making it the only food source for Acanthaspis quinquespinosa. From this habitat, Acanthaspis siva, Acanthaspis pedestris, Ectomocoris cardiger, Allaeocranum quadrisignatum, and other species of Acanthaspis were collected. On the bark of Tamarinadus indica, Acanthaspis quniquespinosa and another reduviid, Acanthaspis siva, were discovered (Sahayaraj 2007). The increased prey population of carabid beetles (18–33), ants (18–27), caterpillars (13–19), plant bugs (15–26), and grasshoppers (20–40) had a favourable link with the population of Irantha armipe (Ambrose 1980). First record of a myrmecophiline and coprophagous Reduviid is Lophocephala guerini (Ambrose and Livingstone 1979). However, later research revealed that Lophocephala gueriniis is an entomophagous species that hunts termites in fermented cow dung (Murugan 1988). The formicine ant that lives in the fissures is called Anoplolepis longipes. During the feeding march, they are observed to gather around the adults and nymphs of L. guerini and to accompany them back and forth (Ambrose 1980). Only Sri Lanka and India were known to contain this unusual species. Only a few individuals of Lophocephala guerini were found during our recent survey for Reduviidae in Pakistan’s district Swabi of Khyber Pakhtunkhwa Province. The anatomy, pronotal region, and lack of paramere in the Lophocephala guerini make it unique, but so are its ethology (dung-feeding and myrmecophily) (Shah et al. 2022). To make identification and subsequent research on this beetle easier, we redescribe this species in this article along with digital photographs of the habitat, male genitalia, and significant morphological features. In the scrub jungles and semiarid regions, Coranus vitelllnus, a multivoltine, entomosuccivorous polyphagous alate (both sexes), can be found. It can be found in many nooks and hidden microhabitats, such as those found beneath stones. Despite being discovered beneath the stones, the nymphal instars are never discovered with the adults, unlike in several other reduviids. This species has been reported to co-occur with various kinds of carabid and Tenebreonid beetles (Ambrose 1980). Underneath stones, Rhinocoris kumarii inhabits areas of hiding. Under the stones, nymphal instars can also be spotted, although never with adults. There have not been any other insects discovered yet that may be considered members of the same species. Never are the grownups seen in couples. There is only one adult or one nymph in a specific microhabitat, indicating that there is no sign of congregational activity (Ambrose 1980). Rhinocorls marginatus is a Harpactorine species that lives in concealed microhabitats such cracks and stones. It is alate (both sexes), sanguineous, entomosuccivorous, polyphagous, crepuscular, and multivoltine. The same microhabitat as the adults is also home to the nymphal instars, although no parental care has been seen. The same microhabitats are also home to many kinds of beetles (Carbidae and Tenebreonidae), a common yellow scorpion (Buthus sp.), and

4.2 Association with Tropics

103

occasionally poisonous reptiles like Echis carinatus. Even though the nymphs are reported to gather (up to five), this beetle is very rarely found in pairs, and no more than two adults are ever found in the same microhabitat at once (Ambrose 1980). Black, entomosuccivorous, polyphagous, alate (both sexes), crepuscular, and multivoltine describe Rhinocoris fuscipes. Tamil Nadu’s scrub jungles, semiarid regions, and tropical rainforests are all home to reduviid. This harpactorine species has been discovered in hiding places, like under rocks. The Pechiparaidam location in the Kanyakumari district is where it was found. Chandrapuram, Malumichampatti, and Kangayampalayam are semiarid regions, as is the site of the Forest College in Coimbatore. Under the stones, nymphal instars can also be spotted, although never with adults. No specific microhabitat has so far been discovered with cohabitants. They are never discovered in pairs, and only one bug has ever been discovered in a single microhabitat. There are lone nymphs hidden beneath the stones (Ambrose 1980). A violaceous black, crepuscular, polyphagous, entomosuccivorous, and alate (both sexes) reduviid called Sphedanolestes aterrlmus can be found in the Azhagarmalai Tropical Rainforest as well as the higher altitude Maruthamalai Scrub Jungle, where tropical rainforest conditions are prevalent. Sphedanolestes aterrlmus nymphs and adults have both been found in the waste material beneath Tamarindus indices. Adults and nymphal instars of numerous species of blattids, as well as nymphal instars of another alate tropical rainforest reduviid called Sycanus ater, are discovered to coexist with this species. However, at Maruthamalai, a single adult female was discovered on the leaf of a higher-elevation plant, after removing the nymph of this species from beneath a stone. Typically, Sphedanolestes aterrimus is found in pairs. A single microhabitat can house up to nine adults (including males and females) and four nymphal instars (Ambrose 1980). Ectomocoris tibialis is an entomosuccivorous, polyphagous, brachypterous (both sexes), univoltine reduviid that can be found in semiarid and scrub jungle regions but not in tropical rain forests. It is a violaceous black crepuscular species. Ectomocoris tibialis inhabits areas that provide cover, including those found under stones. Along with the adults, many nymphal instar stages can also be detected. Common yellow scorpions (Buthus sp.), many species of Carabid and Tenebreonid beetles, another entomosuccivorous micropterous reduviid Acanthaspls pedestris, and a brachypterous can occasionally be seen. Ectomocoris tibialis and reduviid Catamiarus brevlpennis have also been observed living together. Pairs of Ectomocoris tibialis are frequently observed. Up to six adults (four males and two females) can live in one microhabitat, according to records. According to a previous report (Ambrose 1987), these bugs go from one habitat to another in pursuit of Camponotine ants, which are their main source of food. As a result, the population dynamics of the nearby Camponotine ants greatly influence the population dynamics of this beetle in a particular microhabitat (Ambrose 1980). therasii Violaceous black, crepuscular, Neohaematorrhophus entomosuccivorous, multivoltine, and micropterous (males alate) reduviid are found in scrub jungles and seroiarid. Neohaematorrhophus therasii can be found in hiding places like cracks and under stones. It can be found widely throughout the

104

4

General Ecology of Insect Predators

diverse ecotones of scrub jungles and semiarid regions but is never present in tropical rain forests. Along with the adults, nymphal instars are also present. This species frequently coexists with carabid and tenebreonid beetles as well as the common yellow scorpion (Buthus sp.). On rare occasions, this species has been discovered among venomous reptiles like Echis carinatus. This species closely resembles a certain type of carabid beetle in terms of size, colour, and posture, making identification more challenging. Additionally, when disturbed, the collector constantly discovers them motionless, which is a distinctive ethological trait of importance in this particular species of reduviid. It rarely appears in pairs. So yet, only two adults at most have been observed in a single microhabitat. The males are rarely seen under stones because they are alate (Ambrose 1980). The assassin bug Acanthaspis quinquespinosa is alate, warningly coloured, crepuscular, entomosuccivorous, polyphagous, and multivoltine. It is found in peninsular India’s semiarid regions, scrub jungles, tropical evergreen forests, and agroecosystems. Acanthaspis quinquespinosa is widespread in tropical rainforests, scrub jungles, semiarid zones, and agroecosystems, according to extensive collection in Tamil Nadu (cotton and groundnut crops). Only at lower elevations, where the predators Camponotus compressus, Odentotermes obesus, carabid beetles, and Chrotogonus sp. are dominant, was the bug discovered. Acanthaspis quinquespinosa, Odentotermes obesus, Chrotogonus sp., and carabid beetles were found beneath stones in scrub jungles. On the bark of Tamarinadus indica, Acanthaspis quinquespinosa, and another reduviid, Acanthaspis siva, were discovered in semiarid regions (Sahayaraj 2007). Reduviid surveys in tropical rainforests show that there are 118 species overall, spread across 22 genera and 7 divisions. With 11 genera and 52 species, Euagorasaria predominated, followed by Harpactoraria (4 genera, 41 species), Polididusaria (4 genera, 9 species), Coranusaria (6 species in the sole representative genus Coranus), Sycanaria (1 genus, 5 species), Rhaphidosomaria (1 genus, 4 species), and Panthousaria (1 genus, 1 species) (Sivarama Krishnan 2009). In oil palm fields, Sycanus dichotomus are frequently seen at ground level (Ahmad et al. 2020).

4.2.7

Anthecoridae

Around the world, agricultural systems have significant natural adversaries that are predatory bugs from the family Anthocoridae. Anthocoris nemoralis is one species that was introduced to North America from elsewhere in the world (Horton et al. 2004). It developed a population and adapted to the surrounding environment. In Tetraphleps abdulghanii, Temnostethus pusillus, and Tetraphleps gracilis (Anthocorinae), there has been observation of embryonic diapause; in Lyctocoris campestris and some species of Xylocoris (Lyctocorinae), there has been observation of continuous development throughout all seasons (a homodynamic seasonal cycle) (Saulich and Musolin 2009). The scientists also noted that many temperatezone anthocorid species are multivoltine, producing numerous generations annually—up to eight in certain cases. As you move further north, the number of

4.2 Association with Tropics

105

generations normally drops to one every year. Temperature and day length play a major role in determining how multivoltine organisms evolve during the seasons. All multivoltine anthocorids of the temperate zone that have been investigated so far exhibit a long-day photoperiodic response, in which the females reproduce during long days but go into diapause during short days. The photoperiodic response rapidly deteriorates as one moves south; some populations, particularly at higher temperatures, do not enter diapause even under short-day conditions. Anthocorids’ ability to end diapause is poorly understood; however, some species need to be treated with low temperatures for a few weeks before oviposition can begin. Alary and colour polymorphism are uncommon in the family, and there is no evidence that they are affected by the seasons or the environment. Additionally, Anthocoridae has never been associated with pronounced seasonal migrations or aggregation behaviour. Only Tetraphleps abdulghanii has experienced summer diapause, which appears to be extremely unusual for the family. Another seasonal adaptation uncommon for Heteroptera is the seasonal shift of host plants, which is known in some populations of Anthocoris nemorum and Anthocoris nemoralis (Saulich and Musolin 2009). Hibernation occurs in Elatophilus nigricornis (Hemiptera, Anthocoridae) at the stage of the fertilised female. The females are then in a quiescent state so that when chosen, they can lay in the lab simultaneously (Fabre et al. 2000). Furthermore, Elatophilus nigricornis eggs are dispersed throughout the needles of maritime pine in their natural environment. The five stages of nymphs and adults are located in the cracks in the bark.

4.2.8

Dermaptera

The maritime or seaside earwig, Anisolabis maritima (Anisolabididae), is an insect that can be found globally along temperate and tropical coastal areas. Anisolabis maritima, a bug found under driftwood in large concentrations around the world, exhibited aggregation behaviour that was influenced by time, sex, body size, and shelter accessibility. Overall, the findings show that earwigs were more likely to cohabitate during the daylight (12 h) than during the night. Earwigs were still more likely to coexist with two shelters than one, regardless of the time (12 h vs. 24 h). As earwigs with two shelters at 12 h are more likely to cohabitate than those with one shelter at 24 h, there was a significant interaction between time and the number of shelters present. Time did not significantly affect sex relationships (Hack and Iyengar 2017). Cohabitation was different for men and women. Females were more likely to live with a male than a female, but they were less likely than males to live with conspecifics. It is interesting to see that men were equally likely to live with other men or women (Hack and Iyengar 2017). Additionally, authors noted that relative size has an impact on cohabitation since earwigs are less likely to live together if their sizes differ. Size did not affect time, shelter, or sex in any way. Larger individuals of both sexes were generally more likely to occupy the shelter at 24 h

106

4

General Ecology of Insect Predators

among different-sized intra-sexual pairings, and there were no sex differences in this size-based advantage (Hack and Iyengar 2017).

4.3

Intra- and Inter-Specific Relations

In light of predators and their prey, inter- and intra-tropical connections are more significant. Intra-specific predation, or the act of killing and devouring a member of the same species, is a serious and pervasive phenomenon that, until recently, did not receive the attention it rightfully deserved. Insect cannibalism is a typical occurrence. Odonata, Orthoptera, Thysanoptera, Hemiptera, Trichopteran, Lepidoptera, Diptera, Neuropteran, Coleoptera, and Hymenoptera all have extensive records of it. It is a significant element in the biology of many animals and may have an impact on behaviour, life history, competition for mates, and population structure. Adult females of Adalia bipunctata exhibited a stronger resistance to cannibalising eggs than males in the absence of Acyrthosiphon pisum. Cannibalism was more likely to occur in eggs, early larvae, and famished larvae than in older larvae or larvae that had been well-fed. Cannibalism of eggs and larvae was inversely correlated with Acyrthosiphon pisum abundance. In terms of larval growth and survival, eggs outperformed Acyrthosiphon pisum. Inter-specific predation between the coccinellids, Adalia bipunctata, Adalia decempunctata, Coccinella septempunctata, and Coccinella undecempunctata occurred in the absence of Acyrthosiphon pisum, but not equally. Coccinella septempunctata larvae were more likely to die after consuming a few eggs from Adalia bipunctata than vice versa, and both larvae and adults of the two species were reluctant to ingest conspecific eggs painted with a water extract of the other species’ eggs. The findings showed that cannibalism is adaptive in that it increases the likelihood of survival and happens most frequently when aphid prey is in short supply. Coccinellids have varied degrees of defence against predation by other species (Agarwala and Dixon 1992). The predator Nabis pseudoferus exhibits sit-and-wait behaviour and is a generalist. Since they consume both herbivores and other predators, which is known as “intra-guild predation”, generalist arthropod predators are often bitrophic (occupying the third and fourth levels simultaneously) (IGP). The lack of prey (both phytophagous and/or other predatory species) necessitates intra-specific predation or cannibalism because Nabis pseudoferus is a non-omnivore (FernandezMaldonado et al. 2017). The biological control agent Mallada basalis (Chrysopidae) was researched to determine the correlations between feeding density, starving time, food type and quantity, temperature, humidity, and photoperiod and larval cannibalism. The frequency of larval cannibalism rose when feeding density, temperature, or photoperiod increased, whereas it decreased as prey availability increased. When food (Corcyra cephalonica eggs, Lipaphis erysimi nymphs, or an artificial diet) was scarce, total cannibalism mortality was highest. When the feeding density was two larvae per petri dish, the overall mortality rate from cannibalism was lowest. Humidity and hunger duration had little effect on the overall mortality from cannibalism.

4.3 Intra- and Inter-Specific Relations

107

According to the findings, the four main variables that influenced cannibalism in Mallada basalis were feeding density, food volume, temperature, and photoperiod. For the mass rearing of Mallada basalis, these findings offer vital new information (Ye and Li 2020). These relationships have been studied using a variety of tools. Despite the variety of methods for studying the niche, Rahman et al. (2021) used carbon and nitrogen stable isotopes to examine changes in the nutritional niche of six ground beetle species (Carabidae) in response to disturbances from planned fires and grazing by reintroduced bison in 20 tallgrass plains. The use of serological tests or DNA analysis of gut content can help determine a person’s or a group’s preferred types of food under particular circumstances and over a particular time period. In multispecies communities, the relative dietary contributions of various food sources can be determined through the measurement of C and N stable isotopes. Stable isotopes are used less frequently in terrestrial insects than in other systems, and they are specifically used less frequently in agroecosystems. In recent years, trophic linkages and the ecological functions of a number of arthropods in crops have been studied using C13 and N15 studies. The heavy isotopes C13 and N15 of carbon and nitrogen have a tendency to be stored. While the process of C13 build-up is still unknown, N14 is eliminated. Stable isotope analysis thus makes use of variations in the ratio of heavy to light stable isotopes, such as those deposited in tissues (C13 or C12 and N15 or N14). As a result, the consumer’s tissues have an isotopic enrichment in response to its diet. N15 is utilised to determine the trophic position of the organism since it accumulates more quickly in every trophic transition than C13, whereas C13 reveals the primary basal food source of various species when its isotopic concentration changes (Morente and Ruano 2022). Three mirid predators, Dicyphus errans, Dicyphus bolivarirg, and Dicyphus cerastii, were studied for their competitive interactions with Nesidiocoris tenuis. The population expansion of Nesidiocoris tenuis. in the greenhouse was successfully suppressed by pre-establishment of hetero-specific mirid species; it was decreased by more than 90% when compared to plants without competing predators. In the absence of extra-guild prey, additional laboratory experiments showed reciprocal intra-guild predation between these species. In contrast to Macrolophus pygmaeus, Dicyphus adults preyed on Nesidiocoris tenuis. nymphs. Females of Nesidiocoris tenuis. preferred plants in olfactometer bioassays that had previously been exposed to heterospecifics but not conspecifics, indicating that this mirid does not shun competition. The three Dicyphus species may make attractive candidates for preventive releases in tomato crops, according to these findings, because of their potent ability to effectively control Nesidiocoris tenuis. population growth (Mouratidis et al. 2022). Filial cannibalism has been documented in a variety of fish species where male parents provide care, and it is often interpreted as a reaction to the high energy demands of brood defence and the reduced availability of food during the care period. Rhinocoris tristis, an assassin bug, was used as the subject of Thomas and Manica’s (2003) investigation into filial cannibalism in insects. In this species, males cannibalise some of the young in the brood while guarding the eggs of various

108

4

General Ecology of Insect Predators

females. In the field and the lab, we kept an eye on male guardians. The eggs closest to the edge of the brood, which were more likely to have been parasitised by wasps, were frequently eaten by males. However, in the absence of parasites, cannibalism continued in the lab. The quantity of cannibalised eggs was correlated with the duration of care and overall brood size, suggesting that males utilise eggs as a substitute food source. The fact that males in the field did not lose weight while guarding, while being unable to feed effectively, furthered this conclusion. In a lab experiment, males were also seen to adopt broods, although they did not consume more eggs from adopted than from their own broods.

4.4

Niche

Numerous studies show that the majority of predators are either general or specialised predators. The habitat-use of sympatric lady beetles varies at scales of landscapes, individual agricultural fields, and single plants, according to Snyder (2009). Additionally, seasonal and daily activity patterns vary among species. By encouraging various predator species to target various subgroups of the prey population, these spatiotemporal disparities in habitat utilisation should have complementary effects on prey. The author also emphasised that inter-specific predation of lady beetles was rarely observed to be a very disruptive factor, possibly because differences in spatiotemporal niches prevented inter-specific interactions. For predatory arthropods such as Hymenoptera’s Odontoponera denticulata, Solenopsis sp., Coleoptera’s Coccinella transversalis, Micraspis discolour, and Lophyra intermedia, soybean can also serve as a refugee by providing a niche for them (Anggraini et al. 2021). According to Sahayaraj (2014), the morphological and physiological adaptations of reduviid predators to predation are the reason for their success in every ecosystem and trophic niche. Heteropterans are capable of handling huge prey, and they can occupy predatory niches that are otherwise only accessible to much larger chewing arthropods (Cohen 1990). It was emphasised that two predators had not cohabited or shared a common food source, which would have altered morphological and physiological adaptations. Several studies demonstrate that reduviid predators favoured using stones as niches, which also improved the likelihood of copulation (Tomson et al. 2017). Research on the reduviid fauna in Tamil Nadu, India’s several districts strongly suggests that each species might be linked to a certain microhabitat. Most non-tibiaroliate subfamilies, including the Harpactorinae, Stenopodinae, Emesinae, Tribelocephalinae, Holoptilinae, and Saicinae, are characterised by alate forms that are invariably arboreal. These subfamilies’ alate species are mostly known to be endemic to tropical rainforests (Murugan 1988). Additionally, it was noted that among the tibiaroliate group, the alate species are predominately found in tropical rainforests, as are other species such as Pirates affinis, Haematorrhophus marginatus, Edocla spp., etc., where the males are alate and the females apterous. The majority of species that are commonplace, like Acanthaspis siva, Acanthaspis trimaculata, Acanthaspis rama, Acanthaspis tergemina, Acanthaspis sexguttata,

4.4 Niche

109

etc., are alate. It is interesting to note that, aside from a few stray species like the Sycanus, which is exclusive to Tropical Rainforests, many of the alate widespread species are represented in practically all the light trap samples made in the agroecosystems. Most frequently found in litter, the Salyavatinae prey on blattids and other litter fauna. However, apterous species are those that live underground, such the Nudiscutella frontispina. Questions concerning the processes allowing the coexistence of such species-rich assemblages are raised by the generalist feeding behaviour of the carabid beetle and its similar environmental needs. Compared to the daily routines, nutritional preferences, and plant climbing behavior of the two carabid species, Bembidion quadrimaculatum and Phyla obtusa, in a laboratory setting. Our findings point to temporal niche differentiation at the nychthemeron scale (a period of 24 consecutive h), with one species being more diurnal and the other more nocturnal, as well as spatial differentiation in their habitat use at the plant stratum scale, even though no obvious difference in trophic preference was observed. Intra-specific variation shows that behavioural plasticity in these two carabid species may be a mediator of micro-scale spatiotemporal niche divergence. We hypothesise that in intensively managed agricultural areas, carabid beetles may have a high potential for adaptation due to their behavioural plasticity. Therefore, even if the two species’ basic (possible) trophic niches are likely to overlap, their actual trophic niches are likely to be different, which will support the coexistence of the two species. On the one hand, species plasticity hypothesises that the extinction of a particular species may be offset by the expansion of the niches of the remaining species. On the other hand, species plasticity that permits niche differentiation may lead to a fine complementarity of species that would result in a positive link between biodiversity and overall community functional efficiency. Comparing species niche-widths in communities with different species richness is necessary to evaluate the functional flexibility of the community and to more precisely predict the plasticity of behavioural preferences (Kamenova et al. 2015). In a different study, González Macé et al. (2019) investigated intra-specific changes in the diet of two of the most prevalent predatory arthropods in grasslands, the ground beetle Harpalus rufipes and the wolf spider Trochosa ruricola, using stable isotope and fatty acid analyses. The findings demonstrate that the diet of Harpalus rufipes varied substantially with the richness of plant species, ingesting more plant material, possibly seeds, at plots with high plant species diversity and in the presence of grasses and tiny herbs. In contrast, Harpalus rufipes ingested more animal prey when there were legumes present, most likely aphids and/or collembolans (González Macé et al. 2019). As a result of disruptions, five out of six species showed expanded trophic niche area and breadth, pointing to a shift to a more general diet that included a greater variety of foods (Rahman et al. 2021).

110

4.5

4

General Ecology of Insect Predators

Future Areas

. Inter- and intra-glued predations of various predators should be undertaken for the better utilisation as biological control agent. . Ecology and biological trait can be given prime importance for invasive or commercial predators. . Since niche concept is poorly understood in most of the predators, entomologists should give priority to study them.

References Agarwala BK, Dixon AF (1992) Laboratory study of cannibalism and interspecific predation in ladybirds. Ecol Entomol 17(4):303–309 Ahmad SN, Masri MMM, Kamarudin N (2020) Assessment on voracity and predation by predator, Sycanus dichotomus STAL. (Hemiptera: Reduviidae) to oil palm bagworm, Pteroma pendula. J Oil Palm Res 32(3):480–487 Al-Ahmadi SS (2016) First record of Dielocroce chobauti (Neuroptera: Nemopteridae) from the Madina Region of Saudi Arabia, including ecological and morphological analyses. J Taibah Uni Sci 10(6):860–867 Alexandria W, Teresa H, Diane W (2019) Predatory hymenopteran assemblages in boreal Alaska: associations with forest composition and post-fire succession. EcoScicnce 26(3):205. https:// doi.org/10.1080/11956860.2018.1564484 Ambrose DP (1980) Bioecology ecophysiology and ethology of reduviids heteroptera of the scrub jungles of Tamil Nadu India. Bharathiar University, India, p 229 Ambrose DP (1987) Biological, behavioural and morphological tools in the biosystematics of Reduviidae (Insecta—Heteroptera—Reduviidae). Proc Anim Sci 96(5):499–508 Ambrose DP, Livingstone D (1979) On the bioecology of Lophocephala guerini Lap. (Reduviidae: Harpactorinae) a coprophagous reduviid from the Palghat Gap, India. J Nat Hist 13:581–588 Anggraini E, Irsan C, Gunawan B (2021) Phytophagous insects and predatory arthropods in soybean and zinnia. Biodiversitas 22(3):1405–1414 Aspöck H (1992) The Neuropteroidea of Europe: a review of present knowledge (Insecta). In: Canard M, Aspöck H, Mansell MW (eds) Current research in neuropterology. Proceedings of the Fourth International Symposium on Neuropterology, 24–27 June 1991, Bagnères-deLuchon, Haute-Garonne, France. Privately printed, Toulouse, pp 43–56. 414 pp Burgio G, Ferrari R, Pozzati M, Boriani L (2004) The role of ecological compensation areas on predator populations: an analysis on biodiversity and phenology of Coccinellidae (Coleoptera) on non-crop plants within hedgerows in Northern Italy. Bull Insectol 57:1–10 Burgio G, Ferrari R, Boriani L, Pozzati M, van Lenteren J (2006) The role of ecological infrastructures on Coccinellidae (Coleoptera) and other predators in weedy field margins within northern Italy agroecosystems. Bull Insectol 59(1):59 Busch AK, Wham BE, Tooker JF (2021) Life history, biology, and distribution of Pterostichus melanarius (Coleoptera: Carabidae) in North America. Environ Entomol 50(6):1257–1266 Carayon J (1961) Quelques remarques sur les Hemipteres-Heteropteres: leur importance comme insectes auxiliaires et les possibilites de leur utilisation dans la lutte biologique. Entomophaga 6: 133–141 Cividanes FJ (2021) Carabid beetles (Coleoptera: Carabidae) and biological control of agricultural pests in Latin America. Ann Entomol Soc Am 114(2):175–191 Cohen AC (1990) Feeding adaptations of some predaceous Hemiptera. Ann Entomol Soc Am 83(6):1215–1223

References

111

Ehler LE, Miller JC (1978) Biological control in temporary agroecosystems. Entomophaga 23:207– 201 Fabre JP, Menassieu P, Foing JJ, Chalon A (2000) Biology and ecology of Elatophilus nigricornis Zetterstedt (Hemiptera Anthocoridae) predator of Matsucoccus feytaudi Ducasse (Homoptera Matsucoccidae) in the South-East of France. Ann For Sci 57(8):777–792 Fernandes FL (2012) Ecology of Vespidae (Hymenoptera) predators in Coffea arabica plantations. Sociobiology 59(4):1269–1280 Fernandez-Maldonado FJ, Gallego JR, Valencia A, Gamez M, Varga Z, Garay J, Cabello T (2017) Cannibalism: do risks of fighting and reprisal reduce predatory rates? Commun Ecol 18(1): 87–96 Frizzi F, Masoni A, Santedicola M, Servini M, Simoncini N, Palmieri J, Santini G (2022) Intraspecific relationships and nest mound shape are affected by habitat features in introduced populations of the red wood ant Formica paralugubris. Insects 13(2):198 González Macé O, Ebeling A, Eisenhauer N, Cesarz S, Scheu S (2019) Variations in trophic niches of generalist predators with plant community composition as indicated by stable isotopes and fatty acids. Soil Organ 91(2):45–59. https://doi.org/10.25674/so91204 Hack NL, Iyengar VK (2017) Big wigs and small wigs: time, sex, size and shelter affect cohabitation in the maritime earwig (Anisolabis maritima). PLoS One 12(10):e0185754. https://doi.org/ 10.1371/journal.pone.0185754 Hammond RB, Stinner BR (1987) Soybean foliage insects in conservation tillage systems: effects of tillage, previous cropping history, and soil insecticide application. Environ Entomol 16:524– 531 Horton DR, Lewis TM, Broers DA (2004) Ecological and geographic range expansion of the introduced predator Anthocoris nemoralis (Heteroptera: Anthocoridae) in North America: potential for nontarget effects? Am Entomol 50(1):18–30 Irwin ME, Shepard M (1980) Sampling predaceous Hemiptera on soybean. In: Kogan M, Herzog DC (eds) Sampling methods in soybean entomology. Springer, New York, Heidelberg, Berlin, pp 505–531. 587 pp Kamenova S, Tougeron K, Cateine M, Marie A, Plantegenest M (2015) Behaviour-driven microscale niche differentiation in carabid beetles. Entomol Experiment Appl 155(1):39–46 Kerzhner M (1986) Miridae (Heteroptera). I: a new genus, some new species, and notes on synonymy. J N Y Entomol Soc 94:180–193 Killington FJ (1936) A monograph of the British Neuroptera. Ray Society Koschel H (1971) Zur Kenntnis der Raubwanze Himacerus apterus F. (Heteroptera, Nabidae). Teil I. II. Z Angew Entomol 68:1–24. 113–37 Messina FJ (1982) Comparative biology of the goldenrod leaf beetles. Trirhabda virgata Leconte and T. borealis Blake (Coleoptera: Chrysomelidae). Coleopt Bull 36(2):255–269 Milanez JM (1984) Ocorrência de artrOpondos em urn sistema de consOrcio feijdo-milho, comparado aos respectivos monocultivos. Resumos IX Congr. Bras. Entomol. Londrina. Brasil. Soc. Entomol. Bras, London, p 38 Morente M, Ruano F (2022) Understanding the trophic relationships amongst arthropods in olive grove by δN15 and δC13 stable isotope analysis. J Appl Entomol 146:372. https://doi.org/10. 1111/jen.12986 Mouratidis A, Leman A, Poelman EH (2022) Dicyphus predatory bugs pre-established on tomato plants reduce Nesidiocoris tenuis population growth. J Pest Sci 95:1659–1670. https://doi.org/ 10.1007/s10340-022-01482-8 Murugan C (1988) Biosystematics and ecophysiology of the tibiaroliate assassin bugs heteroptera reduviidae of Southern India. Ph.D. thesis, Bharathiar University, p 279 Myers JG (1925) Biological notes on Arachnocoris albotnaculatus Scott (Hemiptera: Nabidae). J N Y Entomol Soc 33:136–145 Ohl M (2004) A new wasp-mimicing [sic] species of the genus Euclimacia from Thailand (Neuroptera: Mantispidae). Denisia 13:193–196

112

4

General Ecology of Insect Predators

Ohl M (2011) Aboard a spider--a complex developmental strategy fossilized in amber. Naturwissenschaften 98:453 Pericart J (1987) Hémiptères Nabidae d’Europe occidentale et du Maghreb. Faune de France, vol 71. Fédération Française des Sociétés de Sciences Naturelles, Paris. 185 pp Rahman AU, Jones HP, Hosler SC, Geddes S, Nelson M, Barber NA (2021) Disturbance-induced trophic niche shifts in ground beetles (Coleoptera: Carabidae) in restored grasslands. Environ Entomol 50(5):1075–1087 Rodríguez-Vélez JM, Uribe-Mu CA, Sarmiento-Cordero MA, Rodríguez-Vélez B, Cruz-Rodríguez JA, Contreras-Ramos A, Huerta-Martínez FM, Arredondo-Bernal HC (2021) Relationships among Melanaphis sacchari, its predator insect communities, and weather conditions in sorghum crops at Colima, Mexico. Southwest Entomol 46(1):47–58. https://doi.org/10.3958/059. 046.0104 Roy H, Migeon A (2010) Ladybeetles (Coccinellidae). Chapter 8.4. BioRisk 4:293 Sahayaraj K (2007) Ecotypic variation in the biology of Acanthaspis quinquespinosa Fabricius 1781 (Hemiptera: Reduviidae: Reduviinae) from peninsular India. Egypt J Biol 9:53–59 Sahayaraj K (2014) Reduviids and their merits in biological control. In Basic and applied aspects of biopesticides. Springer, New Delhi, pp 195–214 Saulich AK, Musolin DL (2009) Seasonal development and ecology of anthocorids (Heteroptera, Anthocoridae). Entomol Rev 89(5):501–528 Shah SIA, Ahmad A, Saeed R, Cai W (2022) Redescription of a rare assassin bug Lophocephala guerini Laporte (Hemiptera: Heteroptera: Reduviidae: Harpactorinae) from Pakistan. Heteroptera Poloniae Acta Faunistica 16:73–82 Sivarama Krishnan S (2009) Biosystematics of chosen harpactorine assassin bugs Insecta Hemiptera Reduviidae. Ph.D. thesis. MS University, Tiruenveli, India, p 119 Sloggett JJ, Zeilstra I (2020) Geographic variation in the habitat preference of a scarce predatory insect: evolutionary and conservation perspectives. Ecol Entomol 45(3):386–395 Snyder WE (2009) Coccinellids in diverse communities: which niche fits? Biol Control 51(2): 323–335 Souma J, Nozaki T, Otsui K, Ishikawa T (2022) New records of the damsel bug Alloeorhynchus reinhardi Kerzhner & Günther, 1999 (Hemiptera, Heteroptera, Nabidae) from Japan. Check List 18(2):261–264 Southwood TRE (1977a) Habitat, the templet for ecological strategies? J Anim Ecol 46:337–365 Southwood TRE (1977b) The relevance of population dynamic theory to pest status. In Origins of pest, parasite, disease and weed problems. 18th Svmp. Br. Ecol. Soc. Bangor, Wales, 1976. Blackwell Science, Oxford, pp 35–54 Stoner A (1973) Incidence of Wesmaelia pendula (Hymenoptera: Braconidae). A parasite of male Nabis’s species in Arizona. Ann Entomol Soc Am 66(2):471–473 Sunil J, Chandish RB (2013) Syrphid predators for biological control of aphids. J Biol Control 27(3):151–170 Thomas LK, Manica A (2003) Filial cannibalism in an assassin bug. Anim Behav 66(2):205–210 Tomson M, Sahayaraj K, Kumar V, Avery PB, McKenzie CL, Osborne LS (2017) Mass rearing and augmentative biological control evaluation of Rhynocoris fuscipes (Hemiptera: Reduviidae) against multiple pests of cotton. Pest Manag Sci 73(8):1743–1752 Vennison SJ, Ambrose DP (1991) Biology and behavior of Ectomocoris xavierei Vennison & Ambrose 1990 (Heteroptera Reduviidae Piratinae). Trop Zool 4(2):251–258 Veríssimo BA, Auad AM, Oliveira CM, Paiva IG (2021) Seasonality of predatory insects (Diptera: Syrphidae and Asilidae) in pasture monoculture and silvopastoral systems from Southeast Brazil. Int J Trop Insect Sci 41(1):861–872 Ye J, Li J (2020) Factors affecting cannibalism by Mallada basalis. Biocont Sci Technol 30(5): 442–450 Withycombe CL (1923) Notes on the biology of some British Neuroptera (Planipennia). Tran Royal Entomol Soc London 70(3–4):501–594 Withycombe CL (1925) Some aspects of the biology and morphology of the Neuroptera with special reference to the immature stages and their possible phylogenetic significance. Tran Royal Entomol Soc London 72(3–4):303–411

5

Egg Biology of Insect Predators

Contents 5.1 General Structure of Insect Eggs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Incubation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3 Description of Eggs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.1 Mantidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.2 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.3 Coccinellids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.4 Carabide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.5 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.6 Neuropteran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.7 Reduviid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.8 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.9 Anthocorids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.10 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.11 Thysanoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4 Egg-Laying Pattern and Protecting Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.1 Plant Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.2 Volatile and Non-volatile Organic Compounds (VOCs) . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.3 Protecting Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.4 Plants and Their Phenology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.5 Other Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.6 Inter-, Intra-guild Predation and Cannibalism (IGP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5 Egg Dumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6 Future Focus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.1

113 114 115 115 116 117 117 119 120 120 123 124 124 124 125 125 125 128 131 134 139 141 141 142

General Structure of Insect Eggs

Life begins as an autonomous egg in the majority of insects. Ovarian reproduction is the term for this process. Each egg is created within the female’s genital system and subsequently expelled from her body through an ovipositor, a part of her external

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_5

113

114

5 Egg Biology of Insect Predators

genitalia that resembles a tube, saw, or blade. The female’s body produces eggs through a process known as öogenesis, and the egg-laying process is known as oviposition. Each insect species produces eggs that are spherical, ovate, conical, sausage-shaped, barrel-shaped, or torpedo-shaped, as well as genetically distinct and frequently physically unusual. In spite of this, each egg only contains one living cell, the female gamete, regardless of its size or shape. The vitelline membrane is the name of the cell membrane of an egg. Its phospholipid bilayer structure is comparable to that of the majority of other animal membranes. It encircles the entire egg cell’s contents, the majority of which is made up of yolk (food for the soon-to-develop embryo). The cytoplasm of the cell is often distributed in two locations: in a thin band just within the vitelline membrane (where it is known as periplasm) and in diffuse strands that run throughout the yolk (cytoplasmic reticulum). The haploid nucleus of the egg cell is located inside the yolk, typically near one end of the egg. The öosome (a region of increased optical density) may be seen as a black area in the more translucent yolk near the opposite end. The relationship between the nucleus and the öosome determines the anterior/ posterior polarity of the egg. In the majority of insects, auxiliary glands in the female’s reproductive system create a protective protein “shell” around the egg before oviposition. This egg shell, known as the chorion, is frequently decorated with microscopic ridges or grooves that could only be apparent under an electron microscope’s extreme magnification. Aeropyles, which are minute pores, allow for the exchange of oxygen and carbon dioxide during respiration while losing only a small amount of water. During fertilisation, sperm enters through a unique aperture called the micropyle, which is located close to the anterior end of the chorion. Micropyles According to Iossa (2022), micropyles in insect eggs are gaps in the chorion that sperm can enter to fertilise the egg. Micropyles are a variety of structures that exhibit striking differences in number, spatial configuration, and physical structure among extant insect groups. Although they are fairly universal among insects, they have not gotten much attention. It is conceivable that some of the variability displayed by micropyles is adaptive, supporting other egg structures throughout embryo development as they are essential morphological characteristics of an immobile life stage. Consequently, even though micropyles’ main job is to fertilise eggs, they may also contribute to embryo development and be influenced by both sexual and natural selection.

5.2

Incubation

Because it directly affects offspring survival and fitness, oviposition behaviour is key to studies of insect population dynamics. The number of eggs laid in a patch, how egg and cluster sizes are determined, and how the ovipositing female assesses batch quality are crucial elements of oviposition. It is especially important for insects whose newly hatched progeny are unable to travel long distances that they have the

5.3 Description of Eggs

115

ability to assess quality. Such progeny are restricted to flowering plants that grow underground and offer nectar and pollen as a food supply for the female’s chosen offspring. As a result, a key element in the evolution of oviposition behaviour is the relationship between female oviposition preference and larval performance. These insects must consume in order for consecutive batches of eggs to mature (Stearns 2000; Heimpel and Rosenheim 1998). Fitness is based on the pattern of resource distribution to somatic and immediate reproduction, which determines their survival and future reproduction. This pattern of allocation in many of these synovigenic insects is controlled by ovarian dynamics (Papaj 2000). When pregnant females repeatedly come across low-quality food sources or no oviposition sites, they stop oogenesis and reabsorb oocytes to continue looking for high-quality food sources. When favourable circumstances arise, egg production is restored. Insects that feed on transient resources can adjust to spatial and temporal variations in resource availability thanks to these two opposing processes, both of which occur very quickly. It may be possible to explain differences in predator abundance among crops by considering the predators’ ovipositional preferences for particular crops.

5.3

Description of Eggs

5.3.1

Mantidae

The Harpagides produce a long, narrow egg case that is often cream in colour and adheres to stems and twigs that are more or less horizontal. The eggs are covered in a very thin layer of froth and arranged symmetrically along a central axis. On the outside, it is glossy and smooth. Theopropus elegans and Humenopus bicornis are both dedicated moms; a captivity of the former used to remain constantly above its egg-case. And twice I took the latter right next to her eggs (Shelford 1903). The ootheca, or egg case, is a complex construction that female praying mantises form during oviposition to support and protect eggs from ambient conditions and natural adversaries, according to Brannoch et al. (2017). The frothy secretions of the female vaginal accessory glands, which eventually harden when in contact with the air system, are what make up the ootheca itself. Mantodean oothecas stand out from other dictyopterans due to their substantial architectural and cryptic variation as well as variation in the mechanical qualities of their constituent parts, which are primarily calcium- and protein-based substances. At the chemical, microscopic, and macroscopic structural levels, mantodean ootheca continue to be largely unexplored. Acanthopoidea, the polymorphic earless praying mantises, served as a recent example of how ootheca characteristics could be used to distinguish higher-level taxa. Galepsus (=Lygdamia) lenticularis oothecae were typically found clinging to flat substrates like long-stemmed grasses or sticks when they were gathered in the field. As shown by Stagmatoptera supplicaria, none of the species in the Mantidae family have oothecae that are oval or covered with the typical foamy sheath (Brannoch et al. 2017). The dorsally flattened, oblong Galepsus lenticularis oothecae are typically light to dark brown in hue. The white eclosion sack-like structures on the greyish,

116

5 Egg Biology of Insect Predators

dorsally flattened area of emergence are a telltale sign of hatched oothecae. The ootheca parameters were measured from the first egg chamber to the last egg chamber and did not include the residual process. The region of emergence, which did not include the residual process, was quantified as the length of the ootheca. Othecae were dorsally cut along their length and examined under a microscope to count the number of eggs per ootheca. Dissection of the residual process revealed that there were no egg chambers. The eggs were set in rectangular rows, each with two or three eggs, next to one another. The 18 female Galepsus lenticularis captivereared and 9 field-collected individuals produced a total of 42 oothecae. Nine of the 42 oothecae generated by the 9 field-collected females—which were mated with field-collected males in captivity—were fertilised and hatched. The 19 oothecae that the field-collected females produced never hatched (unhatched). In the terrariums, 18 unpaired females lay 14 unfertilised oothecae; no nymphs developed from these unfertilised oothecae. The possibility of parthenogenesis, which has been documented in other mantid species like Coptopteryx viridis (Coptopterygidea) (Cukier et al. 1979), Miomantis paykulli (Adair 1924), and the Springbok mantis, Miomantis caffra (Mantidae), was also investigated, so breeding with the captivereared females was not done (Walker and Holwell 2016). Nine out of the 18 captivereared females that were not mated gave birth to oothecae at some point in their lives. Two unfertilised oothecae were laid by each of these five females. In the instance of unfertilised ootheca, the pre-oviposition phase lasted 53 days (mean female age of 214 days). The two unfertilised oothecae laid their eggs 20 days apart (mean female age of 235 days). A female could live for 50 days after depositing her last unfertilised ootheca (mean female age of 285 days). The oothecae had a length that varied from 18.9 to 30.0 mm. The quantity of eggs per ootheca varies among the various oothecae species. Hatched and unhatched oothecae had an average of 50.2 and 59.2 eggs per ootheca, compared to 36.6 eggs in unfertilised oothecae (Greyvenstein et al. 2020).

5.3.2

Hymenoptera

According to Lomholdt (1984), female Lindenius feed on minute insects from the orders Diptera, Hemiptera, and Hymenoptera. The egg is laid on one of the first prey creatures to enter the cell and is positioned on the ventral side of the prey beneath its head. In 1–5 days, the larva consumes its entire prey. The carcasses of the prey are spread across the cocoon. Adult digger wasps visit a variety of flowering plants, including those in the Apiaceae, Lamiaceae, Asteraceae, Euphorbiaceae, and many more families (Kazenas 2001). The white bean-shaped eggs of the digger wasp species Ampulex compressa ranged in length from 2.2 to 3.0 mm, in width from 0.66 to 0.72 mm, and in weight from 345 to 832 μg (Gnatzy et al. 2018).

5.3 Description of Eggs

5.3.3

117

Coccinellids

Beginning in the spring or early summer, female lady beetles can lay anywhere between 20 and 1000 eggs during a 1- to 3-month period. Eggs are typically laid close to their food, such as aphids, and are frequently found in little clusters in safe places on leaves and stems. Numerous lady beetle species lay spindle-shaped, tiny (approximately 1 mm), cream, yellow, or orange eggs that are perched vertically on leaf surfaces or bark surfaces. They resemble the Colorado potato beetle and Mexican bean beetle, but are often smaller. It is crucial to set up a substrate made up of three to four crumpled white or brown paper towels when working in a lab. The emergence from the chorion occurs either via a rupture at one end of the egg, such as in the case of yellow to red, oval or spindle-shaped eggs (e.g. Chlocorus species) (Ahmad 1970.) Eggs deposited by the Scymnus (Neopullus) sinuanodulus hatch in about 10 days after being laid in the spring. Eggs are oval shape with dimensions of 0.57 × 6 × 0.06 mm on average and 0.27 and 0.02 mm on average. When young they are orange, as it grew older, it is duller. Before hatching, it frequently has an iridescent sheen (produced by the chorion being separated from the embryo, creating an air space). Chorion is naked, covered in a membranous layer, and has shallow depressions that resemble dimples on its exposed surface but not on its surface that is linked to the substrate. Four to 11 (usually 7–9) cups placed centrally in an irregular ring, each encircled by 15–20 semi-circular walls, and projecting in the shape of tubes on the exposed apical pole. Chorion is clearly white and frequently iridescent under light after hatching (Lu et al. 2002). Both Propylea quatuordecimpunctata and Harmonia axyridis (Coleoptera: Coccinellidae) species favoured a hydrophilic surface as the oviposition site, which can be associated with their superior performance as larvae and adults on these substrates as opposed to hydrophobic ones. Both species of ladybirds’ egg glue has the ability to wet hydrophobic surfaces, including those on many plant leaves and surfaces heavily covered in 3D wax. In Propylea quatuordecimpunctata, but not in Harmonia axyridis, the choice of the oviposition site is significantly influenced by the surface roughness. The preference for smooth surfaces for oviposition in Propylea quatuordecimpunctata may be explained by the superior performance of larvae on smooth platforms than rough ones. With the exception of very high asperity sizes or large trichomes, the egg glue of both species can adapt to artificial and natural surfaces with a variety of asperity sizes while faithfully mimicking their structure (Salerno et al. 2022).

5.3.4

Carabide

Female carabide will lay between 30 and 600 oval eggs individually at a suitable location, either beneath the soil surface or in a layer of plant debris. Because immature larvae have less mobility to acquire food and because of their comparatively fragile bodies, predators are particularly attracted to them, protected egg

118

5 Egg Biology of Insect Predators

locations are crucial. In some species that have tiny litter sizes, parental care has been seen in the form of guarding the eggs and hiding seeds. Between 4 August and 1 September of 2008, 16 eggs were collected. The Carabus (=Oreocarabus) ghilianii female placed each egg independently and buried it in the substrate. Egg production per female ranged from 1 to 9. Six of the eggs did not hatch and began to form fungus, while one was immediately conserved in Scheerpeltz for the research of chorion. As a result, they were preserved in Scheerpeltz so that researchers could study them and nine of them finally hatchet. Their embryonic development took them between 10 and 17 days to complete. As an egg develops, its size grows, going from 3.5 mm in length and 2.5 mm in width in the early stages to 5 and 2.7 mm soon before hatching. The egg colour varies slightly as well; it begins as a greyish-white ivory tone at the time of laying and gradually turns more cream till hatching. A rough estimate of the density of chorion reticulation is 10.5 cells per 0.1 mm. The average cell length is 8.8 m, while the average width of ridges is 2 μm. Common species found on Honshu Island, Japan, include Harpalus griseus, Harpalus eous, Harpalus tridens, Synuchus cycloderus, and Carabus procerulus. In the three species of Harpalus, those caught between June and August had immature ovaries and few, developed eggs after September. The eggs were between 2.1 and 2.4 mm in size, with a 0.17–0.20 egg-to-body length ratio. Synuchus cycloderus individuals collected between May and July had immature ovaries, whereas those captured in the middle of October and later had developed eggs. No one was apprehended in August or September. There were 75.9 mature eggs on average. The size of the eggs and their proportion to body length were reduced. Carabus procerulus had immature ovaries from June to July, whereas individuals caught in September and later had eggs. No one was apprehended in August. The size of the eggs was very large despite the low average quantity of ovarian eggs. In comparison to the other four species, Synuchus cycloderus mature eggs were less than half the size and had an average number that was more than ten times higher. With Synuchus cycloderus showing the “many-small egg type” and the other four species being the “few-large egg type”, respectively (Shibuya et al. 2020). The quantity and size of eggs are influenced by parental care for kids (Ito 1959). Some ground beetles have also been discovered to provide for their parental brood. For instance, Pterostichini guard their eggs after oviposition and dig a place for egg deposition (Thiele 1977). One egg is laid by Carabus procerulus (few-large egg) in a gap created by the tip of the abdomen (Sota 2000). Consequently, we must consider both egg kinds and parental care when examining the reproductive strategy of ground beetles. Kolesnikov and Karamyan (2019) claim that Pterostichus anthracinus engaged in reproductive activity in the lab for approximately 4 months, from 10 May to 30 August. During the breeding season, females constructed two different types of nests for their egg clutches. The first variety was built beneath chunks of stone or bark and had a 20-mm-diameter opening at the top. It was pyramidal in shape. The second type of nest was buried between 20 and 30 mm beneath the surface of the ground and lacked a hole at the top. The egg chamber had a 20 mm diameter. The

5.3 Description of Eggs

119

females of Pterostichus anthracinus produced 132 clutches of eggs in the laboratory from 2015 to 2018. The majority of egg clutches were laid in June. Each clutch contained between 17 and 46 eggs, with a diameter of about 5–7 m and a mean SE of 25.25. A female kept watch over each of these clutches of eggs, either sitting on the eggs or loitering nearby. Nearly 40% of the females had multiple children throughout the mating season. Pterostichus anthracinus females typically produced 2 clutches, 9.2% produced 3, and 2.3% either 4 or 5. Embryos developed from eggs in a period of 4–8 days. The interval between clutches, when a female laid more than one, varied from 5 to 32 days (mean = 14.53). The autumn-breeding Pterostichus melanarius develops irrespective of photoperiod. Mid-May to June sees the emergence of new adults, while mid-July to August sees a peak in population due to mating season. Egg production is limited by temperature, but by early August, females are pregnant. Mid-July to September is when ovulation takes place. In total, females typically lay 130 eggs, 2–12 of which are laid parallel to one another. Larger eggs provide more resources and increase larvae survival, therefore there is a trade-off between egg size and egg number. Due to their inability to travel great distances in search of supplies, carabid larvae must carefully choose their oviposition sites. No matter if the setting is artificial or natural, female Pterostichus melanarius prefer to lay her eggs there. More than 70% of the time in oviposition choice studies, females favoured locations with more man-made structures or vegetation. For oviposition, females avoided full light and favoured a damp substrate. Egg-laying preferences for moist soil and shade may be a reflection of the larvae’s and eggs’ sensitivity to desiccation. At the time of oviposition, eggs are translucent white and range in size from 1.1 to 1.4 mm; however, before hatching, they darken (Busch et al. 2021).

5.3.5

Syrphids

There is not a lot of research about oviposition site, egg deposition, egg structures, and hatching in the literature. Typically placed singly on the underside of leaves supporting aphid colonies are the white syrphid eggs (1.0 × 0.5 mm). Generally, eggs are 1/25 in. (1 mm) or less in length, oblong, slightly curved, grey to white, and of any colour. The top and sides are convex, but the underside is flattened. Frequently, eggs are found alone next to nourishment for the developing larvae. Some species’ eggs’ identities can be identified by the surface sculpting (Werner and Chandler 1995). Eggs are 1 mm or less in length, oblong or elongate oval-shaped, grey to white, and slightly curved. The top and sides are convex, but the underside is flattened. Frequently, eggs are found alone next to nourishment for the developing larvae. Syrphid larvae have different appearances depending on the habitat in which they eat. Around aphid colonies, syrphid fly eggs are frequently discovered, providing the larvae with an immediate source of food. They have a sculpting design on the chorion. Parallel, thin white strips make up the sculpture. Instead of continuing, these stripes are broken up by depressions that seem to be made up of numerous tiny white longitudinal patches. The stripes are raised above the ground.

120

5 Egg Biology of Insect Predators

Such elevations are numerous, parallel, and completely cover the area (Joshi and Ballal 2013). Fecundity typically increases during the first week of egg laying, peaks, and then gradually declines until death. The females survived for up to 6 weeks, and the hatching rate for the eggs was between 80% and 100%.

5.3.6

Neuropteran

Eggs are typically elongated oval, but less frequently nearly spherical. Eggs that have just been laid are a bright green colour that is almost whitish in hue. A raised knob or plate is typically used to indicate the clearly defined micropyle. Myrmeleontidae and Ascalaphidae eggs appear to have a micropylar plate at both ends; however, this is unusual for the order. Some osmylids and nemopterids have pronounced papillae, and the chorion typically exhibits at least traces of areolate, pentagonal, or hexagonal sculpturing (New 2003). The egg’s apical pole is where the micropylar region is located. It consists of a circular area with 30 semi-circular channels, each of which corresponds to a micropylar orifice (Mazzini 1976). The Chrysoperla zastrowi sillemi’s round, light green eggs are placed singly at the end of long, silky stalks, and turn grey as they approach hatching.

5.3.7

Reduviid

Stride (1956) described the eggs of Phonoctonus fasciatus, Phonoctonus subimpictus, and Phonoctonus lutescens. He asserts that the eggs of Phonoctonus fasciatus and Phonoctonus subimpictus are elongated, curved, and basally considerably inflated. The chorion has numerous shallow depressions that are submerged, but they are partially hidden by a covering that resembles varnish that covers the egg’s surface and adheres it to the substrate. This varnish turns mucilaginous and bloated when moistened. The egg’s upper pole has an operculum and three rows of chorionic processes (Stride 1956). But compared to the other two species, Phonoctonus lutescens has larger eggs. Phonoctonus lutescens lack any darker markings and are a homogeneous pale amber colour. The extensive net-like chorionic collar extensions and opercular outgrowths that distinguish harpactorine eggs from the eggs of other reduviid subfamilies help to identify these clustered, spumaline-coated eggs. The collar extensions can be rather lengthy with a honeycomb structure that ends in filamentous apices, or they can be quite short with a spongy structure. The egg’s body is clearly defined by follicular growth patterns, and its anterior end is distinguished by a distinctive collar that contains the spermatic groove, aero-micropylar apertures, and a sealing bar. Aeromicropylar systems, the inner aerostatic layer, the chorionic collar extensions, and opercular outgrowths work together to effectively deliver atmospheric oxygen to the interior of the eggs (Haridass 1986). Triatomines and reduviines lay the next most eggs, followed by harpactorines. The fewest eggs are laid by stenopodines (George 1988). Rihirbus trochantericus produced eggs both singly and in clusters, with each

5.3 Description of Eggs

121

female producing 26 eggs in 3–7 clusters over the course of a 13-day incubation period (Bhat et al. 2013). The highest amount of spumaline has been observed to be exuded by Endochus species. More spumaline is secreted by Endochus cinqalensis than by Endochus inornatus. Each batch of eggs is deposited in the dark brown spumaline in nearly a triangle pattern by Brassivola hystrix and Sycanus species, which have also been observed to secrete rather substantial amounts of spumaline that covers practically the whole body of the egg. Both Sycanus species, S. versicolor and Sycanus collaris, lay eggs that are cylindrical and longer in shape. The spumaline covering extends almost to the tip of the chorionic collar, and the eggs are arranged in parallel rows vertically. In comparison to Brassivola hystrix, Sycanus spp. lay a lot more eggs per batch. The egg of Platerus pilcheri is a little bit elongate, the spumaline is quite lavishly secreted, covering the eggs almost up to the chorionic collar, and the eggs are oriented in a linear pattern in each batch, similar to that of Sycanus spp. (George 1988). All three Rhynocoris species (Rhynocoris marqinatus, Rhynocoris kumarii, and Rhynocoris fuscipes) deposit their eggs in a vertical, linear pattern. Each egg is adhered to the other and to the substrate by a moderate quantity of spumaline, and they all deposit more than 100 eggs per batch. The oviposition techniques of Irantha armipes and a Coranus species are comparable in that the spumaline is dispersed over the substrate, typically on leaves, and the amount of spumaline secreted is adequate to secure the eggs to the leaves with some spillover. The eggs are longitudinally less elongate in Neohaematorophus therasii, Coranus atricapillus, and Coranus wolffi than in the other species mentioned above. Although the eggs of Cydnocoris qilvus are also ovate in shape, they are arranged in the shape of a florette, with a central egg deposited erect and 5–6 eggs surrounding it that are slantingly arranged with their cephalic ends facing outward. Each egg’s bottom and lateral sides are where the sparse spumaline deposit is located, pointing towards the basal region (George 1988). Harpactor pyqmaeus and Harpactor nilqiriensis both lay eggs that are tenuously attached to leaves by their caudal end and have very little spumaline. As previously mentioned, the eggs of Endochus sp. and Euaqoras plaqiatus deviate from the typical cylindrical shape found among the eggs of the vertical deposition pattern. The primary axis is aligned with the convexity thanks to the oblique flattening of the egg’s bottom in this instance. The eggs are occasionally discontinuous and are found linearly adhered to the leaf blades by their bottoms. Although there is very little spumaline, it is enough to stick the egg. The egg of Euaqoras plaqiatus is quite similar to the egg of Macracanthopsis nodipes, with a few minor differences, the Coranus spiniscutis egg. With the exception of their cephalic ends, Coranus obscurus and Rhapftidosoma atkinsoni are more cylindrically elongate, horizontally bonded to the substratum, and grouped in a linear pattern. The eggs of all six species of Emesinae that have been researched demonstrate that each species’ spumaline is deposited in varied amounts, giving each species’ eggs a distinctive surface pattern (George 1988).

122

5 Egg Biology of Insect Predators

Fig. 5.1 (a) Cydnocoris gilvus, (b) E. bicolor, (c) Irantha armipes, (d) Endochus sp., (e) R. trochantericus, and (f) S. signatus eggs

Only members of 9 out of the 24 subfamilies have data on the morphology of nymphs and/or eggs, as reported by Bugaj-Nawrocka et al. (2022). Many are only represented by a single report, though: Eggs of Neohaematorrhophus therasii, and Ectrichodiinae. Other examples include Emesinae, Apiomerus crassipes, Arilus gallus, Endochus migratorius, E. umbrinus, Irantha armipes, Manicocoris rufipes, Pselliopus barber, Repipta, Rhynocoris kumarii, Sinea complexa, S. spinipes, Sphedanolestes minusculus, Sycanus pyrrhomelas, Dipetalogaster maximum, Linshcosteus karupus, and Panstrongylus geniculatus (Bugaj-Nawrocka et al. 2022). The number of eggs laid by ectrichodines, peiratines, and salyavatines is average (Ambrose and Ganesh Kumar 2016). They were also informed of the following conclusion: A few species of reduviids’ oviposition behaviour, as well as the description and variety of their eggs as well as some reduviids’ typical oviposition pattern were documented (Harpactorinae). In reduviids, there are five main forms of oviposition (Fig. 5.1): 1. 2. 3. 4.

Each egg is separately cemented to the substratum (Holoptilinae). Eggs are buried deeply in the soil (Peiratinae and Stenopoainae). Eggs are loosely scattered around erratically without any pattern (Ectrichodinae). Eggs are laid in clusters and cemented to each other and the substratum (Harpactorinae).

5.3 Description of Eggs

123

5. Eggs clusters but not longitudinally cemented to each other but rather cemented to the substratum (Reduviinae, Salyavainae, and Triatominae). Sphedanolestes variabilis laid both solitary light brown eggs and tiny clusters of five to ten eggs. Sphedanolestes variabilis laid brown eggs that were cement-fixed to the substrate basally. They also reported the presence of Sphedanolestes minusculus (63.33 eggs/female), Sphedanolestes pubinotum (54.0 eggs/female), Sphedanolestes himalayensis (74.8 eggs/female), Sphedanolestes variabilis (34.33 eggs/female), and Sphedanolestes signatus (15.33 eggs/female) (Ambrose et al. 2009). Rihirbus trochantericus laid 67.50 eggs per female and produced elongate, spherical, dark brown eggs with a flower-like opercular morphology that were attached basally to the substrate (Bhat et al. 2013). To keep the eggs from drying out, a Sycanus annulicornis egg was vertically attached to the substrate and smeared with what looked like white glue. Sycanus annulicornis produces less eggs when raised with Tenebrio molitor (97 eggs/female) than when raised with Crocidolomia pavonana (114.4 eggs/female).

5.3.8

Nabidae

When first oviposited, the eggs are white; as they mature, they turn yellow. There are two black marks on the egg before it hatches. These are the nymph’s developing eyes (Ojeda-Peña 1971). About 8 days in the summer and 12 days in the winter pass during the incubation period (Ojeda-Peña 1971). The only visible part of the egg while it is in place is the reticulate cap, which is cylindrical in shape, slightly carved in the contour, and seemingly specific to each species. Similar to how a cork fits in a vial, it slips into and over the end of the egg. When the egg hatches, it is simply pulled outward, where it hangs hung by a fine thread. Scanning electron microscopy was used to examine the surface morphology and internal chorionic structure of the eggs of Nabis pseudoferus pseudoferus, Nabis occidentalis, Nabis punctatus, and Nabis rugosus. The eggs are jar-shaped, with a short “collar” at the front that is closed by an operculum. The chorion is arranged into an exterior layer of about 4–5 μm that is separated from the inner surface by a “pillars” layer of about 0.5–1 μm, with the exception of the collar and operculum areas. The chorion has tiny internal channels near the collar that stand in for the aeropyles; even within the same species, the number of these aeropyles varies greatly. The operculum is made up of air-filled, confined areas in contact with the aeropyles and the “pillars” layer that together create the respiratory system’s model. The traits and general appearance of Nabis pseudoferus pseudoferus, Nabis occidentalis, and Nabis rugosus eggs are extremely similar (Elisabetta and Maria 1998).

124

5.3.9

5 Egg Biology of Insect Predators

Anthocorids

Big-eyed bugs have been found to prefer the undersides of leaves and stems to the upper surfaces of leaves, petioles, and apex when ovipositing. The egg is cylindrical, somewhat incurved, and long. It has a sculptured operculum with follicular pits at one end that is uniformly rounded and somewhat inflated at the other. A typical egg is around 0.38 mm long, 0.11 mm wide, and 0.08 mm in diameter. The egg changes throughout development from colourless on the first day to vermilion on the day of hatching. The operculum is propelled upward during hatching and often only has one point of attachment, much like a hood. A single female could lay anywhere between 0 and 14 eggs each day on average. A female could lay anywhere between 50 and 88 eggs in her lifetime (Rajasekhara and Chatterji 1970). The egg cover of Orius sauteri is white and clearly visible, and the eggs are short and eggplantshaped.

5.3.10 Pentatomidae Ellipsoid-shaped eggs with average (SD) dimensions of 0.91 mm in height and 0.80 mm in breadth are placed in clusters. Initial colour is opaque, followed by a transition to silver or maroon with a shining surface. The operculum is surrounded by an average of 8.25 (0.90) micropylar processes, which typically curve outward. White micropylar processes end in a spherical structure that ranges in colour from white to black.

5.3.11 Thysanoptera There are numerous families of predatory thrips available in agroecosystems. Ten predatory Phlaeothripidae species in five genera, including Aleurodothrips fasciapennis, were discovered during studies in three northern Thai districts. All through the year, these species aided pest management efforts. In apple and almond orchards, respectively, Scolothrips sexmaculatus (Thripidae) and Leptothrips mali (Phlaeothripidae) are recognised as essential biological control agents. On the other hand, the majority of predatory thrips species are members of the Aeolothripidae family. The majority of aeolothripid species are native to temperate regions and are generalist facultative predators of small arthropods, whereas a small number are specialised and native to the tropics. Aeolothripidae family member Franklinothrips species eat a wide variety of insects. The tropics are full of these swiftly moving, predatory thrips that resemble ants. Franklinothrips vespiformis and Franklinothrips orizabensis are considered to represent significant biological controllers on a global scale. There are 16 additional species of Franklinothrips, including Franklinothrips atlas and Franklinothrips megalops (both found in Africa), Franklinothrips basseti and Franklinothrips variegatus (both found in Australia), Franklinothrips brunneicornis (found in New

5.4 Egg-Laying Pattern and Protecting Mechanism

125

Caledonia), Franklinothrips fulgidus and Franklinothrips lineatus (both found in Brazil), and Franklinothrips rarosae (found in the Philippines [Arizona, California, Colorado, Florida, and Texas]). They lay individual eggs, which can be identified by their yellow-green protrusion, inside the tissue of the leaf. Eggs are translucent white, kidney-shaped, and 0.4 × 0.1 mm in size (Hussain et al. 2022).

5.4

Egg-Laying Pattern and Protecting Mechanism

When neuropterans lay several eggs, they do so relatively far apart from one another. Some insects use poisonous compounds, hair, spines, or hard egg shells to ward off predators.

5.4.1

Plant Chemistry

Foraging and oviposition behaviour can also be influenced by plant chemistry (also known as secondary plant metabolites or allelochemicals). Few studies have compared the performance of syrphid larvae feeding on the same species of aphid but from different host plants. Instead, the majority of studies in the literature have concentrated on the effects of host plant chemistry on the suitability of aphid prey for overall performance and subsequent fecundity. The same aphid species was present on both Brassica napus and white mustard plants, but Vanhaelen et al. (2001) showed that E. balteatus females significantly prefer to oviposit on white mustard plants (Sinapis alba containing high glucosinolate [GLS] levels) rather than oilseed rape plants (Brassica napus containing low GLS levels) (Myzus persicae). Recent research by Almohamad et al. (2007) demonstrated that ovipositing E. balteatus females favoured potato plants Solanum tuberosum over Black Nightshade plants Solanum nigrum afflicted with the same aphid species. These aphid–host interaction is due to the emission of volatile chemicals, such as E-(β)-farnesene (EβF), which may be the cause of these aphids’ preferred oviposition sites.

5.4.2

Volatile and Non-volatile Organic Compounds (VOCs)

Kairomones are chemical cues that help organisms that can sense them interact with other species. These attractants might be single chemicals, mixes of herbivoreinduced plant volatiles (HIPVs), or chemicals produced by herbivores themselves, such as pheromones, which operate as a medium for intraspecific communication. During their foraging behaviour, or the location of oviposition sites and feeding resources in nature, natural enemies listen in on kairomones.

5.4.2.1 Coccinellids Around the world, aphids are recognised to have a natural enemy in ladybird beetles (coccinellids) (Hodek and Honek 1996). Although it has been demonstrated that

126

5 Egg Biology of Insect Predators

plants, prey (such as aphids), and products of the prey (such as honeydew) can act as cues for oviposition in coccinellids. Additionally, specific volatile and non-volatile organic compounds (VOCs) like salicylic acid, o-coumaric acid, protocatechuic acid (Smith et al. 1973), guaiacol, and resorcinol (Smith and Williams 1976), quercetin are responsible for the oviposition of Coleomegilla maculatalengi (Riddick et al. 2018b); taxifolin, naringenin for Harmonia axyridis (Alhmedi et al. 2010). In both lab and field tests, chemical cues used by Harmonia axyridis (Coleoptera: Coccinellidae) in both host location and oviposition were examined. The two volatiles that notably attracted gravid H. axyridis females were limonene and E-caryophyllene out of the five that were examined in a four-arm olfactometer: (E)-E-farnesene, E-pinene, E-caryophyllene, and cis-3-hexen-1-ol. The oviposition of H. axyridis on plants was similarly accelerated by these two substances. Furthermore, testing of the attraction of Harmonia axyridis to limonene in the field with controlled-release dispensers (Alhmedi et al. 2010).

5.4.2.2 Syrphids Hoverflies, also known as syrphids, are well known for helping to control aphids in both unmanaged and managed environments around the world. Only the larval stages are predatory; adults visit flowers to gather pollen for egg maturation and nectar for energy, making them pollinators. Syrphid females regulated the rate of oviposition in response to the density and dispersion of their aphids prey, although this reaction varied according on the species of syrphid. Tricosane, tricosane + tetracosane + pentacosane + hexacosane + octacosane mix, hexacosane (Shonouda et al. 1998), (Z)-3-hexenol, and (E)-β-farnesene are all used to promote oviposition in Syrphidae predators like Metasyrphus corolla and Episyrphus balteatus (Verheggen et al. 2008). 5.4.2.3 Chrysopidae L-tryptophan and a combination of acetic acid, methyl salicylate, and phenylacetaldehyde are also effective in enhancing Chrysoperla carnea oviposition (Bakthavatsalam et al. 2007), acetic acid + methyl salicylate + phenylacetaldehyde mixture (Koczor et al. 2015). Methyl salicylate is essential for causing Chrysoperla rufilabris oviposition in screen-house conditions (Salamanca et al. 2017). The same function is controlled by (Z)-3-hexenyl acetate, and linalool (3E)-4,8-dimethyl1,3,7-nonatriene for Chrysopa phyllochroma (Xu et al. 2015). Because there are different amounts of volatile and non-volatile organic compounds (VOCs), the average number of eggs or egg clutches produced in test/control treatments (egg production ratio, or EPR) of Coccinellids, syrphids, and chrysopids varies (Table 5.1).

Harmonia axyridis Metasyrphus corollae Episyrphus balteatus Chrysoperla carnea Chrysoperla rufilabris Chrysopa phyllochroma

Coleomegilla maculata

Predators Coleomegilla maculata lengi

Smith and Williams (1976) Smith and Williams (1976) Riddick et al. (2018a, b) Riddick et al. (2018a, b) Alhmedi et al. (2010) Shonouda et al. (1998) Verheggen et al. (2008) Bakthavatsalam and Singh (1996), Bakthavatsalam et al. (2007) Salamanca et al. (2017) Xu et al. (2015)

Concentrations 1% 0.04 mg/cm2 0.004 mg/mL 0.004 mg/mL 0.10 μg/μL 2.5 mg/mL 0.40 μg/μL 33.3 mg/mL 1.0 mg/mL 5 μL

VOCs Guaiacol Salicylic acid Quercetin Naringenin Limonene Tricosane + tetracosane + pentacosane + hexacosane + octacosane mix (E)-β-farnesene

L-tryptophan

Methyl salicylate

Linalool

Table 5.1 Due to the varying quantities of volatile and non-volatile chemical substances (VOCs), coccinellids, syrphids, and chrysopids egg production ratio (EPR = the average number of eggs or egg clutches produced in test/control treatments)

5.4 Egg-Laying Pattern and Protecting Mechanism 127

128

5.4.3

5 Egg Biology of Insect Predators

Protecting Mechanism

5.4.3.1 Site Selection The qualities of the plant and the availability of prey regularly affect the oviposition decisions of predators that lay their eggs on plants. A few indications are used by insect predators to decide if a location is appropriate for oviposition. Prey-mediated cues entail reacting to prey densities or semio-chemicals released by prey, and this is true of many of the researched cues. Predaceous insects choose oviposition sites based on plant traits in addition to prey considerations, and in some situations plant characteristics can outweigh cues acquired from prey (Griffin and Yeargan 2002a, b). Although they cannot rely on the escape and aggressive actions of mobile stages, a variety of morphological and chemical defensive mechanisms have evolved to protect insect eggs from predation and parasitism. When the risk of progeny mortality is significant, predatory insects always prioritise their capacity to identify any natural adversaries prior to oviposition and to choose foliage free of opponents. All ovipositing females can distinguish between different plant species, plant types, artificial substrates, and whether or not a plant is plagued with pests. Insects frequently possess the ability to choose egg-laying locations in order to improve the survival rate of their progeny. This behaviour can be influenced by a number of variables, including the presence of natural enemies or competitors, as well as site characteristics like the availability of food resources, illumination intensity, temperature, site size, shape, or colour. Additionally, plant species and its physical traits or tissues have a favourable or negative impact on how predatory insects lay their eggs. 5.4.3.1.1 Anthocorid Only the white operculum was visible because the female of Orius indicus placed eggs just beneath the epidermis onto the flower stalk (peduncle) of Cajana cajan (Rajasekhara and Chatterji 1970). Orius insidiosus (Heteroptera: Anthocoridae) is an example of an anthocorid that exhibits distinct oviposition preferences for particular plant species and plant-related settings (Askari and Stern 1972; Isenhour and Yeargan 1982; Seagraves and Lundgren 2010; Pumariño and Alomar 2012). Most anthocorids exhibit certain behaviours and cues that lead to oviposition, but it is yet unclear how polycultures affect a population’s capacity for reproduction and prey suppression. Orius insidiosus in soybeans showed apparent preferences for oviposition sites; it oviposited most frequently in leaf petioles. Additionally, stratification of oviposition sites was found along the soybean plant’s vertical side (Isenhour and Kenneth 1982). Orius sauteri (Heteroptera: Anthocoridae), a polyphagous predator that serves as a natural adversary, did experience a significant impact from the plant species on single-female overall fertility. Along with Leonurus artemisia, Trifolium repens, Vicia villosa, Vicia sativa, Vicia glabrescens, Viola Philippica, Glycine javanica, Vigna radiate, Vigna angularis, Arachis Hypogaea, Phaseolus vulgaris, and Camellia sinensis, the list of plants also includes Camellia sinensis (Zhang et al. 2021). The following details the fecundity and egg-laying seasons (Fig. 5.2). Zhang et al. (2021) used Camellia sinensis as a reference, but found that for the examined plant species,

5.4 Egg-Laying Pattern and Protecting Mechanism

129

Fig. 5.2 Reproductive capacity of Orius sauteri adults in the presence of various plants (after Zhang et al. 2021)

the results on single-female total fecundity and oviposition period varied significantly. When compared to tea plants, Orius sauteri adults deposited more eggs in Vigna angularis.

5.4.3.2 Diptera Aphidophagous midge eggs and new-born larvae are susceptible to intra-guild predation, desiccation, and hunger. Eggs might be set on stalks, be covered in faeces, protected by armour or oil, or be placed near a food source in an oviposition site that is less exposed. Review finds that mosquitoes had access to a tremendous quantity of work. The midge that eats aphids in contrast to other microhabitats, Aphidoletes aphidimyza chose to lay its eggs on potato plants, which had a high density of trichomes. 5.4.3.3 Reduviids Haridass (1985) claims that species of ground-dwelling piratine as Piratus affinis, Ectomocoris tibialis, Ectomocoris ochropierus, and Catamiarus breoipennis use their ovipositors, which resemble plates, to lay their eggs in the ground. The pregnant woman adopts a slanting position with her head and thorax lifted while making side-to-side, twisting, and downward thrusting movements with just the tip of her abdomen contacting the ground. The exposed portion of the egg is next covered with tiny sand and dirt fragments by the hind legs. Similar Ectrichodiinae like Haematorrhophus niqrooiolaceus, Guionius nigripennis, Ectrychotes pilicornis, and other species lay their eggs in groups. The females of the earlier species dig slanting tunnels (7–8 cm) in the ground to lay eggs loosely, in contrast to the latter two species that glue the eggs in the cracks of tree bark. After oviposition, they take their time filling the tunnel with the mud they dug up with their fore and rear legs before pressing the hole shut with their bellies. These Ectrichodiinae structures are significantly smaller and more stub-like than the plate-like ovipositors of the species that lay their eggs in the ground or in cracks. All harpactorines lay

130

5 Egg Biology of Insect Predators

clusters of many eggs that are adhered with cement to the bark of trees or bushes, the undersides of boulders, or other hard surfaces (Haridass 1985). The gravid females of Agriosphodrus dohrni typically attach a cluster of many eggs on the stems of trees when in the field. The females glue the eggs in vertical but oblique rows, working from the edges of the egg pile to its centre. Each egg is linked to the substrate and to the egg that was laid before it, giving the whole egg mass a unique shape. Such egg masses are then covered by the females’ abundant auxiliary gland secretions, transforming them into something resembling an ootheca. The females will select the spot specifically to prevent the freshly hatched nymphs from becoming crowded there (Luo et al. 2010). Miller (1956) first described the anthropomorphic behaviour of reduviids, and Odhiambo (1959) noticed it in Rhinocoris albopunctatus. Nyiira (1970) material on Rhinocoris albopunctatus supports earlier findings. The male parent of Rhinocoris albopunctatus cared for the eggs until the nymphs emerged. The male Rhinocoris albopunctatus guard regularly walked over and around the eggs while consuming insects nearby. In some cases, no feeding was seen over the whole brooding period or until the nymphs had fully emerged. Eggs may be protected from parasites and predators by brooding. Only one Rhinocoris albopunctatus female was discovered to be linked with an egg cluster in the presence of a male out of the 103 clusters of eggs that were seen. These eggs took 5–6 days to hatch, so perhaps the mother had just finished laying them. Out of the 103 clusters, 93 were guarded, 10 were discovered without guards, and 8 of these 10 had an average of 70–78% parasitism. The two remaining clusters naturally hatched. Three scenarios could account for this phenomenon: The eggs were either. (a) Left unattended because of delayed hatching. (b) The male guard noticed the parasitism and left. (c) The eggs in the eight clusters were parasitised because they had not been watched over.

5.4.3.4 Nabidae Under greenhouse conditions, paired-choice experiments were used to ascertain Nabis roseipennis’ ovipositional preferences between a preferred standard (soybean) and four crops (corn, tomato, tobacco, and squash). Soybean contained noticeably more eggs or egg groups than corn or tomato. Only squash was favoured over soybeans for oviposition (Pfannenstiel and Yeargan 1998). In soybean, Nabis roseipennis was seen to have apparent preferences for oviposition sites; it oviposit most frequently in the median leaflet petioles. Additionally, stratification of oviposition sites was found along the soybean plant’s vertical side (Isenhour and Kenneth 1982). Adult female Nabis capsiformis can deposit up to 200 eggs during the summer months, which are placed into the stems of low-growing plants (Krey and Renkema 2018).

5.4 Egg-Laying Pattern and Protecting Mechanism

131

5.4.3.5 Syrphids Oviposition sites vary greatly and are influenced by both the quantity and location of eggs laid. Although certain species of syrphid lay eggs in batches far from the colony or even on uninfected plants, syrphid eggs are often laid singly, either adjacent to or within aphid colonies (Chambers 1988). Additionally, it has been noted that female syrphids can oviposit in response to species-specific aphid colonies. According to Malcolm (1992) and Sadeghi and Gilbert (2000), the type of food and its nutritional value affect how well larvae perform. When they come across any aphid colonies, they come around and lay eggs and are frequently observed hovering over the blooms in search of nectar and pollen. Example: On Brassica (Brassica rapa oleifera), Aphidophagous Hoverflies, Eupeodes luniger, Ischiodon scutellaris, and Episyrphus balteatus mostly preferred to deposit their eggs near green-peach aphid colonies (Myzus persicae): Ischiodon scutellaris (36.22%), Eupeodes luniger (19.5%), and Episyrp (Jamali et al. 2018). Eggs are frequently laid on plants with or without aphids in certain melanostomine species and in all Platycheirus species with the exception of Platycheirus scutatus. Eggs of the latter species are laid in batches of two to four rather than singly (Gilbert 1986). The plants with the most aphids produced the greatest syrphid Episyrphus balteatus responses, as determined by oviposition over a 5-day period (Sutherland et al. 2001). The eggs are placed single, in close proximity to the aphid colony, or in groups next to each other on the surface and stem of the leaf (Joshi and Ballal 2013).

5.4.4

Plants and Their Phenology

The acceptance of an aphid–host plant as a site for oviposition has also been demonstrated to be influenced by some physical plant traits, such as the presence of trichomes. Field research has shown that ovipositing Episyrphus balteatus, females did not prefer nettle (Urtica dioica) contaminated with Microlophium carnosum (Sadeghi and Gilbert 2000). It is discovered that other host plant variables, such as floral characteristics and colour, have a significant impact on seeking and oviposition behaviour. Coleoptera Plant chemical combinations have an impact on oviposition. One such example is that a polyphagous carabid Pterostichus melanarius lays more eggs in a Brussels sprout intercropped with barley than in a Brussels sprout alone (Trefas and van Lenteren 2008). Chinese kale, cotton, cucumber, eggplant, and sweet potato are five of the common host plant species of Bemisia tabaci. The effect of these hosts on oviposition preference, offspring performance, and adult performance of the whitefly predator Serangium japonicum is studied. The smallest and biggest attachment forces for Serangium japonicum were supported by Chinese kale and eggplant, respectively (Feng-Luan et al. 2020). Diptera According to this research, Aphidoletes aphidimyza (Cecidomyiidae) females prefer to lay their eggs in areas with a lot of trichomes because coccinellid

132

5 Egg Biology of Insect Predators

predation is less likely there (Lucas and Brodeur 1999). The fitness of hoverfly larvae, pupae, and adults as well as the quantity of eggs laid by the two first aphid species Acyrthosiphon pisum, Megoura viciae, and Aphis fabae were much higher in another hoverfly species, Episyrphus balteatus (Syrphidae) (Almohamad et al. 2007). Anthocoridae Laboratory tests were conducted to determine which plants Orius insidiosus (Anthocoridae) females preferred and accepted, including pole beans, soybeans, redroot pigweed, velvetleaf, green foxtail, orchard grass, buffalo grass, smooth brome, redtop grass, blue grama, and tall fescue. Orius insidiosus did not differentiate between nodes or petioles of various lengths, preferring to lay its eggs on the petiole and leaflet petioles of pole beans. The acceptance of these plants varied greatly, with pole beans being the most palatable and nearly no eggs being placed on velvetleaf, despite the fact that all broadleaved plants were suited for egg development (Lundgren and Fergen 2006). Thinner and less pubescent leaves, which allow the offspring to feed on them and promote survival, were selected by female Orius insidiosus as oviposition locations (Lundgren et al. 2008). For oviposition, Orius strigicollis favoured the seams of the pods, particularly those at the tops of the kidney bean pods (KBPs). Females did not lay eggs when the KBP tail pieces were not present, according to trials including choice and non-choice. Although there were no appreciable differences in the rates of egg hatching on the various KBP components, the tip seam required much less time for females to lay their eggs. On the tip seam, the insect can take advantage of support points and gain leverage for ovipositor insertion, which reduces oviposition time (Yu et al. 2021). Mantidae In a study, Orthodera novaezealandiae ootheca was discovered on trees between 200 and 3400 mm above the ground, with a median height of 1055 mm and a mean height of 1167 mm, respectively. Of the ootheca discovered on trees, 80.3% were at light levels of 3000 lux or more. The majority of oothecas were facing north or north-west, with the median (11 oothecas) having a 340° aspect. Sophora microphylla, Carmichaelia sp., Pseudopanax crassifolius, and Cordyline australis were the four plants that received the most oothecas, accounting for 28%, 20%, 18%, and 12% of the total oothecas, respectively. On tree limbs 50 mm in diameter or smaller, more than half (56.6%) of all oothecas were discovered (Bowie and Bowie 2003). Miomantis caffra frequently lays eggs within or on top of buildings and other structures (Fea 2011). The most common plant species for putting oothecas is Lantana camara, where the Giant Asian Mantis Hierodula membranacea oviposits and lays its eggs. Oothecas were positioned on delicate branches at a mean height of 2.10 m (0.9–3.5 m) above the ground (Balakrishnan 2012). When plant cover is denser, the European mantis, Mantis, will deposit more egg cases (oothecas). Additionally, the height of egg deposition increases with the height of the adjacent plants. The conduct that was seen could guarantee that developing offspring are properly isolated (Kajzer-Bonk 2020).

5.4 Egg-Laying Pattern and Protecting Mechanism

133

Syrphid Heringia calcarata was the most prevalent syrphid fly in both orchards, and a comparison of the number of syrphid eggs on woolly apple aphid colonies revealed there was no statistically significant interaction between hover fly species and orchard. While Eupeodes americanus and Syrphus rectus eggs were discovered in woolly apple aphid, rosy apple aphid, and spirea aphid colonies when all three aphid species were temporally sympatric; Heringia calcarata eggs were never discovered in rosy apple aphid or spirea aphid colonies (Short and Bergh 2004). Nabidae Eggs were laid at comparable rates on the remaining plant structures by Nabis americoferus, although eggs were laid on petioles substantially more frequently than anyplace else on the plants. Eggs were laid at comparable rates in the two treatments on florets, internodes, and leaves (Pumariño et al. 2011). Petioles and petioles were the only structures where the treatment with both bugs had significantly more eggs than the Nabis americoferus—only treatment. The microstructure of the leaf surface or other portions’ wetness can significantly affect predator feeding behaviour in addition to plant phenology, which is a well-known factor. There is a vast amount of literature available on this topic. However, nothing is known about its impacts on predator oviposition preference, which is important for arthropod fitness at the population level. Serangium japonicum, a whitefly Coccinellid predator, preferred oviposition on Chinese kale and eggplant leaf discs because the former possessed epicuticular wax crystals and the latter had stellate trichomes (Yao et al. 2021). Neuropteran Female Chysoperla carnea lays more eggs on cauliflower (67 and 41 eggs/female for broccoli and kohlrabi, respectively) than on broccoli (72 and 38 eggs/female for broccoli and kohlrabi, respectively). Chysoperla carnea produces more eggs on leaves in cabbage (31/female), compared to stems (13/female) and other plant parts (5/female) (Reddy et al. 2004). Chrysoperla rufilabris’ oviposition is negatively impacted by the presence of aphids (5 and 4.5, respectively) (Frechette and Coderre 2000). The eggs (clutches) are frequently found on the upper side, underside, tips, or margins of the leaves of various plants. They are laid as single (one egg), group, or cluster clutches (5–40 grouped eggs with separated pedicels or creating a bouquet) (Monserrat 2016). Chrysoperla lucasina, Chrysoperla pallid, Pseudomallada prasinus, Cunctochrysa baetica, and Chrysoperla mutata all preferred to lay their clutches on the upper side of the leaves as well as the leaf’s upper side, which are frequently spiny in Mediterranean vegetation (Alcalá et al. 2019). Single clutches were the most common.

134

5.4.5

5 Egg Biology of Insect Predators

Other Factors

5.4.5.1 Natural or Laboratory Hosts 5.4.5.1.1 Coleoptera Only a small portion of prey species actually behave as “required meals,” which is necessary to support juvenile growth and adult reproduction, despite the fact that many predatory insects appear to be opportunistic generalists in their prey choice. Evans et al. (1999) conducted an experiment with two aphidophagous lady beetles, Coccinella septempunctata and Coccinella transversoguttata, produced eggs when fed a diet of aphids and weevils, which are both necessary prey (alternative prey). In line with expectations, female predators laid more eggs when alfalfa weevil larvae were added to their diet of sparse pea aphids. When given only sugar or weevils, the predators did not lay any eggs. However, once aphids were introduced to the diet, Coccinella transversoguttata females (but not Coccinella septempunctata) produced eggs more frequently when they had previously fed on weevils rather than sugar. When given both weevils in excess and sugar, females of both species also produced eggs in modest quantities, although this diet sustained a lower rate of egg production than did a diet of weevils in excess with a small number of aphids. Although Coccinella septempunctata and the alfalfa weevil have a longer history of connection than Coccinella transversoguttata does, the former species was not more successful in utilising this alternate prey to promote reproduction. The propensity of generalist predators like adult lady beetles to consume both secondary and primary prey likely greatly increases their capacity to take advantage of transient and sporadic opportunities as they search for optimal places for reproduction. Females of aphidophagous ladybirds must be adapted to exploit prey that varies greatly in their occurrence and abundance over both space and time. The searching behaviour of female ladybirds clearly reflects the ephemeral nature of local aphid populations. Adult ladybirds often do not remain long in any given location, but instead appear to move frequently between sites and habitats throughout the breeding season. Anthocoris nemorum, Anthocoris nemoralis, and Anthocoris confusus all lay their eggs near high densities of prey (Sigsgaard 2005). Ugine et al. (2019) examined the impact of prey nutritional composition on beetle propensity for herbivory using seven spotted lady beetles (Coccinella septempunctata). Ugine et al. (2019) demonstrate that beetles fed only prey exhibit normal growth and development but completely lose fitness (spermatogenic failure), which is recovered through herbivory and the addition of phytosterols and cholesterol. Furthermore, we demonstrate that lady beetles have a sterol-specific hunger that is state-dependent and that they compensate for their sterol deficiency by eating on leaves. These findings disclose a selection mechanism (sterol nutrition) that leads predatory species to omnivory and show that predators balance their nutrient intake through herbivory when prey quality is low. Not only can prey nutrition affect predatory insects’ oviposition, but fertilisers with varying levels of nitrogen content also affect the crops grown. This subsequently has an impact on the coccinellids’ egg-laying capacity. For instance,

5.4 Egg-Laying Pattern and Protecting Mechanism

135

Hippodamia variegata was utilised as the predator in trials to gauge the predator’s growth, reproduction, and predation rate using colonies of Aphis gossypii as its prey. A two-sex, age-stage life table was created after the biological parameters were computed. Due to the low N content of the aphids, larval development was prolonged and aphid consumption increased at the low level of N fertilisation. However, female longevity was highest in this treatment, probably as a result of a decreased reproduction rate, and there was no compensatory eating of aphids by adults, as found in larval stages, potentially as a result of spillover effects of subpar nutrition during the larval stage. The high N treatment did not produce the highest intrinsic rate of increase despite feeding on aphids with the highest N content (r). As a result, the medium N treatment (160 ppm N) produced the maximum aphid consumption rate, r value, and aphid to egg conversion efficiency (Hosseini et al. 2019). Hemptinne and colleagues in 2022 recently reported and proposed that ladybird predators of aphids lay larger eggs than ladybird predators of coccids, according to research controlling for female mass. Only the sort of prey consumed affects this distinction; phylogenetic relatedness has no bearing. Because neonatal larvae do not need to hunt for, capture, and subdue prey, we propose that ladybird predators of coccids lay smaller eggs. When phylogeny is taken into account, the reproductive investment of both species of ladybirds is comparable to their body mass. Understanding the dynamics of the food web requires an understanding of how prey has shaped the life histories of predators. From an applied standpoint, biological control methods should be guided by and made more effective by using this fine evolutionary tuning of prey–predator relationships. Chrysopidae Chysoperla carnea’s egg period varied depending on the type of insect host, from 2.35 days for Aphis craccivora on groundnut to 3.45 days for Aphis gossypii on cotton (Than et al. 1999). The egg period of Chysoperla carnea is also influenced by the host insect’s life stages. The fertility of Chrysoperla mutate is affected by different Lipaphis erysimi densities (2, 4, 8, and 12 per day). Fecundity is a prey density based on the number of eggs laid per female each day at rates of 84, 175, 295, and 468 (Abdulhay 2021). Table 5.2 shows the impact of different plants and host preys on the egg period of Chysoperla carnea and Chrysoperla zastrowi sillemi. The feeding habits of Chrysoperla zastrowi sillemi larvae on different hosts were assessed and the results showed that Chrysoperla zastrowi sillemi is one of the finest biological control agents for consuming lepidopteran eggs and sucking pests. The results showed that Corcyra cephalonica eggs had the highest host preference of first instar larvae (grub) (24.000%) and the lowest host preference of Ceroplastes cirripediformis (10.00%). The maximum host preference of second and third instar larvae (grub) was found in Lipaphis erysimi (21.50%) and (24.00%), with a minimum in the case of Ceroplastes cirripediformis (12.50%) and (24.00%) (10.00%) (Beerendra and Ganguli 2022).

136

5 Egg Biology of Insect Predators

Table 5.2 Influence of different preys and their host plants on the fecundity of Chrysoperla spp. Prey(s) Life stage offered Corcyra cephalonica Eggs Corcyra cephalonica Not specified Earias vitella Eggs Helicoverpa armigera Eggs Earias vilela Larvae Helicove1pa armigera Larvae Spodoptera litura Not specified Aphis craccivora Not specified Plants Aphis gossypii Cotton Aphis gossypii Okra Aphis gossypii Guava Aphis craccivora Cowpea Aphis craccivora Groundnut Chrysoperla zastrowisillemi Unsterilised eggs Corcyra cephalonica Corcyra cephalonica Sterilised eggs Brevicoryne brassicae Eggs Eggs Aphis gossypii Aphis craccivora Eggs Uroleucon compositae Eggs

Egg period (days) 3.0 569 2.8 2.5 3.3 3.l 359 299

References Than et al. (1999) Khanzada et al. (2018) Than et al. (1999) Than et al. (1999) Than et al. (1999) Than et al. (1999) Khanzada et al. (2018) Khanzada et al. (2018)

3.5 3.1 3.1 2.4 2.3

Than et al. (1999) Than et al. (1999) Than et al. (1999) Than et al. (1999) Than et al. (1999)

371.6 338.8 153.8 281.4 262.2 113.4

Mounika et al. (2021) Mounika et al. (2021) Mounika et al. (2021) Mounika et al. (2021) Mounika et al. (2021) Mounika et al. (2021)

Additionally, Table 5.2 shows the impact of preys on the egg-laying duration of several predatory insects. Table 5.2 demonstrates how various preys have minimal to moderate effects on predatory insect traits like egg-laying capacity. Because the majority of predators are generic predators that consume a variety of food at the field level in order to survive in nature. Reduviids Zelus exsanguis adults were laying their eggs on tree leaves (West and DeLong 1956). Table 5.3 provides the effect of preys on egg production. Following ideas about reduviids were developed through several studies (Ambrose and Ganesh Kumar 2016). Crowding reduced body weight, increasing rearing space increased fecundity, flooding of eggs decreased hatchability and lengthened incubation, soil moisture affected incubation period, hatchability, and egg mortality, and mating with multiple males of different ages resulted in higher fecundity and hatchability as well as a shorter incubation period. Anthocoridae Orius spp. are frequently utilised in biological control methods to eradicate various pests all over the world because they are quick-moving, energetic, and have a greater host search effectiveness than other species. In many agronomic systems, Orius strigicollis (Anthocoridae) is a significant native natural predator of a variety of soft-bodied insect pests like thrips, aphids, and mites. They choose thinner

5.4 Egg-Laying Pattern and Protecting Mechanism

137

Table 5.3 Influence of various food sources on the incubation period (days) and oviposition (eggs/ female) of different predator families Predator Syrphid Eupeodes frequens Episyrphus balteatus Coleoptera–Coccinellidae Coccinella septempunctata Lygaeidae Anagasta kuehniella Aphis gossypii Caliothrips phaseoli Anthecoridae Orius insidiosus Orius insidiosus Orius insidiosus Blaptostethus pallescens Reduviidae Sycanus falleni Sycanus falleni Sycanus falleni Sycanus falleni Sycanus falleni Sycanus falleni Sycanus falleni Sycanus falleni Endochus albomaculatus Panthous bimaculatus Epidaus bicolor Sphedanolestes signatus Euagoras plagiatus Irantha armipes

Life stage offered

Egg period

References

Brevicoryne brassicae Brevicoryne brassicae

3.5 3.8

Singh et al. (2020) Singh et al. (2020)

Aphid young ones

255.0

Sarwar and Saqib (2010)

Eggs Eggs Eggs

5.0 6.9 5.1

Mendes et al. (2002) Mendes et al. (2002) Mendes et al. (2002)

Anagasta kuehniella Caliothrips phaseoli Aphis gossypii Corcyra cephalonica eggs

5.0 5.1 7.0 5.8

Mendes et al. (2002) Mendes et al. (2002) Mendes et al. (2002) Jose and Subramanian (2020)

Pieris rapae Spodoptera litura Plutela xylostela Corcyra cephalonica Pieris rapae larvae Spodoptera litura larvae Plutela xylostela larvae Corcyra cephalonica larvae Field collected predator

17.6 17.9 18.2 18.6 243.35 165.08 148.89 113.89

Truong et al. (2020) Truong et al. (2020) Truong et al. (2020) Truong et al. (2020) Truong et al. (2020) Truong et al. (2020) Truong et al. (2020) Truong et al. (2020)

181

Srikumar et al. (2014)

Field collected predator Field collected predator Field collected predator Field collected predator Field collected predator

465 203 53.4 203 88

Srikumar et al. (2014) Srikumar et al. (2014) Srikumar et al. (2014) Srikumar et al. (2014) Srikumar et al. (2014)

epidermal plant surfaces where the vesicular and cellular tissues are favourable for nymph survival and development (Lundgren and Fergen 2006; Lundgren et al. 2008). The biggest percentage of eggs was laid on the seam of the tail, which had more laid on it than on the seam of the head (Yu et al. 2021). Additionally, the time it took each female to deposit one egg on the tip was on average 28.7% less than the time it took to do so on the centre.

138

5 Egg Biology of Insect Predators

Pentatomidae DeClercq and Degheele (1997) looked at how mating status affected body weight, oviposition, egg load, and predation in the predatory stinkbug Podisus maculiventris. Numerous studies simply highlight the prey type while ignoring the macromolecular composition. However, the research done by Shapiro and Legaspi (2006) shows the females of the colony-reared Podisus maculiventris species were fed one of the following species of natural or artificial prey: wax moth, Galleria mellonella, yellow mealworm, Tenebrio molitor, beet armyworm, Spodoptera exigua, fall armyworm, Spodoptera frugiperda, cabbage looper, Trichoplusia ni. Over intervals of 7, 15, and 22 days, fresh weights and compositions of lipid, protein, and yolk protein were compared. By trial length or prey species, fresh weights and protein revealed no significant differences. Total lipid content, which ranged from 5.3% to 15.5% of mean fresh weight, was the most significant characteristic in connection to time and species of prey. When fed autumn armyworm at 15 and 22 days, female Podisus maculiventris total lipid content differed significantly by prey species, week, and only then. Females fed yellow mealworm had the highest lipid amounts, whereas those fed cabbage looper and beet armyworm had the lowest lipid values. Yolk protein concentration varied over time in females that consumed wax moths or beet armyworms, but it did not correspond with cumulative oviposition. A quantitative indicator of dietary quality, lipid content in female predators may change inversely with reproductive potential or egg load. Lygaeidae Eggs are shaped like a hot dog. They are normally laid down singly and horizontally on plant leaves or stems, and they range in colour from white to peach. Soon after the egg is laid, reddish eyespots appear at the tip. They can be distinguished from other insect eggs by their two red eyespots. The typical length and width of a G. punctipes egg are 0.9 and 0.38 mm, respectively (1/28 in. × 1/66 in.). They have longitudinal striations and are white, yellowish, or tan in colour. In a lab setting, the average time it took for eggs to hatch was around 10 days. Orius insidiosus preferred to lay eggs on the pole bean (213 egg/female), followed by pigweed (147 egg/female), soybean (111 egg/female), and velvetleaf (8 egg/female). However, Orius insidiosus strongly preferred pole beans as a location for oviposition (Lundgren and Fergen 2006). Lundgren and Fergen (2006) also demonstrated a strong preference for laying eggs on specific plant components, with the petiole and leaflet petioles of pole beans being the most preferred places. There were no clear favourites for particular nodes along the stem. Orius insidiosus failed to differentiate between leaflet petioles of various lengths. Miridae A significant natural enemy of the eggs and juvenile nymphs of the rice leaffolder Cnaphalocrocis medinalis (Crambidae), rice stem borer Chilo suppressalis (Crambidae), and pink rice borer Sesamia inferens (Noctuidae) is Cyrtorhinus lividipennis (Heteroptera: Miridae) (Zhu et al. 2014). Different preys have an impact on Dicyphus errans’ fertility (Arvaniti et al. 2021). Diptera The specialised aphid predator Aphidoletes aphidimyza (Diptera: Cecidomyiidae) is frequently employed in biological control efforts. The majority

5.4 Egg-Laying Pattern and Protecting Mechanism

139

(98.6%) of the Aphidoletes aphidimyza lay their eggs either singly (18.8%), in compact masses (12.6%), or in loose batches (68.6%) on the underside of the potato leaves. Furthermore, comparing the trichome density, the minimum number of Aphidoletes aphidimyza females lay eggs (12.3 ± 4.7) on leaves with low trichome density, and more females lay eggs on leaves with high trichome density (Lucas and Brodeur 1999). According to Titi 1972), Aphidoletes aphidimyza laid more eggs as a function of aphid density. These elements improve the new-born larvae’s ability to eat. In another study, two unique preys Aphis gossypii and Myzus persicae demonstrated that Aphidoletes aphidimyza larvae favoured the third nymphal stage of both species over adults and that Aphidoletes aphidimyza larvae ingested Aphis gossypii more than Myzus persicae. When all four types of prey were examined, the Manly’s index for nymphs of Aphis gossypii and Myzus persicae and adults of Aphis gossypii and Myzus persicae was 0.379, 0.235, 0.208, and 0.176, respectively. As the most favoured food of both species, nymphs were utilised to evaluate the reproductive capabilities of the predator under microcosmic conditions. While feeding third-instar nymphs of Aphis gossypii, Aphidoletes aphidimyza had considerably longer adult lifespans (female: 7.62 0.15, male: 7.42 0.23 days), higher fertility (93.75 2.94 offspring per female), and higher intrinsic rates of increase (0.175 0.009 d1). Finally, our research showed that Aphis gossypii third-instar nymphs make the best meal for Aphidoletes aphidimyza mass rearing (Madahi et al. 2019). Syrphid Oviposition sites vary greatly and are influenced by both the quantity and location of eggs laid. Syrphid eggs are frequently laid individually, either nearby or inside aphid colonies, though some species may even lay their eggs on plants that are not afflicted. Eupeodes frequens and Episyrphus balteatus’ incubation periods were unaffected by the cabbage aphid, Brevicoryne brassicae. Different prey types have an impact on things like oviposition and the incubation period (Table 5.3).

5.4.6

Inter-, Intra-guild Predation and Cannibalism (IGP)

Predator–prey or predator–predator interactions are common among a food web’s potential interactions (Krey et al. 2021). In this sense, cannibalism is the act of preying on members of the same species, and it can play a crucial part in maintaining a population’s age structure and stability. Negative intra-specific and inter-specific interactions among predators have been characterised as cannibalism and intra-guild predation, respectively. It is important to note that these interactions could be seen as pillars of communities, depending on how they turn out (Oliveira et al. 2022). Another aspect influencing female oviposition is inter-guilding. The interactions between the hoverfly species Episyrphus balteatus and Metasyrphus corollae and the ladybird, Propylea japonica, were both asymmetric and symmetric. Second, we investigated whether hoverfly oviposition preference was associated with the danger of intra-guild predation posed by the presence of a ladybird larva in an aphid colony in order to test the preference and performance hypothesis for hoverflies. The two

140

5 Egg Biology of Insect Predators

hoverflies’ oviposition preferences in relation to larval performance varied by species: females of the hoverfly Episyrphus balteatus tended to adjust their egg-laying rate in response to the presence of ladybird larvae, whereas females of the hoverfly Metasyrphus corollae responded weakly in a similar manner (Putra et al. 2009). Another study found that the availability of prey (aphids), the safety from predators provided by “hiding” in the colony, and the impact of inter-specific and intra-specific competition all influenced female oviposition (Sentis et al. 2012). Omnivorous predators can survive in agricultural system during prey shortage and overcome rapid spatial and temporal changes in prey availability. Intra-guild predation (IGP) on an inter-specific competitor and cannibalism are observed among those general predators and can induce anti-predator behaviour as a response to the potential risk. The perceived risk of intra-guild/inter-guild predation and cannibalism on general predators’ eggs could influence the oviposition behaviour of females. If two species do not utilise space in the same way, females may assemble their eggs in a constrained area in reaction to the presence of intra-guild predators, which would result in egg aggregation. Females, on the other hand, might use a variety of odd oviposition places at danger of cannibalism. Eggs would therefore be more dispersed when cannibals are a potential threat and there are few other food sources available (Dumont et al. 2021). In a study by Dumont et al. (2021), it was shown that female Macrolophus pygmaeus (Miridae) (=M. caliginosus) are sensitive to both the risk of intra-guild predation by Nesidiocoris tenuis (Miridae) and the risk of cannibalism on their eggs. According to the findings, female Macrolophus pygmaeus preferentially placed her eggs on lower leaves. This tendency is highlighted by the possibility of Nesidiocoris tenuis males engaging in intra-guild rivalry or predation. Artificial Substrates The success of insect populations is significantly influenced by the choice of oviposition sites. The use of synthetic substrates for the entrapped bug Macrolophus caliginosus to lay eggs and produce embryos is assessed. As a synthetic oviposition substrate, a dental cotton roll covered in stretched parafilm and moistened with water showed promising results. It got an average of 1.2 eggs per day from each female, with a yield of 36.4% larvae and 15% adults. Its hardness was comparable to the softest section of the tobacco plant. On host plants, oviposition averaged three eggs per day per female under the same circumstances. Spraying an inula viscosa leaf extract solution over a substrate caused Macrolophus caliginosus to lay an average of 1.6 eggs per female each day. On an artificial substrate, egg laying and subsequent development to the adult stage were possible (Constant et al. 1996). Gelatin coated with parafilm was employed by Castañé and Zalom (1994) as an oviposition substrate for Orius sinsidiosus. For Nesidiocoris tenuis (Heteroptera: Miridae) and Orius laevigatus, De Puysseleyr et al. (2012, 2014) used a parafilm dome (1.5 cm diameter, 1 cm high) utilising a food encapsulation device as an oviposition platform. As soon as the eggs hatch, the parafilm covering needs to be cut to let the nymphs out. However, the majority of the nymphs are stuck to the parafilm’s surface and unable to move. In light of this, while the parafilm can be

5.6 Future Focus

141

manually sliced in small-scale rearing, a large-scale rearing system cannot use this method (De Puysseleyr et al. 2014).

5.5

Egg Dumping

Egg dumpers are females who lay eggs in the care of their conspecifics. Egg dumping is a reliable reproductive technique that decreases risks for practitioners and may even boost fecundity. Even while insect egg dumpers have the potential to be social parasites that prey on egg recipients’ maternal behaviour, dumping is more likely to be a viable reproductive option when the costs to egg recipients are low and possible hosts have little to no resistance against egg dumping incursions. These requirements are satisfied by insects that solely protect their eggs or by insects whose eggs hatch into self-sufficient, precocial offspring that require little more than parental protection. When this is the true, natural and/or kin selection favours egg dumping as a technique through which dumpers can reduce parental risks and increase fertility, while egg recipients can improve offspring survival by reducing predation (Tallamy 2005). Given that dumpers often deposit their eggs on the outer edges of the clutch, where mortality is higher, egg dumping may be advantageous for the receiver (Tallamy 2005). As a result, defences against egg dropping in lace bugs, Gargaphia solani, are not very sophisticated (Tallamy and Horton 1990). Insects that are income breeders or synovigenic have extremely few or no eggs in their ovaries when they first emerge. For each new batch of eggs to hatch, these insects need to eat. Fitness is based on the pattern of resource distribution to somatic and immediate reproduction, which determines their survival and future reproduction. This pattern of allocation is affected by ovarian dynamics in several of these synovigenic insects. Gravid females stop oogenesis and resorb oocytes to fuel the search for high-quality patches of prey when they repeatedly come across poor quality or no oviposition sites. When favourable circumstances arise, egg production is restored. Insects that use ephemeral resources can adjust to regional and temporal changes in resource availability thanks to these two opposing processes, which both occur very quickly (see Ferrer et al. 2011). When starved for more than 24 h, females of the ladybird beetle species Adalia bipunctata and Adalia decempunctata lay a single sterile egg, which they immediately consume. These eggs are the mature oocytes present in the oviducts at the time of famine. This behaviour is part of a strategy used by ladybird beetles to recover energy used during reproduction and shares some parallels with egg dumping observed in herbivorous insects (Ferrer et al. 2011).

5.6

Future Focus

In order to release the right predators into a crop, it is important to understand the seasonal specific oviposition of predatory insects. It is also important to thoroughly research the defensive and offensive significance of the materials used to cover eggs

142

5 Egg Biology of Insect Predators

utilising high throughput technology. Research on inter- and intra-guild is necessary. Biology of predation and eggs can be studied.

References Abdulhay HS (2021) Numerical response of predator Chrysoperla mutata MacLachlan to varying densities of (Kalt.) Lipaphis erysimi. Indian J Ecol 48(3):745–747 Adair EW (1924) On parthenogenesis in Miomanris savigny Saussure. Bull Soc Enromol Egypte, Cairo 8:104–148 Ahmad R (1970) Studies in West Pakistan on the biology of one Nimulid species and two Coccinellid species (Coleoptera) that attack scale insects (Homoptera: Coccoidea). Bull Entomol Res 1970(60):5–16 Alcalá HR, Campos M, Ruano F (2019) Late summer oviposition of green lacewings (Neuroptera: Chrysopidae) on olive groves and adjacent trees. Environ Entomol 48(3):506–513 Alhmedi A, Haubruge E, Francis F (2010) Identification of limonene as a potential kairomone of the harlequin ladybird Harmonia axyridis (Coleoptera: Coccinellidae). Eur J Entomol 107:541–548 Almohamad R, Verheggen FJ, Francis F, Haubruge E (2007) Predatory hoverflies select their oviposition site according to aphid host plant and aphid species. Entomol Exp Appl 125(1): 13–21 Ambrose DP, Ganesh Kumar A (2016) Reduviid predators. In: Omkar (ed) Ecofriendly pest management for food security. Academic, New York, pp 217–257 Ambrose DP, Sebasti Rajan XJ, Nagarajan K, Jeba Singh V, Krishnan SS (2009) Biology, behaviour and functional response of Sphedanolestes variabilis distant (Insecta: Hemiptera: Reduviidae: Harpactorinae), a potential predator of lepidopteran pests. Entomol Croat 13(2):33– 34 Arvaniti KA, Kordas NA, Fantinou AA, Perdikis DC (2021) Impact of prey supply levels on growth performance and optimization of the mass rearing of an omnivorous mirid predator. J Pest Sci 94(3):947–958 Askari A, Stern VM (1972) Biology and feeding habits of Orius tristicolor (Hemiptera: Anthocoridae). Ann Entomol Soc Am 65(1):96–100 Bakthavatsalam N, Singh SP (1996) L-tryptophan as an ovipositional attractant for Chrysoperla carnea (Stephens) (Neuroptera: Chrysopidae). J Biol Control 10:21–27 Bakthavatsalam N, Tandon PL, Patil SB, Hugar B, Hosamani A (2007) Kairomone formulations as reinforcing agents for increasing abundance of Chrysoperla carnea (Stephens) in cotton ecosystem. J Biol Control 21:1–8 Balakrishnan P (2012) Ambush and oviposition site selection by Giant Asian Mantis Hierodula membranacea Burmeister (Mantodea: Mantidae) in tropical wet evergreen forests, Western Ghats, India. J Trop Asian Entomol 13:1–19 Beerendra PN, Ganguli J (2022) Host preference of green lace wing, Chrysoperla zastrowi sillemi (Esben-Petersen) (Chrysopidae: Neuroptera) fed on various hosts. Pharma Innov J SP 11 (9):1753–1756 Bhat PS, Srikumar KK, Raviprasad TN, Vanitha K, Rebijith KB, Asokan R (2013) Biology, behavior, functional response and molecular characterization of Rihirbus trochantericus Stal var. luteous (Hemiptera: Reduviidae: Harpactorinae) a potential predator of Helopeltis spp. (Hemiptera: Miridae). Entomol News 123(4):264–277 Bowie MK, Bowie MH (2003) Where does the New Zealand praying mantis, Orthodera novaezealandiae (Colenso) (Mantodea: Mantidae), deposit its oothecae? N Z Entomol 26 (1):3–5 Brannoch SK, Wieland F, Rivera J, Klass KD, Béthoux O, Svenson GJ (2017) Manual of praying mantis morphology, nomenclature, and practices (Insecta, Mantodea). ZooKeys 696:1–100

References

143

Bugaj-Nawrocka A, Danielczyk A, Sułkowska I, Chłond D (2022) Psyttala horrida (Stål, 1865) (Hemiptera: Reduviidae: Reduviinae)—a morphological study of eggs and nymphs. Insects 13(11):1014 Busch AK, Wham BE, Tooker JF (2021) Life history, biology, and distribution of Pterostichus melanarius (Coleoptera: Carabidae) in North America. Environ Entomol 50(6):1257–1266 Castañé C, Zalom F (1994) Artificial oviposition substrate for rearing Orius insidiosus (Hemiptera, Anthocoridae). Biol Control 4:88–91 Chambers RJ (1988) Syrphidae. In: Minks AK, Harrewijn P (eds) Aphids, their biology, natural enemies, and control. Elsevier, Amsterdam, pp 259–270 Constant B, Grenier S, Bonnot G (1996) Artificial substrate for egg laying and embryonic development by the predatory BugMacrolophus caliginosus (Heteroptera: Miridae). Biol Control 7(2):140–147 Cukier M, Guerrero GA, Maggese MC (1979) Parthenogenesis in Coptopteryx viridis Gigliotos (1915) (Dyctioptera, mantidae). Biol Bull 157(3):445–452 De Puysseleyr V, De Man S, Höfte M, De Clercq P (2012) Plantless rearing of the zoophytophagous bug Nesidiocoris tenuis. BioControl 58:205–213 De Puysseleyr V, Höfte M, De Clercq P (2014) Continuous rearing of the predatory anthocorid Orius laevigatus without plant materials. J Appl Entomol 138:45–51 DeClercq P, Degheele D (1997) Effects of mating status on body weight, oviposition, egg load, and predation in the predatory stinkbug Podisus maculiventris (Heteroptera: Pentatomidae). Ann Entomol Soc Am 90:121–127 Dumont F, Lucas É, Alomar O (2021) Oviposition behavior of the mirid Macrolophus pygmaeus under risk of intraguild predation and cannibalism. Insect Sci 28(1):224–230 Elisabetta C, Maria CR (1998) Egg structures of four Nabis species (Rhynchota: Nabidae). Int J Insect Morphol Embryol 27(2):95–102 Titi AE (1972) Einflüsse von Beutedichte und Morphologie der Wirtspflanze auf die Eiablage von Aphidoletes aphidimyza (Rond.) (Diptera: Itonididae). Zeitschrift für Angewandte Entomologie 72(1–4):400–415 Evans E, Stevenson A, Richards D (1999) Essential versus alternative foods of insect predators: benefits of a mixed diet. Oecologia 121:107–112. https://doi.org/10.1007/s004420050911 Fea MP (2011) Reproductive ecology and impact of the invasive praying mantis Miomantis caffra. Doctoral dissertation, University of Auckland Feng-Luan Y, Lin S, Wang L-X, Mei W-J, Monticelli LS, Zheng Y, Desneux N, He Y-X, Weng Q-Y (2020) Oviposition preference and adult performance of the whitefly predator Serangium japonicum (Coleoptera: Coccinellidae): effect of leaf microstructure associated with ladybeetle attachment ability. Pest Manag Sci 77(1):113–125 Ferrer A, Corbani AC, Dixon AF, Hemptinne JL (2011) Egg dumping by predatory insects. Physiol Entomol 36(3):290–293 Frechette B, Coderre D (2000) Oviposition strategy of the green lacewing Chrysoperla rufilabris (Neuroptera: Chrysopidae) in response to extraguild prey availability. Eur J Entomol 97(4):507– 510 George AR (1988) Bursal glands of Reduviidae (Insecta, Heteroptera). Proc Anim Sci 97:211–222 Gilbert F (1986) Hoverflies. Naturalists’ handbook 5. Cambridge University Press, Cambridge Gnatzy W, Volknandt W, Dzwoneck A (2018) Egg-laying behavior and morphological and chemical characterization of egg surface and egg attachment glue of the digger wasp Ampulex compressa (Hymenoptera, Ampulicidae). Arthropod Struct Dev 47(1):74–81 Greyvenstein B, Du Plessis H, Moulin N, Van den Berg J (2020) Distribution of Galepsus spp. in Southern Africa and life history of Galepsus lenticularis (Mantodea: Tarachodidae). Insects 11 (2):119 Griffin ML, Yeargan KV (2002a) Factors potentially affecting oviposition site selection by the lady beetle Coleomegilla maculata (Coleoptera: Coccinellidae). Environ Entomol 31(1):112–119

144

5 Egg Biology of Insect Predators

Griffin ML, Yeargan KV (2002b) Oviposition site selection by the spotted lady beetle Coleomegilla maculata (Coleoptera: Coccinellidae): choices among plant species. Environ Entomol 31 (1):107–111 Haridass ET (1985) Ultrastructure of the eggs of Reduviidae: I. Eggs of Piratinae (Insecta— Heteroptera). Proc Anim Sci 94(5):533–545 Haridass ET (1986) Ultrastructure of the eggs of Reduviidae: II eggs of Harpactorinae (Insecta: Heteroptera). Proc Indian Acad Sci (Anim Sci) 95:237–246 Heimpel GE, Rosenheim JA (1998) Egg limitation in parasitoids: a review of the evidence and a case study. Biol Control 11(2):160–168 Hodek I, Honek A (1996) Ecology of Coccinellidae. Kluwer Academic Publishers, Dordrecht Hosseini A, Hosseini M, Michaud JP, Awal MM, Ghadamyari M (2019) Life history responses of Hippodamia variegata (Coleoptera: Coccinellidae) to changes in the nutritional content of its prey, Aphis gossypii (Hemiptera: Aphididae), mediated by nitrogen fertilization. Biol Control 130:27–33 Hussain M, Wang Z, Arthurs SP, Gao J, Ye F, Chen L, Mao R (2022) A review of Franklinothrips vespiformis (Thysanoptera: Aeolothripidae): life history, distribution, and prospects as a biological control agent. Insects 13(2):108 Iossa G (2022) The ecological function of insect egg micropyles. Funct Ecol 36(5):1113–1123 Isenhour DJ, Kenneth VY (1982) Oviposition sites of Orius insidiosus (Say) and Nabis spp. in soybean (Hemiptera: Anthocoridae and Nabidae). J Kansas Entomol Soc 55:65–72 Isenhour DJ, Yeargan KV (1982) Oviposition sites of Orius insidiosus (Say) and Nabis spp. in soybean (Hemiptera: Anthocoridae and Nabidae). J Kansas Entomol Soc:65–72 Ito Y (1959) Comparative ecology, 1st edn. Iwanami Shoten, Tokyo, 366 pp [in Japanese] Jamali RA, Memom N, Shah MA, Khan K, Ansari A (2018) Prevalence of aphidophagous hoverflies (Syrphidae: Syrphinae) in relation to their prey, green aphids (Myzus persicae) on brassica (Brassica rapa oliefera) in Dadu. J Anim Plant Sci 28(5):1447–1456 Jose A, Subramanian M (2020) Biology of anthocorid predator, Blaptostethus pallescens Poppius (Heteroptera: Anthocoridae). Entomon 45(2):149–152 Joshi S, Ballal CR (2013) Syrphid predators for biological control of aphids. J Biol Control 27(3): 151–170 Kajzer-Bonk J (2020) Factors affecting egg deposition of the European mantis. Folia Biol 68(2): 45–49 Kazenas VL (2001) Fauna and biology of sphecid wasps (Hymenoptera, Sphecidae) of Kazakhstan and Central Asia. Kazgos INTI, Almaty, p 333 Khanzada KK, Chandio RH, Nahiyoon RA, Siddiqui AA, Jat MI, Mastoi SM, Mastoi PM (2018) Rearing of Chrysoperla carnea (Stephens) against two laboratory hosts and one natural host. J Entomol Zool Stud 6(1):980–983 Koczor S, Knudsen G, Hatleli L, Sentkiralyi F, Toth M (2015) Manipulation of oviposition and overwintering site choice of common green lacewings with synthetic lure (Neuroptera: Chrysopidae). J Appl Entomol 139:201–206 Kolesnikov FN, Karamyan AN (2019) Parental care and offspring survival in Pterostichus anthracinus (Coleoptera: Carabidae): an experimental study. Eur J Entomol 116:33–41 Krey KL, Renkema JM (2018) Pale Damsel Bug Nabis capsiformis Germar (Insecta: Hemiptera: Nabidae). EDIS (4) Krey KL, Smith OM, Chapman EG et al (2021) Prey and predator biodiversity mediate aphid consumption by generalists. Biol Control 160:104650. https://doi.org/10.1016/j.biocontrol. 2021.104650 Lomholdt O (1984) The Sphecidae (Hymenoptera) of Fennoscandia and Denmark. Fauna Entomologica Scandinavica. 4. 2nd ed. Leiden, Copenhagen: E.J. Brill Scandinavian Science Press, pp 452 Lu W, Souphanya P, Montgomery ME (2002) Descriptions of immature stages of Scymnus (Neopullus) sinuanodulus Yu and Yao (Coleoptera: Coccinellidae) with notes on life history. Coleopt Bull 56(1):127–141

References

145

Lucas É, Brodeur J (1999) Oviposition site selection by the predatory midge Aphidoletes aphidimyza (Diptera: Cecidomyiidae). Environ Entomol 28(4):622–627 Lundgren JG, Fergen JK (2006) The oviposition behavior of the predator Orius insidiosus: acceptability and preference for different plants. BioControl 51(2):217–227 Lundgren JG, Fergen JK, Riedell WE (2008) The influence of plant anatomy on oviposition and reproductive success of the omnivorous bug Orius insidiosus. Anim Behav 75:1495–1502 Luo X, Zhou D, Li H, Cheng W, Cai W (2010) Taxonomic and bionomic notes on Agriosphodrus dohrni (Signoret) (Hemiptera: Reduviidae: Harpactorinae). Zootaxa 2358(1):57–67 Madahi K, Sahragard A, Hosseini R, Baniameri V (2019) Prey-stage preference and comparing reproductive performance of Aphidoletes aphidimyza (Diptera: Ceccidomyiidae) feeding on Aphis gossypii (Hemiptera: Aphididae) and Myzus persicae. J Econ Entomol 112(3):1073–1080 Malcolm SB (1992) Prey defence and predator foraging. In: Natural enemies: the population biology of predators, parasites and diseases. Wiley, New York, pp 458–475 Mazzini M (1976) Fine structure of the insect micropyle—III. Ultrastructure of the egg of Chrysopa carnea Steph. (Neuroptera: Chrysopidae). Int J Insect Morphol Embryol 5(4–5):273–278 Mendes SM, Bueno VH, Argolo VM, Silveira LCP (2002) Type of prey influences biology and consumption rate of Orius insidiosus (Say) (Hemiptera, Anthocoridae). Revista Brasileira de Entomologia 46:99–103 Miller NCE (1956) The biology of the Heteroptera, 1st edn. Leornard Hill, London, p 162 Monserrat VJ (2016) Los crisópidos de la Península Ibérica y Baleares (Insecta, Neuropterida, Neuroptera: Chrysopidae). Graellsia 72:e037 Mounika K, Gosalwad SS, Neharkar PS (2021) Biology of Chrysoperla zastrowi sillemi (EsbenPetersen) on different hosts. Pharma Innov J 10(8):671–674 New TR (2003) The Neuroptera. In: The Neuroptera of Malesia. Brill, Leiden, pp 5–21 Nyiira ZM (1970) The biology and behavior of Rhinocoris albopunctatus (Hemiptera: Reduviidae). Ann Entomol Soc Am 63(5):1224–1227 Odhiambo TR (1959) An account of parental care in Rhinocoris albopilosus Signoret (HemipteraHeteroptera: Reduviidae) with notes on its life history. Proc Roy Entomol Soc Lond A 34:175– 185 Ojeda-Peña D (1971) Biologia y hábitos de Nabis capsiformis Germar (Hemip.: Nabidae). Sociedad Entomológica de Perú 14:297–303 Oliveira CM, Silva-Torres CS, Torres JB, De La Pava N (2022) Cannibalism and intraguild predation between endemic and introduced coccidophagous lady beetles (Coleoptera: Coccinellidae). Austral Entomol 61(2):225–235 Papaj DR (2000) Ovarian dynamics and host use. Annu Rev Entomol 45(1):423–448 Pfannenstiel RS, Yeargan KV (1998) Ovipositional preference and distribution of eggs in selected field and vegetable crops by Nabis roseipennis (Hemiptera: Nabidae). J Entomol Sci 33(1):82– 89 Pumariño L, Alomar O (2012) The role of omnivory in the conservation of predators: Orius majusculus (Heteroptera: Anthocoridae) on sweet alyssum. Biol Control 62(1):24–28 Pumariño L, Alomar O, Lundgren JG (2011) The influence of intraguild competitors on reproductive decisions by two predatory Heteroptera, Orius insidiosus (Anthocoridae) and Nabis americoferus (Nabidae). Biocontrol Sci Tech 21(11):1321–1330 Putra NS, Yasuda H, Sato S (2009) Oviposition preference of two hoverfly species in response to risk of intraguild predation. Appl Entomol Zool 44(1):29–36 Rajasekhara K, Chatterji S (1970) Biology of Orius indicus (Hemiptera: Anthocoridae), a predator of Taeniothrips nigricornis (Thysanoptera). Ann Entomol Soc Am 63(2):364–367 Reddy GVP, Tabone E, Smith MT (2004) Mediation of host selection and oviposition behavior in the diamondback moth Plutella xylostella and its predator Chrysoperla carnea by chemical cues from cole crops. Biol Control 29(2):270–277 Riddick EW, Wu Z, Eller FJ, Berhow MA (2018a) Do bioflavonoids in Juniperus virginiana heartwood stimulate oviposition in the ladybird Coleomegilla maculata? Int J Insect Sci 10:1– 13

146

5 Egg Biology of Insect Predators

Riddick EW, Wu Z, Eller FJ, Berhow MA (2018b) Utilization of quercetin as an oviposition stimulant by lab cultured Coleomegilla maculata in the presence of conspecifics and a tissue substrate. Insects 9:77 Sadeghi HF, Gilbert F (2000) Oviposition preferences of aphidophagous hoverflies. Ecol Entomol 25(1):91–100 Salamanca J, Souza B, Lundgren JG, Rodriguez-Saona C (2017) From laboratory to field: electroantennographic and behavioral responsiveness of two insect predators to methyl salicylate. Chemoecology 27:51–63 Salerno G, Rebora M, Piersanti S, Büscher TH, Gorb EV, Gorb SN (2022) Oviposition site selection and attachment ability of Propylea quatuordecimpunctata and Harmonia axyridis from the egg to the adult stage. Physiol Entomol 47(1):20–37 Sarwar M, Saqib SM (2010) Rearing of predatory seven spotted ladybird beetle Coccinella septempunctata L. (Coleoptera: Coccinellidae) on natural and artificial diets under laboratory conditions. Pak J Zool 42(1):47–51 Seagraves MP, Lundgren JG (2010) Oviposition response by Orius insidiosus (Hemiptera: Anthocoridae) to plant quality and prey availability. Biol Control 55(3):174–177 Sentis A, Lucas É, Vickery WL (2012) Prey abundance, intraguild predators, ants and the optimal egg-laying strategy of a furtive predator. J Insect Behav 25(6):529–542 Shapiro JP, Legaspi JC (2006) Assessing biochemical fitness of predator Podisus maculiventris (Heteroptera: Pentatomidae) in relation to food quality: effects of five species of prey. Ann Entomol Soc Am 99(2):321–326 Shelford R (1903) Bionomical notes on some Bornean Mantidae. Fortschr Zool 7(4):293–304 Shibuya S, Kiritani K, Fukuda K (2020) Life history traits of five ground beetle (Coleoptera: Carabidae) species common in Honshu Island, Japan. Acta Zool Acad Sci Hung 66 (Suppl):185–198 Shonouda ML, Bombosch S, Shalaby AM, Osman SI (1998) Biological and chemical characterization of a kairomone excreted by the bean aphids, Aphis fabae Scop. (Hom, Aphididae) and its effect on the predator Metasyrphus corollae Fabr. II. Behavioural response of the predator M. corollae to the aphid kairomone. J Appl Entomol 122:25–28 Short BD, Bergh JC (2004) Feeding and egg distribution studies of Heringia calcarata (Diptera: Syrphidae), a specialized predator of woolly apple aphid (Homoptera: Eriosomatidae) in Virginia apple orchards. J Econ Entomol 97(3):813–819 Sigsgaard L (2005) Oviposition preference of Anthocoris nemoralis and A. nemorum (Heteropera: Anthocoridae) on pear leaves affected by leaf damage, honeydew and prey. Biocontrol Sci Tech 15(2):139151 Singh P, Thakur M, Sharma KC (2020) Larval feeding capacity and pollination efficiency of the aphidophagous syrphids, Eupeodes frequens (Matsmura) and Episyrphus balteatus (De Geer) (Diptera: Syrphidae) on the cabbage aphid (Brevicoryne brassicae L.) (Homoptera: Aphididae) on mustard crop. Egypt J Biol Pest Control 30:105. https://doi.org/10.1186/s41938-02000300-6 Smith BC, Williams RR (1976) Temperature relations of adult Coleomegilla maculata lengi and CM medialis (Coleoptera, Coccinellidae) and responses to ovipositional stimulants. Can Entomol 108(9):925–930 Smith BC, Starratt AN, Bodnaryk RP (1973) Oviposition responses of Coleomegilla maculata lengi (Coleoptera: Coccinellidae) to the wood and extracts of Juniperus virginiana and to various chemicals. Ann Entomol Soc Am 66:452–456 Sota T (2000) Carabus in four seasons: life history evolution and species diversity. Kyoto University Press, Kyoto, 247 pp [in Japanese] Srikumar KK, Bhat PS, Raviprasad TN, Vanitha K, Saroj PL, Ambrose DP (2014) Biology and Behavior of Six Species of Reduviids (Hemiptera: Reduviidae: Harpactorinae) in a Cashew Ecosystem. J Agri Urban Entomol 30(1):65–81 Stearns SC (2000) Life history evolution: successes, limitations, and prospects. Naturwissenschaften 87(11):476–486

References

147

Stride GO (1956) On the biology of certain West African species of Phonoctonus (Hemiptera, Reduviidae), mimetic predators of the Pyrrhocoridae. J Entomol Soc South Afr 19(1):52–69 Sutherland JP, Sullivan MS, Poppy GM (2001) Oviposition behaviour and host colony size discrimination in Episyrphus balteatus (Diptera: Syrphidae). Bull Entomol Res 91(5):411–417 Tallamy DW (2005) Egg dumping in insects. Annu Rev Entomol 50:347–370 Tallamy DW, Horton LA (1990) Costs and benefits of the egg-dumping alternatives in Gargaphia lace bugs (Hemiptera: Tingidae). Anim Behav 39:352–359 Than V, Murali Baskaran RK, Mahadev AN (1999) Biology and predatory potential of green lacewing (Chrysoperla carnea) (Neuroptera: Chrysopidae) on different insect hosts. Indian J Agric Sci 69(7):502–505 Thiele H-U (1977) Carabid beetles in their environments. Springer-Verlag, Berlin, 369 pp. https:// doi.org/10.1007/978-3-642-81154-8 Trefas H, van Lenteren JC (2008) Egg-laying-site preferences of Pterostichus melanarius in monoand intercrops. Bull Insectol 61(2):225–231 Truong XL, Pham HP, Thai TNL (2020) Biology and predatory ability of the reduviid Sycanus falleni Stal (Heteroptera: Reduviidae: Harpactorinae) fed on four different preys in laboratory conditions. J Asia Pac Entomol 23(4):1188–1193 Ugine TA, Krasnoff SB, Grebenok RJ, Behmer ST, Losey JE (2019) Prey nutrient content creates omnivores out of predators. Ecol Lett 22(2):275–283 Vanhaelen N, Haubruge É, Gaspar C, Francis F (2001) Oviposition preferences of Episyrphus balteatus. Med Fac Landbouww Univ Gent 66(2a):269–275 Verheggen FJ, Arnaud L, Bartram S, Gohy M, Haubruge E (2008) Aphid and plant volatiles induce oviposition in an aphidophagous hoverfly. J Chem Ecol 34:301–307 Walker LA, Holwell GI (2016) Sexual cannibalism in a facultative parthenogen: The springbok mantis (Miomantis caffra). Behav Ecol. 27(3):851–856. https://doi.org/10.1093/beheco/arv221 Werner FG, Chandler DS (1995) Anthicidae (Insecta, Coleoptera). Fauna of New Zealand, 34 West AS, DeLong B (1956) Notes on the biology and laboratory reaving of a predatory insect, Zelus exsanguis (Stahl) (Hemiptera: ReduYiidae). In: 86th annual report of the Entomological Society of Ontario, 1955, pp 97–101 Xu X, Cai X, Bian L, Luo Z, Xin Z, Chen Z (2015) Electrophysiological and behavioral responses of Chrysopa phyllochroma (Neuroptera: Chrysopidae) to plant volatiles. Environ Entomol 44: 1425–1433 Yao FL, Lin S, Wang LX, Mei WJ, Monticelli LS, Zheng Y, Weng QY (2021) Oviposition preference and adult performance of the whitefly predator Serangium japonicum (Coleoptera: Coccinellidae): effect of leaf microstructure associated with ladybeetle attachment ability. Pest Manag Sci 77(1):113–125 Yu C, Huang J, Ren X, Fernández-Grandon GM, Li X, Hafeez M, Lu Y (2021) The predatory bug Orius strigicollis shows a preference for egg-laying sites based on plant topography. PeerJ 9: e11818 Zhang R, Ji D, Zhang Q, Jin L (2021) Evaluation of eleven plant species as potential banker plants to support predatory Orius sauteri in tea plant systems. Insects 12(2):162 Zhu P, Lu Z, Heong K, Chen G, Zheng X, Xu H, Gurr GM (2014) Selection of nectar plants for use in ecological engineering to promote biological control of rice pests by the predatory bug, Cyrtorhinus lividipennis, (Heteroptera: Miridae). PLoS One 9(9):e108669

6

Insect Predators Immature Stages Biology

Contents 6.1 6.2 6.3 6.4

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Neuropteran Larva . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Lygeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hemiptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.1 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.2 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.3 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.4 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.5 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.6 Geocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5 Syrphidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.1 Larval General Characters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.2 Pupa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.3 Eclosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.4 Development Periods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.5 Survival . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5.6 Larval Feeding Behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.6 Odonata . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.7 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.8 Coccinelidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.9 Thysanoptera (Thrips) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.10 Mantodea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.11 Tri-Trophic Interactions of Bt Toxins–Preys–Predator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.12 Future Area of Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

150 150 153 154 154 156 158 158 164 166 166 167 168 168 169 169 170 170 171 173 174 175 175 177 177

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_6

149

150

6.1

6

Insect Predators Immature Stages Biology

Introduction

Getting accurate information on natural enemies’ biology and interactions with herbivorous insects is one of the first tasks before they can be used effectively as mass-raised biological control agents. Each predator may have different methods depending on the field conditions, such as the usage of microhabitats or adjusting their activity cycles, to cope with stress conditions like drought. To determine this predator’s applicability as a biological control agent, it is crucial to understand its habits of survival and diet. According to published research, both biotic (related prey’s and predators; plant type) and abiotic (temperature, humidity, water availability, light, wind velocity) circumstances are crucial for the development of various embryonic stages of predators. There are two general laws that can be used to characterise the relationship between body size, fitness, and temperature in animals, particularly in insects. 1. According to the first rule of “hotter is smaller”, ectotherms that mature at higher 290 temperatures will be smaller as adults (Angilletta and Dunham 2003; Kingsolver and Huey 2008), and our findings on wing length are consistent with this hypothesis. 2. According to the second “Bigger is better” rule, people with larger bodies tend to perform better and be more fit than people with smaller bodies, as seen by higher rates of survival, fecundity, and successful mating. This rule says that the size of bodily components may have a detrimental impact on the rise in temperature. There is another rule, though, according to some writers, who contend that organisms that have adapted to high optimum temperatures can also have improved fitness at high temperatures due to a reduced generation time. Interestingly, given the three larval stages developed more quickly with rising temperatures, our findings also provide support to this theory. In regard to diverse biotic requirements, notably natural or laboratory hosts, this chapter attempts to provide information about biological features (young ones developmental periods, their survival rate/mortality) of various predators.

6.2

Neuropteran Larva

Larvae of coniopterygids are predatory. Adults are also mostly predators; however, they can eat honeydew. A wide range of slow-moving arthropod prey, including as mites, scale insects, insect eggs, coccids, aphids, and phylloxerae, have been documented as being consumed by adults and larvae. Instead of herbs, the majority of coniopterygid species are found near woody vegetation (trees or shrubs). It is clear from plant “host” data that at least some coniopterygid species have preferences for particular plant species, which may indicate that these species have arthropod prey preferences that are related to their host plant preferences. Even in temperate climates, many coniopterygid species have multivoltine life cycles. Some

6.2 Neuropteran Larva

151

coniopterygid species may be useful for the biological control of mites and sessile homopterans due to their small size, quick generation time, and feeding preferences. Hypermetamorphic mantispid larvae are present. In other words, the larva in its first instar is elongated and actively seeks out dormant prey. The second and third instars are lethargic, scarabaeiform, and feed on the prey that was discovered in the first instar. With each subfamily comes a different type of larval diet. Symphrasinae larvae have been discovered in solitary wasp cells, Polybiinae wasp nests, and underground scarab pupae. Mantispinae larvae appear to be forced to eat a variety of spider eggs. They use one of two methods to enter the spider egg sacs: either they board an adult spider and then enter the egg sac as it is developing, or they enter the egg sac straight after it has formed. Hemerobiid larvae are predators that feed on sessile homopterans, which are softbodied insect prey (e.g. aphids, whiteflies, and scales). Some larvae appear to have a particular affinity for certain types of prey. Adults seem to be predominantly predators, though some might also eat non-animals. The majority of families have three larval phases, whereas Ithotze fusca and a few Coniopterygidae have been reported to have up to five larval stages each. An American dilarid has been known to go through up to 12 larval stages; however, this could have been caused by artificial raising practices. Larvae have a variety of shapes and behaviours. Although most are aggressive predators, subterranean ithonid larvae appear to eat decomposing plant matter (Gallard 1932). The egg sacs of certain spiders and the nests of social Hymenoptera are hosts to second and third instar larvae of the Mantispidae family, which resemble “grub-like” parasites. The sole berothid (a Nearctic Lomamyia) whose entire life cycle has been fully documented is a termitophile with grub-like larvae in the second instar. Sisyrid larvae eat exclusively on fresh water sponges and possibly sometimes on bryozoans using their extraordinarily long, thin jaws. Sisyrids are the only lacewing larvae with ventral abdominal gills in their older instars. The Psychopsidae and some Berothidae share this habitat with elongate osmylid larvae, which can be found in moist areas along streams or under bark. These larvae also have long needle-like jaws. On plants, Coniopterygid, Hemerobiid, and Chrysopid larvae are commonly seen. Larvae of Myrrneleontoid species can be found in tree bark or litter, and they are frequently hidden. Osmylops and Myiodactylus larvae are flattened, disc-shaped, and hide among vegetation with their jaws open wider than 180°. In sandy soil, myrmeleontid larvae are most common. An aggressive, predatory campodeiform larva with noticeable, projecting jaws is the typical form. Rarely, in families like the parasitic Mantispidae and subterranean Ithonidae that have more sedentary feeding habits, a more scarabaeiform larva can be found. Many Chrysopidae and some ground-dwelling larvae have tangled initial filaments, while others have long, curled hairs or setae that are occasionally gathered on elevated tubercles. All of these features trap material on the dorsal side, assisting in hiding. Many Myrmeleontoidea have distinct lateral processes on the thorax and abdomen, and their arrangement has significant taxonomic relevance. Particularly on the body, Ascalaphidae have elaborately decorated papillae (dolichasters). There have only two tarsal claws and I-segmented tarsi. The hind tibia and tarsus are

152

6

Insect Predators Immature Stages Biology

sometimes combined into a single component in Myrmeleontoidea groups, presumably making backward movement more effective. A sucking tube is formed by the mandible and maxilla fitting together. The maxillary palp is non-existent, and the maxilla is a stylet. Sisyridae lack the typically well-defined labial palps. Except in Coniopterygidae, the labrum is decreased. “Suction discs”, which occasionally have hooked spicules, may be the end of the abdomen. Since the midgut and the hindgut are divided, the larva does not pass any faeces until pupation (New 2004). A hypermorphic larva of mantispids is an elongated, first-instar larva that actively pursues dormant prey. The second and third instars are lethargic, scarabaeiform, and feed on the prey that was discovered in the first instar. With each subfamily comes a different type of larval diet. Symphrasinae larvae have been discovered in solitary wasp cells, Polybiinae wasp nests, and underground scarab pupae. Mantispinae larvae appear to be forced to consume a variety of spider eggs (Hoffman and Brushwein 1989, 1990). Berothid biology is not well understood. Apparently obligate termitophiles, the larvae of Lomamyia laiipennis and Lomamyia longicollis live in the galleries of termites belonging to the species Reticulitermes (Isoptera: Rhinotermitidae). The first instar stages find and enter the termite colony’s gallery system, where they paralyse and consume individual termites. It appears that a toxin is administered into prey with a fast thrust of the straight larval jaws, immobilising it. Only after immobilisation does feeding take place. Only the first and third instar stages are interestingly active and feeding. Larvae in their second instar are dormant and have morphologically diminished legs and mouthparts. While some Chrysoperla larvae have small setae on the body and do not carry any debris, others, including Ceraeochrysa and Leucochrysa, have lengthy thoracic and abdominal setae on tuberacles and carry debris and prey skins all over their bodies. The second kind of larvae is known as trash-bearers, or “bichodelixo” in Portuguese. Before developing into small, rounded, whitish, silken pupal cells on leaf surfaces, larvae go through three instars. Insect eggs as well as soft-bodied adult and larval insects are consumed by larvae. There could be a single generation or multiple. Similar to antlions, larval ascalaphids are powerful predators with long sickleshaped jaws. All ascalaphid larvae seem to be ambush-style, sit-and-wait predators. Some species have been shown to conceal themselves with environmental dust such that they blend in with the substrate (e.g. Ululodes). Other species’ larvae, such as those of Ascaloptynx, Ascalobyas, and Haploglenius, reside in the leaf litter or on the leaves and branches of trees (Henry et al. 1992). Chrysoperla zastrowi arabica’s biology on various preys was captured by Naruka and Ameta (2015). Green lacewing Chrysoperla zastrowi sillemi was fed unsterilised and sterile eggs from Corcyra cephalonica, Aphis craccivora, Aphis gossypii, Brevicoryne brassicae, and Uroleucon compositae while being examined for its biology. On various hosts, Chrysoperla zastrowi sillemi’s total larval life span varied from 9.0 to 11.2 days. The larvae raised on unsterilised Corcyra cephalonica eggs with the first, second, and third instars of Chrysoperla zastrowi sillemi as 2.6, 2.8, and 3.6 days, respectively, were found to have the shortest larval duration of 9.0 days. Sterilised Corcyra cephalonica eggs came next, while Uroleucon

6.3 Lygeidae

153

compositae had the longest duration. Unsterilised Corcyra cephalonica eggs had the highest proportion of Chrysoperla zastrowi sillemi larvae pupating (89.1), followed by sterilised eggs (86.5), and Uroleucon compositae had the lowest percentage (Uroleucon compositae) (74.5). Chrysoperla zastrowi sillemi’s total developmental period was much shorter on Corcyra cephalonica unsterilised eggs and longer on Uroleucon compositae (Mounika et al. 2021). Chrysoperla zastrowi sillemi was found on unsterilised eggs of Corcyra cephalonica, and its pre-oviposition duration was significantly shorter (4.5 days) and its oviposition period was significantly longer (29.6 days). Chrysoperla carnea pre-oviposition period was previously reported by Geethalakshmi et al. (2000) to be 4.0 days when raised on Corcyra cephalonica. Chrysoperla zastrowi sillemi had the highest fecundity when raised on Corcyra cephalonica unsterilised eggs (371.6 eggs), and Uroleucon compositae had the lowest fertility (113.4 eggs). When raised on Corcyra cephalonica eggs, Kubavat et al. (2017) reported 352.9 eggs per female of Chrysoperla zastrowi sillemi. Aphis craccivora was found to have 260.7 eggs, 250.8 eggs by Vivek et al. (2013), 274.7 eggs by Naruka and Ameta (2015), and 250.8 eggs by Nandan et al. (2014).

6.3

Lygeidae

Numerous species of big-eyed bugs are important commercial polyphagous predators. Geocoris nymphs resemble adults but are smaller and lack wings. Some species’ nymphs look like small greyish adults, while others may have a bluishpurple to red tint. Nymphal stages behave and eat similarly to adults; however, they typically consume smaller animals. Young instars are small and simple to miss. Before becoming winged adults, nymphs that hatch from eggs go through five juvenile stages of development. They get bigger, get wing pads, and shed their skin with each new instar. Temperature affects juvenile stage growth, and studies have shown that some species grow more slowly in colder climates. Nymphal development takes place in the laboratory over the course of 30 and 60 days at temperatures of 25 and 20, respectively. Georgia’s cotton fields have recently become more and more home to the species Geocoris floridanus. Although Georgia and Florida seem to be the main distribution hubs, it can be found throughout the southern United States, from East Texas through Florida and farther north to Washington, D.C. A typical geocorid in appearance, Geocoris floridanus is most frequently found on the ground or on low-growing foliage. Small size, agile activity, the ability to quickly vanish in debris and dirt crevices, overall dull coloration, and apparent similarity to other southeastern United States species, particularly Geocoris punctipes, may cause the abundance of this species to be underestimated (Torres et al. 2004). When given maize earworm eggs, Geocoris floridanus nymphs developed more quickly than when given Spodoptera exigua beet armyworm larvae. In comparison to bugs given Spodoptera exigua larvae, nymphs fed maize earworm (Heliothis zea) eggs developed through each instar around 1 day faster, and the entire length of the

154

6

Insect Predators Immature Stages Biology

nymphal period was 4 days shorter. The sex ratio of emerging adults (number of females/total of adults) was 0.45 and 0.5 for nymphs fed Heliothis zea egg and Spodoptera exigua larvae, respectively, showing no difference between prey treatments. The average death rate for Geocoris floridanus nymphs in both prey treatments was rather low: for nymphs fed corn earworm eggs, it was 19.7%, and for nymphs fed beet armyworm neonates, it was 21.2%. Only the first three instars of Geocoris floridanus nymphs died in both feeding treatments. Additionally, there was no discernible difference in prey treatments or for the entire nymphal cycle in stagespecific nymphal mortality (Torres et al. 2004). The eggs of Sitotroga cerealella that had been exposed to UV light and the nymphs of Phenacoccus solenopsis were given to the young of Geocoris superbus. The Geocoris superbus had an incubation time of 10 days, nymphal durations of 7, 5.2, 5.2, 5.6, and 8.2 days, and a total nymphal period of 31.2 days. There was a 9-day pre-oviposition interval. The stages of development for males and females were quite similar. Hatching percentage was 87. In the lab, the sex ratio was practically evenly balanced at 1.03:1.00 (female: male) (Varshney and Ballal 2017). Geocoris erythrocephalus nymphs resemble adults but are smaller and lack wings. While nymphal stages often eat smaller prey Brevicoryne brassicae, they exhibit similar behaviour and feeding patterns to adults. Five nymphalin stars have been experienced by bugs in 30–35 (mean 32.4) days. I, II, III, IV, and V instar nymphs lived for an average of 4.40–0.52 days, 8.50–0.53 days, 6.70–0.48 days, 7.60–0.52 days, and 5.20–0.42 days, respectively (Rajan et al. 2018). The current findings are consistent with those of Ramirez et al. (2011), who observed that nymphs emerge and develop over a period of 3–4 weeks.

6.4

Hemiptera

6.4.1

Nabidae

Myers investigated the biological notes on Arachnocoris albomaculatus (Hemiptera; Nabidae) (1925). The tomato leafminer, Tuta absoluta, has a natural enemy in the predatory bug Nabis pseudoferus (Lepidoptera: Gelechiidae). Pagasa fusca nymphs resemble ants in certain ways (Harris 1928). First moults are experienced by Nabis ferus nymphs; the first four instars take an average of 3 days apiece, while the last instar takes about 6 days (Harris 1928). The average nymphal duration of Nabis alternatus was 16.1 days under laboratory circumstances (28 °C, 59% RH, and 15 h of light each day), with a total nymphal mortality of 67.8% (Perkins and Watson 1972). Research on the Pacific damsel bug, Nabis kinbergii, was done in a lab setting (27–28 °C). When provided with Australian crop mirid (Sidnia kinbergi), pea aphid (Acyrthosiphon pisum), and diamondback moth (Plutella xylostella), respectively, the time it took for development from egg to adult was 20.9, 22.5, and 20.9 days (Siddique and Chapman 1987). Using laboratory settings of 26–28 °C, 60–70% RH, and 15 h of photophase each day, Tropiconabis capsiformis (Nabidae) was researched. The average nymphal

6.4 Hemiptera

155

duration was 18 days for males and 22.4 days for females, resulting in a generation time of 51.6 days (Hormchan et al. 1976). Nabis capsiformis’ biological features were examined in a lab setting at 26 °C, 65°RH, and a photoperiod of 16:8 (L:D) h. The overall nymphal period was 7.2 days, while the complete developmental time was 14 days. Of Nabis capsiformis immatures, 23.93% reached the adult stage. In Nabis capsiformis, the developed immatures were 29.63% (Fathipour and Jafari 2008). Numerous aphid species, various hemipteran species, lepidopteran eggs, and larvae like tomato pinworm are also essential prey items for Nabis pseudoferus. By feeding on Ephestia kuehniella egg, Madadi et al. (2016) investigated the life table parameters of Nabis pseudoferus in the laboratory circumstances of 25 °C, 65° RH, and 16:8 (L: D) photoperiod. According to the findings, the mean lifespan of eggs, first, second, third, fourth, and fifth nymphs, as well as adults, was 11.2, 3.6, 4.0, 3.5, 3.9, 7.4, and 33.3 days, respectively. Male and female pre-adult periods lasted 65.4 and 69.6 days, respectively. Male and female first to fifth instar nymph Nabis pseudoferus eggs had a survival rate of 1, 0.8679, 0.8616, 0.5472, 0.283, and 0.2453 correspondingly. The rates of birth, mortality, and survival for this predator were calculated to be 0.017, 0.983, and 0.061, respectively (Madadi et al. 2016). In greenhouse tomatoes, Tuta absoluta is a pest that can be augmented with Nabis pseudoferus. When Trichogramma brassicae parasitism and egg age increased, the contents of the eggs’ proteins, lipids, and glycogen all decreased while the contents of carbohydrates increased. These changes were reflected in the nutritional makeup of adult predators. These effects were assessed by Mohammadpour et al. (2020) in a laboratory setting. Nymphs could not survive on Trichogramma brassicae parasitised eggs that were older than 24–48 h, although juvenile survival on healthy eggs was not different. In comparison to a diet of healthy eggs of the same age, growth was slowed down, the pre-oviposition phase was prolonged, and lifelong fertility was decreased in nabids fed eggs that were 24 or 48 h old and parasitised (Mohammadpour et al. 2020). An important insect pest of vegetable crops is the damsel bug, Nabis pseudoferus (Namidae), a generalist predator of small arthropods. When feasting on the eggs and first- and fourth-instar nymphs of the invasive South American tomato pinworm, Tuta absoluta, Mahdavi et al. (2020) described the predation and development of several Nabis pseudoferus life stages (Lepidoptera: Gelechiidae). Mahdavi et al. (2020) also made comparisons with the cotton aphid, Aphis gossypii, a common victim (Hemiptera: Aphididae). Due to their long lifespan, huge size, and high energy needs for reproduction, females in all life stages examined had the greatest predation rate on all tested species. Predator oviposition rates were most frequently laid by fourth-instar Tuta absoluta, which was its preferred prey. These findings demonstrate that Nabis pseudoferus can be regarded as a significant indigenous natural enemy for sustainable pest control strategies against Tuta absoluta in newly invaded areas given the zoophytophagy of most of the life stages of other mirid species of tomato and the lower propensity of Nabis pseudoferus to feed on plants (Mahdavi et al. 2020).

156

6

Insect Predators Immature Stages Biology

The development of insects is greatly influenced by the various seasons. The Laboratory studies have established this fact. For instance, under controlled laboratory conditions, the predator Nabis consimilis life cycle was studied during the summer (26.1±0.5 °C; HR 51.4±0.9) and winter (20.4±1.0 °C; HR 60.3±2.7). In comparison to winter, summer has a stronger capacity for development, oviposition, pre-oviposition, oviposition, and post-oviposition. The I, II, III, IV, and V instars had an incubation period, viability, and length of 7.9 days, 41.7%, 4.6 days, 3.6 days, 3.2 days, 2.9 days, and 4.1 days, respectively, in the summer and 13 days, 48.2%, 6 days, 5.2 days, 4.8 days, 5.3 days, and 6.4 days, respectively, in the winter. The biggest mortality rates occur in the summer in the II and III instars and in the winter in the IV and V instar. In the nymphal instars, cumulative mortality was higher in the summer than in the winter. Similar sex ratios could be seen at both locations (Salcedo et al. 2020).

6.4.2

Pentatomidae

Predaceous species are a secondary characteristic of pentatomids, and they belong to the subfamily Asopinae. The asopines’ life cycle is comparable to that of phytophagous pentatomids. Adults appear, begin to feed and reproduce utilising various food, and some species can survive for up to 3 months. They find and identify their prey through tactile, chemical, and visual signals. Similar to phytophagous species, males initiate courtship through antennal movement during mating. Eggs are placed in large numbers on various plant sections that adults use or on surrounding buildings. Different species, people, or clutches from the same mother lay different numbers of eggs in each. Five nymphal instars are present. First instars merely require moisture to survive and are not predatory. They must consume prey to survive after the second instar. When nymphs first emerge, they are sociable and have a tendency to attack in packs. As they develop, however, they become less gregarious and spread out in search of food. The influence of food and abiotic conditions (temperature and humidity) impacts the expression of these qualities, and these traits include incubation period, nymphal development time, pre-reproductive period, and longevity (Jocélia Grazia et al. 2015). Alcaeorrhynchus, Apateticus, Euthyrhynchus, Brontocoris, Perillus, Podisus (Podisus maculiventris and Podisus nigrispinus), Stiretrus, Supputius, and Tylospilus are some of the genera that include predatory insects. One of the most prevalent asopine species in the Neotropics is Podisus nigrispinus (Pentatomidae), with reports of its prevalence in various South and Central American nations. Independent of the instar and sex of the predator, the mean development periods of Podisus nigrispinus fed with third instar cotton leaf worm larvae (4.25 days), fifth instar cotton leaf worm larvae (4.1 days), or Tenebrio molitor larvae (4.1 days) were comparable (Lemos et al. 2003). Conflicting results among studies using the same prey type, according to De Clercq et al. (1998), may be due to variations in the studied strains’ geographic origins, their adaptation to the

6.4 Hemiptera

157

Table 6.1 Influences of Spodoptera frugiperda, Helicoverpa armigera, Tenebrio molitor and Oecophylla smaragdina on Eocanthecona furcellata nymphal periods (days) with references Predator Eocanthecona furcellata-male Eocanthecona furcellatafemale Eocanthecona furcellata

Eocanthecona furcellata

Preys Spodoptera frugiperda Spodoptera frugiperda Helicoverpa armigera Tenebrio molitor Oecophylla smaragdina

Nymphal developmental period (days) First Second Third Fourth Fifth 2.4 3.3 3.3 4.0 4.4 2.5

3.5

4.0

4.2

5.1



4.3

3.6

3.8

4.8



3.7

3.7

4.3

7.3

4.1

3.4

3.0

3.1

5.2

References Sravika et al. (2020) Sravika et al. (2020) Simonato et al. (2020)

Rustam and Gani (2019)

diet under laboratory conditions, or, alternatively, the prey’s diet’s composition may have an impact on the predator’s life history parameters. A South American polymorphic species known as Stiretrus decemguttatus (Pentatomidae: Asopinae) preys on Chrysomelidae. On the Marajó Island, Brazil, Stiretrus decemguttatus is a significant predator of the cassidine beetle species Botanochara sedecimpustulata and Zatrephina lineata (Coleoptera: Cassidinae). The development time increases from the first to the fourth instars, varying between 2.4, 2.6, 2.6, and 3.3, which were relatively minor increases in comparison to the fifth instar’s 6-day increase (Paleari 2013). Eocanthecona furcellata is a pentatomid bug that is widespread in the eastern region, particularly in south-east Asian nations like India, China, Indonesia, Taiwan, and Japan. It is an effective generalist predator of Lepidoptera, Coleoptera, and Heteroptera, and it has been frequently seen in cotton, chickpea, groundnut, and vegetable crops. Eocanthecona furcellata feeds on Oecophylla smaragdina, which has five instars, with the first nymphal stage being 4.1 days, the second lasting 3.4 days, the third lasting 3.0 days, the fourth being 3.1 days, and the fifth being 5.3 days (Rustam and Gani 2019). Eocanthecona furcellata developmental cycles are influenced by prey (Table 6.1). Andrallus spinidens is a widely distributed species that is found in the Oriental (China, Taiwan, Philippines, Japan, Malaysia, Indonesia, Vietnam, Bangladesh), Australasian (Australia, New Guinea), and Afrotropical regions. It has also been found in parts of Europe and North America, as well as in India in the states of Jammu and Kashmir, Karnataka, Uttaranchal, and West Bengal. The majority of Pentatomidae (Hemiptera: Asopinae) stinkbug species feed on other insects, such as sawfly larvae (Hymenoptera: Symphyta). Although the Argidae and Pergidae sawfly larvae contain poisonous peptides, it is unclear whether or not they are protected from stinkbugs. In a laboratory setting, interactions between Picromerus bidens and four sawfly species—Arge ochropus, Arge pagana, Lophyrotoma zonalis, and Allantus rufocinctus—at the last larval instars were

158

6

Insect Predators Immature Stages Biology

examined. Allantus rufocinctus larvae did not survive, based on their interactions, whereas most or all of the larvae of the other sawfly species did and were still alive 48 h later. Only 6–20 s were spent during the average feeding time of an argid or pergid larva before some bugs removed their rostrums and rapidly withdrew. Arge pagana larvae that were fully grown were less likely to be attacked. The examined Argidae and Pergidae are probably well protected by Picromerus bidens by strong internal anti-feedants, with defensive body movements and a big body size playing a supporting role (Boevé 2021).

6.4.3

Miridae

I–V of Deraeocoris nebulosus although instars were commonly observed on the same plant, we were able to clarify seasonal history through periodic sampling from several landscape plantings. Mid- to late-May saw the emergence of the first instars. Instars I through IV were present by early June, while instars 11 through 111 dominated. Between late-June and late-July, nymphs of the second generation began to emerge. Mid- to late-August saw a predominance of instars IV–V. Early third-generation instars proliferated in large numbers in our September samples. On 11 October, instars IV–V were still prevalent at the weekly sampling location (Wheeler et al. 1975). In the Valencia region (East Coast of Spain), the omnivorous predator Dicyphus maroccanus (Hemiptera: Miridae) was first observed in tomato fields in 2009. Dicyphus maroccanus nymphs emerged 12.2 days (n = 153) following oviposition. On tomato plants, male and female immature Dicyphus maroccanus fed Ephestia kuehniella (Lepidoptera: Pyralidae) eggs developed over the course of about 19.5 days, with an immature survival rate of 85% from the first instar to the adult stage. With no discernible variations between the sexes, Dicyphus maroccanus nymphs required the consumption of 267 and 312 eggs, respectively, to mature into adult male and female stages. When fed only on tomato plants, Dicyphus maroccanus was unable to complete juvenile development and the majority of nymphs perished between the second and third nymphal instars (Abbas et al. 2014). Nesidiocoris volucer, a tropical native mirid, is an essential biological pest control for whiteflies and other insect pests on tomato crops under greenhouses in France (Marquereau et al. 2022).

6.4.4

Reduviids

Ectomocirs xaviereii (Heteroptera Reduviidae Piratinae) (Vennison and Ambrose 1990): Black overall with brownish ochre antennae, femora and tibia apices, and rostrum tip; except in the first stage luteous, the second and seventh abdominal

6.4 Hemiptera

159

segments have oval spots on their dorsums. The abdomen is sanguineous in the early stages and is lengthy and delicately pilose. The declivous anteocular part and the upright postocular portion of the head are separated by a transverse sulcus, and the base of the anteocular portion has an ancient sulcus that gets smaller as it gets closer to the apex. Antennas placed close to compound eyes; rostrum slightly curved and its tip touching the prosternal groove; sparsely longly pilose; neck evident; pedicel slender; second flagellar segment longest; finely sparsely and longly pilose; antenniferous tubercles 1 at base of each antenna noticeable. Legs multi-coloured; fore and mid tibiae with fossula spongiosa covering nearly two-thirds of their under surface; pronotum monochrome, transverse behind the middle; anterior lobe very much sculptured longitudinally; anterolateral and posterolateral pronotal angles obtuse; long and finely pilose; tarsus three segmented; densely longly pilose. All instars have an abdomen that is longer than it is wide, with the exception of the first stage, which has oval fasciae on the dorsum of the second and seventh abdominal segments. All nymphal stages have six lateral spots on the connexivum that are delicately and longly pilose (Vennison and Ambrose 1991). The biology and behaviour of Ectomocoris xaviereii were also investigated by the authors. Post-embryonic development: Compared to Coranus nodulosus (Sahayaraj and Ambrose 1993) and Coranus vitellinus, the egg incubation period, the overall nymphal developmental duration, and adult longevity of both sexes were the lowest (Ambrose and Livingstone 1985). The overall stadial period from I to V instar lasted for 22.1 days. The stadial period of I, II, III, IV, and V instars, respectively, lasted for 3.9, 4.2, 3.9, 4.3, and 6.4 days. The fifth instar of Coranus spiniscutis had the longest stadium, and the third instar had the shortest (Claver and Reegan 2010). The line diagrams of the Alleocranum quadrisignatus (Fig. 6.1) and Endochus umbrinus (Fig. 6.2) at the egg, first instar, second instar, third instar, fourth instar, and fifth instar stages are shown below. When Cydnocoris gilvus nymphs first emerged from their eggs, they were delicate and quickly got browned. They then began to feed, preferring to eat little, slow-moving larvae. The I, II, III, IV, and V instars of Cydnocoris gilvus had developmental lifespans of 9.5, 6.3, 4.5, 6.0, and 11.0 days, respectively. Compared to other instars, the initial instar’s survival percentage was lower. Females developed more quickly (36.7 days) than males (38.0 days) (Srikumar et al. 2014). Panthous bimaculatus also showed similar morphological features during its nymphal development (Fig. 6.3). This assassin insect has traits with other piratine reduviids, including having the longest fifth nymphal stage and an earlier male emergence. Additionally, the reasons for nymphal mortality in this beetle are comparable to those documented for other piratine reduviids (Vennison and Ambrose 1991). The findings demonstrate that different prey species, such as Pieris rapae, Spodoptera litura, Plutela xylostella, and Corcyra cephalonica, had different effects on the reduviid Sycanus falleni raised in a laboratory, but not on egg development, hatching eggs, the development stage of III, IV, and V nymphal instars, life cycle, male and female longevity, or the number of eggs laid. However, when the adults were raised by four different preys, the female reduviid Sycanus falleni lived longer than the male. Reduviid Sycanus falleni

160

6

Insect Predators Immature Stages Biology

Fig. 6.1 Line diagram of Alleocranum quadrisignatus egg, first instar, second instar, third instar, fourth instar, and fifth instar

Fig. 6.2 Line diagram of Endochus umbrinus egg, first instar, second instar, third instar, fourth instar, and fifth instar

6.4 Hemiptera

161

Fig. 6.3 Micro-photographs showing Panthous bimaculatus, first instar (a), second instar (b), third instar (c), fourth instar (d), and fifth instar (e) nymphs Table 6.2 Influences of different preys on nymphal developmental periods (days) and Sycanus annulicornis male and female weight (mg) Life traits Incubation period Total nymphal period (days) Male longevity (days) Male body weight (mg) Female body weight (mg)

Crocidolomia pavonana 20.3 74.0

Tenebrio molitor 14.1 80.1

81.0

44.0

140.6

122.5

309.9

251.8

Plutella xylostella

Corcyra cephalonica

118.26

130.48

83.47

63.99

598.1

627.0

life cycle and total developmental stage as a nymph were much longer when raised by Corcyra cephalonica larvae than when raised by Pieris rapae, Spodoptera litura, and Plutela xylostella larvae (Truong et al. 2020). A reduviid predator called Sycanus reclinatus lives in southern India’s tropical evergreen woods. After 22 days from emergence, females begin to lay clusters of brown eggs. The eggs open in 14–23 days, and the pale-orange nymphs transform deep-orange in about 1 h. From I instar through adulthood, the whole stadial period spans from 61 to 90 days at 32 °C (Vennison and Ambrose 1992). Male and female Sycanus reclinatus lifespans range from 5 to 54 days and 5 to 50 days, respectively. The gender ratio is marginally in favour of men (Vennison and Ambrose 1992). According to Sahid and Natawigena (2018), Crocidolomia pavonana (Lepidoptera: Crambidae) is an appropriate prey species for Sycanus annulicornis nymph rearing (Table 6.2). First instar Sycanus dichotomus nymphs had a mean lifespan of 16.7 days when fed Plutella xylostella and 24.4 days when fed Corcyra cephalonica. When fed with Corcyra cephalonica and Plutella xylostella, the mortality at this stage was 17.25% and 14.4%, respectively. It took the second instar less time to moult into the third instar. When fed with Corcyra cephalonica, it only took 16.9 days, and when fed with Plutella xylostella, it only took 15.8 days. Except for the body, the colours were similar to those of the early instar. When fed with Corcyra cephalonica, first and second instar nymphs had greater body weights (Zulkefli et al. 2004). With 83.5 and 87.6 days, respectively, the male and female Sycanus dichotomus adults who were fed Plutella xylostella lived longer. For the male, the range was 3–167 days, while

162

6

Insect Predators Immature Stages Biology

for the female; it was 5–153 days. Adult Sycanus males and females fed Corcyra cephalonica lived 64 and 61.9 years longer, respectively (Zulkefli et al. 2004). By 21.3% and 2.8%, respectively, compared to Tenebrio molitor (Coleoptera: Tenebrionidae) (115 days) and Corcyra cephalonica (Lepidoptera: Pyralidae) (98 days) larvae alone, Sycanus dichotomus total nymphal stage duration (95 days) fed on a combination of both larvae was significantly lower. The average life span of Sycanus dichotomus that consumed Corcyra cephalonica larvae was significantly shorter—154 days—than that of Sycanus dichotomus that consumed Tenebrio molitor larvae or both—184 days and 164 days, respectively. Tenebrio molitor larvae came in second because they remained active for the longest amount of time during the fifth nymphal stage, which allowed the predator to develop at a slower rate overall and live longer. With the exception of the fifth nymphal stage, Sycanus dichotomus growth and development after the third nymphal stage had significantly slowed down compared to those of other food sources in terms of body weight and femur length. However, due to the shorter nymphal stage duration and higher body weight, Corcyra cephalonica represents a favoured diet for the first and second nymphal stages, but it was not a good food source for the third nymphal stage to adults. Early nymphal stages of Corcyra cephalonica were found to be easily consumed because their outer skin is silky soft and easy to penetrate by the younger predator than Tenebrio molitor larvae (Ahmad and Kamarudin 2016). When Sycanus dichotomus was fed both insect larvae, their weight and other body parts increased, and the adults lived noticeably longer and in better condition (Ahmad and Kamarudin 2016). Sphedanolestes variabilis produced single light brown eggs as well as little clusters of 5–10 eggs. As with other Sphedanolestes species, Sphedanolestes variabilis eggs were placed individually as well as in tiny clusters of 5–10. About 34.3 eggs were deposited by Sphedanolestes variabilis. When compared to Sphedanolestes himalayensis (74.8 eggs), Sphedanolestes pubinotum (54.0 eggs), and Sphedanolestes minusculus (63.3 eggs), Sphedanolestes variabilis had a higher fecundity than Sphedanolestes signatus (15.3 eggs) (Ambrose et al. 2009). Prior to hatching, the fertilised eggs took on a crimson hue while the unfertilised eggs shrank. 6.9 days were spent in the incubation stage. Nymphs in the I, II, III, IV, and V instars had stadial lengths of 6.8, 6.8, 6.6, 8.4, and 8.7 days, 19.7 days, and 19.7 days, respectively. Adult males and females of Sphedanolestes variabilis had a sex ratio of 1:1 and lived for 93.8 and 102.8 days, respectively (Ambrose et al. 2009). It was less than the 66.70 days of its sister species Sphedanolestes pubinotum and nearly as short as the 46.9 days of Sphedanolestes signatus, Sphedanolestes himalayensis, and Sphedanolestes minusclus (50.4 days). The nymphal instars had a 60% survival rate due to abnormal hatching and moulting, which resulted in a 40% nymphal death rate from I to V instar. Sphedanolestes variabilis had a lower nymphal mortality rate (89.3%) than Sphedanolestes pubinotum, but a higher nymphal mortality rate than Sphedanolestes minusculus (21.1%) and Sphedanolestes himalayensis (13.0%) (Ambrose et al. 2009). Teneral nymphs were frail when they first emerged from the egg, but 3–4 h later, their cuticles toughened. After that, they began to eat, demonstrating a preference for

6.4 Hemiptera

163

little, lame larvae. The reduviid species’ nymphal instars varied widely in size and shape as they grew. First instar development took Euagoras plagiatus (5.2 days) about 2 days less time than it did for the other species. The initial instars of Endochus albomaculatus (15.0 days) and Irantha armipes (12.2 days) exhibited the slowest rates of development. Sphedanolestes signatus (6.8 days), Euagoras plagiatus (7.8 days), and Epidaus bicolor (8.2 days) all developed substantially more quickly than the other species during the second instar stage. When compared to Euagoras plagiatus (7.0 days) and E. bicolor, Panthous bimaculatus (11.2 days) had a much longer developmental period (7.6 days). As fourth instars, Panthous bimaculatus (9.8 days) and Endochus albomaculatus (10.1 days) matured noticeably more slowly than other species. None of the six reduviid species significantly differed in the time from the fifth instar to adult emergence. Between the first nymphal instar and adult emergence, the total postembryonic developmental period varied greatly among species, ranging from 36.8 days for Euagoras plagiatus to 57.5 days for Endochus albomaculatus. In comparison to Epidaus bicolor, Sphedanolestes signatus, Irantha armipes, Panthous bimaculatus, or Endochus albomaculatus, Euagoras plagiatus nymphal development was substantially quicker. Sex ratios in Irantha armipes, Sphedanolestes signatus, and P. bimaculatus were skewed toward females, whereas those in Euagoras plagiatus, Epidaus bicolor, and Endochus albomaculatus were skewed toward males (Srikumar et al. 2014). The reduviid species’ cumulative survival rates varied considerably. Endochus albomaculatus (13.2%), Epidaus bicolor (15.0%), and Irantha armipes (22.2%) had the lowest rates of survival, while Euagoras plagiatus (54.2%), Sphedanolestes signatus (48.2%), and Panthous bimaculatus (47.8%) had the highest rates. The six reduviid nymphal instars’ rates of survival differed greatly. Euagoras plagiatus (59.5%), Panthous bimaculatus (53.2%), and Sphedanolestes signatus (52.4%) had the highest first instar survival rates. Endochus albomaculatus (28.1%) and Epidaus bicolor (25.2%) had considerably lower second-instar survival rates than Euagoras plagiatus (63.8%) and Sphedanolestes signatus (67.6%), respectively. Comparable to Sphedanolestes signatus (68.0%) and Euagoras plagiatus (74.6%), I. armipes (52.8%) and Panthous bimaculatus (53.6%) had lower third instar survival rates. Sphedanolestes signatus (74.2%) and Euagoras plagiatus (78.4%) had considerably higher fourth instar survival rates. Euagoras plagiatus had a considerably greater fifth instar survival rate (79.8%) than the other species (Srikumar et al. 2014). The dark brown, elongate eggs of Coranus spiniscutis have characteristic rounded opercula and are often placed singly or sporadically in tiny clusters on plant foliage. As described in Edocla slateri (Vennison and Ambrose 1986), Acanthespis philomanmariae and Coranus soosaii, and in the laboratory, eggs were located on tissue paper, the bottom and sides of the culture container, or on the cotton swabs (Vennison 1989). After each mating, a female produced an average of 16.9 eggs. Over the course of her lifetime, a female will typically lay between 150 and 220 eggs (mean 173.7). In Coranus spiniscutis, the pre-oviposition period was 6.2 days long, and the index of oviposition days was 60.3 days. Incubation and hatching: Under laboratory settings, the incubation period lasts roughly 3–5 days

164

6

Insect Predators Immature Stages Biology

(mean 4.7 days) from June to September, and 5–6 days from October to November (Claver and Reegan 2010). Males outlived females in terms of adult longevity and sex ratio. Female adult life expectancy was 83.7 days and male adult life expectancy was 74.5 days. The sex ratio was skewed towards men (1:0.8). Nymphal mortality: According to Coranus vitellinus (Ambrose and Livingstone 1985), the first instar had the greatest rate of nymphal mortality at 20%, followed by the second instar at (4%). No one died in the later stages (III, IV, and V instars). Not only has the prey played a significant role in the development of a predator, but the predator’s food has also been crucial. For instance, Corcyra cephalonica, a laboratory host of Rhynocoris marginatus, was raised in four different categories: 1. 2. 3. 4.

Corcyra cephalonica reared on jower medium fed group (JFC). Corcyra cephalonica reared on rice medium fed group (RFC). Corcyra cephalonica of sorghum medium fed group (SFC). Corcyra cephalonica.

The results show that the JFC nymphal instars had the shortest overall stadial period, which was 38.5 days. Sorghum medium fed group came in second place with 40.1 days, followed by rice medium fed group with 41.0 days, and wheat medium fed group with 42.0 days. They were not, though, statistically significant. The short total developmental period seen in the JFC group may be due to the predator’s ability to feed on people with fever (19.37 preys/predator) with less energy expended. Rhynocoris marginatus developed over a total of 38.82 days, spending 89.05 days on Spodoptera litura larvae and 100.97 days on Corcyra cephalonicae. Rhynocoris marginatus nymphs in our experiment developed on average during the course of roughly 40 days, which is 0.4 days less than George (1999) observed for Rhynocoris marginatus nymphs reared alone on Corcyra cephalonica. The variations are most likely caused by the various trial designs and food varieties. For instance, Ambrose et al. (2013) discovered that Rhynocoris marginatus’s development period was significantly influenced by the type of prey. According to this study, group rearing was important to accelerate nymphal development. The first stadium among the five was the shortest, with the exception of JFC, while the fifth stadium was the longest (Shahayaraj and Sathiamoorthi 2002).

6.4.5

Anthocoridae

Orius are omnivorous predators that naturally exist in a variety of field crops (Hemiptera: Anthocoridae). Their nymphs and adults attack lepidopteran eggs and young larvae in addition to a variety of tiny arthropods like thrips, aphids, mites, and whiteflies. Orius sauteri eggs from adults that had consumed Megalurothrips usitatus (Thysanoptera: Thripidae) had a high hatch rate, and the ensuing nymphs had great survival rates at 26 °C, according to Liu et al. (2018) research. Orius sauteri adults exhibited a lengthy oviposition period while feeding on

6.4 Hemiptera

165

Megalurothrips usitatus. Orius sauteri eggs had a 59.7% hatch rate. Orius sauteri nymphs’ first instar mortality was higher than their fourth or fifth instar mortality rates. Orius sauteri’s average development period was 15.9 days (Liu et al. 2018). Aphis craccivora, Planococcus citri (Hemiptera: Pseudococcidae), and Ephestia kuehniella were used to raise Oruis albidipennis (Lepidoptera: Pyralidae). When Oruis albidipennis fed on eggs of Ephestia kuehniella or nymphs of Aphis craccivora, their total embryonic period was much shorter than when they fed on nymphs of the citrus mealybug, Planococcus citri. For Ephestia kuehniella eggs or Aphis craccivora nymphs, it was 2.3 or 2.8 days, respectively. In the event of feeding the predator on Planococcus citri nymphs, which reached 6.1 days in length, the total embryonic time was noticeably slower. When insects were permitted to feed on Aphis craccivora nymphs 2.1 as opposed to Ephestia kuehniella eggs 3.0 or Planococcus citri nymphs 3.5, it revealed that the developmental time for the predator’s first nymphal instar was much shorter. In comparison to insects allowed to feed on Planococcus citri nymphs, the developmental time for the second nymphal instar was much longer while feeding on Aphis craccivora nymphs 1.5 and Ephestia kuehniella eggs 2.1 without any significant differences between them. The third nymphal instar’s developmental time was shown to be much shorter when insects were fed Aphis craccivora nymphs at instar 2.1 as opposed to Ephestia kuehniella eggs at instar 2.7 or Planococcus citri nymphs at instar 3.5, with a significant difference between all treatments. When eating on Aphis craccivora nymphs (1.58) and Ephestia kuehniella eggs (2.3), the developmental time for the fourth nymphal instar was much reduced. In contrast, when insects were permitted to feed on Planococcus citri nymphs, the developmental time increased to 3.9. When insects were fed Aphis craccivora nymphs 2.5 instead of Ephestia kuehniella eggs 2.4 or Planococcus citri nymphs 3.6, the predator’s fifth nymphal instar developed substantially faster. There was a significant difference between all treatments. Oruis albidipennis data revealed that the predator’s total developmental nymphal instars period was significantly shorter when it fed on Aphis craccivora nymphs (11.9) compared to insects allowed to feed on Ephestia kuehniella eggs (15.3) or Planococcus citri nymphs (18.5), with a significant difference between all preys (Amer et al. 2021). Amphiareus constrictus, a tiny pirate insect (Hemiptera: Anthocoridae) whose biology has received less attention, was successfully raised in a lab for more than 20 generations. Its biology, life table characteristics, and predatory potential were studied. In addition to the adult, live photos are used to describe the egg and all five nymphal instars, which can be used as supplementary characters for identification. At various temperatures, the biological parameters were examined. Hatching and nymphal survival to adult stage were determined to be optimal at rearing temperatures of 25 and 30 °C (88.5% and 75%, respectively). At 25 °C, life table metrics like the hypothetical F2 females and the net reproduction rate (R0), precise intrinsic rate of increase (rm), and finite rate of increase, were greater. A nymph could eat 32.8 Corcyra cephalonica eggs in total. A total of 179.0 and 388.5 Corcyra cephalonica eggs were devoured by the male and female adults, respectively. The standardisation of a straightforward production procedure for A. constrictus using

166

6

Insect Predators Immature Stages Biology

the eggs of the rice moth Corcyra cephalonica was made possible by the data gathered from studies on biology, fertility, and predatory capacity. According to studies on Tuta absoluta predatory capabilities, a nymph could eat up to 154.8 eggs while she was still a nymph. Adult males and females may each eat 1280.5 eggs, or 1435 eggs total. The mass-raised Amphiareus constrictus can be tested in the field against the brown plant hopper Nilaparvata lugens and the tomato pinworm TT absoluta (Ballal et al. 2019).

6.4.6

Geocoridae

As a control diet, Geocoris ochropterus was given pupae of Oecophylla smaragdina, Bombyx mori, and Aphis gossypii. Results demonstrated that on all diets examined, Geocoris ochropterus nymphs matured into adults. Oecophylla smaragdina, Bombyx mori, and Aphis gossypii were fed for a total of 35.1 days, 35.9 days, and 36.0 days, respectively, throughout the typical development period. Except for females raised on ant pupae, food had little effect on the size of several body parts such head width, body length, forewing length, and fresh body weight of adults. Adults who consumed the various examined diets showed no changes in the sex ratio (Ngoc Bao Chau et al. 2021).

6.5

Syrphidae

Similar to other Cyclorrhaphous Diptera, the larvae of Syrphus luniger, Syrphus balteatus, Syrphus ribesii, Catabomba pyrastri, Sphaerophoria flavicauda, Sphaerophoria scripta, and Platychirus scutatus pass through three phases that are clearly separated by two moults (Bhatia 1939). The majority of syrphids typically develop between 20.10 and 25.80 days at a temperature of 26.2 °C (Joshi et al. 1999). Ischiodon scutellaris pupal stage was considerably shorter (5.7 days) than Episyrphus balteatus pupal stage (8.7 days). However, Ischiodon scutellaris and Episyrphus balteatus larvae of the third instar lasted the longest (4.70 and 4.4 days, respectively), followed by those of the second instar (3.5 and 3.5 days, respectively), and finally those of the first instar (2.7 and 2.6 days, respectively) (Faheem et al. 2019). According to review of literatures, Eupeodes corollae preys on at least 64 Aphididae species as well as a few Lepidoptera and Thysanoptera species. The larval stage’s typical developmental period was 5.6 days. Only 3 people died during the second instar out of the 100 who emerged. The pupal stage, which was survived by 86 people, lasted an average of 6.1 days. Therefore, Eupeodes corollae pre-adult stage lasted on average of 13.7 days. The length of the pre-adult stage did not significantly differ between males and females (Lillo et al. 2021). When Myzus persicae was the prey, the pre-adult stage’s length of Eupeodes corollae was significantly shorter than when other aphid species were employed (Pu et al. 2019). The size, quantity, and kind of prey may also have an impact on the length

6.5 Syrphidae

167

of the life cycle of Eupeodes corollae (Putra and Yasuda 2006; Almohamad et al. 2009).

6.5.1

Larval General Characters

The general characteristics of syrphid larva, which can be used to distinguish the family Syrphidae, are as follows: The head and mouth are not particularly noticeable, and in the majority of species, they are retractile. There may be obtrusive antennae and retractile parallel mouth hooks. Additionally, there are tiny, laterally extending mouth spines that are sclerotised. The abdomen is divided into nine wrinkly segments that would be difficult to identify if not for the 12 setae, or spines, that are typically noticeable and clearly situated. The setae are referred to as median, dorsal, dorso-lateral, lateral, and two ventro-laterals laterally from the meson. The entire exoskeleton may be covered in many microspines. Although present, rectal gills are barely noticeable. Two of the caudal spiracles are connected and essentially united. There may also be intra-spiracular nodules, setae, and lamellae in addition to the three slits on each spiracle (Peterson 1960). Tenhumberg and Poehling (1995) used a variety of aphids collected from various areas throughout Germany to observe the normal development of several Syrphids. The head and mouth are not particularly noticeable, and in the majority of species, they are retractile. There may be obtrusive antennae and retractile parallel mouth hooks. Additionally, there are tiny, laterally extending mouth spines that are sclerotised (Sunil and Chandish 2013). Most of the previous research did not differentiate between the larval stages. This is most likely because moulted skins are incredibly fragile, transparent, and cling to the surface on which they are laid. Syrphid larva’s unusual moulting behaviour. Early winter is not the time to find exuviae. Additionally shed are tracheal exuviae and bucco-pharangeal armature exuviae. The anterior end of the moulted skin has a short longitudinal slit through which the larva has emerged. The larva passes the faeces and empties the contents of its stomach before each moulting. The larva does not move around before moulting. At that time, it occasionally feeds and releases a colourless fluid to prepare for moulting (Joshi et al. 2023). The Allograpta exotica species is widely distributed throughout the New World, from the United States south to Argentina and it has also been imported to Hawai’i. When given Aphis craccivora, Allograpta exotica’s development took an average of 2.0, 7.4, and 5.7 days for eggs, larvae, and pupae, and 15.04 days from egg to adult emergence. Larvae developed during an average period of 7.4 days, with variations occurring every 5–8 days. The first and second larval stadiums (L1 and L2) lasted for 2 days each, but the third stadium (L3) required 3–4 days. Last but not least, the pupae developed for an average period of 5.7 days, ranging from 5 to 6 days. When the age-specific survival rate was calculated, it was shown that eggs had the highest mortality rate, followed by first- and second-instar larvae (Arcaya et al. 2017). The development and reproduction of the hoverfly Eupeodes corollae (Diptera: Syrphidae) were documented by Zheng et al. (2019).

168

6.5.2

6

Insect Predators Immature Stages Biology

Pupa

The larvae secrete a gooey liquid through their mouths before pupating. The larvae move to the ground to pupate. Either an oval or a teardrop-shaped puparium is used. The ventral section is flattened, while the dorsal and lateral regions are inflated. At the caudal end are the posterior spiracles. P. serratus and I. scutellaris preferred to pupate on cotton pads in the lab, among other materials like corrugated sheet, tissue paper, cowpea leaf, cotton pad, and muslin cloth, possibly because syrphid larvae need a dry surface for pupation and seek cover when doing so. Cotton pads satisfy both of these basic needs (PDBC 1998). Hoverfly predator Sphaerophoria rueppellii is typically found in Mediterranean crops. Many workers noted a secondary impact of prey availability on development. When food is scarce, more energy is devoted to body upkeep, which causes larval development to take longer and pupae to grow smaller. The biological characteristics of several Syrphid predators, such as their larval duration, pupal period, and adult period (days), were reviewed by Joshi and Ballal in 1999. Table 6.3 summarises the information.

6.5.3

Eclosion

The full metamorphosis of the Syrphidae includes the egg, three larval stages, the puparium, and the adult. Most of the previous research did not differentiate between the larval stages. This is most likely because moulted skins are incredibly fragile, transparent, and cling to the surface on which they are laid. Syrphid larva’s unusual mudding. Early winter is not the time to find exuviae. Additionally shed are tracheal exuviae and bucco-pharangeal armature exuviae. The front end of the moulted skin

Table 6.3 Influences of different preys on larval, pupal and adult periods (days) of various Syriphids Syriphid Paragus serratus Paragus yerburiensi Paragus tibiali Betasyrphus serarius Dideopsis aegrota Ischiodon scutellaris

Episyrphus nubilipennis Episyrphus balteatus Syrphus pyrastri Eupeodes corollae

Prey(s) Aphis craccivora Aphis craccivora T. aurantii T. aurantii Aphis craccivora T. aurantii Aphis craccivora B. brassicae T. aurantii T. aurantii B. brassicae L. erysimi B. brassicae

Larval period 9–19 8–9 6–7 7–8 11–12 9–10

Pupal period 8–9 7–8 9–10 8–9 10–11 10–11

Adult period 14–15 10–11 1–6 2–6 14 1–6

9–12

13–14

6–9

9–10 8–12 10–16 10

10–11 8–9 10–18 10–11

2–7 5–14 20–22 12–28

6.5 Syrphidae

169

has a small longitudinal slit through which the larva has emerged. The larva passes its faces and empties its stomach before each moulting. The larva is dormant before it moults. At that time, it occasionally feeds and releases a colourless fluid to prepare for moulting (Sunil and Chandish 2013).

6.5.4

Development Periods

The larvae of the Eupeodes corollae were 1.00-mm long and had hair that was a little fluffy. Each larva feeds 998 Aphis craccivora for the course of the growth stage, which lasted 8 days. Before the third day, there was limited aphid eating by early instars, and less than 80 aphids were attacked per day. Beginning on the fourth day after hatching, predatory behaviour intensified, reaching a climax on the seventh day when 308 aphids were eaten by predators (Zheng et al. 2019). The larger, older syrphid larvae generally ingested more aphids; this shows that in order for adult syrphids to be efficient biological control agents, they must lay their eggs on lettuce early in the crop cycle and aphid populations’ establishment. Syrphid larval developmental stages are displayed in Table 6.4 along with references. According to data, the entire amount of time for development was between 10 and 14 days.

6.5.5

Survival

In contrast to older larvae, newly enclosed syrphid larvae frequently consumed the eggs and larvae of other species, regardless of the presence of aphids. For instance, Table 6.4 First, second, and third larval developmental periods (days) of syrphids reared on various hosts with references

Syrphidae Eupeodes frequens

Episyrphus balteatus

Eupeodes corollae Scaeva albomaculata Ischiodon scutellaris

Preys Brevicoryne brassicae Aphis fabae

Larval developmental period (days) First Second Third Total 4.2 3.4 4.8 12.4 5.0

3.5

5.5

14.0

Brevicoryne brassicae Schizaphis graminum Aphis pomi Aphis pomi

4.8

3.3

5.6

13.7

2.6

3.53

4.4

3.40

4.60

4.90

11.3 12.3

Macrosiphum rosae Schizaphis graminum

4.5

2.5

2.7

10.0

2.7

3.5

4.7

10.4

References Singh et al. (2020) Verma et al. (2005) Singh et al. (2020) Faheem et al. (2019) Jalilian (2015b) Jalilian et al. (2016) Jalilian (2015a) Faheem et al. (2019)

170

6

Insect Predators Immature Stages Biology

Pseudodorus clavatus initial instars were capable of attacking and feeding on aphids (Aphis spiraecola) that were many times bigger than themselves. When aphid shed exuviae were contacted during casting, larvae were occasionally observed to attack them, indicating a contact response to elements of the host cuticle comparable to those reported for Aphidiidae. Early instars occasionally perished as a result of having their mouthparts covered in Toxoptera citricida cornicle secretions; an occurring that was not seen in larvae fed Aphis spiraecola (Belliure and Michaud 2001).

6.5.6

Larval Feeding Behaviour

The larvae of Episyrphus balteatus, I. scutellaris, and Episyrphus confrater wriggled down the leaf midrib in search of prey before changing their direction, according to studies on the searching behaviour of syrphid larvae. The larvae crawl forward while simultaneously swaying side to side. After fixing its rear end to the substrate, the larva lifts and stretches the rest of its body before moving sideways in a semi-circular motion. Up until the prey is encountered, these movements are continued (Kumar et al. 1996). Joshi et al. (1999) made similar observations about B. fletcheri, B. linga, D. aegrota, P. serratus, and P. yerburiensis. They discovered that the initial instar stages following eclosion stayed stationary for 5–10 min before beginning to look for prey. When there was no prey around, it ate the unhatched eggs (Sunil and Chandish 2013). In the first 2 instars and again, 1–2 days before pupation, Episyrphus balteatus larvae were said to feed relatively little. At average summer temperatures, over 70% of the total food was reportedly consumed between days 7 and 10 of larval life. Peak voracity individuals made up the majority of syrphid populations when aphid densities were at their highest. Female syrphids stop producing eggs in older aphid colonies or when third instar larvae are present because younger larvae might not find enough food to finish their larval development during this time. As a result, at times of peak aphid abundance, one third-instar larva consumes about 90 aphids every day. Each larva consumes between 1.8 and 19.8 aphids per day at syrphid densities between 0.02 and 0.22 larvae per stalk. Based on that, it is anticipated that the syrphids will eat, on average, 50% of the aphids present. The daily biomass consumed by all 4 syrphid species peaked in the third instar, at least 1 day before pupation. However, temperature and the appropriateness of various aphid species as prey had a significant impact on this. Potential candidates for the biological management of pests like aphids include predatory syrphid larvae. The third larval stage, or third instar, is typically the hungriest.

6.6

Odonata

The Odonata’s aquatic larval stage lasts the longest (months, occasionally years). Significant mass and size growth takes place at this stage. Odonatan larvae are significant intermediate predators in aquatic food webs, eating a variety of tiny

6.7 Dermaptera

171

creatures, including conspecifics and other odonates, and serving as food for larger predators like fish. Larvae moult around ten times, with the last moult resulting in the metamorphosis into the flying, terrestrial adult stage. The entire duration of the F0 larval instar had significant intra-specific variation, according to Okude et al. (2021). We analysed the lengths of these embryonic phases and found significant differences in stage 1’s duration within and between Odonata species. Stages 2 and 3’s durations, on the other hand, varied less between and within species. As F0 instar larvae of Anax imperator and Asiagomphus pryeri, numerous river-dwelling Odonata species, including those in the Calopterygidae, Gomphidae, Cordulegastridae, Chlorogomphidae, and Macromiidae, are known to overwinter. Notably, in these species, stage 2 and stage 3 durations were very short and consistent across species, at 3–21 days and 1–8 days, respectively, despite the fact that the total durations of the F0 larval instar were frequently and obviously substantial. Because only a small variety of food was provided in this study, the length of the larval instar under natural conditions (particularly prior to entering stage 2 of the F-0 instar) may differ from our findings. The results did show that the larvae can mature into adults without consuming any food after they reach stage 2 of the F-0 instar. According to research on Ischnura senegalensis, the durations after the start of morphological change often showed little fluctuation, indicating that the development of metamorphosis after stage 2 proceeds slowly and irreversibly (Okude et al. 2021).

6.7

Dermaptera

Chelisoches morio nymphal developmental phases were examined in a lab setting in India. In the pockets of the leaf axils, Mhelisoches morio eggs were laid in masses of 38–89, with a typical yield of 156 eggs per female. The durations of the incubation phase, 4 nymphal instars, and adult stage were, respectively, 6–7, 6–10, 7–11, 9–19, 13–22, and 22–114. Male to female ratio was one to three (Abraham and Kurian 1973). In biological control campaigns against the cotton boll weevil Anthonomus grandis grandis, Euborellia annulipes (Anisolabididae) (Curculionidae), Euborellia annulipes eggs incubated for 1.6 times longer at 25 °C than at 30 °C, although having somewhat poorer survival rates there (Table 6.1). Similar findings were made by Melamed-Madjar (1971), who showed that more nymphs will enclose at a temperature of 20 °C than at a temperature of 32 °C, a finding that is consistent with the increased survival rate observed at the lower temperature in this study. For Euborellia annulipes, Koppenhöfer (1995) at 25° C and Rankin et al. (1995) at 28 °C reported egg incubation periods of 8 days. According to Knabke and Grigarick, this time might range from 6 to 17 days when this insect develops at temperatures between 20 and 29 °C (Bharadwaj (1966)). According to Rankin et al. (1995), the survival rate of Euborellia annulipes eggs at 28 °C was 78%, which is comparable to the rates we saw in our study with insects raised at 30 °C. This suggests that in this temperature range, Euborellia annulipes exhibits comparable egg viability (28–30 ° C). According to instar, sex, and temperature, Euborellia annulipes development

172

6

Insect Predators Immature Stages Biology

Table 6.5 Impact of preys Puccinia polysora (PP), Spodoptera frugiperda (SF), pollen (PO) on the nymphal developmental period (days) and survival rate (%) of Doru luteipes Food(s) Puccinia polysora Spodoptera frugiperda Puccinia polysora + Spodoptera frugiperda Pollen Control

Total nymphal period (days) 39.5 59.0 33.5 33.2 30.2

Survival rate (%) 30.2 56.6 72.9 55.2 73.9

time varies (Table 6.5), but it tends to go up with each succeeding instar at both temperatures and for both sexes. Rankin et al. (1995) also noticed this fact. Despite the fact that individuals in our population varied in terms of development time when compared to those of Bharadwaj (1966) and Koppenhöfer (1995), the relative development times of instars were comparable, with the fifth instar having the longest duration regardless of sex or temperature. Rankin and colleagues discovered a comparable length (1995). The variations in nutrition (Neiswander 1944; Toft and Wise 1999), temperature (Rankin et al. 1995), and experimental settings may account for the differences seen in these studies. Only in the fifth instar did differences in growth times between males and females become noticeable, with males developing more slowly than females. Males took slightly longer to fully mature to adulthood than females, 89 days at 25 °C to 53 days at 30 °C compared to 85 days at 25 °C to 49 days at 30 °C for males. The omnivorous predator Doru luteipes consumes pollen, uredospores of Puccinia polysora, and eggs of Spodoptera frugiperda from the maize plant. When many food sources, including eggs of Spodoptera frugiperda, uredospores of Puccinia polysora, and pollen are present simultaneously in the same plant, it may have an impact on the foraging behaviour and predatory capacity of Doru luteipes. It was discovered how long Doru luteipes survived in the nymph stage and which food sources they preferred—often found on maize plants—to eat. Newly born nymphs were fed only Puccinia polysora uredospores, Spodoptera frugiperda eggs, maize pollen, a combination of uredospores + eggs, and artificial food to test the survival of Doru luteipes (control). In a different experiment, Doru luteipes nymphs and adults who had fasted for 24 or 48 h were individually released in the middle of a container with four diets: corn pollen, eggs from Spodoptera frugiperda, uredospores from Puccinia polysora, and artificial diet. They were kept for 10 min to assess the food preference and feeding time. Low nymphal survival (8%) resulted from feeding only Spodoptera frugiperda eggs; however 58.3% of nymphs were able to survive when Puccinia polysora uredospores and Spodoptera frugiperda eggs were combined. Doru luteipes favoured feeding at night, and nymphs and adults made the same substantial amounts of dietary and pollen choices, with adults spending more time doing so. These results show that Doru luteipes trophic preferences are significant for understanding how it contributes to maize pest insect and fungal disease control (Table 6.5).

6.8 Coccinelidae

173

From hatching to adulthood, the European earwig Forficula auricularia (Dermaptera: Forficulidae) has four larval moults, according to the majority of studies. However, Tourneur et al. (2020) described the use of observational and quantitative methods to identify the occurrence of a second intermediate moult that takes place right away after egg hatching. Feeding with various phases of Plutella xylostella had an impact on the growth of each instar of Euborellia annulipes (Lepidoptera: Plutellidae). The development of the second instar was higher on an artificial diet (10 days) compared to the other feeding treatments; the third instar stages displayed a lower stage duration when they fed on fourth instar stages (7 days) and higher on 1-day-old pupae (10 days); the fourth instar stages developed in a lower developmental period. Euborellia annulipes nymphs in the first instar showed the lowest developmental period when fed either on larvae or 3-day-old pupae (8 days); The entire nymphal stage of Euborellia annulipes took 45 days to mature in the shortest case when nymphs were fed on fourth-instar stages, and 56 days in the case of artificial diet and pupae that were 1-day-old. The survival rate of Euborellia annulipes nymphs fed on thirdday pupae and fourth-instar stages was 90%, compared to 75% and 85% for artificial diets, respectively. According to the fourth instar stages, 1-day-old pupae, 3-day-old pupae, and artificial nutrition, the sex ratio of Euborellia annulipes was 0.56, 0.73, 0.61, and 0.63, respectively (Nunes et al. 2022).

6.8

Coccinelidae

Aphids, scales, thrips, mealybugs, and other soft-bodied insects that sucking insect pests feed on are among the most significant coccinellid predators. There are numerous habitats where the coccinellid beetles are seen. Cheilomenes sexmaculata, Coccinella septempunctata, Chilocorus nigritus, Adalia bipunctata, and Harmonia axyridis are several ladybirds (Coleoptera: Coccinellidae) that have been thoroughly studied, have a wide variety of prey, and are consequently of interest to people all over the world. Due to their sexual dimorphism, higher capacity for reproduction, ease of availability, ease of laboratory maintenance, and shorter life cycles, these predators make superior models for studying the aspects of life history (Ghanim et al. 2021). Adalia decempunctata, the ten-spotted lady beetle, is a species that originated in the Palaearctic and has since been found in Europe, North Africa, and West Asia. A prevalent predator in both agroecosystems and the wild world is Adalia decempunctata. Its biology and life cycle were investigated on the eggs of Ephestia kuehniella, Aphis fabae, and nymph/adult hosts of Aphis gossypii. According to the study’s findings, Adalia decempunctata immature developmental period (IDP) was the smallest when it fed on Ephestia kuehniella eggs (18.3 days) and the longest when it fed on Aphis fabae eggs (21.8 days) (Mojib-Haghghadam et al. 2018). The mustard aphid, Lipaphis erysimi (Hemiptera: Aphididae), on broccoli was fed to the Propylea dissecta life stages, and under laboratory conditions, many life features were recorded. The incubation time was 4.0 days, and the females laid their

174

6

Insect Predators Immature Stages Biology

eggs in groups. There were 10.9 and 4.3 days of mean pupal and larval durations, respectively. The whole development time was 19.2 days on average (Boopathi et al. 2020). On the biological parameters of the predator Oenopia conglobata, different meals including varied combinations of the flour moth egg Ephestia kuehniella (E), bee pollen (P), Artemia urmiana cysts (A), Sitotroga cerealella (S), and lyophilised artificial food (L) were assessed. The LEP diet produced the adults with the quickest developmental time (13.6 days) and the heaviest weight (8.9 mg) of the three studied diets (lyophilised artificial diet, Ephestia kuehniella eggs and bee pollen). While the LEPS diet (lyophilised artificial feed, Ephestia kuehniella eggs, bee pollen, and eggs of Sitotroga cerealella) had the highest larval survival rate (55.93%) (Vahmani et al. 2022).

6.9

Thysanoptera (Thrips)

About 6500 species of the Thysanoptera (thrips), which are found all over the world, are among the smallest flying insects. Due to their ability to spread through the plant trade and serve as plant vectors for spoviruses, which result in significant agricultural losses, a number of thrips species are important on a global scale. About 300 species of thrips have evolved a predatory lifestyle; the majority are detritivores (mostly fungal feeders) and herbivores (feeders of flowers, fruits, and leaves). The larval stage includes two instars. The third antennal segment of the initial instars, which have just emerged, is around 3.5–4.5 times longer than it is wide. The mesothorax and abdomen segments III–VII turn red after consuming food for 1 or 2 days. The second instars have a recognisable hump on their backs. Additionally, the prothorax and head take on the same red colouring as the mesothorax. The fore tibia and tarsus of the third antennal segment’s second instar are dark, and it is roughly seven to eight times longer than wide. Seven segmented, closely united, seven segmented antennae are present on both instars. Only the femora have the red hypodermal pigments. Pupae are found inside a white silk cocoon that the caterpillar made underneath the leaves. Pre-pupal, pupal stage 1 and pupal stage 2 are the three phases of the pupa, which is red in colour. Well-developed wing buds are shorter in the pre-pupal stage (show non-obvious movement, prepared for cocoon construction). Only in pre-pupa is the pupal skin of the appendages segmented. The antennal sheaths reach the abdomen but not the metathorax (pupa 1) (pupa 2). Additionally, while both the anterior and posterior wing buds reach abdominal segment V, the posterior wing buds only reach abdominal segment III (pupa 1) (pupa 2). Legs and the hind tibiotarsus are longer than the pterothorax (pupa 1) and shorter than the pterothorax (pupa 1) (pupa 2). The larval stage includes two instars. The third antennal segment of the recently emerged Franklinothrips vespiformis (Thysanoptera: Aeolothripidae) first instars is roughly 3.5–4.5 times longer than broad. The mesothorax and abdomen segments III–VII turn red after consuming food for 1 or 2 days. The second instars of Franklinothrips vespiformis have a recognisable hump-back. Additionally, the prothorax and head take on the same red colouring as the mesothorax. The fore tibia and

6.11

Tri-Trophic Interactions of Bt Toxins–Preys–Predator

175

tarsus of the third antennal segment’s second instar are dark, and it is roughly seven to eight times longer than wide. Seven segmented, closely united, seven segmented antennae are present on both instars. The Franklinothrips vespiformis femora are the sole part of the insect that has red hypodermal pigments (Hussain et al. 2022).

6.10

Mantodea

Before reaching maturity, mantids, hemi-metabolic insects, moult six to nine times. Mantids have a life span of 3–8 months; however, they can live up to a year in captivity. However, the longevity of mantids varies depending on the species and is regulated by a number of abiotic factors including temperature, the availability of prey, and humidity. In this study, 45 of the 63 neonatal Harpagomantis tricolour nymphs that hatched from the 12 various oothecae survived to adulthood (14 males and 31 females). The average time between mating and the development of an ootheca was 11.82 days, while the actual act of copulation lasted about 6 h. About 20 weeks passed during an ootheca’s incubation period (143 days). Harpagomantis tricolour had an average hatch rate of 31% and an average survival rate of about 68%. Although the sex ratio varied across the different oothecae, the mean sex ratio (M:F) was 1.5. Two of the oothecae produced solely male offspring, while the other two only female offspring (Greyvenstein et al. 2021).

6.11

Tri-Trophic Interactions of Bt Toxins–Preys–Predator

Insect pests must be controlled by natural enemies such predators, parasitoids, and entomopathogens in order to preserve the stability of agroecosystems. However, since the release of transgenic Bt crops, this strategy of control has been used more frequently, raising concerns about potential negative effects on natural enemies. These creatures may be impacted both directly and indirectly, thus it is important to monitor and research how these organisms interact with GM plants. In this section, the Bacillus thuringiensis bacterium and its potential effects on nontarget creatures, particularly natural enemies, are briefly discussed. Initially, Spodoptera littoralis (Lepidoptera: Noctuidae) larvae infected with Bt were attacked and eaten by adults of Coccinella undecimpunctata (Coleoptera), according to Mahmoud (1992). When fed on infected Spodoptera littoralis, Coccinella undecimpunctata larvae showed no negative consequences and eventually matured into adults that were able to reproduce and lay eggs. The nymphs and adults of the pentatomid predator, Podisus nigrispinus, were tested for their reactions to Bt-infected larvae of Plutella xylostella as prey by Carvalho et al. (2012), and they found that Bt had no influence on their biological parameters compared to the control. Magalhaes et al. (2015) looked into how the commercial product “Agree”—a mix of Bt vars. Kurstaki and aizawai—affected Podisus nigrispinus’ biological characteristics after it consumed Plutella xylostella larvae that were infected. They discovered that during the nymphal development, the infected larvae

176

6

Insect Predators Immature Stages Biology

were devoured more frequently than the healthy ones, with no impact on nymphal survival or population growth metrics. However, Carvalho et al. (2018) found that raising Podisus nigrispinus nymphs on Tenebrio molitor (Coleoptera: Tenebrionidae) larvae treated with Bt had an impact on the oviposition times and total quantity of eggs laid by the predator’s female. The predator Geocoris punctipes was given two different types of prey: Spodoptera exigua larvae (BAW), which carried the Bt Cry1Ac toxin to the next trophic level, and Helicoverpa zea eggs (CEW), which were free of the toxin. For each treatment, the percentage of Geocoris punctipes that survived the nymphal stages varied, ranging from 0% to 100% per cage. During the experimental periods, the Weld’s fluctuating environmental circumstances undoubtedly had an impact on survival variability. Additionally, across the 2 seasons, 12 cages were destroyed by Solenopsis invicta ants, leading to the removal of these replications from analysis. No differences in nymphal survival were found between Bt- and non-Bt cotton, cotton as the primary host, or between preys, according to statistical analysis. Nymph survival did differ dramatically across years, though. Overall, nymphal survival was lower in 2004 than it was in 2003, but nymphs lived as well independent of the type of prey or cotton utilised. Newly born big-eyed bug nymphs showed identical developmental time, survival, sex ratio, and adult weight when confined on Bt or non-Bt cotton. By preying on herbivorous insects that consume cotton, Geocoris pallidipennis, a significant natural enemy insect in cotton fields, is able to consume the Bt protein generated in GM cotton. However, it is yet uncertain if GM cotton poses a threat to Geocoris pallidipennis. Here, we assessed how Cry1Ac/1Ab protein expression in Bt cotton affected Geocoris pallidipennis nymphs and adults. In addition to being found in the midgut of Geocoris pallidipennis nymphs and adults that feed on Bt-fed Helicoverpa armigera, Cry1Ac protein was also found in the midgut of the cotton bollworm, Helicoverpa armigera, after it consumed Bt cotton. However, Geocoris pallidipennis survival rate, growth, development, and fertility were not negatively impacted. Additionally, the genes involved in immunological function, detoxification, antioxidant activity, and nutritional utilisation did not alter in expression in response to Cry1Ac exposure. Finally, we demonstrated that neither Geocoris pallidipennis nymphs nor adults could bind Cry1Ac to the proteins found on brush border membrane vesicles (BBMV). In conclusion, these findings suggest that there is no risk associated with transgenic Cry1Ac/1Ab cotton for the insect predator Geocoris pallidipennis (Zhang et al. 2022). Live aphids (10) (Brevicoryne spp., Hemiptera: Aphididae) were used to feed Galepsus (Lygdamia) lenticularis’ first and second instars, whereas live crickets (2) (Acheta sp., Orthoptera: Gryllidae) of various sizes (nymphal instars, or pinheads) were used to feed nymph 40.3% of eggs were laid on average per batch. A total of 76 of the 192 new-born nymphs that hatched made it to the second instar, and 48 (63%) of them finished the rest of their life cycle. Only 11 (40.0%) of the eggs that hatched survived to adulthood. Twenty-one weeks were needed on average from hatch to adulthood (148 days). While most nymphs matured into adults after seven moults, there were a few documented exceptions. Ten moults were necessary for six

References

177

people to attain adulthood (three males and three females). One male only required four moults to reach adulthood. Although there were no statistically significant variations between male and female development times or the length of an instar for each sex, there was a difference between female and male adult longevity. Males lived for an average of 167 days from first instar to death, while females lasted an average of 253 days. Once they reached adulthood, males and females survived for 93 and 26 days, respectively. For the most part, males and females had comparable mean durations per instar.

6.12

Future Area of Research

Since there are so few publications on mentodea, we advise focusing on this subject. Research on the effects of various natural, laboratory, and fictitious host foods on the biology of predatory insects should be conducted. Research on the biology of regionally specific predators and/or globally specific predators should be prioritised by various funding agencies. Research on the effects of genetically modified crops on the biology of predatory insects should be concentrated.

References Abbas S, Pérez-Hedo M, Colazza S et al (2014) The predatory mirid Dicyphus maroccanus as a new potential biological control agent in tomato crops. BioControl 59(5):565–574 Abraham VA, Kurian C (1973) Chelisoches moris F. (Forficulidae: Dermaptera), a predator on eggs and early instar grubs of the red palm weevil Rhynchophorus ferrugineus F. (Curculionidae: Coleoptera). J Plant Crop 1:147–152 Ahmad SH, Kamarudin N (2016) Growth and longevity of the insect predator, Sycanus dichotomus Stål (Hemiptera: Reduviidae) fed on live insect larvae. J Oil Palm Res 28:471–478 Almohamad R, Verheggen F, Haubruge É (2009) Searching and oviposition behavior of aphidophagous hover flies (Diptera: Syrphidae): a review. Biotechnol Agron Soc Environ 13: 467–481 Ambrose DP, Livingstone D (1985) Development of Coranus vitellinus Distant (Hemiptera: Reduviidae: Harpactorinae). J Soil Biol Ecol 5(1):65–71 Ambrose DP, Sebasti Rajan XJ, Nagarajan K et al (2009) Biology, behaviour and functional response of Sphedanolestes variabilis Distant (Insecta: Hemiptera: Reduviidae: Harpactorinae), a potential predator of lepidopteran pests. Entomol Croat 13(2):33–44 Ambrose DP, Nagarajan K, Kumar AG (2013) Interaction of reduviid predator, Rhynocoris marginatus (Fabricius) (Hemiptera: Reduviidae) with its prey teak skeletonizer, Eutectona machaeralis Walker (Lepidoptera: Pyralidae) as revealed through functional response. J Entomol Res 37(1):55–59 Amer ME, Abdel-Razak SI, El-Sobky HF (2021) Influences of some insect pests as prays on biology and consumption rate of predator, Oruis albidipennis (Reuter) (Hemiptera, Anthocoridae) under laboratory conditions. J Plant Prot Pathol 12(1):37–42 Angilletta MJ Jr, Dunham AE (2003) The temperature-size rule in ectotherms: simple evolutionary explanations may not be general. Am Nat 162(3):332–342 Arcaya E, Pérez-Bañón C, Mengual X et al (2017) Life table and predation rates of the syrphid fly Allograpta exotica, a control agent of the cowpea aphid Aphis craccivora. Biol Control 115:74– 84

178

6

Insect Predators Immature Stages Biology

Ballal CR, Varshney R, Joshi S (2019) Morphology, biology and predation capacity of Amphiareus constrictus (Stål) (Hemiptera: Anthocoridae). Neotrop Entomol 48(4):668–677 Belliure B, Michaud JP (2001) Biology and behavior of Pseudodorus clavatus (Diptera: Syrphidae), an important predator of citrus aphids. Ann Entomol Soc Am 94(1):91–96 Bharadwaj RK (1966) Observations on the bionomics of Euborellia annulipes (Dermaptera: Labiduridae). Ann Entomol Soc Am 59(3):441–450 Bhatia ML (1939) Biology, morphology and anatomy of aphidophagous syrphid larvae. Parasitology 31(1):78–120 Boevé JL (2021) Some sawfly larvae survive predator-prey interactions with pentatomid Picromerus bidens. Sci Nat 108(2):1–8 Boopathi T, Singh SB, Dutta SK et al (2020) Biology, predatory potential, life table, and field evaluation of Propylea dissecta (Coleoptera: Coccinellidae), against Lipaphis erysimi (Hemiptera: Aphididae) on broccoli. J Econ Entomol 113(1):88–97 Carvalho VFP, Vacari AM, Pomari AF et al (2012) Interaction between Podisus nigrispinus and Bacillus thuringiensis. Environ Entomol 41:1454–1461 Carvalho GASD, Martins DJ, Brito IMCD et al (2018) Can Bacillus thuringiensis affect the biological variables of natural enemies of Lepidoptera? Arq Inst Biol 85:1–6 Claver MA, Reegan AD (2010) Biology and mating behaviour of Coranus spiniscutis Reuter (Hemiptera: Reduviidae), a key predator of rice Gandhi bug Leptocorisa varicornis Fabricius. J Biopest 3(2):437 De Clercq P, Merlevede F, Tirry L (1998) Unnatural prey and artificial diets for rearing Podisus maculiventris (Heteroptera: Pentatomidae). Biol Control 12(2):137–142 Faheem M, Saeed S, Sajjad A et al (2019) Biological parameters of two syrphid fly species Ischiodon scutellaris (Fabricius) and Episyrphus balteatus (DeGeer) and their predatory potential on wheat aphid Schizaphis graminum (Rondani) at different temperatures. Egypt J Biol Pest Control 29:2. https://doi.org/10.1186/s41938-019-0105-0 Fathipour Y, Jafari A (2008) Biology of Nabis capsiformis (Het., Nabidae) preying upon i> Creontiades pallidus (Het., Miridae) in laboratory conditions. JWSS-Isfahan Univ Technol 12(43):157–166 Gallard L (1932) Notes on the feeding habits of the brown moth-lacewing, Ithone fusca. Austr Nat 8:168–170 Geethalakshmi L, Muthukrishnan N, Chandrasekaran M et al (2000) Chrysopids biology on Corcyra cephalonica and feeding potential on different host insects. Ann Plant Prot Sci 8(2): 132–135 George PJE (1999) Biology and life table studies of the reduviid Rhynocoris marginatus (Fabrcius) (Heteroptera: Reduviidae) on three lepidopteran insect pests. J Biol Control 13:33–38 Ghanim AA, El-Serafy HA, Hassan HA et al (2021) Biological aspects of two Coccinellid predatory insects reared on artificial diets and natural preys under constant temperature. J Plant Prot Pathol 12(1):19–22 Grazia J, Panizzi AR et al (2015) Stink Bugs (Pentatomidae). In: Panizzi AR, Grazia J (eds) True bugs (Heteroptera) of the neotropics. Springer, Dordrecht, pp 681–756. https://doi.org/10.1007/ 978-94-017-9861-7_2 Greyvenstein BM (2020) Distribution of Mantodea in South Africa and biological studies of selected species. Doctor of Philosophy in Environmental Sciences at the North-West University Greyvenstein B, du Plessis H, Van den Berg J (2021) Life history of the false flower mantid (Harpagomantis tricolor Linnaeus, 1758) (Mantodea: Galinthiadidae) and its distribution in southern Africa. J Orthoptera Res 30:17 Harris HM (1928) A monographic study of the hemipterous family Nabidae as it occurs in North America (1928). Retrospective Theses and Dissertations 14758. https://lib.dr.iastate.edu/rtd/14 758 Henry CS, Penny ND, Adams PA (1992) The neuropteroid orders of Central America (Neuroptera and Megaloptera), pp 432–458. In: Quintero D, Aiello A (eds) Insects of Panama and Mesoamerica. Oxford University Press, Oxford, p 692

References

179

Hoffman KM, Brushwein JR (1989) Species of spiders (Araneae) associated with the immature stages of Mantispa pulchella (Neuroptera, Mantispidae). J Arachnol 17:7–14 Hoffman KM, Brushwein JR (1990) Spider (Araneae) taxa associated with the immature stages of Mantispa interrupta (Neuroptera: Mantispidae). Entomol News 101(1):23–28 Hormchan P, Schuster MF, Hepner LW (1976) Biology of Tropiconabis capsiformis. Ann Entomol Soc Am 69(6):1016–1018 Hussain M, Wang Z, Arthurs SP (2022) A review of Franklinothrips vespiformis (Thysanoptera: Aeolothripidae): life history, distribution, and prospects as a biological control agent. Insects 13(2):108 Jalilian F (2015a) Development and feeding capacity of Scaeva albomaculata (Macqaurt) (Diptera: Syrphidae) fed with rose aphid, Macrosiphum rosae (Homoptera: Aphididae). Biol Forum 7(1): 1377–1381 Jalilian F (2015b) Biology and larval feeding rate of Episyrphus balteatus (Dip.: Syrphidae) on Aphis pomi (Hom.: Aphididae) at laboratory conditions. Biol Forum 7(1):1395–1399 Jalilian F, Karimpour Y, Aramideh S (2016) Investigation on some biological characteristics of Eupeodes corollae (Dip.: Syrphidae) on Aphis pomi (Hom: Aphididae) in vitro. J Entomol Zool Stud 4:432–435 Joshi S, Venkatesan T, Ballal CR et al (1999) Comparative biology and predatory efficiency of six syrphids on Aphis craccivora Koch. Pest Manag Hortic Ecosyst 5(1):1–6 Kingsolver JG, Huey RB (2008) Size, temperature, and fitness: three rules. Evol Ecol Res 10(2): 251–268 Joshi SS, David KJ, Sachin K (2023) Syrphid Predators (Diptera: Brachycera). In: Insect Predators in Pest Management. CRC Press, pp 229–243 Koppenhöfer AM (1995) Bionomics of the earwig species Euborellia annulipes in Western Kenya (Dermaptera: Carcinophoridae). Entomol Gen 20:81–85 Kubavat AK, Jethva DM, Wadaskar PS (2017) Biology and feeding potential of Chrysoperla zastrowisillemi (Esben Peterson) on eggs of Corcyra cephalonica (Stainton). Environ Ecol 35(3A):1948–1952 Lemos WP, Ramalho FS, Serrão JE et al (2003) Effects of diet on development of Podisus nigrispinus (Dallas) (Het., Pentatomidae), a predator of the cotton leafworm. J Appl Entomol 127(7):389–395 Lillo I, Perez-Bañón C, Rojo S (2021) Life cycle, population parameters, and predation rate of the hover fly Eupeodes corollae fed on the aphid Myzus persicae. Entomol Exp Appl 169(11): 1027–1038 Liu P, Jia W, Zheng X et al (2018) Predation functional response and life table parameters of Orius sauteri (Hemiptera: Anthocoridae) feeding on Megalurothrips usitatus (Thysanoptera: Thripidae). Fla Entomol 101(2):254–259 Madadi H, Ahmadi Z, Heshmati A (2016) Biology and life table parameters of predatory bugs Nabis pseduferus (Hemiptera: Nabidae). In: Proceedings of 22 College of Agriculture and Natural Resources, University of Tehran, Karaj, Iran, p 584 Magalhaes GO, Vacari AM, De Bortoli CP et al (2015) Interaction between Bacillus thuringiensis insecticides and Podisus nigrispinus (Hem.: Pentatomodae), a predator of Plutella xylostella (Lep.: Plutellidae). Neotrop Entomol 44(5):521–527 Mahdavi TS, Madadi H, Biondi A (2020) Predation and reproduction of the generalist predator Nabis pseudoferus preying on Tuta absoluta. Entomol Exp Appl 168(10):732–741 Mahmoud BA (1992) The role of some Spodoptera littoralis parasitoids and predators as distributers of bacterial and viral insecticides. M.Sc. thesis, Fac. Agric., Cairo University, p 105 Marquereau L, Cottineau JS, Fontaine O et al (2022) Life history parameters and predation capacities of Nesidiocoris volucer: a new biological control agent for tomato crop. Bull Entomol Res 112:1–9. https://doi.org/10.1017/S0007485321001164 Melamed-Madjar V (1971) Bionomics and ecology of the Earwig Anisolabis (Euborellia) annulipes Luc. (Labiduridae-Dermaptera) in Israel 1. Z Angew Entomol 69(1–4):170–176

180

6

Insect Predators Immature Stages Biology

Mohammadpour M, Hosseini M, Michaud JP et al (2020) The life history of Nabis pseudoferus feeding on Tuta absoluta eggs is mediated by egg age and parasitism status. Biol Control 151: 104–401 Mojib-Haghghadam Z, Sendi JJ, Zibaee A et al (2018) Suitability of Aphis gossypii Glover, Aphis fabae Scop. and Ephestia kuehniella Zeller eggs for the biology and life-table parameters of Adalia decempunctata (L.) (Coleoptera: Coccinellidae). Arch Biol Sci 70(4):737–747 Mounika K, Gosalwad SS, Neharkar PS (2021) Biology of Chrysoperla zastrowi sillemi (EsbenPetersen) on different hosts. Pharm Innov J 10(8):671–674 Myers JG (1925) Biological notes on Arachnocoris albomaculatus Scott (Hemiptera; Nabidae). J N Y Entomol Soc 33(3):136–146 Nandan N, Korat DM, Dabhi MR (2014) Influence of different host insects (prey) on biological parameters of Chrysoperla zastrowisillemi (Esben-Peterson). Insect Environ 20(2):40–44 Naruka P, Ameta OP (2015) Biology of Chrysoperla zastrowi arabica on different prey. Indian J Appl Entomol 29(1):61–65 Neiswander RB (1944) Insect pests of strawberries in Ohio. Bulletin. Ohio Agricultural Experiment Station, Wooster, p 651 New TR (2004) Insecta: Neuropteroidea. Freshwater invertebrates of the Malaysian Region. Academy of Sciences Malaysia, Kuala Lumpur, pp 491–500 Ngoc Bao Chau N, Quynh Phuong Anh N, Le Nhu TT et al (2021) Ant and silkworm pupae as convenient diets for the development and reproduction of big-eyed bug Geocoris ochropterus, vol 24. Geocoridae, Hemiptera, p 131 Nunes GDS, Dantas TAV, Figueiredo WRS et al (2022) The life history of Euborellia annulipes (Lucas). (Dermaptera: Anisolabididae) fed on larvae and pupae of Plutella xylostella (L.) (Lepidoptera: Plutellidae). Turk J Zool 46(2):175–185 Okude G, Fukatsu T, Futahashi R (2021) Comprehensive comparative morphology and developmental staging of final instar larvae toward metamorphosis in the insect order Odonata. Sci Rep 11:5164. https://doi.org/10.1038/s41598-021-84639-2 Paleari LM (2013) Developmental biology, polymorphism and ecological aspects of Stiretrus decemguttatus (Hemiptera, Pentatomidae), an important predator of cassidine beetles. Revista Brasileira de Entomologia 57(1):75–83 PDBC (1998) Annual report, 1997-98. Project Directorate of Biological Control. ICAR, Bangalore, p 217 Perkins PV, Watson TF (1972) Biology of Nabis alternatus (Hemiptera: Nabidae). Ann Entomol Soc Am 65(1):54–57 Peterson A (1960) Larvae of insects. An introduction to nearctic species. Part II. Columbus, Ohio, p 416 Pu D, Zheng Z, Liu H et al (2019) Development and reproduction of the hover fly Eupeodes corollae (Diptera: Syrphidae). J Earth Sci Environ Stud 4:654–660 Putra NS, Yasuda H (2006) Effects of prey species and its density on larval performance of two species of hover fly larvae, Episyrphus balteatus de Geer and Eupeodes corollae Fabricius (Diptera: Syrphidae). Appl Entomol Zool 41:389–397 Rajan SL, Sree Latha E, Sathish R (2018) Biology of big-eyed bug, Geocoris erythrocephalus (Lepeletier & Serville) on cabbage aphid, Brevicoryne brassicae (L.). Int J Curr Microbiol App Sci 7(7):3301–3305 Ramirez R, Patterson R, Carbon C (2011) Beneficial true bugs: big-eyed bugs. Utah State University Extension and Utah Plant Pest Diagnostic Laboratory, ENT-159-11PR Rankin SM, Palmer JO, Larocque L et al (1995) Life history characteristics of ringlegged earwig (Dermaptera: Labiduridae): emphasis on ovarian development. Ann Entomol Soc Am 88(6): 887–893 Rustam R, Gani MA (2019) Biology of Eocanthecona furcellata (Wolff) (Hemiptera: Pentatomidae) Predator Nettle Caterpillar Setora nitens Walker Origin from Riau. In: IOP conference series: earth and environmental science, vol 347, no 1. IOP Publishing, p 012009

References

181

Sahid A, Natawigena WD (2018) Laboratory rearing of Sycanus annulicornis (Hemiptera: Reduviidae) on two species of prey: differences in its biology and efficiency as a predator of the nettle caterpillar pest Setothosea asigna (Lepidoptera: Limacodidae). Eur J Entomol 115: 208–216 Salcedo C, Iannacone J, Alvariño L (2020) Ciclo biológico de Nabis consimilis Reuter, 1912 (Heteroptera: Nabidae) bajo condiciones de laboratorio. Revista Electrónica de Divulgación de Metodologías emergentes en el desarrollo de las STEM 2(2):3–21 Sahayaraj K, Ambrose DP (1993) Biology and predatory potential of Coranus nodulosus Ambrose & Sahayaraj on Dysdercus cingulatus Fab. and Oxycarenus hyalinipennis Costa(Heteroptera: Reduviidae). Hexapoda 5:16–22 Shahayaraj K, Sathiamoorthi P (2002) Influence of different diets of Corcyra cephalonica on life history of a reduviid predator Rhynocoris marginatus (Fab.). J Cent Eur Agric 3(1):53–62 Siddique AB, Chapman RB (1987) Effect of prey type and quantity on the reproduction, development, and survival of Pacific damsel bug, Nabis kinbergii Reuter (Hemiptera: Nabidae). N Z J Zool 14(3):343–349 Simonato J, de Oliveira HN, Grigolli JF (2020) Potential of predator Podisus nigrispinus Dallas (Hemiptera: Pentatomidae) in the control of Helicoverpa armigera Hübner (Lepidoptera: Noctuidae). Embrapa Agropecuária Oeste-Artigo em periódico indexado (ALICE) 12:204 Singh P, Thakur M, Sharma KC et al (2020) Larval feeding capacity and pollination efficiency of the aphidophagous syrphids, Eupeodes frequens (Matsmura) and Episyrphus balteatus (De Geer) (Diptera: Syrphidae) on the cabbage aphid (Brevicoryne brassicae L.) (Homoptera: Aphididae) on mustard crop. Egypt J Biol Pest Control 30:105. https://doi.org/10.1186/s41938020-00300-6 Sravika A, Shylesha AN, Jagadeesh KS (2020) Biology and potential of pentatomid predator, Eocanthecona furcellata (Wolff) (Hemiptera: Pentatomidae) on fall armyworm, Spodoptera frugiperda (Smith). J Biol Control 34(1):26–29 Srikumar KK, Bhat PS, Raviprasad TN et al (2014) Biology, behaviour, and functional response of Cydnocoris gilvus Brum. (Heteroptera: Reduviidae: Harpactorinae) a predator of tea mosquito bug (Helopeltis antonii Sign.) on cashew in India. J Threat Taxa 6(6):5864–5870 Sunil J, Chandish RB (2013) Syrphid predators for biological control of aphids. J Biol Control 27(3):151–170 Tenhumberg B, Poehling HM (1995) Syrphids as natural enemies of cereal aphids in Germany: aspects of their biology and efficacy in different years and regions. Agric Ecosyst Environ 52(1): 39–43 Toft S, Wise DH (1999) Growth, development, and survival of a generalist predator fed single- and mixed-species diets of different quality. Oecologia 119(2):191–197 Torres JB, Silva-Torres CS, Ruberson JR (2004) Effect of two prey types on life-history characteristics and predation rate of Geocoris floridanus (Heteroptera: Geocoridae). Environ Entomol 33(4):964–974 Tourneur JC, Cole C, Meunier J (2020) The first of five moults of Forficula auricularia L. (Dermaptera: Forficulidae). Can Entomol 152(6):783–789 Truong XL, Pham HP, Thai TNL (2020) Biology and predatory ability of the reduviid Sycanus falleni Stal (Heteroptera: Reduviidae: Harpactorinae) fed on four different preys in laboratory conditions. J Asia Pac Entomol 23(4):1188–1193 Vahmani A, Shirvani A, Rashki M (2022) Biological parameters of Oenopia conglobata (Linnaeus) (Coleoptera: Coccinellidae) feeding on different diets. Int J Trop Insect Sci 42:2241 Varshney R, Ballal CR (2017) Biology and feeding potential of Geocoris superbus Montandon (Heteroptera: Geocoridae), a predator of mealybug. J Entomol Zool Stud 5(3):520–524 Vennison SJ (1989) Bioecology, ecophysiology and ethology of chosen assassin bugs. Ph.D thesis, Madurai Kamaraj University, Madurai Vennison SJ, Ambrose DP (1986) Bioecology of a dimorphic assassin bug, Edocla slateri Distant (Heteroptera: Reduviidae). Entomon 11(4):255–258

182

6

Insect Predators Immature Stages Biology

Vennison SJ, Ambrose DP (1990) Biology and behaviour of Sphedanolestes signatus Distant (Insecta: Heteroptera: Reduviidae) a potential predator of Helopeltis antonii Signoret. Uttar Pradesh J Zoology 10(1):30–43 Vennison SJ, Ambrose DP (1991) Biology and behavior of Ectomocoris xaviereii Vennison & Ambrose 1990 (Heteroptera Reduviidae Piratinae). Trop Zool 4(2):251–258 Vennison SJ, Ambrose DP (1992) Biology, behaviour and biocontrol efficiency of a Reduviid predator, Sycanus reclinatus Dohrn (Heteroptera: Reduviidae) from southern India. Mitteilungen aus dem Museum für Naturkunde in Berlin. Zoologisches Museum und Institut für Spezielle Zoologie (Berlin) 68(1):143–156 Verma JS, Sharma KC, Sood A et al (2005) Biology and predatory potential of syrphid predators on Aphis fabae infesting Solanum nugrum (L.). J Entomol Res 29:39–41 Vivek S, Paul B, Pandi GGP et al (2013) Biology and predatory potential of green lacewing, Chrysoperla sp. (carnea-group) on different aphid species. Ann Plant Prot Sci 21(1):9–12 Wheeler AG Jr, Stinner BR, Henry TJ (1975) Biology and nymphal stages of Deraeocoris nebulosus (Hemiptera: Miridae), a predator of arthropod pests on ornamentals. Ann Entomol Soc Am 68(6):1063–1068 Zhang M, Ma Y, Luo J et al (2022) Transgenic insect-resistant Bt cotton expressing Cry1Ac/1Ab does not harm the insect predator Geocoris pallidipennis. Ecotoxicol Environ Saf 230:113129 Zheng Z, Liu H, Wang X et al (2019) Development and reproduction of the hoverfly Eupeodes corollae (Diptera: Syrphidae). J Earth Sci Environ Stud 4(4):654–660 Zulkefli M, Norman K, Basri MW (2004) Life cycle of Sycanus dichotomus (Hemiptera: Pentatomidae) a common predator of bagworm in oil palm. J Oil Palm Res 14:50–56

7

Predatory Insects: Adults Biology of Various Orders

Contents 7.1 7.2

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Coleopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2.1 Plant Varieties/Cultivars on Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3 Mantodea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.5 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6 7.6. Hemipteran Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6.1 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6.2 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6.3 Miridae Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6.4 Anthecoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.6.5 Lygaeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.7 Lacewings (Neuropteran) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.8 Miscellaneous Predators (Orthopera, Lepidotpera, Others) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.9 GM Crops or Bt Proteins on the Biology of Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.10 Future Area of Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

7.1

183 184 190 190 192 195 196 196 203 205 206 207 209 211 213 214 215

Introduction

It is crucial to describe the development, stage structure, fecundity, and predation rate of a pest’s predators used for a biological control programme to be successful. In this process, accurate data analysis of the life history and predation rates is crucial. The life table is a crucial part of this knowledge. A complete depiction of the growth, survival, and fecundity of a population can only be found in life tables. Although predatory arthropods have been demonstrated to be affected by the quality of their prey in the past, surprisingly little is known about the precise nutrients that cause these effects.

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_7

183

184

7

Predatory Insects: Adults Biology of Various Orders

Life tables are essential tools for population ecology and pest management, making it possible to mass cultivate and collect insects for biocontrol, extract pheromones, and research area like toxicity. Analysing fertility, development, and survival and their impacts on the population growth rate and life table is essential to understanding population ecology and the large-scale production of predatory insects. In this chapter, we collected life table data for a variety of predators and conducted an analysis using the two-sex, age-stage life tables to account for both sex and stage differentiation.

7.2

Coleopterans

Merely recently, adult Coccinella transversalis were initially yellowish without any markings and subsequently changed to dazzling yellow or warm buff with black dots. The adult ladybird beetle had an oval, convex body shape, and was medium in size. It was red and black in colour. Anterolateral corners of the pronotum are light cream. The scutellum is dark. Bright carmine red, orange, or yellow elytra with an erratic black stripe along the sutural line, an oval subscutellar spot, a big trilobed spot on the humeral callus, a transverse band at the apical third that does not reach the lateral margin, and three smaller apical spots, of which one sutural and two laterals and both are typically fused to form a transverse marking (Shukla and Jadhav 2014). The four larval phases that Coccinella septempunctata larvae underwent each took 2.5, 1.91, 2.0, and 4 days, respectively. Under laboratory conditions, Coccinella septempunctata had an extremely brief adult life span that, on average, lasted between 6 and 8 days. Both adults and larvae fed in different ways on aphids, and they each ingested on average 1203.556 of them during the course of their entire lives, which lasted 17.19 days (Akram et al. 1996). Coccinella transversalis adult emergence rates ranged from 57% to 100%. The present findings are thus only partially confirmed. When raised on Lipaphis erysimi, Aphis craccivora, and Myzus persicae, the pre-oviposition duration ranged from 4 to 7, 4 to 8, and 4 to 7 days, with an average of 5.50, 5.22, and 5.7 days, respectively. When raised on L. erysimi, Coccinella undeceimpunctata’s pre-oviposition duration was reported by Solangi et al. (2007) to be 4.1 days. It is supporting the current findings in part. When raised on Aphis craccivora, Lipaphis erysimi, and M. persicae, the oviposition period varying from 24 to 39 (mean = 28), 21 to 28 (mean = 24), and 23 to 32 (mean = 27.2) days, respectively. This somewhat confirms the findings of Tank and Korat (2007), who found that Coccinella sexmaculata’s oviposition period was 16.1 days when raised on A. gossypii. When raised on L. erysimi, Coccinella undeceimpunctata had mean oviposition duration of 37.7 days, according to Solangi et al. (2007). When raised on Aphis craccivora, Lipaphis erysimi, and M. persicae, the post-oviposition duration varied from 3 to 8 days, 3 to 6 days, and 3 to 6 days, respectively, with an average of 5.4, 4.6, and 4.2 days. When raised on L. erysimi, Coccinella undecempunctata’s post-oviposition time was 4 days (Solangi et al. 2007). The egg-laying capacity of female beetles raised in laboratories on Aphis craccivora, Lipaphis erysimi, and Myzus persicae

7.2 Coleopterans

185

varying from 234 to 467 (mean = 376.5 eggs/female), 319 to 423 (mean = 364.9 eggs/female), and 325 to 433 eggs, respectively. The difference in fecundity may be the result of the prey species used for rearing as well as differing rearing settings. Cheilomenes sexmaculata, a predatory coccinellid that is extensively dispersed in India, Iran, Australia, and other areas of the Oriental region, is one of the probable aphid predators. Aphis craccivora, Aphi gossypii, Rhopalosiphum maidis, and Lipaphis erysimi all had mean incubation periods of 3.2, 3.3, 3.4, and 4.2 days, respectively (Pandi et al. 2012). Similarly, when raised on various aphid hosts, instar and pupa development times differed greatly. In contrast to larvae fed with other aphid hosts, such as Aphis craccivora, Rhopalosiphum maidis, and Lipaphis erysimi, which completed their life cycles in 9.6, 9.8, and 12.4 days, respectively, larvae fed with Aphis craccivora completed their life cycle in just 8 days. Cheilomenes sexmaculata’s total developmental period was determined to be longest on Lipaphis erysimi and shortest when fed with Aphis craccivora (13.7 days) (20.4 days). Male longevity ranged from 26.8 days for Lipaphis erysimi to 41.68 days for Aphis craccivora. The female also lived for a short time on Lipaphis erysimi (34.6 days) and for a considerable time on Aphis craccivora (48.2 days). Irrespective of the aphid hosts, females lasted longer (Pandi et al. 2012). When fed with Aphis craccivora, Aphi gossypii, Rhopalosiphum maidis, and Lipaphis erysimi, respectively, the mean female fecundity was determined to be 856.8, 692.6, 677.6, and 333.6 eggs. With Aphis craccivora, egg hatchability was found to be 67.2%, with Aphi gossypii, 63.4%, with R. maidis, and 39.8% with Lipaphis erysimi. On Aphis craccivora, the oviposition period was shown to be longer. The results of the current study showed that Cheilomenes sexmaculata’s life characteristics were significantly influenced by its aphid hosts, with Aphis craccivora showing the species’ best overall performance. Aphis craccivora’s shorter developmental period suggests that it was the most suitable aphid host for larval development. The highest survival was reported on A. craccivora and Aphi gossypii compared to Lipaphis erysimi (Pandi et al. 2012). This result is consistent with Patel and Vyas (1984) observation that Cheilomenes sexmaculata larvae can eat more Aphis craccivora per day. According to Solangi et al. (2007), Cheilomenes sexmaculata larvae and adults were voracious feeders on cotton aphid Aphis craccivora, alfalfa aphid Therioaphis trifollil, and maize leaf aphid R. maidis. Females ingested significantly more prey per day than did men, measured by the number of prey individuals consumed daily (Roy and Sinha 2002). The fact that Aphis craccivora and Aphi gossypii are suitable hosts for Cheilomenes sexmaculata suggests that these aphids do not sequester harmful substances from their host plants. Cheilomenes sexmaculata’s reduced life span when bred on Lipaphis erysimi compared to other prey species revealed the presence of potentially toxic compounds or alkaloids (Okamoto 1966). As previously documented for other Coleoptera, Lipaphis erysimi possesses a distinct fragrance that is likely the result of substances sequestered from its host plant. It is also probable that allyl isothiocyanates sequestered from its host plant are partially to blame for its decreased suitability as prey (Noble et al. 2002). Additionally, Cheilomenes sexmaculata grows more slowly when fed Lipaphis erysimi (Joshi et al. 1999). When female Cheilomenes sexmaculata were fed on Aphis craccivora,

186

7

Predatory Insects: Adults Biology of Various Orders

the longest reproductive and shortest non-reproductive periods were recorded, demonstrating the nutritional appropriateness. On Aphis craccivora, prolonged oviposition time had also been noted (Rajmohan and Jayaraj 1974). Earlier research had also discovered that Cheilomenes sexmaculata reproduced successfully on Aphis craccivora (Islam and Haque 1978). In addition, Mari et al. (2004) found that Cheilomenes sexmaculata and Cheilomene undecimpunctata, when raised on Therioaphis trifolii, had mean fecundities of 602.3 and 761.6 eggs, respectively. When raised on Aphis craccivora, Lipaphis erysimi, and M. persicae, male adult longevity ranged from 26 to 38 (mean = 31.6), 21 to 40 (mean = 30.1), and 22 to 47 (mean = 29.1) days, respectively. The average female longevity was 39.1, 33.9, and 37.1 days, with ranges of 34 to 50, 29 to 41, and 32 to 43 days. When raised on the alfalfa aphid, T. trifolii, Mari et al. (2004) found that the mean lifespan of male Cheilomenes sexmaculata and Cheilomene undecimpunctata was 29.7 and 50.7 days, respectively, whereas the mean lifespan of females was 34.9 and 56.7 days when Coccinella septempunctata was raised on Lipaphis erysimi (Khursheed et al. 2006). Moreover, it was mentioned by Shukla and Jadhav (2014) that the average lifespan for male was 15.2 days and for female it was 20.2 days. Furthermore, 36.6% and 56.6% of men and women emerged on average, respectively. About 17.7% of pupal deaths were average. The sex ratio of men to women was 1:1.5. It revealed that the predatory beetle’s adult sex ratio was skewed towards females (Shukla and Jadhav 2014). Another study used two types of aphid to examine the effects of food quality on the biology of the staphylinid generalist predator Tachyporus hypnorum (Rhopalosiphum padi and Sitobion avenae). For comparison, fruit flies (Drosophila melanogaster) were utilised as the prey. According to reports, Tachyporus hypnorum feeds mostly on aphids and exhibits a strong predilection for them. This study demonstrated that neither adult females nor Tachyporus hypnorum larvae view aphids as high-quality prey when compared to fruit flies. Aphid use by larvae was lower than that of adult females. Both types of aphids had an impact on reproduction; Sitobion avenae lowered fecundity whereas Rhopalosiphum padi in comparison to a fruit fly diet, padi decreased egg hatching success and lengthened the time spent in the egg stage. Pure feeds of both Sitobion avenae and Rhopalosiphum padi caused substantial larval mortality in Tachyporus hypnorum larvae. Larvae and adults had different rankings of the three prey categories based on various fitness metrics. The two aphid species had about the same overall value according to the adult fitness metrics, but the larval fitness factors showed the same rankings for all three prey types: Drosophila melanogaster > Sitobion avenae > Rhopalosiphum padi. Both aphid diets significantly decreased larval stage survival, but this did not apply to the adult females (Kyneb and Toft 2004). Coccinella septempunctata, also known as the seven-spotted lady beetle, is a sizable, extremely successful invasive species that has recently spread widely over North America. The relative small, introduced species Propylea quatuordecimpunctata, for example, has spread less quickly and extensively across North America than have Coccinella septempunctata and Harmonia axyridis. High fecundity associated with large size has been suggested as a possible key factor

7.2 Coleopterans

187

promoting predominance of invasive lady beetles, including both Coccinella septempunctata and Harmonia axyridis. The I, II, III, and IV larval stages of Coccinella septempunctata, lasted on average 2.9, 4.69, 5.4, and 7.5 days, respectively, whereas the pupal stage lasted 4.9 days (Sattar et al. 2008). Averaging 13.1 days, Coccinella septempunetata larvae reared on Acyrthosiphon pisum, pea aphids, developed substantially more quickly than those reared on Rhopalosiphum maidis, corn leaf aphids, did (16 days). Coccinella septempunetata adults developed from larvae raised on Acyrthosiphon pisum. Adults raised on Acyrthosiphon pisum were smaller and lighter than Rhopalosiphum maidis. Although larval prey had no effect on the Nearctic Hippodamia variegate or Propylea quatuordecimpunetata’s developmental times, the adult Propylea quatuordecimpunetata was heavier and larger when raised on Acyrthosiphon pisum. The first three Helicoverpa armigera instars dried up in 3 days at 23 °C. For Nearctic populations of these predators, Acyrthosiphon pisum is a very suitable larval diet; the redistribution programme releases in Acyrthosiphon pisum-infested alfalfa. These coccinellids can grow on Rhopalosiphum maidis in corn can use O. nubilaalis, but first instars cannot from as a substitute food source (Obrycki and Orr 1990). A significant predator of the red spider mite, Oligonychus coffeae, which infests tea, is the staphylinid beetle Oligota pygmaea. Oligota pygmaea biology, life cycle, and predatory effectiveness were investigated in a lab setting. Developmental phases lasted for an average of 23.0 days from egg to adult emergence, or 3.2, 5.7, and 12.5 days for eggs, larvae, and pupae, respectively. Each female produced an average of 400.5 eggs throughout the course of her lifetime after a pre-ovipositional period of 2.9 days. The average lifespan of an adult Oligota pygmaea was 54.1 days. Adult females lived for an average of 58.8 days compared to adult men’s 49.4 days. According to studies, the life table of this species is characterised by an intrinsic rate of natural population increase (r) of 0.118 days, net reproduction rates (R0) of 243.693 eggs/female, gross reproduction rates (mx) of 245.313 eggs/female, generation times (T) of 46.575 days, doubling times (DT) of 5.874 days, and a finite rate of increase of 1.125 days. By randomly selecting 25 tea leaves from each experimental block that was cultivated under the prevailing field conditions, Oligota pygmaea and its prey, Oligota coffeae, were counted and their seasonal abundance was tracked. The peak of Oligota pygmaea’s population coincided with the species’ overabundance in the tea fields. Weather conditions like high temperatures, low relative humidity, and few daylight hours have a negative impact on Oligota pygmaea populations. Oligota pygmaea larvae in their first through third instars ate 31.0–133.2 mite eggs per day. Oligota pygmaea larvae in their third instar ate on average of 133.2 eggs, 46.4 hexapod larvae, 39.6 nymphs, and 11.4 adults each day (Perumalsamy et al. 2009). The ladybird beetle Harmonia dimidiata was raised at temperatures of 15, 20, and 25 °C, and the developmental days and predation rates for each temperature were recorded and compared. The beetles were kept at room conditions while being fed Aphis gossypii. For beetles housed at 15, 20, and 25 °C, the net reproductive rates (R0) were 147.4, 98.7, and 62.5 offspring, respectively. In addition, the means,

188

7

Predatory Insects: Adults Biology of Various Orders

variances, and standard errors of the population parameters were estimated using the jackknife and bootstrap approaches. The bootstrap technique yielded sample means of R0 and other population parameters that fit a normal distribution, while the jackknife method produced biologically nonsensical zero values for R0. For beetles housed at 15, 20, and 25 °C, the net predation rates were 10,963, 13,050, and 7492 aphids, respectively. We combined the finite rate with the predation rate to create the finite predation rate in order to compare predation potential in its entirety (Yu et al. 2013). Further, findings demonstrated that Harmonia dimidiata is a more effective biological control agent for Aphis gossypii at 20 and 25 °C than at 15 °C when both the growth rate and the predation rate were taken into account. However, Adalia decempunctata lived the longest after consuming Aphis gossypii. The Adalia decempunctata beetles that fed on Aphis gossypii had the longest oviposition time (88.7 days), whereas those that fed on Aphis fabae had the shortest oviposition period (77.7 days). Females reared on Aphis fabae (1876.8 eggs/female) had a lower fecundity than those reared on Aphis gossypii (2382.6 eggs/female) and Ephestia kuehniella eggs (2405.1 eggs/female), which had greater fecundity rates. Adalia decempunctata’s intrinsic rates of growth were, respectively, 0.177, 0.171, and 0.155 day-1 when feeding on Aphis gossypii, Ephestia kuehniella eggs, and Aphis fabae. The highest finite rates of increase when raised on Aphis gossypii and Ephestia kuehniella, respectively, were 1.193 and 1.187 day-1. Adalia decempunctata’s mean generation time for Aphis fabae was much longer than for other hosts (41.40 days). The findings of this study demonstrated that the rate of energy reserves in emerging females was significantly influenced by the quality of the host. We come to the conclusion that all three host species might be viewed as necessary prey. Adalia decempunctata adults are affected by these hosts’ influence on the larval developmental phase and reproductive success; however, the outcomes of the biochemical experiments indicated that Aphis gossypii and Ephestia kuehniella eggs were the favoured hosts (Mojib-Haghghadam et al. 2018). Additionally, Golizadeh and Jafari-Behi (2012) discovered effects of three different aphid species (Macrosiphum rosae, Aphis fabae, and Aphis gossypii) on the biological characteristics of Hippodamia variegata. According to these findings, Hippodamia variegata fed on Aphis gossypii had the least total developmental period (15.2 days), whereas those fed on Aphis fabae had the longest (18.9 days). Adalia decempunctata’s adult weight (mg) is similarly impacted by its preys (Fig. 7.1) (Mojib-Haghghadam et al. 2018). Aphis fabae, also known as the variegated lady beetle, or Hippodamia variegata, is one of its main coccinellid predators. This coccinellid is widely dispersed throughout much of the Palearctic region and is a significant aphidophagous predator in Europe. The beetle has been observed in Australia feeding on 12 different aphid species and one psyllid species that attack different crops, weeds, and ornamental plants. Hippodamia variegata life table parameters were examined using conventional female age-specific life tables of individuals raised on a variety of aphid species, including Dysaphis craraegi, Aphis gossypii, Myzus persicae, Brevicoryne brassicae, and Rhopalosiphum padi, as prey on various crops under various environmental conditions. Of the 123 Hippodamia variegata eggs that were initially

7.2 Coleopterans

189

Fig. 7.1 Aphis gossypii, Ephestia kuehniella eggs, and Aphis fabae effects on Adalia decempunctata adult weight (mg)

gathered for the life table study, 96 successfully hatched, and 81 of these matured into adults. Hippodamia variegata has a pre-adult mortality rate of 34.1%. About 16.3 days were required for the development of all pre-adult stages. Hippodamia variegata male adults had an average lifespan of 62.38 days, which was a lot longer than that of the female adults (44.93) (Farhadi et al. 2011). Authors further reported that, the first copulation occurred within 24 h of emerging and occurred repeatedly throughout adulthood. About 3.4 days after emergence, on average, mated females started ovipositing (pre-oviposition period). Total 19.6 days made up the entire pre-oviposition phase. The average female fecundity was 1139.2 eggs (Farhadi et al. 2011). When given Lipaphis erysimi, the mustard aphid, Propylea dissecta males and females lived an average of 32.2 and 33.8 days, respectively (Hemiptera: Aphididae). The times before oviposition, during oviposition, and following oviposition were 7.3, 28.0, and 4.1 days, respectively. For eggs and the first instar, the likelihood of the age-stage survival rate (Sxj) was 1.0. Age-stage survival rate likelihood varied from 0.0 to 1.0 for male, female, pupa, second through fourth instars, and Sxj. The fact that some phases overlap suggests that individual developmental rates varied more widely. The ranges for age-specific survival rates (lx), age-specific fecundities (mx), age-specific net fecundity (lxmx), age-stage-specific fecundities (fx), SAD%, age-specific life expectancy (Ex), and age-specific reproductive value (Rx) were, respectively, 0.03–1.00, 0.00–4.57, 0.00–0.97, 0.00–8.67, and 0.00–7.13 (Boopathi et al. 2020). Recently, Ghanim et al. (2021) found that the average female fecundity of Hypodamia tridecimpunctata and Chilomenus propinquaisis was 150.7 and 193.4 eggs when these predators were reared on artificial diet (AD1-dried fish 10.5, pollen grains 4.1, dry powdered aphids 4.6, yeast (powdered) 3.5, sucrose 61, dry powdered drone honey bee 4.5, royal jelly (capsules) 4.2, stereptophenicol 1.7, and multivitamins and minerals 5.9%). When this predator was raised on artificial meals, the acquired data showed that the average fecundity of Hypodamia tridecimpunctata

190

7

Predatory Insects: Adults Biology of Various Orders

and Chilomenus propinquaisis females recorded 189.7 and 250.8 eggs, respectively (AD2). It can be concluded that the artificial diets (AD2-dried fish 11.9, pollen grains 4.6, dry powdered aphids 5.5, yeast (powdered) 2.9, sucrose 57, dried yolk of eggs 6.3, royal jelly (capsules) 3.7, stereptophenicol 1.5, and multi-vitamins and minerals 6.1%) were very effective for rearing Hypodamia tridecimpunctata and Chilomenus propinquaisis (AD1). The information gathered demonstrated that each species’ females laid more eggs when fed artificial meals (AD2). On a LEPS diet (lyophilised artificial diet, Ephestia kuehniella eggs, bee pollen and eggs of Sitotroga cerealella), Oenopia conglobata had the lowest projected fecundity (83.6 eggs). Though much shorter than those of LEPA (lyophilised artificial diet, Ephestia kuehniella eggs, bee pollen, and cysts of Artemia urmiana) and LEPS, the oviposition duration (17.6 days) and pre-oviposition period (LEP-12.7 days) for LEP were significantly longer than those of LEPA and LEPS. On the LEP, Oenopia conglobata adults lived for 31.0 fewer days than they did on the other two diets. More eggs hatched into larvae each day on LEP compared to other diets (6.9). In conclusion, a food rich in a variety of nutrients can prolong Oenopia conglobata’s life, but as the proportion of Ephestia eggs rises, the predator’s reproductive characteristics are positively impacted (Vahmani et al. 2022).

7.2.1

Plant Varieties/Cultivars on Biology

On the growth, reproduction, and survival of Propylaea japonica, the impacts of three cotton cultivars with low (ZMZ13), medium (HZ401), and high (M9101) gossypol levels were examined. Between cotton aphids raised on the three distinct cultivars and those fed with Propylaea japonica, no appreciable changes in survival and lifetime fecundity of the plant were found. The development time and adult weight of Propylaea japonica fed M9101 aphids were much shorter than those of the other two cultivars. The elevated fatty acid content of the prey aphids may have contributed to the shortened larval developmental period and increased adult weight of Propylaea japonica fed cotton aphids raised on the high gossypol-containing cultivar. According to our findings, Aphis gossypii were inhibited by high levels of gossypol in the host cotton, while Propylaea japonica, at the third trophic level, grew and developed favourably. This shows that one type of host plant resistance and predator biological control are compatible. The impact of the allelochemical components on both herbivores and entomophagous insects should be taken into account in integrated pest management (Du et al. 2004).

7.3

Mantodea

Approximately 2 weeks following the last moult, mantid adults reach sexual maturity. Males in some mantid species may moult and reach sexual maturity before females do, or they may take longer. Mantodean species frequently exhibit sexual

7.3 Mantodea

191

dimorphism because males are typically smaller than females. According to a recent study, female mantids employ sexual pheromones to entice males, and this is thought to be the main method of long-distance partner attraction. Similar to how female Lepidoptera moths pose when they release sex pheromones, females often arch their abdomens when producing pheromones to attract males. When a male and female are ready to mate, the male hops onto the female’s back and attaches his forelegs to her thorax. The female can eat the male completely or only his head at the end of a mating ritual that can last anywhere from 2 to 8 h. Barry and Wilder (2013) investigated whether the lipid and protein composition of live prey had an impact on the mating preferences, behaviour during reproduction, egg production, and nutrient reserves of adult female praying mantids, Pseudomantis albofimbriata. Despite being fed the same total biomass of food, females on a highprotein diet produced more than twice as many eggs as females on a high-lipid diet. Furthermore, the food that females were fed had a direct impact on the lipid and protein composition of eggs and the female body (i.e. high-lipid content on the highlipid diet). What was even more startling was how the diet therapy affected the number of men attracted to women: while 56 male were attracted to female on the high-protein diet, only one man was on the high-lipid diet. Although it is expected that females that have more eggs will attract more males, the extremeness of this discrepancy is unexpected given that earlier research has revealed that females with as little as a few eggs can draw a large number of males. Further, Barry and Wilder (2013) findings thus reveal that rather than just egg quantity, the quality/nutritional composition of eggs may have an impact on female pheromone production. All other behaviours observed during mating trials, including the frequency of sexual cannibalism, showed no discernible variation. Praying mantids may be predicted to select more protein-biased prey in nature or, if prey choice is limited, to have better reproductive output or population expansion in groups dominated by protein-rich prey, given the favourable effects of prey protein content on mate attraction and egg production (Barry and Wilder 2013). Ephestiasula pictipes, a frequent predator in cashew farms, was investigated in captivity for three seasons (Feb–May, Jun–Sep, and Oct–Jan) while preying on larger wax moth larvae. Comparing Feb–May to Jun–Sep and Oct–Jan, the incubation and nymphal developing periods were shorter. For a span of 70–77 days, the female mantids produced ootheca once every 3–4 days. While there was no significant variation in sex ratio, oviposition interval, or pre-oviposition period between seasons, adult longevity and oviposition period were considerably higher throughout the months of February to May (Vanitha et al. 2016). The beneficial characteristics of Ephestiasula pictipes allow for its mass upbringing and potential development as a successful biocontrol agent in the future. These include high fertility, high fecundity, a shortened life cycle, multivoltine, and the practicality of laboratory rearing throughout the year (Vanitha et al. 2016). The sex ratios of nymphs differed between individual oothecas but were predominantly male biased. Overall, 57% of the nymphs that survived to adulthood were males and 43% were females. However, of the 48 individuals that reached adulthood, 18 (37.5%) were female and 30 (62.5%) males (Greyvenstein et al. 2020).

192

7

Predatory Insects: Adults Biology of Various Orders

For the first through third instar stages, aphids (Brevicoryne spp., Hemiptera: Aphididae) were provided as food for the Mantodean species Harpagomantis tricolour. Thereafter, live crickets (Acheta sp., Orthoptera: Gryllidae) of various sizes (nymphal stages, i.e. pinheads), as well as their developmental period. The findings show that there were no discernible variations between the average duration per instar of males and females. It took the nymphal stage about 20 weeks to complete. Even though this difference was not significant, females needed a longer nymphal stage (145.71 days) than men (128.0 days). In this study, the average Harpagomantis tricolour life cycle lasted 6 months (191.3 days) (Greyvenstein et al. 2021). The study shows that 45 of the 63 neonatal Harpagomantis tricolour nymphs that hatched from the 12 various oothecas completed their life cycles (14 males and 31 females). The act of copulation itself lasted for almost 6 h, and the mean time between mating and the development of an ootheca was 11.82 to 9.51 days. About 20 weeks passed during an ootheca’s incubation period (143 days). Harpagomantis tricolour had an average hatch rate of 31% and an average survival rate of about 68%. The sex ratio varied among the oothecas, but the average (M:F) ratio was 1.5. Two of the oothecas only gave birth to males, while the other two only gave birth to females (Greyvenstein et al. 2021).

7.4

Syrphids

Although they are excellent flyers, adult aphidophagous syrphids are highperformance insects that spend most of the year dormant in cold, damp, or windy conditions. While larvae typically require aphid feeding to complete their development, adult hover flies need honeydew, nectar, or nectar plus pollen to assure reproduction. Adult: The false or spurious vein in the wing that crosses the r-m between R4+5 and M1+2 distinguishes syrphid adults from other dipterans. The R5 cell has been shut. The proopsis is short, the face is thin, and there are no grooves beneath the antennae. Syrphid adults are known as hover flies or flower flies because of their propensity to hover close to sources of pollen and prey. Adults have an abdomen that is black with brightly coloured spots. They frequently resemble honeybees and wasps. The adults are out and about during the day, visiting flowers to eat the pollen and nectar (Joshi et al. 2013). Oviposition Sites: In the 1960s, a beautifully straightforward model of hoverfly oviposition behaviour was developed. The presence of aphid-infested plants is the only requirement for female aphidophagous hoverflies to choose a potentially successful oviposition site because of their considerable mobility. Syrphid females choose an oviposition place deliberately so as to stimulate various hoverfly species’ oviposition responses. Over the past few decades, observations and laboratory and field studies have usually supported this hypothesis. Syrphid eggs typically wind up being laid close to aphid colonies as a result of their oviposition and seeking behaviour, allowing the newly emerged young larvae to find the food sources right away. After finding aphids, syrphid predators might perform a more thorough local search. A female aphid approaches an infested plant in a straight line, hovers, and

7.4 Syrphids

193

moves slowly around plants until it reaches a position opposite and close to an aphid colony, where it hovers for a brief period before landing with the ovipositor extended. This behaviour has been observed in the field and in the lab. The egg is then laid after the ovipositor has been bent ventrally and pulled over the substrate. Oviposition sites vary greatly and are influenced by both the quantity and location of eggs laid. Although some species of syrphid lay their eggs in batches far from the colony or even on uninfected plants, syrphid eggs are frequently laid singly, either nearby or inside aphid colonies. In the latter scenario, immature larvae might eat conspecific eggs to live. Aphids of various kinds have been discovered in the wild with syrphid eggs attached to them, proving how closely aphids and eggs can coexist (Almohamad et al. 2009). Description of Adults According to Bugg et al. (2008), the following aphidophagous syrphids are the most prevalent in California: Allograpta obliqua, the adult of the chevroned Allograpta fly, is shorter than Eupeodes volucris and is 0.85 cm or less in length. The medial stripe is incomplete, and the face is yellow. This species features two oblique yellow spots close to the tip as well as transverse yellow stripes on the abdomen. The 0.9–1.1 cm long, smooth, green larva with a wide, white median stripe. There are obvious breathing tubes. Other species of Allograpta exotica, among others, might also be seen. Eupeodes americanus, an adult American flower fly, resembles S. opinator, which is 0.9–1.2 cm long, has shiny thorax and black vitta (stripes) on its face, including a stripe along the front. Larvae are around 1.1 cm long, salmon brown, yellowish brown or black and white with yellowish white markings. Either segment from 6 to 11 has a transverse rectangular bar, or the dorsal lateral carinas of the larva have a thin line running down each side. The heartline, also known as the dorsal blood artery, is represented by six wedge-shaped black markings with a broad brown rim. This species’ larvae are particularly active, and the early instars feature observable black setae. The adult Eupeodes flower fly (Eupeodes volucris), which is about 0.85–1.00 cm long and has a face that is white yellow with black cheeks and a dark medial stripe, resembles the Scaeva flower fly in appearance. Males also have a small cylinder at the tip of the abdomen. The larva is between 0.9 and 1.4 cm long, fairly spiny, and greenish with soft pink and white, yellow and white, or green and white dorsal streaks. Two lateral narrow, uneven white lines that follow the ridge of the dorso-lateral segmental bristles surround the dorsum. Adult Paragus tibialis are small (0.3–0.5 cm long) and have a rounded posterior abdomen. A wide median black line runs from the antennae to the mouth edge, and the face is bright yellow with yellow on the sides. The female’s abdomen can be either reddish brown or greenish black, as opposed to the males, which is reddish brown. The larvae measure approximately 0.75 cm in length, 2–2.5 mm in width, and 1.5 mm in height. They range in colour from yellow to light yellow-brown. Platycheirus species adults are dark in colour, 1.0–1.1 cm long, and have faint silver or tan patterns on the abdomen. Silver, dark grey, or black describe the face. Only

194 Table 7.1 Influence of various preys on the fecundity of syrphids

7

Predatory Insects: Adults Biology of Various Orders

Syrphids Paragus serratus Paragus yerburiensi Paragus tibiali Betasyrphus serarius Dideopsis aegrota Ischiodon scutellaris

Episyrphus nubilipennis Episyrphus balteatus Syrphus pyrastri Eupeodes corollae

Prey(s) Aphis craccivora Aphis craccivora Toxoptera aurantii Toxoptera aurantii Aphis craccivora Toxoptera aurantii Aphis craccivora Brevicoryne brassicae Toxoptera aurantii Toxoptera aurantii Brevicoryne brassicae Lipaphis erysimi Brevicoryne brassicae

Fecundity 40.4 11.4 11.7 11.6 6.8 3.5 607.8 – 7.6 6.7 400–721 355 704

Data are not available

this species regularly produces parallel, continuous clusters of eggs. The 1.0–1.2 cm long, tan to orange larvae is little. These species are probably found on plants with a lot of aphids (Smith and Chaney 2007). Scaeva pyrastri, an adult Scaeva flower fly, is 1.27 cm in length. The medial vitta are dark and the face is white (stripes). There are six curving, white stripes on the dark abdomen. The light green larvae, measuring 1.2–1.8 cm long, have a white longitudinal dorsal stripe. Adult species of Sphaerophoria are comparable in size to Allograpta obliqua, measuring 0.85 cm in body length with a slender, cylindrical abdomen. A medial black stripe may be present, and the face is either white or yellow (as with Sphaerophoria sulphuripes). The transparent, 1.0 cm length, greenish yellow larvae are of average size. Syrphus opinator adult, the adult form of the western Syrphus fly, is 0.7–1.2 cm long, with a yellow face, two black spots, and two bands that cross the abdomen. The spiky, 1–1.3 cm length, yellow or brown larvae are spiny. Adult species of Toxomerus are tiny, spherical flies with dark lines around the abdomen’s borders. Toxomerus marginatus has a body length of 0.5–0.6 cm, while Toxomerus occidentalis has a body length of 0.6–0.75 cm. The latter species’ males can be identified by their notably larger hind femurs. Toxomerus marginatus has a yellow face, while the female has a dark forehead with lateral yellow stripes. Toxomerus occidentalis has a white face, but its forehead is black. Larvae have a visible gastrointestinal system that is coloured by their food supply, are between 0.4 and 0.45 cm length, transparent, and cream in colour. When extremely few or even no aphids are present, this type is more prone than others to lay eggs on plants (Smith and Chaney 2007). Joshi et al. (2013) analysed information on the biological characteristics of several Syrphid predators, including their larval phase, pupal period, adult period (days), and fecundity (the number of eggs laid by each female). Table 7.1 contains the summary information.

7.5 Dermaptera

195

Syrphid flies Pseudodorus clavatus is one of the most prevalent and significant predators of citrus aphids in Florida, the Caribbean islands, and South America. It is a native of tropical and subtropical areas of the New World. The adult sex ratio of Pseudodorus clavatus at eclosion was 0.485 (male: female), although 50% of females perished before oviposition. At 23 °C, female ovipositing flies had a mean lifespan of 29.8 to 1.9 days and male flies of 16.8 days (Belliure and Michaud 2001). According to Lillo et al. (2021), when given Myzus persica, Eupeodes corollae had a mean lifespan of 23.8 days, with the males living longer than the females. The wheat aphid, Schizaphis graminum, is the food source for Ischiodon scutellaris and Episyrphus balteatus, which are frequently found in central Europe and South Asia (Faheem et al. 2019). Due of the high female reproductive rates and voracious eating capacities of their larvae, both species are effective aphid predators in natural agro-ecosystems. In comparison to the corresponding males (17.2 and 16.2 days), females of both species had greater life spans (20.4 and 22.4 days).

7.5

Dermaptera

When Rhynchophorus ferrugineus infestations are present, the predator Chelisoches morio is regularly spotted in the coconut palms’ crowns. Abraham and Kurian (1973) studied the biology of the forficulid both within a lab and outside in Kerala State of India. The eggs, which were placed by the females in masses of 38–89 inside the leaf axil pockets, weighed an average of 156 apiece. The duration of the incubation period, the four nymphal instars, the adult stage, and the pre-oviposition period was 6–7, 6–10, 7–11, 9–19, 22–114, 20–24, 20–29, and 11–28 days, respectively. Average daily consumption for nymphs and adults was 4.2–6.7 larvae and 5.3–8.5 Rhynchophorus ferrugineus eggs (Abraham and Kurian 1973). In the Philippines, researchers examined the biology of Nala lividipes and Euborellia annulata in a lab and on maize fields. The overall development time for Nala lividipes males and females at room temperature was 18–23 and 20–25 days, respectively. Both species typically had four nymphal instars; however, under stressful circumstances, male Euborellia annulata had five. At different heights on maize plants, both species looked for and consumed egg masses and larvae of the pyralid Ostrinia furnacalis. Several other pest species were also attacked by them. For both species, tunnelling, mating, and maternal behaviour are described. Euborellia annulata and Nala lividipes females produced an average of 320.7 and 200.8 eggs per female. In fields of sorghum, cabbage, soybeans, mung beans (Vigna radiata), and sugarcane, Nala lividipes and Euborellia annulata were discovered. First-instar nymphs of both species that were 2–3 days old had a thermal limit for survival of 40–42 °C and were vulnerable to drought. Additionally sensitive to high humidity (92–100%) was Nala Lividipes (Situmorang and Gabriel 1988). Butnariu et al. (2013) documented Doru lineare, an earwig, providing maternal care (Dermaptera: Forficulidae).

196

7

Predatory Insects: Adults Biology of Various Orders

Tirathaba rufivena is preyed upon by Chelisoches morio (Dermaptera: Chelisochidae), a significant predator of Tirathaba rufivena (Lepidoptera: Pyralidae). Additionally, under controlled laboratory circumstances, Chelisoches morio’s ability to feed on Tirathaba rufivena was investigated. According to a biological study on the Chelisoches morio, adult females typically lay their eggs in masses. An average of 140.17 eggs was laid by each female, and the incubation period was 7.50 days. Four instars made up the nymphal development period. 83.95 days made up the immature phase. Males and females lived 58.60 and 93.55 days longer than each other, respectively. According to the findings of the study on the eating ability of Chelisoches morio using Tirathaba rufivena as food, Chelisoches morio adults could eat, in that order, 11.1, 7.9, 7.1, 6.8, and 5.9 Tirathaba rufivena larvae in the first through fifth instars. Chelisoches morio demonstrated intense predation on this pest at all larval stage, indicating a great potential for its application in Tirathaba rufivena management (Zhong et al. 2016). In its natural environment, the European earwig Forficula auricularia (Forficulidae). We use both man-made and natural shelters to depict the spring field population structure and dynamics in organic apple orchards in southeast France. In a laboratory reproduction investigation, two nymph cohorts could be distinguished. Pre-imaginal survival was 1.74 times higher in the first brood, which had its initial egg-laying event in late November, than it was in the second brood, which had its egg-laying event in early April. With a higher mortality rate in the first brood than the second, the egg phase was the most prone. Nymphal survival rose after the second nymphal instar (N2), with levels exceeding 96%, regardless of brood quantity. From the end of April to the middle of June, N3 were typically seen in orchards in both natural and man-made shelters. The N4 life stage was the most prevalent and the only one present on the majority of sampling dates, particularly beginning in early May. A given year’s abundance of new adults gradually rose starting in June, notably in the artificial shelters. With the help of this descriptive study, phenological models may now be created to reduce earwig mortality brought on by horticultural management methods and raise the degree of predation in either conservative or augmentative strategies (Dib et al. 2017). Male and female Euborellia annulipes lived substantially longer than other species, with the longest lifespan ever recorded on an artificial diet alone (190.4 and 187.1 days for male and female respectively). Compared to females fed Plutella xylostella (Lepidoptera: Plutellidae) life stages, those fed an artificial diet survived at a much greater rate (Fig. 7.2) (Nunes et al. 2022).

7.6

7.6. Hemipteran Predators

7.6.1

Reduviids

According to life table research conducted on Sycancus collaris in a lab setting, the species’ net reproductive rate (R0) was 30.46 female eggs per female. The actual natural increasing rate (rm) was 0.066. Every week, the population expanded by

7.6 7.6. Hemipteran Predators

197

Fig. 7.2 Adult life span (days) of Euborellia annulipes males and females fed with various stages of Plutella xylostella (Lepidoptera: Plutellidae) and synthetic food

1.59, and the predator population doubled every 10.6 days. The length of the nymphal stage was 49.3 days. Females had a longer life expectancy (45.2 days) than males (31.8 days) (George et al. 1998). On the bottom of the rearing bottles, Cydnocoris gilvus lay batches of broadly oval, orange-yellow eggs, each of which was firmly adhered to the substrate. There were a total of 56.3 eggs deposited by the females in 8.7 batches. The batch sizes ranged from 3.3, which was the average lowest, to 11.3, which was the average maximum. Unfertilised eggs shrink after a few days, but fertilised eggs develop into dark reddish chorion with eyespots before hatching. After 8.2 days, the eggs hatched. Cydnocoris gilvus had an 11-day pre-oviposition time and a 3.8-day postoviposition period following the last batch of eggs it laid. The detected oviposition index value was 67.1%. A generation from Cydnocoris gilvus took 45.5 days to complete. At 67.8 females per female and an intrinsic capacity of natural increase (rc) of 0.07, gross reproduction (mx) was achieved. The average generation time (Tc) was 52.3 days (Srikumar et al. 2014a, b). Reduviid predators of Harpactorinae, Sycanus species feed on a variety of insects, including Lepidoptera, Coleoptera, and Diptera larvae and pupae. They have a great ability to control pests. Sycanus versicolor, a ferocious predator of Heliothis armigera and Earias insulana, belongs to the genus Sycanus and has been investigated specifically in terms of biology. Vennison and Ambrose (1992) documented the biology, behaviour, and biocontrol effectiveness of Sycanus reclinatus from Southern India. Male and female Sycanus reclinatus lifespans range from 5 to 54 days and 5 to 50 days, respectively. The gender ratio is marginally in favour of men (Vennison and Ambrose 1992). Adult male Sycanus affinis survived for 86.1 days, whereas females lived for 69.3 days. Females may lay up to 807 eggs at a time, with pre-oviposition and oviposition times averaging 3 and up to 37 days, respectively (Satpathy et al. 1975). Based on laboratory studies carried out in 1971–1972, the biology of Sycanus affinis, a lepidopterous larval predator found in coconut plantations in Orissa, India, is described. The average length of the

198

7

Predatory Insects: Adults Biology of Various Orders

egg, nymphal instars, and overall life cycle was 17.4, 90.1, and 106.9 days, respectively. Sycanus affinis has five nymphal instars (Satpathy et al. 1975). It is described how Sycanus affinis nymphs and adults prey on larvae and pupae, and a list of the eight Lepidoptera species they have attacked in the lab is provided. Termites, aphids, and grasshopper and cockroach nymphs were also targeted. A larva of Acherontia styx that weighed 37.2 times as much as the predator itself could not be overcome by an adult female. Less than one full-grown Corcyr cephalonica larva was devoured every day by the first four nymphal instars, but the fifth-instar nymphs and adults consumed five and one larvae per day, respectively (Satpathy et al. 1975). Corcyra cephalonica, a laboratory host species of Rhynocoris marginatus, was reared in four different groups: 1. 2. 3. 4.

The Corcyra cephalonica of the jower medium fed group (JFC) The Corcyra cephalonica of the rice medium fed group (RFC) The Corcyra cephalonica of the sorghum medium fed group (SFC) The Corcyra cephalonica of the wheat (WFC)

The longest adult longevity was found in the JFC group of Rhynocoris marginatus, with males living an average of 176.4 days and females 186.4 days. Male and female longevity were significantly different from one another (P 0.05). The wheat medium fed group had the shortest adult longevity (147.9 and 135.7 for males and females, respectively), although it was not statistically significant. Rhynocoris marginatus’ pre-oviposition period changed depending on the prey rearing medium. The jower medium fed group (62.7 days) had the lowest pre-oviposition period among the four media, followed by the rice medium fed group (70.8 days), the wheat medium fed group (73.3 days), and the sorghum medium fed group (63.3 days) (68.0 days). Similar to pre-oviposition, higher fecundity was seen in the wheat medium fed group (223.6), followed by sorghum medium fed group (208.3), and lowest in the rice medium fed group (360.9). (208.4). Thus, a diet made up only of Corcyra cephalonica raised on wheat seems to satisfy the ideal needs of Rhynocoris marginatus egg-laying females (Shahayaraj and Sathiamoorthi 2002). Sphedanolestes variabilis’ pre-oviposition period was closer to Rhynocoris fuscipes and shorter than that of other harpactorines, such as Rhynocoris marginatus (33.3 days) and Rhynocoris kumarii (26.0 days) (19.0 days). It was longer than those of the Sphedanolestes signatus (9.3 days), Sphedanolestes pubinotum (11.7 days), Sphedanolestes sp. (8.6 days), and Sphedanolestes minusculus among the Sphedanolestes species (12.6 days). The Sphedanolestes variabilis eggs hatched in 6.9 days in a lab setting, with 71.3% of the eggs hatching between morning (8–11 AM) and afternoon (1–3 PM) hours. According to Sphedanolestes signatus sibling species, Sphedanolestes pubinotum (81.5%), Sphedanolestes himalayensis (77.8%), Sphedanolestes sp. (92.9%), and Sphedanolestes minusculus (95%) the higher hatching percentage is a defining distinguishing trait of harpactorines (Ambrose et al. 2009).

7.6 7.6. Hemipteran Predators

199

In India, Helopeltis antonii (Hemiptera: Miridae) was found being consumed by six species of harpactorine reduviids (Hemiptera: Reduviidae: Harpactorinae). They were Panthous bimaculatus, Irantha armipes, Epidaus bicolor, Euagoras plagiatus, Endochus albomaculatus, and Sphedanolestes signatus. The larger wax moth larvae, Galleria mellonella, were used to grow these species in a lab environment. For these six species, the incubation duration, stadial period, nymphal mortality, fecundity, longevity, and sex ratio were examined. The most eggs were laid by Panthous bimaculatus, which also had a long incubation period. Euagoras plagiatus had the shortest incubation period. Endochus albomaculatus had the lowest nymphal survival rate and Euagoras plagiatus had the highest. The six species showed considerable differences in post-embryonic development from the first nymphal instar to the adult stage. While Euagoras plagiatus, Epidaus bicolor, and Endochus albomaculatus had male-biased sex ratios, Irantha armipes, Sphedanolestes signatus, and Panthous bimaculatus had female-biased sex ratios. The six reduviid species’ adult lifespans, both for males and females, greatly varied. Various species exhibited varying levels of rostral thrusting and prey capture aggression. Only in Epidaus bicolor was post-copulatory cannibalism of males by females noted. The quantity of eggs laid, or batches, also differed greatly between species. Numerous desirable biological traits suggest that Euagoras plagiatus can be successfully massproduced. Panthous bimaculatus (109.8 eggs/batch) laid the most eggs per batch, while Endochus albomaculatus (37.2 eggs/batch) and Epidaus bicolor (39 eggs/ batch) were statistically equal. Compared to Endochus albomaculatus (203 eggs), Epidaus bicolor (181 eggs), and Panthous bimaculatus laid much more eggs (465 eggs). The reduviid egg hatchability ranged from 84% to 99%. In contrast to Epidaus bicolor, Irantha armipes, and Panthous bimaculatus, Euagoras plagiatus (5.4 days), Sphedanolestes signatus (6.0 days), and Endochus albomaculatus (6.4 days) had shorter incubation periods. Panthous bimaculatus’s incubation period was much longer (21.0 days). Adult male and female lifespans also varied significantly from one another. In comparison to Endochus albomaculatus, Epidaus bicolor, Euagoras plagiatus, Irantha armipes, and Panthous bimaculatus, the lifespan of Sphedanolestes signatus was significantly lower for both sexes (Srikumar et al. 2014b). Sycanus is thought to have the potential to act as a biological control agent. There have been reports of the Sycanus falleni in Cambodia, China, Myanmar, and Vietnam. However, nothing is now known about the biological and ecological characteristics of S. falleni. The reduviid S. falleni is a frequent tree-predator in the Vietnamese provinces of Bac Kan, Cao Bang, Ha Tay, Ha Tinh, Hoa Binh, Ninh Binh, Son La, Vinh Phuc, Dak Lak, Kon Tum, Gia Lai, and Lam Dong. It also frequents lowland semi-evergreen forests and agricultural environments. The reduviid Sycanus falleni, which preys on numerous lepidopteran larvae, has the potential to biologically manage nuisance insects in coffee, cotton, and vegetable plants. When comparing the Sycanus dichotomus adult life span, there were significant differences between the food sources: the adult life span was shorter when either Tenebrio molitor or Corcyra cephalonica larvae alone were fed (66 and 56 days, respectively), and it was significantly longer when both larvae were fed together

200

7

Predatory Insects: Adults Biology of Various Orders

(69.4 days) (Ahmad and Kamarudin 2016). The adult predator’s capacity for predation may be improved by the extended lifespan of adults. Food quality influences numerous predators’ ability to survive and develop, as well as how effectively adults function during reproduction. Additionally, Sycanus dichotomus adults that were raised using both larvae tended to weigh more than those that were raised using only Corcyra cephalonica or Tenebrio molitor larvae, by 18.4% and 7.7%, respectively (Ahmad and Kamarudin 2016). Liu et al. (2012) record certain biological traits of Sycanus sichuanensis based on laboratory rearing and field observations. Beginning in early April of the following year, matured nymphs leave their overwintering locations. Late April marks the height of emergence, and by early May, the populations are primarily made up of adults. Early to mid-May is when couples always mate. In late May and June, the females lay their eggs. Under laboratory settings, the gravid females lay eggs in masses made up of 55–70 eggs that are tightly bound together by slime. In oil palm, Sycanus dichotomus (Hemiptera: Reduviidae) is a frequent predator. It is suitable for the biological management of the bagworm since it can attack the larval stage of the worm. The suitability of Tenebrio molitor (Coleoptera: Tenebrionidae) and Corcyra cephalonica (Lepidoptera: Pyralidae) larvae as food sources for Sycanus dichotomus (Hemiptera: Reduviidae), a general insect predator, was assessed by Ahmad and Kamarudin (2016) by examining the effects of various prey consumption on the predator’s growth parameters. The findings revealed that several food sources varied in their suitability: Corcyra cephalonica considerably (P0.01) decreased the developmental time of the nymphal and adult stages compared to Tenebrio molitor larvae and the combination of both larvae by 17.5% and 6.9%, respectively. In addition, compared to late nymphs and adults, Corcyra cephalonica larvae alone constitute a more favourable food source for the development of the first and second nymph instars. However, the union of the two larvae resulted in a nymph that developed in 95 days less time, had a higher body weight, a longer femur, and lived longer as adults (69 days). In contrast to the Corcyra cephalonica larvae treatment, the treatment that fed on Tenebrio molitor larvae and a combination of both larvae had a higher ratio of male adults. Five Reduviidae species have been identified as Anacardium occidentale casheweating predators in India: Sycanus collaris, Sphedanolestes signatus, Endochus inornatus, Irantha armipes, and Occamus typicus. But Sycanus galbanus is also prevalent reduviid in cashew. It was clear that Sycanus galbanus exhibited sexual dimorphism. Males and females generally looked the same from the outside, with a black body and a yellow band on the wing, but they differed in size and shape. Males were relatively short (2.4 cm in length) and slender, with a pointed abdominal base, whilst females were larger (2.8 cm length) and had a circular bulged abdominal base. The adult male longevity (74 days) and whole duration of the male life cycle was 142.9 days, while the female longevity was 81 days and total duration was 150.02, indicating that females outlived males (Nitin et al. 2017). Sycanus collaris, an assassin bug, was known to combat pests in teak, marigolds, and tea. South Asia and South East Asia made up the majority of the distribution. Sycanus collaris should be mass-reared and released in biocontrol programmes, according to studies

7.6 7.6. Hemipteran Predators

201

that showed it has the ability to act as a biological agent to control Lepidopteran caterpillars (such as the teak defoliator, teak skeletonizer, poplar leaf defoliator, cotton leafworm, and black looper). Recently, the biology of the assassin bug Sycanus collaris in relation to the larvae of the leaf armyworm Spodoptera litura and the rice meal moth Corcyra cephalonica was studied in the lab (Rajan et al. 2017). Male and female adults who were fed Corcyra cephalonica on average lived 73.58 and 80.64 3.4 days, respectively. Male and female adults who were given Spodoptera litura on average lived 75.8 and 85.5 days, respectively. Sycanus annulicornis is a predatory reduviid that preys on a variety of insect orders, including Lepidoptera, Coleoptera, and Diptera larvae and pupae. According to laboratory rearing, Crocidolomia pavonana larvae (Lepidoptera: Crambidae) are a suitable prey for Sycanus annulicornis growth and development because their nymphal development is shorter (74.0 days) and adult longevity is greater (81.0 days for males and 64.8 days for females, respectively) than when reared on Tenebrio molitor larvae (Coleoptera: Tenebrionidae (44.0 days for male and 52.6 days for female). To target and eliminate the nettle caterpillar pest Setothosea asigna, the Sycanus annulicornis was used, which was reared by two prey species, Crocidolomia pavonana and Tenebrio molitor (Sahid and Natawigena 2018). The predatory insect Sycanus falleni is crucial for the biological control of pests on rice, soybean, corn, and vegetable plants. It is an omnivore animal with many different types of prey. The effect of feeding Sycanus fell on four distinct prey species, including Pieris rapae, Spodoptera litura, Plutela xylostella, and Corcyra cephalonica, on its biological characteristics. Prey has no role in this parameter, as evidenced by the pre-oviposition periods on Pieris rapae, Spodoptera litura, Plutela xylostela, and Corcyra cephalonica, which were 10.54, 11.42, 12.61, and 12.71, respectively. The most eggs were laid by Pieris rapae, which fed Sycanus falleni, at a rate of 243.35 per female. Spodoptera litura came in second with 165.08 eggs per female, followed by Plutela xylostella with 148.89 eggs per female, and Corcyra cephalonica with 113.89 eggs per female. The laboratory host has not yet confirmed this reduviis’ fertility (Truong et al. 2020). Except for the first instar, which had a significantly longer developmental time, the Sycanus annulicornis nymphs reared on the larvae of Crocidolomia pavonana (Lepidoptera: Crambidae) had a mean total nymphal period of 74.0 days, which was significantly shorter than that of those reared on the larvae of Tenebrio molitor (80.1 days). Adult Sycanus annulicornis raised on Crocidolomia pavonana lived a lot longer than adult Sycanus annulicornis raised on Tenebrio molitor (Coleoptera: Tenebrionidae). The males raised on Tenebrio molitor survived for a shorter time (44.0 days) than the females, but the males raised on Crocidolomia pavonana lived longer (81.0 days) than the females (64.8 days) (52.6 days). When raised on Tenebrio molitor and Crocidolomia pavonana, respectively, the gender ratio of Sycanus annulicornis was female-biased (1: 1.2 and 1: 1.9, respectively), and the various prey species had a substantial impact. According to Sahid and Natawigena (2018), Crocidolomia pavonana is an appropriate prey item for Sycanus annulicornis (Table 7.2).

202

7

Predatory Insects: Adults Biology of Various Orders

Table 7.2 Crocidolomia pavonana and Tenebrio molitor had an impact on Sycanus annulicornis male and female adult longevity (days), fecundity (eggs/female), and fertility (%) Life traits Male longevity (days) Female longevity (days) Fecundity (eggs/female) Fertility (%)

Crocidolomia pavonana 81.0 64.8 114.4 48.5

Tenebrio molitor 44.0 52.6 97.1 70.2

Table 7.3 Impact of Pieris brassicae and Spodoptera litura prey on Rhynocoris marginatus life table parameters Life table parameters Net reproductive rate (R0) Mean generation time (T ) Intrinsic rate of increase Finite rate of increase

Pieris brassicae 112.4 105.9 d 0.044 d-1 1.045 d-1

Spodoptera litura 411.4 93.28 d 0.064 d-1 1.066 d-1

Rhynocoris marginatus is a reliable biocontrol agent for controlling a variety of insect pests. The purpose of the study was to evaluate the life cycle and demographic characteristics of Rhynocoris marginatus feeding on the lepidopterous nuisance insects Spodoptera litura (Noctuidae) and Pieris brassicae (Pieridae). Rhynocoris marginatus developed more quickly when Spodoptera litura larvae were available at all life stages. In comparison to Pieris brassicae, which had 112.4 offspring per individual, Spodoptera litura had a greater net reproductive rate (R0), which was 411.4 offspring per individual. Rhynocoris marginatus’s mean generation time (T) on Spodoptera litura was shorter (93.28 days) than on Pieris brassicae (105.90 days). Rhynocoris marginatus was shown to have greater intrinsic and finite rates of growth when feeding on Spodoptera litura (0.064 and 1.066, respectively) than when feeding on Pieris brassicae (0.044 and 1.045, respectively). In comparison to Pieris brassicae larvae, Rhynocoris marginatus had a higher age-stagespecific survival rate and fecundity when feeding on Spodoptera litura larvae. When compared to Pieris brassicae, Spodoptera litura was the most suitable prey for Rhynocoris marginatus (Arshad et al. 2021). Additionally, Rhynocoris marginatus is suggested as a helpful biocontrol agent against the Spodoptera litura pest in the Arshad et al. (2021) study (Table 7.3). Another study used Rhynocoris marginatus that was raised alongside five orthoperan crop pests, including Diabolocatantops pinguis, Oxya nitidula, Atractomorpha crenulata, Orthacris maindroni, and Trilophidia annulata, as well as a lab host called Corcyra cephalonica (Sahayaraj et al. 2021). The findings show that all prey species sustain the predator’s whole life cycle, including its nymphal development, oviposition, and post-oviposition stages, adult longevity, and fertility. Additionally, the estimated macromolecular composition of the five orthopteran prey species—Phenacoccus solenopsis, Corcyra cephalonica, Dysdercus koenigii, Spodoptera litura, Diabolocatantops pinguis, Oxya nitidula, Atractomorpha crenulata, Orthacris maindroni, and Trilophidia annulata—and the enzymatic

7.6 7.6. Hemipteran Predators

203

profile of the predators raised alongside them were also calculated According to macromolecular, enzymatic, and biological research, Rhynocoris marginatus can successfully prey on Diabolocatantops pinguis (Sahayaraj et al. 2021) (Table 7.4).

7.6.1.1 Reduviids and Storage Pests A predator called Alloeocranum biannulipes is found in dried yam chips and is a potential biological control agent for Dinoderus porcellus (Coleoptera: Bostrichidae). Predator mortality rates, egg laying rates, pre-imaginal development timeframes, and pre-oviposition and oviposition durations were also noted. When Dinoderus porcellus served as a prey item for Alloeocranum biannulipes, the shortest pre-oviposition period (11.0 days) and the longest oviposition period (15.8 days) were observed. The equation y = 0.158 × 2 - 0.4073x + 3.8151 (R2 = 0.571) was used to estimate the daily number of eggs laid by Alloeocranum biannulipes female as a function of the number of Dinoderus porcellus larvae ingested (x). Female Alloeocranum biannulipes had less success turning prey into eggs as prey density increased. The prey density had no impact on the pre-imaginal development times or the rate at which Alloeocranum biannulipes eggs hatched. The survival percentage of Alloeocranum biannulipes nymphs fed with more (2–8) Dinoderus porcellus, as opposed to those that were fed only one larva, was noticeably greater (Loko et al. 2022).

7.6.2

Nabidae

Adult Nabis alternatus lived for around 38 days under laboratory settings (28 °C, 59% RH, and 15 h of light per day). Each female laid an average of 281 eggs over the course of around 30 days. Incubation took, on average, 6.5 days, and about 79% of the eggs hatched (Perkins and Watson 1972). Under laboratory circumstances (26–28 °C, 60–70% RH, and 15 h of photophase per day), Tropiconabis capsiformis (Nabidae) was researched. Adult nabids had a lifespan of 14.9 and 21.6 days for males and females, respectively. Average egg production was 105.3 eggs per female over 15.5 days. The incubation period was 7.6 days, and 78% of the eggs hatched out (Hormchan et al. 1976). When fed Plutella xylostella larvae instead of Sidnia kinbergi or Acyrthosiphon pisum, adult female Nabis kinbergii lived 58.6 days longer on average and produced 794 more eggs. Adult females were able to survive for 2 months on one Acyrthosiphon pisum every 4 days, demonstrating that the minimum nutritional requirements were modest. One Acyrthosiphon pisum was ingested daily until maximum longevity was reached; after this point, there was no further improvement in longevity. Although the rate of intake did not rise proportionately with the quantity of Acyrthosiphon pisum given, the rate of reproduction was favourably connected with the rate of feeding (Siddique and Chapman 1987). Nabis capsiformis’ biological traits were examined in a lab setting with temperature and humidity controls of 26 °C and 65 °RH and a photoperiod of 16:8 (L:D) h. The oviposition period, female longevity, and life span were each 18.4, 27.6, and 44.9 days longer than the egg incubation duration of 7.2 days. The average number

Biological data Nymphal development period Nymphal survival rate Pre-oviposition Oviposition Post-oviposition Adult longevity Lifetime Fecundity Hatchability

Diabolocatantops pinguis 81.0 93.3 17.5 21.9 7.2 43.6 127.6 107.4 96.5

Preys offered Corcyra cephalonica 78.2

89.8 12.7 16.7 6.4 30.0 107.2 125.6 94.2

93.3 18.4 20.6 7.0 46.0 99.6 91.6 95.7

Atractomorpha crenulata 56.6 90.0 15.5 19.6 7.3 42.4 118.2 64.0 93.9

Orthacris maindroni 75.8

90.0 15.1 17.0 7.7 39.8 113.9 47.0 91.3

Trilophidia annulata 74.1

7

83.3 15.7 19.2 6.5 41.4 118.3 84.0 94.2

Oxya nitidula 76.7

Table 7.4 Biological information on the five Orthopteran species—Orthacris maindroni, Diabolocatantops pinguis, Oxya nitidula, Atractomorpha crenulata, and Trilophidia annulata—and Corcyra cephalonica on which Rhynocoris marginatus was raised

204 Predatory Insects: Adults Biology of Various Orders

7.6 7.6. Hemipteran Predators

205

of eggs laid by Nabis capsiformis females during each ovipositional phase was 119.6, which was a notably high oviposition rate for the predator. About 77.8% of the eggs were hatchable (Fathipour and Jafari 2008). According to calculations, the Nabis pseudoferus adult pre-oviposition period, total pre-oviposition period days, and fecundity per female were 18.2, 52.9, 9.1, and 44.8 eggs, respectively. The intrinsic rate of rise (r), net reproductive rate (R0), gross reproductive rate, finite rate of population increase, and mean generation time (T ) of the main life table were estimated at 0.04 day-1, 12.71 eggs/individual, 35.2 eggs/ individual, and 1.0447 day-1, respectively (Madadi et al. 2016). Egyptian cotton leaf worm Spodoptera littoralis (Lepidoptera: Noctuidae) being consumed by Nabis pseudoferus (Hemiptera: Nabidae) at 25 ± 1 °C and 60 ± 5% RH during a 16: 8 light: dark photoperiod. Feeding on Spodoptera littoralis individuals is Nabis pseudoferus. The mortality rate was 33.33%, and females produced 50.92 eggs on average per lifespan. From the life table that was created, it was possible to determine the intrinsic rate of increase (rm = 0.079 females/female/day), net reproduction rate (R0 = 31.00 females/female), mean generation time (T0 = 43.246 days), gross reproductive rates (GRR = 37.992), doubling time (T2 = 8.729), and the finite rate (1.083) of Nabis pseudoferus (Karaca and Efe 2021).

7.6.3

Miridae Biology

Deraeocoris nebulosus, or dults, are drawn to a variety of host plants that sustain huge populations of prey, and they likely ovulate for a number of weeks (Wheeler Jr et al. 1975). For the predatory mirid beetle, Cyrtorhinus lividipennis, the presence of Tagetes erecta, Trida procumbens, Emilia sonchifolia (Compositae), and Sesamum indicum (Pedaliaceae) in bloom increased the survival of adult males and females (Zhu et al. 2014). During the same period, on Ephestia kuehniella (Lepidoptera: Pyralidae) eggs, Dicyphus maroccanus (Hemiptera: Miridae) the females produced 50.8 (n = 11) nymphs during their 15.8-day life period. Between days 8 and 12, when females were most fertile, they produced about 10 nymphs every day. The progeny’s sex ratio was 75.6 (percent of produced females). Dicyphus maroccanus had a net reproduction rate of 34.5 female eggs per female and an intrinsic growth rate of 0.0868 females per female per day. In comparison to the generation doubling time DT, the generation time T was 40.5 days (Abbas et al. 2014). The average amount of time spent engaging in predatory behaviour when B. tabaci, Tetranychus sp., Tetranychus parvispinus, or Tetranychus vaporariorum was the prey was 64.1, 80.4, 41.8, and 85.6% for Nesidiocoris volucer females and 54.5, 56.1, 24.6, and 75.8% for Nesidiocoris volucer males. On the amount of preys killed, a substantial relationship between sex and prey type was seen. The average number of prey killed by male and female Nesidiocoris volucer did not differ significantly for any prey, with the exception that females killed significantly more Tetranychus sp. than did males. Compared to other prey, Tetranychus parvispinus was killed substantially less frequently by both sexes. Regardless of the prey type,

206

7

Predatory Insects: Adults Biology of Various Orders

agar and the leaf disc were probed. Nesidiocoris volucer nymphs that were fed on Tetranychus sp. or Tetranychus parvispinus adults died before maturing into adults and presumably without prey. The survival rates for nymphs fed with B. tabaci, instars of Tetranychus parvispinus, Tetranychus vaporariorum, or E. kuehniella were 61.2, 66.7, 77.6, and 87.6%, respectively. The mean development period did not significantly differ between male and female nymphs regardless of diet. For Nesidiocoris volucer nymphs fed with E. kuehniella eggs, Tetranychus vaporariorum nymphs, B. tabaci nymphs, and Tetranychus parvispinus nymphal stages, the mean development time greatly varies. The longest development durations had the lowest survival rates for Nesidiocoris volucer nymphs. The interplay of sex and prey type had a considerable impact on the lifetime of Nesidiocoris volucer adults. Female Nesidiocoris volucers survived noticeably longer when given E. kuehniella or B. tabaci compared to the other species studied and male Nesidiocoris volucers fed Tetranychus vaporariorum. Both males and females had lower lifespans when Tetranychus parvispinus was fed to them or no prey was available to them. When fed B. tabaci or E. kuehniella, females lived noticeably longer than males (Marquereau et al. 2022).

7.6.4

Anthecoridae

The diet of the generalist predator Orius insidiosus affects crucial biological characteristics including reproduction and predation. Anagasta kuehniella eggs, Anagasta kuehniella eggs and pollen, Frankliniella occidentalis nymphs and adults, Frankliniella occidentalis nymphs and adults and pollen, Anagasta kuehniella eggs, Frankliniella occidentalis nymphs and adults, and Anagasta kuehniella eggs were tested. The pre-oviposition duration was shortest when the meal contained only Anagasta kuehniella eggs, pollen with prey (Anagasta kuehniella and/or Frankliniella occidentalis), or both of the prey species together. The highest values for oviposition period length (50.1, 48.0, 46.3, and 46.1 days), daily fecundity (3.8, 3.9, 4.0, and 4.2 eggs/female/day), total fecundity (190.3, 187.7, 185.2, and 193.6 eggs/female), and longevity (52.1, 49.9, 48.7, and 48.0 days) were observed in animals fed exclusively on Anagasta kuehniella eggs. When supplied as a sole food source or as a supplement, pollen had no effect on the predator’s performance in any way. Anagasta kuehniella eggs, which are easily generated in great quantities, are a crucial discovery because they provide ideal fake prey for the bulk production of Orius insidiosus (Calixto et al. 2013). According to Blaptostethus pallescens, the average fecundity was 150 eggs, and the average feeding potential was 100 eggs for nymphs and 630 eggs for adults (Ballal et al. 2003). Ballal et al. (2013) studied the biology and feeding abilities of Xylocoris flavipes. The fecundity was reported to be 61.96. The calculated values for the approximate rate of rise (rc), precise intrinsic rate of increase (rm), and finite rate of increase were 0.058, 0.045, and 1.14, respectively (r). Orius sauteri adult males lived considerably shorter lives than females did. Orius sauteri having a sex ratio of 2.8 females to one male with a mean pre-oviposition

7.6 7.6. Hemipteran Predators

207

time of 2.1 days, a fecundity of 95.4 eggs for every female, a female adult life expectancy of 21.1 days, and a male adult life expectancy of 9.7 days. Orius sauteri lay 5–8 eggs/day at its peak oviposition, then 1–4 eggs/day till it died. Under the experimental circumstances, Orius sauteri produced 51.1 offspring per individual, 24.6 days, 0.16 days/year, and 1.2 days/year in terms of net reproductive rate (R0), mean generation time (T ), intrinsic rate of increase (r), and finite rate of growth, respectively (Liu et al. 2018). Later, another findings demonstrated that Oruis albidipennis females and males fed on Aphis craccivora nymphs lived longer than those fed on Ephestia kuehniella eggs, which lasted 20.5 and 7.1 days and did not differ significantly from one another. When it consumed Planococcus citri nymphs, Oruis albidipennis male and female longevity terns lived only 16.6 and 5.0 days, respectively. Females fed Ephestia kuehniella eggs had the maximum lifetime fecundity, 112.7–1.45 eggs/female, with no discernible difference from those fed Aphis craccivora nymphs, 96.04–1.09. Fecundity dramatically decreased to 55.422.18 eggs/female when females consumed Planococcus citri nymphs (Amer et al. 2021).

7.6.5

Lygaeidae

A genus of insects in the Geocoridae family with the common name “large eyed bugs” is Geocoris. These types of insects are frequently referred to as large eyed bugs because of their recognisable enormous, prominent, widely spaced eyes on both sides of their heads. Geocoris species are regarded as among the most significant predators in cotton, maize, alfalfa, soybean, straw berry, pea nut, and many other crops. They are also known as “big-eyed bugs”. In both agroecosystems and the natural world, geocoris species are significant predators. Sunflower seeds, eggs, and nymphs of the great milkweed bug, Oncopeltus fasciatus, were used to effectively rear the western big-eyed bug, Geocoris pollens (Yokoyama 1980). Hagler (2004) asserts that big-eyed bugs spend the winter as adults. The life cycle of any insect is influenced by temperature. Single-egg lays on leaves or stems take about a week to hatch. Five nymphal instars, each lasting 4–6 days, are present in big-eyed bugs. Nymphs and adults are both predatory. A female can produce up to 300 eggs during her lifetime as an adult, and adults typically live for around one month. During a crop growth season, multiple generations may take place. Here, we go into further detail on four different species of Geocoris, including Geocoris punctipes, Geocoris floridanu, Geocoris superbus, and Geocoris erythrocephalus. Geocoris punctipes The Geocoris punctipes, a beetle with large eyes, is a common predator in a number of important cropping systems in the United States. Geocoris punctipes, a predator that preys on a variety of pests in cotton, soybeans, and other crops, is energetic and aggressive in both its nymphal and adult stages. Despite considerable relationships between photoperiod and population, photoperiod significantly influenced the length of embryonic development but not population. Since the variance failed to exhibit any discernible patterns in connection to photoperiod,

208

7

Predatory Insects: Adults Biology of Various Orders

we believe it is more likely to be an artefact of the experiment than a characteristic of the phenology of the bugs. The photoperiod and population had a substantial impact on how long the pre-oviposition periods lasted during the 20-day observation period following adult emergence. The photoperiod and population interacted significantly as well. For both groups, the frequency of female oviposition varied with photoperiod. Additionally, female dissections conducted after the experiment’s 20th day found that nearly all of the females who had not ovulated by that day had significantly smaller ovaries, consistent with the diapause syndrome. The diversity of the pre-oviposition periods decreased and the mean duration also decreased with longer day lengths. Longer days encouraged earlier and more synchronised oviposition as a result. This variation denotes a graded photoperiodic response, which may be depicted as varied intensities of diapause. The pre-oviposition period for the Georgia population was shorter than that for the Kentucky group in all instances where there were significant differences. For the Georgia bugs as opposed to the Kentucky bugs, photoperiod had a more significant impact on the pre-oviposition period (Ruberson et al. 2001). Geocoris floridanu Geocoris floridanu females were given Heliothis zea, and they successfully reproduced. One female was lost on her ninth day of adulthood, thus she was excluded from the analysis. 19 female insects were utilised in the analysis of adult reproduction and predation as well as the calculation of the fertility life table parameters for insects fed maize earworm. Four females were excluded from the research because they produced fewer than 15 eggs, and two of those four females also produced nonviable eggs. Females feeding Spodoptera exigua larvae produced a variety of outcomes. Ultimately, 16 of the 20 females who had been fed beet armyworm larvae were used. The first egg batch was laid by predator females given corn earworm, beet armyworm, and beet armyworm alone, respectively, 5, 5.5, and 11 days after adult emergence. When compared to females fed only beet armyworm, females of corn earworm and beet armyworm often began oviposition earlier. Although there was a wide range in the viability of the eggs (between 23% and 100%), there was no difference in incubation time or viability across treatments. The prey items significantly affected the fertility of Geocoris floridanu. The female corn earworms produced the most eggs, followed by beet armyworm-corn earworm and beet armyworm alone. For both daily egg production and the reproductive season, treatments were given in a similar order. Female beet armyworm-corn earworms and female beet armyworms laid the next-highest number of eggs each day, respectively. Female corn earworms continued to lay eggs for a longer period of time, just like beet armyworm-corn earworm and beet armyworm females did. Female offspring (mx) production peaked in all treatments between the second and third weeks, but it was numerically larger in females fed corn earworm (2.39 females [mx]/day) and beet armyworm-corn earworm (3.08 females [mx]/day) than in females fed simply beet armyworm (1.1 females [mx]/day) (Torres et al. 2004). Geocoris superbus In addition to Sitotroga cerealella eggs, the adults of Geocoris superbus were fed Helicoverpa armigera eggs or Phenacoccus solenopsis nymphs.

7.7 Lacewings (Neuropteran)

209

The male and female adult Geocoris superbus persisted for 24.8 and 30 days, respectively. The Geocoris superbus had a mean fecundity of 29.4 eggs/female. Egg laying was seen beginning the day following mating and continued until the female Geocoris superbus female died after a pre-oviposition phase (Varshney and Ballal 2017). Geocoris erythrocephalus A significant predatory insect in cotton fields, the big-eyed bug Geocoris pallidipennis can feed on a wide range of pests, including aphids, leafhoppers, cotton bollworm Helicoverpa armigera, and Lepidopteran larvae. Adult Geocoris erythrocephalus are 3/16 inch or less long and possess forewings and hindwings. The forewings are rigid at the base and membranous at the tip. The forewings develop during rest a triangle-shaped pattern behind the pronotum (also known as the shoulders), pointing backward. Adult insects display sexual dimorphism. Males and females are larger than one another. In between the eyes, the front part of the head appears golden yellow in females but brown in males. Females live longer than males after adult formation. The average lifespan for males and females was 10–13 (mean = 11.5) and 7–8 (mean = 7.4), respectively. The ratio of men to women was found to be 1:1.29. Geocoris erythrocephalus completed its life cycle in 38–44 (mean = 40.60) days (Rajan and Sathish 2018). Geocoris erythrocephalus deposits its eggs singly on cottonwick or fabric as an adult (Richa and Ballal 2017). Females begin egg-laying between 2 and 3 days (on average, 2.30–0.48 days). On the cotton wick that is given within the beaker, the adult Geocoris erythrocephalus female deposits her eggs. It suggested that adult females prefer hairiness when laying eggs. In her 3- to 5-day ovipositional cycle, a single female produces 10–21 (14.00) eggs (Rajan and Sathish 2018).

7.7

Lacewings (Neuropteran)

Many horticultural and agricultural cropping systems, including ornamentals, vegetables, fruits, nuts, fibre and feed crops, greenhouse crops, and woodlands, have long been regarded essential habitats for Chrysoperla spp. They are among the most popular and readily available natural enemies globally. Researchers examined the biology and functional responses of the green lacewing, Mallada boninensis, in relation to prey density (Neuroptera: Chrysopidae). Berothid biology is not well understood. A virtually global family of Neuropterans with many different species is known as the Hemerobiidae, or brown lacewings. Due in parts to their enigmatic nocturnal or crepuscular behaviour, adult behaviours are poorly understood. Their typical habitats include the branches and foliage of trees and bushes, but other species—like numerous Micromus spp.—prefer open, grassy areas. The biology of Oliarces clara and moth lacewings were described in the literature by Faulkner (1990). Plesiochrysa ramburi’s initial stock culture was cultivated in a lab after being taken from a field of cassava. Mealybug larvae were transferred to mature, 20–25day-old pumpkin fruit carrying mealybugs, which were then put in the spherical

210

7

Predatory Insects: Adults Biology of Various Orders

plastic boxes and covered with nylon cloth. The Plesiochrysa ramburi adults were fed a mixture of honey and yeast at a 1:1 ratio. Eggs were typically placed by adults on the nylon fabric or the inner surfaces of the boxes. According to Plesiochrysa ramburi’s life table parameters, feeding on Pseudococcus jackbeardsleyi resulted in the highest net reproduction number (19.19) and gross reproductive rate (46.01, females/female/generation). On Pseudococcus jackbeardsleyi, the maximum innate capacity for increase (rc), intrinsic rate of natural growth (rm), and finite rate of increase (l) were, respectively, 0.0360, 0.0601, and 1.0367 (females/female/day). Pseudococcus jackbeardsleyi had a cohort generation time (Tc) of 35.6418 (days). On Pseudococcus jackbeardsleyi, the shortest mean generation time (T) and doubling time (DT) were 21.3351 days and 5.0050 days, respectively. According to this research, Pseudococcus jackbeardsleyi is the best food for mass-raising Plesiochrysa ramburi to enable releases in IPM (Sattayawong et al. 2016). With a R0 of 6.6669, which was lower than the current findings, Kantha et al. (2006), who recorded the parameters of Plesiochrysa ramburi on several prey species, demonstrated that Maconellicoccus hirsutus was the most suited prey for Plesiochrysa ramburi. The values for rc and l were higher than the current results at 0.0356 and 1.0550, respectively. The Tc (35.3674) was comparable to the current outcomes, nevertheless. One of the key factors used to select the best biological control agent is the intrinsic rate of natural increase (rm), as a high rm implies a high rate of multiplication per day of the natural enemy Mallada desjardinsi (Chrysopidae: Neuroptera) (Vasanthakumar and Babu 2013). The current study’s rm value demonstrated that Plesiochrysa ramburi can proliferate more when fed on Pseudococcus jackbeardsleyi, which was compatible with the species’ highest GRR and quickest doubling times. Dichochrysa members play a significant role as biocontrol agents, whether it is through the preservation of either natural populations or upcoming flood and augmentation releases. Earlier records of Dichochrysa tacta were from Saudi Arabia and South Africa. Six Dichochrysa species have had their biology researched, including Dichochrysa alcestes, Dichochrysa formosana, Dichochrysa ussuriensis, Dichochrysa flavifrons, Dichochrysa zelleri, and Dichochrysa prasina (Pappas et al. 2011), Dichochrysa alcestes, Dichochrysa formosana, and Dichochrysa ussuriensis (Haruyama et al. 2012). Aphis fabae, Aphis punicae, Macrosiphum rosae, and Ephestia kuehniella are significant agricultural pests that seriously harm horticulture plants. Females deposited eggs for 2.56–2.72 days on average, with no discernible variation across different prey species. The percentage of eggs that hatched ranged from roughly 79.1% to 85.2% and was not significantly impacted by the various prey studied. The prey species that the paternal generation was reared on had no significant impact on the progeny’s sex ratio, which ranged from 0.47 to 0.53. The findings demonstrated a substantial relationship between prey species and pre-imaginal development periods, survival, adult longevity, and fertility. In contrast, the prey species had no appreciable impact on the progeny sex ratio, egg hatch, or egg longevity. The most suited prey among the studied species was Ephestia kuehniella eggs, which produced Dichochrysa tacta with the highest fecundity, highest intrinsic rate of increase (rm), highest net reproduction rate (R0), and maximum survival

7.8 Miscellaneous Predators (Orthopera, Lepidotpera, Others)

211

Table 7.5 Chrysoperla zastrowi sillemi on egg hatch, the percentage of larvae that pupated, the incubation period (IP), the larval period (LP), the pupal length (PD), the total developmental period (TDP) (days), and the fecundity (number of eggs/female) on diverse hosts Unsterilised Corcyra eggs Sterilised Corcyra eggs Brevicoryne brassicae Aphis gossypii Aphis craccivora Uroleucon compositae

IP 3.9 4.2 4.1 4.0 3.8 4.4

EH 94.8 91.1 84.2 86.4 89.9 78.8

LP 9.0 9.5 10.5 10.6 10.1 11.2

LP 89.1 86.5 79.4 81.7 78.9 74.5

PP 5.4 6.3 7.5 6.9 7.3 8.1

TDP 18.30 20.0 22.1 21.54 21.21 23.72

FC 371.6 338.8 153.8 281.4 262.2 113.4

rates. Nymphs of Aphis punicae and Macrosiphum rosae were both extremely advantageous prey, however nymphs of Aphis fabae were less adapted and had poorer development, survival, adult longevity, fecundity, intrinsic rate of increase (rm), and reproduction rate (R0). Under controlled laboratory settings, newly emerging Mallada boninensis larvae were fed 20, 30, 40, 50, 60, 70, 80, 90, and 100 fresh Corcyra cephalonica (Lepidoptera: Gelechiidae) eggs. It was found that Mallada boninensis positive consumption rate, development, and fecundity were all significantly influenced by the prey density. It was generally discovered that maximum consumption with the shortest developmental phase, maximum fecundity, and longest extended adult longevity were observed as prey density increased from 20 to 100. When the prey density was increased from 20 to 100, the predatory potential increased significantly. During the first two larval instars, Mallada boninensis’ daily predation rate grew gradually before reaching its maximum. The findings suggested that depending on food density and having a variance in daily consumption, Mallada boninensis feeding potential and developmental period can vary from 6.0 to 11.3 days. The 100 Corcyra egg treatments consumed the most eggs per day (87.88), followed by the 90 egg treatments (79.3 eggs), and the 80 egg treatments (69.8 eggs) (Elango and Sridharan 2022) (Table 7.5). Under varying temperature conditions in the greenhouse, Chrysopa pallens development time and adult longevity did not change significantly between Corcyra cephalonica eggs and Megoura japonica treatments. Chrysopa pallens was cultured on aphid food (Megoura japonica), and I, II, III instar larvae and pupae had respective lifespans of 3.1, 2.7, 3.0, 4.3, and 13.0 days. Male Chrysopa pallens showed a shorter lifespan (7.7 days) when cultured on aphid meal, whereas females lived longer (17.7 days) when fed rice moth eggs (Wang et al. 2021).

7.8

Miscellaneous Predators (Orthopera, Lepidotpera, Others)

Male adult female mimics of Franklinothrips vespiformis are widespread and range in body length from 2.5 to 3.0 mm. The forewing of a female has a rounded apex and is slender and fully winged. The abdomen is anteriorly constricted, and the body is

212

7

Predatory Insects: Adults Biology of Various Orders

black with white stripes on the second and third segments. At segments five or six, the abdomen is at its widest. The antennae, legs, and body are all brown. The abdominal segments II and III and the antennal segments I through III, however, are yellow. Additionally, the distal end of the femur is frequently yellowish, while the anterior borders are brown. Brown legs with a yellowish femur at the distal end. With three paler regions at the base, middle, and subapex, the forewing is brown overall. Pupa: Franklinothrips vespiformis pupae are found inside a white silk cocoon made by the larva beneath the leaves (2e). Pre-pupal, pupal stage 1 and pupal stage 2 are the three phases of the pupa, which is red in colour. Well-developed wing buds are shorter in the pre-pupal stage (show non-obvious movement, prepared for cocoon construction). Only in pre-pupa is the pupal skin of the appendages segmented. Pupa 1: The antennal sheaths reach the abdomen but not the metathorax (pupa 2). Additionally, while both the anterior and posterior wing buds reach abdominal segment V, the posterior wing buds only reach abdominal segment III (pupa 1), (pupa 2). Legs and the hind tibiotarsus are longer than the pterothorax (pupa 1) and shorter than the pterothorax (pupa 1), (pupa 2). Female Adult Female The common myrmici species Franklinothrips vespiformis has a body length of 2.5–3.0 mm. The forewing of a female has a rounded apex and is slender and fully winged. The abdomen is anteriorly constricted, and the body is black with white stripes on the second and third segments. At segments five or six, the abdomen is at its widest. The antennae, legs, and body are all brown. The abdominal segments II and III and the antennal segments I through III, however, are yellow. Additionally, the distal end of the femur is frequently yellowish, while the anterior borders are brown. Brown legs with a yellowish femur at the distal end. Brown forewings with three paler regions on the base, middle, and tip. Franklinothrips vespiformis adult males are uncommon; they are similar in colour to females but smaller and less ant-like in appearance. Males typically have paler wings, longer, darker antennae, and a less narrow waist. Nearly as long as the head, the second and third antennal segments have a lengthy sensory metanotum made of erratic scallops. The eyes are ventrally extended, the skull is wider than long, and the posterior ocelli are bigger than the anterior. The metanotum lacks medial sculpture, the prothorax is thinner toward the base, and the legs are long and slender. Sternites III–VIII has two pairs of posteromarginal setae and one pair of discal setae in a line, and abdominal sternite II has two pairs of discal setae. Franklinothrips vespiformis is often a unisexual species, with both male and females present. Males have not been discovered in populations from Japan, and they seem to be uncommon elsewhere. Franklinothrips vespiformis and other thrips species have been shown to undergo parthenogenesis that was mediated by Wolbachia. Franklinothrips vespiformis males appeared to be produced in response to heat and tetracycline treatments. Although males produced motile sperm that was passed on through the spermatheca, mating had minimal effect on the sex ratios of the offspring. These findings suggest that sperm do not fertilise eggs. Parthenogenesis is widespread in the introduced thrips, probably dispersing more quickly than sexual forms. Franklinothrips vespiformis reproduces through parthenogenesis,

7.9 GM Crops or Bt Proteins on the Biology of Predators

213

laying one egg into a stem, vein, or other soft plant tissue with the help of its serrated ovipositor. This behaviour shows that the species produces viable eggs in this way. Females have an egg-laying capacity of three in an hour and 150–200 throughout their lifetime. Additionally, the females deposit a drop of a yellowish protective fluid on the exposed tips of the eggs, making it difficult to find the eggs. When fed with nymphs and adult of Nilaparvata lugens, adults of Sogatella furcifera, Nephottetix virescens, Scirpophaga incertulas, and Chilo suppressalis, Marasmia patnalis, Hydrellia philippina, and Leptocorisa oratorius, Conocephalus longipennis (Orthoptera: Tettigoniidae) completed its life cycle in 142–182 days (Rubia et al. 1990). One of the frequent generalist predators in rice fields is Metioche vittaticollis. Metioche vittaticollis had an average life span of 40–61 days at 26–28 °C. The eggs took 14.28 days to hatch after being placed individually into the leaf sheath of either rice or weeds. Four nymphal stades were traversed throughout the nymphal stage, which lasted between 27 and 45 days. A female would typically lay 50 eggs in her lifetime. Adult female or male Metioche vittaticollis lived 20–38 days on average. Male and female average lifespans were 29.24 days and 25.00 days, respectively. Males and females that were not mated lived longer than those who were. While there was lower mortality in the third and fourth instar nymphs, the egg and first instar nymph both experienced significant mortality rates of 30% and 25%, respectively. The Metioche vittaticollis adult females lived for 32 days, having a high survival rate among young adults but a declining rate as the cricket became older. The Brown Plant Hopper (BHP) nymph preferred the Metioche vittaticollis females over the males. The stage of prey that Metioche vittaticollis fed on the most was the early nymph stage. When they were offered late instar BPH nymph prey, however, the predatory behaviour decreased. Both male and female crickets ingested less BPH in its maturity stage (Karindah et al. 2017).

7.9

GM Crops or Bt Proteins on the Biology of Predators

By managing their immune systems, detoxication, and antioxidant activity, insects can protect themselves against infections, hazardous substances, and plant-specific metabolites. Glutathione S-transferase (GST), esterase, and cytochrome P450 enzymes (P450s) are involved in the metabolism, detoxification, and detoxication of pesticides (EST). The only enzymes that can mediate the metabolic digestion of all types of pesticides are the P450 ones. Numerous previous studies have examined the ability of Bt toxins to bind to non-target insect BBMV proteins, including Chrysoperla carnea (Rodrigo-Simón et al. 2006). Consumption of Trichoplusia ni on Cry1Ac/Cry2Ab transgenic cotton had no discernible impact on Coleomegilla maculata mortality, growth, development, or fertility (Li et al. 2011). Numerous researches have been conducted up to this time on the effect of GM crops or Bt protein on predator insects. The GM cotton expressing Bt (Cry1Ac) and CpTI (cowpea trypsin inhibitor) proteins did not significantly affect the growth and development of the larvae of the ladybeetles

214

7

Predatory Insects: Adults Biology of Various Orders

Propylea japonica (Shu et al. 2019) and Chrysoperla sinica, as well as the fecundity of the adults, when compared to the control treatment (Liu et al. 2021). The Japanese ladybeetle, Propylaea japonica, was not affected by treatment with transgenic Cry1Ac/Cry2Ab cotton (Zhao et al. 2013). In another study, authors investigated the effects of the GM cotton variety A26-5, which generates Cry1Ac, on the predatory insect Geocoris pallidipennis. Cry1Ac was found in Helicoverpa armigera larvae fed on GM cotton at 36.6 g and 3.4 ng/g fresh weight; these larvae may have passed Cry1Ac to Geocoris pallidipennis (Zhang et al. 2022). Hendriksma et al. (2013) and Jurat-Fuentes and Crickmore (2017) claim that the Cry toxin in Bt cotton can react with the intestinal epithelial tissue of an insect and kill it. As a result, Zhang et al. (2022) assessed the degree of Cry1Ac expression in both GM cotton and the level that Helicoverpa armigera larvae transported to the stomach of Geocoris pallidipennis. Cry1Ac was discovered in Geocoris pallidipennis nymphs and adults that had been raised with Helicoverpa armigera larvae fed on GM cotton at concentrations of 4.8 ng/g and 5.6 ng/g, respectively. Both the Helicoverpa armigera and the Helicoverpa larvae that Geocoris pallidipennis fed did not contain detectable levels of Cry1Ac. After 40 days, both groups of Geocoris pallidipennis had survival rates of more than 75% (82% in the non-GM group and 78% in the GM group). After feeding on Helicoverpa armigera that consumed control cotton leaves and those that consumed GM cotton leaves, Geocoris pallidipennis adults had an average pre-oviposition period of 8.2 days and 8.4 days, respectively. The female Geocoris pallidipennis produced 1–4 eggs on average per day after mating, with a maximum of 50 laid in 20 days.

7.10

Future Area of Research

• Since there are few publications on mantodean fecundity and life table parameters, we advise focusing on these topics. • Future research can focus on the biology of small predatory insects from the Orthoptera, Lepidoptera, and other predators. • Researchers should investigate the effects of diverse natural, laboratory, and synthetic host foods on the fertility and hatchability of predatory insects. Various funding organisations should prioritise studies of the biology of regionally specific predators and/or globally specific adult predators. • It is crucial to establish the constraints for producing adult predators of different insect orders. • The biology of adult predatory insects should be studied in relation to how genetically modified crops affect those crops.

References

215

References Abbas S, Pérez-Hedo M, Colazza S, Urbaneja A (2014) The predatory mirid Dicyphus maroccanus as a new potential biological control agent in tomato crops. BioControl 59(5):565–574 Abraham VA, Kurian C (1973) Chelisoches moris F. (Forficulidae: Dermaptera), a predator on eggs and early instar grubs of the red palm weevil Rhynchophorus ferrugineus F. (Curculionidae: Coleoptera). JPC 1:147–152 Ahmad SH, Kamarudin N (2016) Growth and longevity of the insect predator, Sycanus dichotomus Stal (Hemiptera: Reduviidae) fed on live insect larvae. J Oil Palm Res 28:471–478 Akram W, Akbar S, Mehmood A (1996) Studies on the biology and predatory efficacy of Coccinella septempunctata with special reference to cabbage. Pak Entomol 18:104–106 Almohamad R, Verheggen FJ, Haubrug E (2009) Optimality theory as applied to oviposition predicts that female choice searching and oviposition behavior of Aphidophagous hoverflies (Diptera: Syrphidae): a review. Biotechnol Agron Soc Environ 13(3):467–481 Ambrose DP, Sebasti Rajan XJ, Nagarajan K, Jeba Singh V, Krishnan SS (2009) Biology, behaviour and functional response of Sphedanolestes variabilis Distant (Insecta: Hemiptera: Reduviidae: Harpactorinae), a potential predator of lepidopteran pests. Entomol Croatica 13(2): 33–44 Amer ME, Abdel-Razak SI, El-Sobky HF (2021) Influences of some insect pests as prays on biology and consumption rate of predator, Oruis albidipennis (Reuter) (Hemiptera, Anthocoridae) under laboratory conditions. JPPP 12(1):37–42 Arshad M, Ullah MI, Khan RR, Anjum S, Tahir M, Shamshad A, Rizwan M (2021) Demographic parameters of the reduviid predator, Rhynocoris marginatus (Reduviidae: Hemiptera) fed on two lepidopterous insect pests. BioControl 66(2):227–235 Ballal CR, Singh SP, Poorani J, Gupta T (2003) Biology and rearing requirements of an anthocorid predator, Blaptostethus pallescens poppius (Heteroptera: Anthocoridae). J Biol Control 17:29– 33 Ballal CR, Gupta T, Joshi S (2013) Production and evaluation of the warehouse pirate bug Xylocoris flavipes (Reuter) (Hemiptera: Anthocoridae). In: Proceedings of the International congress on insect science 2013, 14th to 17th February, 2013 at GKVK, Bangalore Barry KL, Wilder SM (2013) Macronutrient intake affects reproduction of a predatory insect. Oikos 122(7):1058–1064 Belliure B, Michaud JP (2001) Biology and behavior of Pseudodorus clavatus (Diptera: Syrphidae), an important predator of citrus aphids. Ann Entomol Soc Am 94(1):91–96 Boopathi T, Singh SB, Dutta SK, Dayal V, Singh AR, Chowdhury S, Lalhruaipuii K (2020) Biology, predatory potential, life table, and field evaluation of Propylea dissecta (Coleoptera: Coccinellidae), against Lipaphis erysimi (Hemiptera: Aphididae) on broccoli. Econ Entomol 113(1):88–97 Bugg RL, Colfer RG, Chaney WE, Smith HA, Cannon J (2008) Flower flies (Syrphidae) and other biological control agents for aphids in vegetable crops. https://doi.org/10.3733/ucanr.8285 Butnariu AR, Pasini A, Reis FS, Bessa E (2013) Maternal care by the earwig Doru lineare Eschs. (Dermaptera: Forficulidae). J Insect Behav 26(5):667–678 Calixto AM, Bueno VHP, Montes FC, Silva AC, Van Lenteren JC (2013) Effect of different diets on reproduction, longevity, and predation capacity of Orius insidiosus (Say) (Hemiptera: Anthocoridae). Biocontrol Sci Tech 23(11):1245–1255 Dib H, Sauphanor B, Capowiez Y (2017) Report on the life history traits of the generalist predator Forficula auricularia (Dermaptera: Forficulidae) in organic apple orchards in southeastern France. Can Entomol 149(1):56–72 Du L, Ge F, Zhu S, Parajulee MN (2004) Effect of cotton cultivar on development and reproduction of Aphis gossypii (Homoptera: Aphididae) and its predator Propylaea japonica (Coleoptera: Coccinellidae). J Econ Entomol 97(4):1278–1283 Elango K, Sridharan S (2022) Effect of prey density on biology and foraging potential of Mallada boninensis (Okamato) (Neuroptera: Chrysopidae). Madras Agric J 108:10–12

216

7

Predatory Insects: Adults Biology of Various Orders

Faheem M, Saeed S, Sajjad A, Razaq M, Ahmad F (2019) Biological parameters of two syrphid fly species Ischiodon scutellaris (Fabricius) and Episyrphus balteatus (DeGeer) and their predatory potential on wheat aphid Schizaphis graminum (Rondani) at different temperatures. Egypt J Biol Pest Control 29(1):1–8 Farhadi R, Allahyari H, Chi H (2011) Life table and predation capacity of Hippodamia variegata (Coleoptera: Coccinellidae) feeding on Aphis fabae (Hemiptera: Aphididae). Biol Control 59(2):83–89 Fathipour Y, Jafari A (2008) Biology of Nabis capsiformis (Het., Nabidae) preying upon Creontiades pallidus (Het., Miridae) in laboratory conditions. JWSS Isfahan Univ Technol 12(43):157–166 Faulkner DK (1990) Current knowledge of the biology of the moth-lacewing Oliarces clara Banks (Insecta: Neuroptera: Ithonidae). In: Mansell MW, Aspock H (eds) Advances in neuropterology. South African Department of Agricultural Development, Pretoria, pp 197–203 George PE, Seenivasagan R, Karuppasamy G (1998) Life table and intrinsic rate of increase of Sycanus collaris Fabricius (Heteroptera: Reduviidae), a Predator of Spodoptera litura (Fabricius) (Lepidoptera: Noctuidae). J Biol Control 12(2):107–111 Ghanim AA, El-Serafy HA, Hassan HA, Mohamed NE, Abu El Matti FB (2021) Biological aspects of two coccinellid predatory insects reared on artificial diets and natural preys under constant temperature. J Plant Prot Pathol 12(1):19–22 Golizadeh A, Jafari-Behi V (2012) Biological traits and life table parameters of variegated lady beetle, Hippodamia variegata (Coleoptera: Coccinellidae) on three aphid species. Appl Entomol Zool 47:199–205 Greyvenstein B, Du Plessis H, Moulin N, Van den Berg J (2020) Distribution of Galepsus spp. in Southern Africa and life history of Galepsus lenticularis (Mantodea: Tarachodidae). Insects 11(2):119 Greyvenstein B, du Plessis H, Van den Berg J (2021) Life history of the false flower mantid (Harpagomantis tricolor Linnaeus, 1758) (Mantodea: Galinthiadidae) and its distribution in southern Africa. J Orthop Res 30(1):17 Hagler J (2004) Geocoris spp. (Heteroptera: Lygaeidae), bigeyed bug. Biological control: a guide to natural enemies in North America, Cornell University. http://nysaes.cornell.edu/ent/biocontrol/ predators/geocoris.html Haruyama N, Miyazaki Y, Nakahira K, Mochizuki A, Nomura M (2012) Developmental time and survival of trash-carrying versus naked green lacewings, with implications for their utility as augmentative biological control agents. Ann Entomol Soc Am 105(6):846–851 Hendriksma HP, Küting SM, Härtel A, Näther AB, Dohrmann I, Dewenter S, Tebbe CC (2013) Effect of stacked insecticidal Cry proteins from maize pollen on nurse bees (Apis mellifera carnica) and their gut bacteria. PLoS One 8:e59589 Hormchan P, Schuster MF, Hepner LW (1976) Biology of Tropiconabis capsiformis. Ann Entomol Soc Am 69(6):1016–1018 Islam MA, Haque ME (1978) The effect of different food on the longevity of Menochilus sexmaculatus (Fabricius) (Coleoptera: Coccinellidae). BJSIR 13(1-4):104–106 Joshi S, Ballal CR, Rao NS (1999) Biotic potential of three coccinellid predators on six different aphid hosts. J Entomol Res 23:1–7 Joshi S, Lokeshwari D, Krishna Kumar NK, Verghese A (2013) A pictorial guide for the identification of some major aphid species. Technical folder No. 1. National Bureau of Agriculturally Important Insects, Bangalore Jurat-Fuentes JL, Crickmore N (2017) Specificity determinants for cry insecticidal proteins: insights from their mode of action. J Invertebr Pathol 142:5–10 Kantha D, Uraichuen S, Suasaard W (2006) Biology and utilization of Plesiochrysa ramburi (Schneider) (Neuroptera: Chrysopidae) for the control of Maconellicoccus hirsutus (Green) (Homoptera: Pseudococcidae). In: KU-KPS conference, Nakhon Pathom (Thailand), 6–7 Dec 2006

References

217

Karaca I, Efe D (2021) Life table of Nabis pseudoferus Rem. (Hemiptera: Nabidae) feeding on Spodoptera littoralis Boisd. (Lepidoptera: Noctuidae) under laboratory conditions Karindah S, Yanuwiadi B, Sulistyowati L (2017) Biology and predatory behavior of Metioche vittaticollis (Stal) (Orthoptera: Gryllidae). JTPP 1(1):1–9 Khursheed R, Hussain B, Ahmad SB, Ashraf M (2006) Biology and feeding potential of Coccinella septempunctata on mustard aphid, Lipaphis erysimi. Int J Zool Res 2(1):30–33 Kyneb A, Toft S (2004) Quality of two aphid species (Rhopalosiphum padi and Sitobion avenae) as food for the generalist predator Tachyporus hypnorum (Col., Staphylinidae). J Appl Entomol 128:658–663 Li Y, Romeis J, Wang P, Peng Y, Shelton AM (2011) A comprehensive assessment of the effects of Bt cotton on Coleomegilla maculata demonstrates no detrimental effects by Cry1Ac and Cry2Ab. PLoS One 6:e22185 Lillo I, Perez Banoon C, Rojo S (2021) Life cycle, population parameters, and predation rate of the hover fly Eupeodes corollae fed on the aphid Myzus persicae. Entomol Exp Appl 169(11): 1027–1038 Liu H, Yu C, Tadauchi O, Cai W (2012) Taxonomic and bionomic notes on Sycanus sichuanensis Hsiao (Hemiptera: Reduviidae: Harpactorinae). Zootaxa 2644(13):47–56 Liu P, Jia W, Zheng X, Zhang L, Sangbaramou R, Tan S, Shi W (2018) Predation functional response and life table parameters of Orius sauteri (Hemiptera: Anthocoridae) feeding on Megalurothrips usitatus (Thysanoptera: Thripidae). Fla Entomol 101(2):254–259 Liu F, Luo J, Zhu X, Zhao C, Niu L, Cui J (2021) Transgenic Cry1Ac/CpTI cotton assessment finds no detrimental effects on the insect predator Chrysoperla sinica. Ecotoxicol Environ Saf 208: 111680 Loko YLE, Toffa J, Gavoedo DM, Kitherian S, Orobiyi A, Tamo M (2022) Effect of population density on oviposition, development, and survival of Alloeocranum biannulipes (Hemiptera: Reduviidae) preying on Dinoderus porcellus (Coleoptera: Bostrichidae). J Basic Appl Zool 83(1):1–8 Madadi H, Ahmadi Z, Heshmati A (2016) Biology and life table parameters of predatory bugs Nabis pseduferus (Hemiptera: Nabidae). In: Proceedings of 22 college of agriculture and natural resources. University of Tehran, Karaj, p 584 Mari JM, Nizamani SM, Lohar MK, Khuhro RD (2004) Biology of Menochilus sexmaculatus Fab. and Coccinella undecimpunctata L. (Coccinellidae: Coleoptera) on Alfalfa Aphid Therioaphis trifolii Monell. J Asia-Pac Entomol 7(3):297–301 Marquereau L, Cottineau JS, Fontaine O, Chiroleu F, Reynaud B, Delatte H (2022) Life history parameters and predation capacities of Nesidiocoris volucer: a new biological control agent for tomato crop. Bull Entomol Res 2022:1–9. https://doi.org/10.1017/S0007485321001164 Mojib-Haghghadam Z, Sendi JJ, Zibaee A, Mohaghagh J (2018) Suitability of Aphis gossypii Glover, Aphis fabae Scop. and Ephestia kuehniella Zeller eggs for the biology and life-table parameters of Adalia decempunctata (L.) (Coleoptera: Coccinellidae). Arch Biol Sci 70(4): 737–747 Nitin KS, Bhat PS, Raviprasad TN, Vanitha K (2017) Biology, behaviour and predatory efficiency of Sycanus galbanus Distant. Hemiptera: Reduviidae: Harpactorinae recorded in Cashew plantations. J Entomol Zool Stud 5:524–530 Noble RRP, Harvey SG, Sams CE (2002) Toxicity of Indian mustard and allylisothiocyanate to masked chafer beetle larvae. Plant Health Progr. https://doi.org/10.1094/PHP-2002-0610-01-RS Nunes GDS, Dantas TAV, Figueiredo WRS, de Oliveira R, Batista JDL, Vacari AM, Bortoli DE (2022) The life history of Euborellia annulipes (Lucas) (Dermaptera: Anisolabididae) fed on larvae and pupae of Plutella xylostella (L.) (Lepidoptera: Plutellidae). Turk J Zool 46(2): 175–185 Obrycki JJ, Orr CJ (1990) Suitability of three prey species for nearctic populations of Coccinella septempunctata, Hippodamia variegata, and Propylea quatuordecimpunctata (Coleoptera: Coccinellidae). J Econ Entomol 83(4):1292–1297

218

7

Predatory Insects: Adults Biology of Various Orders

Okamoto H (1966) Three problems of prey specificity of Aphidophagous coccinellids. In: Hodek I (ed) Ecology of Aphidophagous insects. Academia, The Hague, pp 45–46 Pandi G, Paul B, Vivek S, Shankarganesh K (2012) Feeding potential and biology of Coccinellid predator Cheilomenes sexmaculata (Fabricius) (Coleoptera) on aphid hosts. Indian J Entomol 7(4):388–393 Pappas ML, Broufas GD, Tsarsitalidou OK, Koveos DS (2011) Development and reproduction of the lacewings Dichochrysa flavifrons and Dichochrysa zelleri (Neuroptera: Chrysopidae) fed on two prey species. Ann Entomol Soc Am 104(4):726–732 Patel KG, Vyas HN (1984) Seasonal abundance of green lacewing, Chrysopa (= Chrysoperla) scelestes banks on cotton, green gram and groundnut at Junagadh (Gujarat). Agric Sci Dig 4(1):7–8 Perkins PV, Watson TF (1972) Biology of Nabis alternatus (Hemiptera: Nabidae). Ann Entomol Soc Am 65(1):54–57 Perumalsamy K, Selvasundaram R, Roobakkumar A, Rahman VJ, Babu A, Muraleedharan NN (2009) Life table and predation of Oligota pygmaea (Coleoptera: Staphylinidae) a major predator of the red spider mite, Oligonychus coffeae (Acarina: Tetranychidae) infesting tea. Biol Control 51(1):96–101 Rajan SL, Sathish R (2018) Biology of big-eyed bug, Geocoris erythrocephalus (Lepeletier & Serville) on cabbage aphid, Brevicoryne brassicae (L.). IJCMAS 7(7):3301–3305 Rajan S, Jesu Suneetha N, Sathish R (2017) Biology and predatory behaviour of an assassin bug, Sycanus collaris (Fabricius) on rice meal moth, Corcyra cephalonica (Stainton) and leaf armyworm, Spodoptera litura (Fabricius). Agric Update 12:1181–1186 Rajmohan N, Jayaraj S (1974) Growth and development of the coccinellid, Menochilus sexmaculatus (Fabricius) on four species of aphids. Madras Agric J 61:118–122 Richa V, Ballal CR (2017) Biology and feeding potential of Geocorissuperbus Montandon (Heteroptera: Geocoridae), a predator of mealybug. J Entomol Zool Stud 5(3):520–524 Rodrigo-Simón AA, De Maagd RA, Avilla C, Bakker PL, Molthoff J, González-Zamora JE, Ferré J (2006) Lack of detrimental effects of Bacillus thuringiensis Cry toxins on the insect predator Chrysoperla carnea: a toxicological, histopathological, and biochemical analysis. Appl Environ Microbiol 72:1595–1603 Roy SP, Sinha DK (2002) Predatory efficiency of Cybister confusus (Coleoptera: Dytisccidae) on developmental stages of Carp, Catla catla (Ham.). Entomon 27(1):29–33 Ruberson JR, Yeargan KV, Newton BL (2001) Variation in diapause responses between geographic populations of the predator Geocoris punctipes (Heteroptera: Geocoridae). Ann Entomol Soc Am 94(1):116–122 Rubia EG, Ferrer ER, Shepard BM (1990) Biology and predatory behavior of Conocephalus longipennis (de Haan) (Orthoptera: Tettigoniidae). JTPP 7(1):47–54 Sahayaraj K, Lakshmanan PD, Loko YLE (2021) Nutritional fitness of a reduviid predator Rhynocoris marginatus (fab.) (Hemiptera: Reduviidae) using biological traits and macromolecules of pests. Int J Trop Insect Sci 41(1):611–616 Sahid A, Natawigena WD (2018) Laboratory rearing of Sycanus annulicornis (Hemiptera: Reduviidae) on two species of prey: Differences in its biology and effi ciency as a predator of the nettle caterpillar pest Setothosea asigna (Lepidoptera: Limacodidae). Eur J Entomol 115: 208–216 Satpathy JM, Patnaik NC, Samalo AP (1975) Observations on the biology and habits of Sycanus affinis Reut. (Hemiptera: Reduviidae) and its status as a predator. JBNHS 72(2):589–595 Sattar M, Hamed M, Nadeem S (2008) Biology of Coccinella septempunctata Linn. (Coleoptera: Coccinellidae) and its predatory potential on cotton aphids, Aphis gossypii Glover (Hemiptera: Aphididae). Pak J Zool 40(4):239–242 Sattayawong C, Uraichuen S, Suasa-ard W (2016) Larval preference and performance of the green lacewing, Plesiochrysa ramburi (Schneider) (Neuroptera: Chrysopidae) on three species of cassava mealybugs (Hemiptera: Pseudococcidae). Agric Nat Resour 50(6):460–464

References

219

Shahayaraj K, Sathiamoorthi P (2002) Influence of different diets of Corcyra cephalonica on life history of a reduviid predator Rhynocoris marginatus (Fab.). J Cent Eur Agric 3(1):53–62 Shu Y, Du Y, Wang J (2019) Presence of Cry1Ab in the Bt maize–aphid (Rhopalosiphum maidis)– ladybeetle (Propylea japonica) system has no adverse effects on insect biological parameters. Entomol Exp Appl 167:553–560 Shukla A, Jadhav DS (2014) Biology of Coccinella transversalis (Fabricius) on different aphid species. Bioscan 9(1):17–22 Siddique AB, Chapman RB (1987) Effect of prey type and quantity on the reproduction, development, and survival of Pacific damsel bug, Nabis kinbergii Reuter (Hemiptera: Nabidae). N Z J Zool 14(3):343–349 Situmorang J, Gabriel BP (1988) Biology of two species of predatory earwigs Nala lividipes (Dufour). (Dermaptera: Labiduridae) and Euborellia (Euborellia) annulata (Fabricius) (Dermaptera: Carcinophoridae). Philipp Entomol 7(3):215–238 Smith HA, Chaney WE (2007) A survey of syrphid predators of Nasonovia ribisnigri in organic lettuce on the central coast of California. J Econ Entomol 100(1):39–48 Solangi BK, Hulno MH, Baloch N (2007) Biological parameters and prey consumption by zigzag beetle Menochilus sexmaculatus Fab. against Rhopalosiphum maidis Fitch, Aphis gossypii Glov. and Therioaphis trifolii Monell. Sarhad J Agric 23(4):1097–1101 Srikumar KK, Bhat PS, Raviprasad TN, Vanitha K (2014a) Biology, behaviour and functional response of Cydnocoris gilvus Brum. (Heteroptera: Reduviidae: Harpactorinae) a predator of Tea Mosquito Bug (Helopeltis antonii Sign.) on cashew in India. JoTT 6(6):5864–5870 Srikumar KK, Bhat PS, Raviprasad TN, Vanitha K, Saroj PL, Ambrose DP (2014b) Biology and behavior of six species of reduviids (Hemiptera: Reduviidae: Harpactorinae) in a cashew ecosystem 1. J Agric Urban Entomol 30(1):65–81 Tank BD, Korat DM (2007) Biology of ladybird beetle, Cheilomenes sexmaculata (Fab.) in middle Gujarat conditions. Karnataka J Agric Sci 20(3):634–636 Torres JB, Silva-Torres CS, Ruberson JR (2004) Effect of two prey types on life-history characteristics and predation rate of Geocoris floridanus (Heteroptera: Geocoridae). Environ Entomol 33(4):964–974 Truong XL, Pham HP, Thai TNL (2020) Biology and predatory ability of the reduviid Sycanus falleni Stal (Heteroptera: Reduviidae: Harpactorinae) fed on four different preys in laboratory conditions. J Asia Pac Entomol 23(4):1188–1193 Vahmani A, Shirvani A, Rashki M (2022) Biological parameters of Oenopia conglobata (Linnaeus) (Coleoptera: Coccinellidae) feeding on different diets. Int J Trop Insect Sci 2022:1–7 Vanitha K, Bhat PS, Raviprasad TN, Srikumar KK (2016) Biology and behaviour of Ephestiasula pictipes (Wood-Mason) (Hymenopodidae: Mantodea) under captive breeding. Int J Pest Manage 62(4):308–318 Varshney R, Ballal CR (2017) Biology and feeding potential of Geocoris superbus Montandon (Heteroptera: Geocoridae), a predator of mealybug. J Entomol Zool Stud 5(3):520–524 Vasanthakumar D, Babu A (2013) Life table and efficacy of Mallada desjardinsi (Chrysopidae: Neuroptera), an important predator of tea red spider mite, Oligonychus coffeae (Acari: Tetranychidae). Exp Appl Acarol 61:43–52 Vennison SJ, Ambrose DP (1992) Biology, behaviour and biocontrol efficiency of a reduviid predator, Sycanus reclinatus Dohrn (Heteroptera: Reduviidae) from Southern India. Mitteilungen aus dem Museum für Naturkunde in Berlin. Fortschr Zool 68(1):143–156 Wang J, Li S, Yang J, Guo M, Dai H, Ramirez-Romero R, Wang S (2021) The fitness of mass rearing food on the establishment of Chrysopa pallens in a banker plant system under fluctuating temperature conditions. Insects 12(11):1014 Wheeler AG Jr, Stinner BR, Henry TJ (1975) Biology and nymphal stages of Deraeocoris nebulosus (Hemiptera: Miridae), a predator of arthropod pests on ornamentals. Ann Entomol Soc Am 68(6):1063–1068 Yokoyama VY (1980) Method for rearing Geocoris pallens (hemiptera: lygaeidae), a predator in California cotton. Can Entomol 112(1):1–3

220

7

Predatory Insects: Adults Biology of Various Orders

Yu JZ, Chi H, Chen BH (2013) Comparison of the life tables and predation rates of Harmonia dimidiata (F.) (Coleoptera: Coccinellidae) fed on Aphis gossypii Glover (Hemiptera: Aphididae) at different temperatures. Biol Control 64(1):1–9 Zhang M, Ma Y, Luo J, Ji J, Gao X, Wu C, Cui J (2022) Transgenic insect-resistant Bt cotton expressing Cry1Ac/1Ab does not harm the insect predator Geocoris pallidipennis. Ecotoxicol Environ Saf 230:113–129 Zhao Y, Ma Y, Niu L, Ma W, Mannakkara A, Chen L, Lei C (2013) Bt cotton expressing Cry1Ac/ Cry2Ab or Cry1Ac/epsps does not harm the predator Propylaea japonica through its prey Aphis gossypii. Agric Ecosyst Environ 179:163–167 Zhong B, Lv C, Qin W (2016) Preliminary study on biology and feeding capacity of Chelisoches morio (Fabricius) (Dermaptera: Chelisochidae) on Tirathaba rufivena (Walker). Springerplus 5(1):1–4 Zhu P, Lu Z, Heong K, Chen G, Zheng X, Xu H (2014) Selection of nectar plants for use in ecological engineering to promote biological control of rice pests by the predatory bug, Cyrtorhinus lividipennis, (Heteroptera: Miridae). PLoS ONE 9(9):e108669. https://doi.org/10. 1371/journal.pone.0108669

8

Polymorphisms in Insect Predators

Contents 8.1 General Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3 Wing Polymorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.1 Neuropteran: Lacewings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.2 Coleopteran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.3 Syrphid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.4 Redviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.5 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.6 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.7 Odonates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.8 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.9 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.4 Molecular Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5 Sexual Dimorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5.1 Mantodean . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5.2 Orthoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5.3 Odonate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5.4 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.5.5 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.6 Mimic Polymorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.7 General Reasons for Polymorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.8 Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

8.1

221 222 222 224 225 227 228 233 234 235 235 236 237 239 239 240 240 241 242 243 243 243 244

General Introduction

The behaviour, life history, physiology, and morphology of natural enemies’ victims are all affected in different ways. Scientists and naturalists have long been fascinated by the intra-specific variety that persists in natural populations. Colour polymorphisms, one type of this variation, offer a valuable chance to link genotypic

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_8

221

222

8

Polymorphisms in Insect Predators

and phenotypic variability in an ecological and evolutionary context. Colour polymorphisms are a common phenomenon that has been well documented in species from all branches of life. Additionally, these polymorphisms provide exceptional opportunities to link genotypic and phenotypic variation and promise insights into enduring questions in evolutionary biology, particularly those questions aimed at elucidating phases of speciation and adaptive radiations, because they frequently have a relatively simple genetic basis. This polymorphism serves important roles in predator deterrence, mimicry, and intra-specific sexual signalling. The adults of the majority of predatory insects exhibit polymorphisms, whereas the young do not exhibit this type of change. As a result, the main purpose of this chapter is to introduce adult polymorphism.

8.2

Definitions

For polymorphisms, two different schools of thinking have gained widespread acceptance. The first is polymorphism, which is population variability (Mayer 1969). Another definition of polymorphism is the simultaneous occurrence of two or more unique forms of the same spice in a ratio such that the rarest form cannot be preserved alone through recurrent mutation (Ford 1940).

8.3

Wing Polymorphism

Lin and Lavine (2018) claim that during their juvenile stages, wing polymorphic insects adapt to nutritional and environmental factors in order to either develop into a migratory adult morph with long wings, well-developed flight muscles, and small ovaries or a reproductive morph with short wings and well-developed ovaries. It has been demonstrated that juvenile hormone (JH) generally works to support growth in addition to keeping insects in their juvenile condition during moulting. Although some studies imply that JH is required for the determination of the reproductive morph, which either has short, non-functional wings or no wings at all, its significance in insect wing polymorphisms has proven to be more difficult to generalise. Within the genus, wing polymorphism is prevalent, and individuals within the same species as well as between species can exhibit macropterous, brachypterous, and apterous forms. In order to examine wing alterations in the Heteroptera with more accuracy, Slater (1975) suggested that the major varieties of front wings be classified as follows: 1. Aptery—In this form, there are no front wings at all. Despite several type 11 circumstances that almost always result in total loss, there are no recorded instances of this among the Lygaeidae. The Aradidae have a lot of true aptery. 2. Microptery—In this condition, the wings is reduced to tiny pads that are typically far apart from one another, leaving the abdomen exposed mesally with the clavus and corium joined and the membrane either non-existent or represented by a little

8.3 Wing Polymorphism

3.

4.

5.

6.

7.

223

flap. The metanotum and the back of the mesonotum may be seen in severe microptery. Staphylinoidy—In this condition, the first three abdominal terga are typically not covered by the fore wings, which are equally truncated across the posterior end and have the clavus and corium indistinguishably united into a coriaceous pad. As a result, the wings completely enclose the anterior half of the abdomen, leaving the posterior portion uncovered. Since this condition is so prevalent, it is more likely than not that it serves a functional purpose rather than merely serving as an intermediary stage before microptery. Coleoptery—In any case, the coriaceous region of the wing is prolonged rather than reduced in length overall in this condition. The clavus and corium are combined, with the former frequently becoming significantly larger along the midline. Like the elytra of Coleoptera, the two wings meet along the midline rather than overlapping. Three distinct subcategories can be identified. (a) In which the “claval commissure” is lengthened and the membrane is reduced to a thin flap, leaving the posterior abdominal segments exposed. The clavus and corium form a single coriaceous wing cover that meets the opposite cover evenly down the midline. (b) In which the membrane is only represented by a small mesa1 flap that does not extend posteriorly past the end of the extended lateral segment of the corium and the corium extends posteriorly laterally to reach or nearly reach the end of the abdomen. (c) When the fore wing is made up of an undifferentiated coriaceous cover that meets down the midline and resembles the wing of a normal beetle. The fore wing completely covers or nearly completely covers the abdomen. There is no trace of the membrane. Brachyptery—A condition in which the clavus and corium are shorter than they are in the macropter and may be clearly differentiated or merged. The membrane is somewhat smaller, usually extending to the third abdominal tergum, and typically just having the inner portions of the two membranes overlapping. Submacroptery—In this state, the clavus and corium typically clearly distinguish one another via a claval suture and share an identical appearance to that of a fully macropterous condition. The posterior abdominal tergum is visible due to the membrane’s relative shortness to the macropter, despite the fact that it is normally fully formed and having veins. The membrane frequently has a more prominent taper than in macropters, and in some cases the veins are absent. The back wing could be severely underdeveloped or well developed. Macroptery—The “normal” or “unmodified” wing condition with the clavus and corium distinct, the membrane well-developed and typically encompassing the abdomen, and the membrane of one wing fully overlapping that of the other. Every animal has a well-developed hind wing.

Mimicry An unprotected bug that mimics another insect that is unpleasant or hazardous, like a wasp, bee, or ant, is known as a Batesian mimic. In the Heteroptera, this phenomenon is pervasive; only ant mimicry can be found in no fewer than seven families.

224

8

Polymorphisms in Insect Predators

Miillerian Mimicry To lessen predation as the possible predator learns to avoid a particular colour or form pattern, several unappealing or dangerous species start to resemble one another in colour or structure, or both. Heteroptera are also commonly affected by this condition. The majority of the data is based on taxa that are not closely related to one another and that resembles other species while they are in sympatry but differ from one another when they are in allopatry, making it difficult to experimentally establish documented cases. Mertensian Mimicry Whereas the mimic is at least partially fatal, the model is just mildly unpleasant or harmful. Although this kind of imitation in Heteroptera has not been proven, it most likely occurs. Wasmannian Mimicry Although it is prevalent in the Staphylinidae and similar beetles, Heteroptera have not yet been recorded for Wasmannian mimicry, in which an insect attempts to “fool” the ant with which it is linked. Aggressive Mimicry A predator impersonates a prey species in order to “fool” the latter and increase predation. According to Mclver and Stonedahl (1993), several animals that eat ants may exhibit this kind. The reduviidpyrrhocorid mimetic situation detailed below may seem to be one of these situations, but it is likely an example of Batesian mimicry, therefore this occurrence needs to be carefully researched. Ground-dwelling Poeantius species from Africa were carnivorous on ants, and this was an instance of aggressive mimicry.

8.3.1

Neuropteran: Lacewings

Green lacewing is one of the most important predators of numerous small, softbodied pest species is Chrysoperla externa (Neuroptera). In a lab colony, these green lacewings also had yellow mutant individuals. In order to test the theory that the yellow colour is caused by an autosomal recessive allele, the method of inheritance of the yellow trait was investigated (Bestete et al. 2019). Basic life-history traits, such as the time to hatch and viability of eggs, the length and viability of the larval and pupal stages, the rate of adult emergence and survival, and the fecundity and longevity of females, were observed in both yellow and green morphs. Regarding all life-history traits examined as well as the rate at which their larvae were eaten, the yellow and green morphs could not be distinguished from one another. A homozygous recessive allele lacking sex-linked expression was shown to be the cause of the yellow colour through crossing tests. We draw the conclusion that there is a genetic polymorphism for body ground colour because the allele for yellow colour is present in the laboratory colony at high frequency (Bestete et al. 2019). In a different study, Chrysopa (Chrysoperla) carnea from a population in eastern North America that is currently classified as Chrysoperla plorabunda was grown in a lab and shown to have a radically unique autosomal mutant “blue” (Tauber and Tauber 1971). In other instances, the stony Mediterranean island of Malta has

8.3 Wing Polymorphism

225

revealed a “yellow” variation among green adults (Duelli 1992). The same applied to a different green lacewing. Western North America’s field-collected green Chrysopa oculata female produced offspring, one of which was yellow (Tauber et al. 1976). One of the many Ceraeochrysa species that are found in agroecosystems in the Neotropical region and have the potential to be used in biological control programmes is Ceraeochrysa caligata. Describe the differences from the original description of the pattern of integument markings for the adults and larvae of Ceraeochrysa caligata, paying particular attention to the amount of setae in the tubercles and somatic segments (Viana and Albuquerque 2009). According to Canard and Grimal (1988), Chrysopa perla pre-pupae likewise immediately enter diapause and the latter exhibits adult polymorphism. Reparative diphase may be the cause of the colour change, which has also been observed in Chrysoperla carnea, which is found in central Europe, and Chrysoperla plorabunda, which is found in North America. A lab-grown reddish Chrysoperla plorabunda specimen lays eggs that turn green again after about a week. Additionally, there is some evidence to suggest that temperature has a far greater impact on colour change than it does on the onset, duration, and cessation of reproductive diapause (Duelli et al. 2014). The owl fly Ptyngidricerus venustus is described by Krivokhatsky (2019) as having infra-sub-specific colour polymorphism, which is manifested in the presence of brightly coloured morphs with obsidian-brown wing membranes, typical of most congeners, and morphs with cloudy or hyaline wings with apical bands. The two sexes both exhibit these colour variants. Psectra diptera is polymorphic with respect to the hindwings (brachypterous individuals are likely flightless), and Ptyngidricerus = Omanoidricerus (Hemerobiidae) is recognised as a new synonym (Aspöck et al. 2020).

8.3.2

Coleopteran

The Coccinellidae family of ladybird beetles’ hereditary foundation for colour variations has been largely explained, and field sampling of these insects is not difficult (Sloggett and Honěk 2012). Therefore, examining the evolution and preservation of colour polymorphisms is appropriate for understanding ladybirds. Many ladybird (Coccinellidae) species, including Cheilomenes sexmaculata, Harmonia axyridis, Hippodamia variegate, Adalia bipunctata, Menochilus sexmaculatus, Coelophora quadrivatrata, and Hippodamia variegata, show a significant variation of elytron colour patterns within the same species. Numerous coccinellid species’ polymorphism has been thoroughly studied. Numerous researchers have looked into and detailed the relationships between colour and factors such as likelihood of survival during hibernation, activity levels at various temperatures, reproductive success, and environmental contamination. Ahmad et al. (2020) identified 37 morphs of Hippodamia variegate based on the quantity, size, and shape of these dots. The elytron had somewhere from 3 and 7 patches, which varied in number.

226

8

Polymorphisms in Insect Predators

Menochilus sexmaculatus features a striking combination of black and red in its elytra. In a polymorphic population of Menochilus sexmaculatus in Japan, Kawakami et al. (2018) studied the impact of maternal morph types on the phenotypic expression of the elytral colour morph as well as the effect of temperature on growth from the first instar larva to the pupal stage. From a wild population, females of three different elytral colour morphs were taken, and each female produced three groups of hatchlings that were raised at three different constant temperatures (20 °C, 25 °C, and 30 °C). The phenotypic frequency of F1 adults suggested that genetic rather than environmental variables influenced the elytral morph type. In particular, type A (almost black morph) was the most prevalent in F1 from type A mothers (male: 52.6%; female: 32.3%), while types B (four small-dotted morph), F (four medium-dotted morph), and type G (four larger-dotted morph) were the most prevalent from type B and type G mothers, respectively (male: 56.7%; female: 53.3%) and type B (four small-dotted morph), respectively. As a result, rather than being a result of phenotypic plasticity brought on by growing temperatures, the manifestation of elytral colour polymorphism in the Osaka, Japan population is believed to have a genetic basis dependent on parental morphs. The results of a 9-year field survey were used in another study (Kawakami et al. 2019) to suggest that the frequency of the light morphs increased significantly from the overwintered generation to the spring generation, whereas the frequency of the dark morphs of Cheilomenes sexmaculata increased significantly from the autumn generation to the overwintered generation. Extreme elytral colour polymorphism is also present in the ladybug, Harmonia axyridis, with more than 100 different varieties having been identified by Murakami et al. (2019). An explanation for this type of polymorphism involves the inheritance of 12 fictitious alleles at a single autosomal locus. Studies by Murakami et al. (2019) demonstrate that a number of morphs increase the size of their aggregation during gathering before hibernation. Each of the four colour morphs had a significantly beneficial impact on the size of the aggregation depending on how much of one (or none) of the four was present relative to the expected values. All aggregating members fared better when the aggregation size was higher. These findings imply that grouping with these morphs is advantageous for survival. The great colour diversity in this species is partially explained by the survival advantage it enjoys during hibernation. The ladybird beetle Cheilomenes sexmaculata was studied for changes in fitness and life cycle parameters, including as body size, mortality, fecundity, hatching rate, and mate preference, in addition to temperature and hibernation (Coccinellidae). An effective natural enemy with a global distribution, the colourful Asian ladybird Harmonia axyridis (Coccinellidae) exhibits a variety of elytral patterns and has a high degree of environmental adaptability. To better appreciate the evolutionary importance of elytral pattern diversity, certain studies have reported differences in ecological adaptability among Harmonia axyridis morphotypes, but none have evaluated their predatory performance under different environmental situations (Chen et al. 2019a).

8.3 Wing Polymorphism

227

Although Pterostichus melanarius can travel rather great distances on foot, its wing dimorphism and potential for flight may be essential for rapid range extension. According to studies, Pterostichus melanarius wing dimorphism is inherited through straightforward Mendelian genetics, with brachyptery predominating over macroptery. While macropterous individuals have long hind wings and can fly, brachypterous individuals have small hind wings and are flightless. Brachypterous individuals tend to grow with habitat stability and duration after colonisation, whereas macropterous individuals tend to make up a larger portion of newly established populations (Busch et al. 2021). Morphological Features in Polymorphism Using scanning and transmission electron microscopy, the antennal receptors of the adult male and female ladybird beetle Semiadulia undecimnotata (Coccinellidae) were studied. The Bohm, trichoid, coeloconic, basiconic, and chetiform sensilla morphological classifications were used to classify the twelve different receptor types. Male and female dimorphism in Semiadulia undecimnotata is primarily characterised by the following traits: (a) (b) (c) (d)

Males have 540 sensilla (of all sorts), compared to females’ 500. Males have two types of taste receptors that females do lack. Females have a particular form of mechanoreceptor that males lack. The three sensilla types that are specific to each sex are only found on the inner side of the antennae in both males and females (Jourdan et al. 1995)

8.3.3

Syrphid

Many members of the Syrphidae family of insects imitate aculeate Hymenoptera. The hymenopterans that serve as models for these syrphids, particularly vespid wasps, have somewhat long, many segmented, filiform antennae that are frequently in motion and are typically a visible feature. The majority of syrphids, like all other cyclorrhaphous Diptera, have three-segmented antennae that are typically short and noticeably unobtrusive, like those of, for instance, the non-mimetic Tropidia quadrat (Waldbauer 1970). The cuticular hydrocarbons of the larval myrmicine ant, Myrmica incompleta pupae, and those of the predatory syrphid fly, Microdon albicomatus, are qualitatively indistinguishable. There were found to include 18 hydrocarbon components, including two Z-9-monoenes, n-alkanes, 3-methyl alkanes, 11- and 13-methyl alkanes, one dimethyl alkane (5,17-dimethyl pentacosane), and 18 different types of methyl alkanes (23:1 and 25:1). The same components as the ant pupae were found in the cuticular hydrocarbons of worker M. incompleta, although in different relative abundances. The Microdon larvae used in a radiolabelling experiment using C-acetate showed that the fly biosynthesises its hydrocarbons rather than obtaining them from their prey (Howard et al. 1990). It is possible that Syrphidae and Hymenoptera exhibit mimicry because of their comparable morphologies and colour patterns. The syrphid is a Batesian mimic of the hymenopteran. Field observations

228

8

Polymorphisms in Insect Predators

can either support mimicry in certain instances or indicate that a different hymenopteran is truly the model in other instances. Specific mimicry refers to the detailed resemblance in colour, anatomy, and behaviour to one or a few species of bees or wasps, while non-specific mimicry refers to the more generic and less precise resemblance, frequently to a group of hymenopterans rather than to a single species (Howarth et al. 2000). Many species of syrphid flies mimic wasps and stinging bees as adults. The term “Batesian mimicry” refers to the phenomenon whereby palatable creatures imitate or “mimic” disagreeable counterparts. There are various syrphid species that feed on aphids. Different hover fly species are believed to replicate the colours and patterns of various insects in order to make them more appetising to prospective predators. Hymenoptera, which includes wasps and bees, have unpleasant and/or venomous models that cannot be reproduced without permission. Polidori and colleagues in 2014 observed this Batesian mimicry between the syrphid fly Aneriophora aureorufa and the bumblebee Bombus dahlbomii.

8.3.4

Redviidae

Intriguing morphological phenomena in insects include the development of wing polymorphism and asymmetric male genitalia. These two intriguing physical characteristics are present in the tribe Metapterini of the Emesinae, or thread-legged bugs. Nevertheless, the absence of phylogenetic ideas for Emesinae prevents the advancement of evolutionary analyses of these occurrences. Due to their elongated and superficially fragile body, thread-legged bugs can be distinguished from assassin bugs with ease. There are over 280 known species in the 28 genera that make up the tribe Metapterini. Wygodzinsky (1966) put out the lone phylogenetic hypothesis for the Emesinae tribes, which suggested Deliastini Villiers as the Metapterini’s sister group. However, this proposal has never been put to the test using cladistic methods. Recent studies based on proleg and genitalia character sets reveal that Metapterini may not be monophyletic. We created a morphological dataset of 138 characters, which includes external morphological characters, specific details of prolegs, and genitalia for both sexes for Metapterini. This dataset was used to conduct cladistical analysis on 43 species of Metapterini, 55 terminals, 24 genera (85.7% of the generic diversity), and 12 out-groups. By include Bergemesa Wygodzinsky, Palacus Dohrn, and Stalemesa Wygodzinsky—all now ascribed to Deliastini—Metapterini was rediscovered as paraphyletic. According to Wygodzinsky, Gardena (Emesini) was discovered as the sister group of Metapterini + Deliastini (1966). Based on these findings, we rename Deliastini syn. n. to Metapterini sensu n. and suggest two new genera: Bacata and Valkyriella for Ghilianella borgmeieri and three Andean species previously classified as Liaghinella. Males and females were both completely winged in the ancestor of Metapterini sensu n., with two independent evolutionary transitions to the apterous and brachypterous conditions. This is indicated by the ancestral state reconstruction of wing polymorphism. The examination of the male

8.3 Wing Polymorphism

229

genitalia’s symmetry reveals that Metapterini had two independent emergences of asymmetrical male genitalia as well as an ancestor with symmetrical male genitalia. Rhynocoris marginatus is a polymorphic reduviid predator that eats insect pests and inhabits tropical rainforests, semi-arid regions, scrub jungles, and agroecosystems. There are three distinct morphologies of it (Ambrose and Livingstone 1988): 1. With black connexivum (niger) 2. With red connexivum (sanguineous) 3. With black and red banded connexivum (nigrosanguineous) The niger morph had the highest fecundity and longevity, followed by the nigrosanguineous and sanguineous morphs. For the niger (N), nigrosanguineous (NS), and sanguineous (S) morphs, the intrinsic rates of natural growth were 0.034, 0.028, and 0.026 per female per day, respectively. In the course of the cohort generation, the population multiplication of the niger morph was noticeably higher than that of the nigrosanguineous and sanguineous morphs. In all morphs, the life expectancy (ex) of freshly deposited eggs declined with age (George 2000). Under controlled laboratory settings, biological control potential and molecular profiles were taken from the partners N male + N female, N male + S female, N male + NS female, and S + NS female in order to determine the impact of Rhynocoris marginatus polymorphism on the biology. Offspring born to S+NS female partners swiftly developed, having the lowest body weight, size, and egg hatching percentage, the highest relative growth rate in both first instars and adults, the highest survival rate (83%), and a sex ratio that is heavily slanted towards men (0.23). Third instars to adults, the prey deprivation period was often lengthened from 24 to 72 h, increasing the predatory rate. According to molecular analysis investigations, N adult morphs had a high total free amino acid content. There was no distinguishing DNA pattern among the Rhynocoris marginatus variants. The approaching time was also reduced in the later instars and adults of the polymorphic Rhynocoris marginatus, N male + N female, N male + S female, N male + NS female, NS male + NS female, and S + NS. The biological control potential is highly evident in 96 h of starvation N+N adults, who showed a minimum approaching time, high handling time, and predatory rate. Low weight increase was seen among the experimental morphs N + N and minimal weight gain was seen among the SNS morphs. The offspring of Rhynocoris marginatus S + NS, NS + NS, and N + NS partners had the lowest total nymphal developmental period, the lowest body weight, and the highest survival rate (Petchidurai et al. 2019). Insecticides that are used to control insect pests have an impact on non-target biological controls like Rhynocoris marginatus. In order to assess the adaptive polymorphic resistance of the morphs of Rhynocoris marginatus morphs to the toxic effects of three insecticides—dimethoate, methylparathion, and quinalphos— each in five concentrations—laboratory studies were carried out. According to the 50% lethal concentration (LC50) values, endosulfan, quinalphos, and methylparathion were the three insecticides that were the most toxic to all three

230

8

Polymorphisms in Insect Predators

Fig. 8.1 Three polymorphic forms of Ectomocoris xavierei: Alate male (1), Brachypterous male (2) and Brachypterous female (3) Fig. 8.2 Polymorphic forms of Ectomocoris tibialis. brachypterous male (BB), alate male (AM)

morphs. The sanguineous was the most sensitive of the three morphs, while the niger was the most resilient (George and Ambrose 2001). Assassin bug Ectomocoris xavierei is polymorphic. Ectomocoris xavierei is a rare example of a reduviid in which the males are both alate and brachypterous whereas the females are just brachypterous. 12.67:1:0.61 days following imaginal moult, Ectomocoris xavierei spawns dull white, elongate oval eggs. Each litter has 7.7 eggs and eggs are placed separately. A maximum of 12 eggs can be laid every day by Ectomocoris xavierei, which lays its eggs in the soil and does so in 2–4 min. The eggs take 23–27 days to hatch, and the pinkish red nymphs turn black with yellow patterns in 45–60 min. From hatching through imaginal moult, the nymphal stage lasts 93.6 days. There is a description of the nymphal instars as well as an identification key. Since females have a biased sex ratio and live longer than males (Vennison and Ambrose 1991). Ectomocoris xavierei exhibits alary polymorphism (Fig. 8.1). Additionally, Ectomocoris tibialis exhibited wing polymorphism. According to Sahayaraj (1991), this reduviid has three polymorphic forms: a brachypterous male (BB), an alate male (AM), and a brachypterous female (BF) (Fig. 8.2). Additionally, Ectomocoris tibialis exhibits complete violaceous black, red, and violaceous black with light red corium males (Fig. 8.3). Male and female morphs differ noticeably, according to morphometric assessments. Alate males married with brachypterous

8.3 Wing Polymorphism

231

Fig. 8.3 Line diagram of Ectomocoris tibialis entire violaceous black male (a), entire red male (b) and violaceous black male with pale red corium (c)

females produced children with larger bodies than brachypterous males mated with brachypterous females. The former groups have greater fertility than the latter. Even though the brachypterous female and alate male produced more eggs, the fecundity index is lower. In comparison to the nymphal phases of brachypterous and alate parents, the nymphal period of brachypterous parents had longer stadial duration. Nymphal mortality was higher in the former group (31%) than in the latter group (26%). All of the morphs successfully interbred, although the offspring did not follow the Mendelian ratio. There are four distinct morphs of the polymorphic assassin insect known as Acanthspis siva. Eggs that are maroon in hue are deposited in groups without any cement. In about 13 and 29 days, the eggs begin to hatch into pale, fuscous, and testaceous nymphs. From initial instar to adult, the stadial phase lasts between 71 and 113 days. Males and females live longer as adults. For two generations of laboratoryraised bugs, the sex ratio is 0.6:1. Breeding research demonstrated its polymorphism (Ambrose and Livingstone 1987). Rihirbus discrepans has clear sexual dimorphism in its body structure and coloration. Males compared to females are significantly smaller and slimmer. Males’ abdomens are not laterally dilated and have almost parallel sides, whereas females’ abdomens are clearly laterally dilated. Pronotal humeral angles vary slightly among individuals, with those from China having more rounded angles. In this species, the colour patterns are extremely varied (Gao et al. 2015). The ectrichodiines are distinguished by their highly frequent wing polymorphism (>50% of the known species) and their iridescent or vivid red and black colour. Only females can reduce their wings in the New World. About half of the species that are known exhibit one or more kinds of wing reduction. All known male members of the genus Zirta are macropterous, while all known female members are either macropterous or submacropterou. Pygolampis, Pygolampis breviptera, is known to commonly exhibit wing polymorphism (Okuda and Chen 2021). The outward sexual

232

8

Polymorphisms in Insect Predators

Fig. 8.4 Neohaematorrhophus therasii entire violaceous black male (a), entire red (except abdomen and central prothorcic spot) male (b), and violaceous black male with pale red corium (c)

dimorphism in Reduvius hirticornis is minimal; females tend to be larger than males and have smaller eyes and ocelli. Females have at least the basal segment naked; males exhibit pubescence on all antennal segments. A midventral keel could be created by compressing sternite 2 laterally. The ectrichodiines are distinguished by their highly frequent wing polymorphism (>50% of the known species) and their iridescent or vivid red and black colour. Only females can reduce their wings in the New World. About half of the species that are known exhibit one or more kinds of wing reduction. All known male members of the genus Zirta are macropterous, while all known female members are either macropterous or submacropterou. Pygolampis, Pygolampis breviptera, is known to commonly exhibit wing polymorphism (Okuda and Chen 2021). The outward sexual dimorphism in Reduvius hirticornis is minimal; females tend to be larger than males and have smaller eyes and ocelli. Females have at least the basal segment naked; males exhibit pubescence on all antennal segments. A midventral keel could be created by compressing sternite 2 laterally. A recently described reduviid found in Tamil Nadu’s scrub-jungles and semi-arid regions is called Neohaematorrhophus therasii. This reduviid exhibits wing polymorphism as well as sexual dimorphism (Figs. 8.3 and 8.4). Ambrose and Livingstone (1991) researched the biology of Neohaematorrhophus therasii and described the violaceous black male and female. The polymorphic forms in this reduviid, however, were described by Sahayaraj (1991) as follows: entire violaceous black male, entire violaceous black female, complete red male (apart from the abdomen and central prothorcic patch), and violaceous black male with pale red corium (Figs. 8.4 and 8.5) No discernible colour differences were discovered in the size of the adults that emerged from the progeny, according to the data. The youngness of a black father and a black mother lived the longest as adults (85 days), while the children of a red father and a red mother lived the least amount of time as adults (48 days). The pre-oviposition era and adult longevity did not correlate. In comparison to the other

8.3 Wing Polymorphism

233

Fig. 8.5 Neohaematorrhophus therasii entire violaceous black female (a), and entire red female (b)

three morph combinations, the eggs from the black father and red mother were larger (47 eggs/female). The children of a black father and a red mother, as well as a black father and a black mother, resembled their parents (40:60 females and males). This haphazard appearance of several morphs had nothing to do with environment or genetic nutrition. This form of feedback is a prime illustration of the ontogenetic evaluation-equivalent non-mutational, non-adaptive trends in evaluation. Additionally, it was noticed that a reduviid’s colouring changes when it is eating on a certain prey. For instance, after being fed Dysdercus superstitiosus for several weeks, nymphs and adults of Phonoctonus subimpictus showed no tendency to produce any red integument pigment. In the same conditions, red and white predators displayed a modest loss of the red pigment over the course of many weeks, which was likely caused by ageing (Stride 1956).

8.3.5

Nabidae

The macropterous and brachypterous morphs of the damsel bug Alloeorhynchus reinhardi (Hemiptera, Heteroptera, Nabidae, Prostemmatinae, Prostemmatini) have been observed in China and Korea. Due to their wing polymorphism, individuals of the Nabis (Austronabis) genus must frequently be identified by comparing the male and female genitalia in order to achieve an accurate identification (Cornelis et al. 2021). Even though the wings of many adult Nabidae are fully grown (macropterous), several species exhibit significant wing polymorphism (microptery and brachyptery). Wing reduction might be little, as in the case of Nabis brevis, or significant, as in the case of Phorticus brevipennis (Lattin 1989). The brachypterous male of Nabis ashworthi is a sordid light brown colour overall, with the exception of

234

8

Polymorphisms in Insect Predators

the middle stripe running from the apex of the clypeus to the pronotum, the lateral angles of the scutellum, the dorsal surface of the abdomen, and the lateral band running from the pre-ocular region to the tip of the abdomen, which are all brown. Similar to the male, brachypterous female are slightly longer and wider in the middle of the abdomen (Cornelis et al. 2016).

8.3.6

Pentatomidae

The Stiretrus decemguttatus nymphs’ colour patterns were categorised into three phenotypes (Hemiptera: Pentatomidae), with types C1–D1 and C2–D2 being the most prevalent and involving warning colours. The head, thorax, wings, legs, and abdominal markings of type C1–D1 are all black, while the rest of the abdomen is red or reddish-orange. Only the presence of two reddish marks on the thorax and a possible all-red head distinguish Type C2–D2 from Type C1–D1. The third phenotype’s nymphs are less common and entirely orange. Adults with the colour pattern eight were produced from nymphs with this pattern that were collected in the field. While a pair with a specific pattern may create children with individuals of that pattern, even from the same egg clutch, a pair with a different pattern may produce offspring with individuals of that pattern (Paleari 2013). A South American polymorphic species known as Stiretrus decemguttatus (Pentatomidae, Asopinae) preys on Chrysomelidae. Aspects of the ecology, polymorphism, and developmental biology of the significant cassidine beetle predator Stiretrus decemguttatus (Hemiptera, Pentatomidae). On the Marajó Island, Brazil, Stiretrus decemguttatus is a significant predator of the cassidine beetle species Botanochara sedecimpustulata and Zatrephina lineata (Coleoptera, Cassidinae). It preys primarily on late-instar stages but attacks people at all phases of development. Its life cycle took 44 days in the lab, including the six days it took for the egg to hatch and the 16 and 22 days it took for it to reach adulthood. One generation (T) lasted for 13 days, with a 0.25 intrinsic population growth rate (r). The life cycle of Zatrephina lineata appears to be more affected by Stiretrus decemguttatus than the other two prey species, according to these data. However, no preference over species of cassidine was evident in the lab. Based on combinations of three fundamental sets of colour marks, Stiretrus decemguttatus adults can display up to 17 different colour patterns. Some of them may be mimetic rings since they mimic chrysomelid marks connected to Ipomoea asarifolia (Convolvulaceae). Three colour patterns were seen in nymphs, however none of them were connected to a particular adult colour pattern (Paleari 2013). A significant cassidine beetle predator from South America is Stiretrus decemguttatus (Pentatomidae). Numerous individuals were studied, and their colour patterns were categorised into 17 morphs based on whether they had uniform (without spots) or varied patterns. While the latter are both metallic and non-metallic, adults that are variegated have different patterns, such as round or elongated maculae on the dorsal surface. Individuals will typically have eight to ten maculae visible, whereas those with circular patterns always have ten maculae

8.3 Wing Polymorphism

235

visible. This depends on how far the elongated maculae have fused. In both the field and the lab, the most typical patterns were 2, 4, 10, and 11. The most uncommon, 8 and 9, were first discovered in 1991 and 1994, respectively, and appear to be phenotypic more prevalent in males (Paleari 2013). Brugnera et al. (2019) fully depict the colour variability of Oplomus catena adults and suggest four colour patterns.

8.3.7

Odonates

There are reports of polymorphism in both male and female insects. But regrettably, it only happens in female odonates. Numerous Odonata species frequently exhibit female-limited polychromatism. According to Sánchez-Guillén et al. (2005), Ischnura elegans possesses three different colour morphs, including an androchrome (a colour that resembles that of a man) and two other gynochrome brown morphs called infuscans and rufescensobsoleta. It was demonstrated molecularly in the damselflies Ischnura graellsii (Cordero 1990; Andrés et al. 2000) and Ischnura elegans (Banham 1990). Colour polymorphism is commonly present in odonates, however it is more frequently found in damselflies (Zygoptera) than dragonflies (Anisoptera). According to Sánchez-Guillén et al. (2005), a single-locus, two-allele autosomal genetic polymorphism is consistent with the genetic regulation of the development of the diverse forms in several species of damselflies. In female-limited colour polymorphism, only females exhibit different colour variations. The coexistence of one or more female-like forms and one or more male-like forms (andromorph) is the most frequent pattern in odonates (heteromorphs). Here, several varieties of female pigmentation are compared to adult male colouring using the terms “heteromorph” and “andromorph” (Fincke et al. 2005). The terms andromorph, androform, androchrome, and androchromotype, as well as isochrome, are used in odonate literature. Individuals with andromorphism may differ from ordinary non-andromorphic people not only in terms of colour but also in terms of conduct (Joshi et al. 2020).

8.3.8

Anthocoridae

In this family, wing polymorphism is typical and frequently correlated with the cryptic habit (Lattin 1999). However, it has since been discovered that colour polymorphism and alary are extremely uncommon in the family and have never been linked to seasonal or environmental factors. There are two alary morphs documented for male Orius retamae and three for females (Wagner 1952). Only a very small number of species, especially those that are found in sheltered settings, contain apterous variants (Lattin 1999). Several instances of wing polymorphism have been documented, despite the fact that the majority of species in the family only have one (alate) morph. For instance, there are two morphs of bugs belonging to the

236

8

Polymorphisms in Insect Predators

genera Brachystekes, Elatophilus, Temnostethus, and Xylocoris: brachypterous and macropterous (Saulich and Musolin 2009). Gamberale-Stille et al. (2010) suggest that the greater benefit of crypsis in larvae and pale adults, as these coincide with a low activity phase in the life cycle, may help maintain this colour polymorphism between the different life stages, whereas in spring, the situation changes and an aposematic strategy is more advantageous as the individual needs to move around to reproduce. Due to the lack of evidence for either the seasonal feature of alary polymorphism or its dependency on external ecological conditions, such relationships are known for many other heteropteran families (Saulich and Musolin 2009). Nidicola marginata exhibits wing polymorphism (Carpintero and Dellapé 2012). Paleari (2013) observed polymorphism in a South American Asopinae predator called Stiretrus decemguttatus (Hemiptera, Pentatomidae), which feeds on Chrysomelidae, during the same year. The observed trends of hundreds of adults could be categorised. Seventeen variants that display uniform (without spots) or variegated patterns were created. The first includes metallic and non-metallic colours, whereas variegated adults show a variety of patterns, such as round or lengthened the dorsal region of the maculae. Depending on how much the bones have fused individuals with eight to ten elongated maculae will present, while those with circular patterns certainly do. The rarest phenotypes, 8 and 9, which were recently discovered in 1991 and 1994, respectively, appear to be ones that are more prevalent in male (Paleari 2013). Colour polymorphism in Pentatomides has been observed in the nymphs and adults as well as the eggs. Podisus maculiventris produces polymorphic eggs, which each female can individually colour from light yellowish-white to dark brown or black (Abram et al. 2015). It is interesting to note that light eggs are typically laid on leaf undersides while dark eggs are typically found on leaf tops. It has been proposed that the environment’s uneven solar radiation intensities are the cause of this paradox. In other words, UV-absorbing pigmentation is required on leaf tops to prevent eggs put there from being exposed to harmful solar radiation, while UV-absorbent pigments in leaves protect pale eggs laid on the bottom (Gutschick 1999). Abram et al. (2015) tested that theory and discovered that darker eggs were more likely than pale eggs to survive prolonged UV radiation exposure.

8.3.9

Hymenoptera

Formica exsecta, which tends aphids on Juniperus, Picea, and other trees, is mostly aphidicolous but is also predatory. The head and thorax of Formica exsecta workers vary in colour, and they are bicolored. This colour can range from entirely black to completely pale (yellow or red), with a number of intermediate phases. The pigment is deposited on the head and thorax in spots that vary in size and shape. This species is a widespread trans-Palearctic species that may be found in a variety of natural habitats, from forest-tundra in the north to steppes in the south. As a result, it inhabits regions with a distinct gradient in environmental circumstances. Throughout its

8.4 Molecular Mechanisms

237

range, it favours open biotopes. There is a link between colouring and environmental gradient adaptation. Putyatina et al. (2022) discovered that although a pale (less melanised) head and three thoracic segments frequently co-occur with paler other sections of the body; the coloration of different regions of the body varies independently to a great extent. They also demonstrate how there are considerable differences amongst complexes in the link between colour pattern and body size. They contend that it is possible to use these colour variations to investigate how ant populations differ spatiotemporally and how they adapt to their microhabitats. Future research can easily quantify the diversity in colour patterns using the suggested scheme. Hlaváček et al. (2022) suggest eight coexisting species of the Oriental hornet (Vespa orientalis, Hymenoptera: Vespidae), which we call a Batesian–Müllerian mimicry ring. We looked at their phenology, natural history, and geographic distribution to uncover broad ecological patterns. The “model-first” theory states that Müllerian mimics of this ring appeared earlier in a season, while Batesian mimics appeared later. Volucella zonaria, a Batesian mimic (Diptera: Syrphidae), is thought to exhibit allopatry with its model and, maybe, a less exact likeness to a different model, the European hornet, as a result of temperature-driven range extension (Vespa crabro: Hymenoptera: Vespidae). It appears that the colour variations of the polymorphic species Cryptocheilus alternatus (Hymenoptera: Vespidae), Delta unguiculatum (Hymenoptera: Vespidae), Rhynchium oculatum (Hymenoptera: Vespidae), and Scolia erythrocephala (Hymenoptera: Scoliidae) exhibit distinctive geographic distribution patterns. This may be caused by sympatry with alternative models Based on observations of the suggested Oriental hornet mimicry ring, general coevolution patterns of models and mimics in heterogeneous and temporally dynamic contexts are examined.

8.4

Molecular Mechanisms

In ecology and evolutionary biology, it is crucial to comprehend the mechanisms underlying genetic polymorphisms within and between species. In insects, polymorphic melanism is a striking phenotype that results from particular genotypes and may play a key role in assortative sexual selection-driven speciation processes. With significant consequences for a range of behavioural, physiological, and reproductive performance that would be implicated in evolutionary processes, pigmentation is a prominent and highly variable characteristic of insect physiology. Melanin is a pigment that is found in all animals. It plays a role in the development of adaptive colour patterns, ultraviolet radiation protection, and even immune responses to diseases and parasites. Melanism in insects is thought to be caused by the melanin pathway at the molecular level. A number of enzyme reactions are involved in the complicated biochemical process of melanin production. Tyrosine is converted to DOPA (dihydroxyphenylalanine) by the enzyme tyrosine hydroxylase (TH), which is subsequently used to make DOPA melanin. As an alternative, DOPA decarboxylase can further transform DOPA into dopamine (DDC). Dopamine is finally

238

8

Polymorphisms in Insect Predators

transformed to dopamine melanin by phenol oxidases (POs), which also break down dopamine into its quinones. In 1993, Roehrdanz’s co-workers utilised the polymerase chain reaction using arbitrary primers to analyse DNA polymorphisms in predatory Coccinellids (RAPDPCR). Since the early twentieth century, the Asian multi-coloured ladybird Harmonia axyridis (Coccinellidae) has been used in field and greenhouse crops for biological control of insect pests. Similar to most holometabolous insects, Harmonia axyridis exhibits simple thermal melanism in the pupal stage and no variation in phenotype during adulthood; however, the species’ highly diverse elytral colour morphs, which are largely formed by various patterns of melanin deposition, make it a useful model for research on melanin synthesis and deposition in insect elytra. Additionally, Harmonia axyridis phenotype varies significantly with the seasons, and assortative mating based on the succinic and melanic phenotypes was discovered in the population. The Asian multi-coloured ladybird (Harmonia axyridis), a perfect holometabolous insect for investigations of melanisation due to their very variable elytra colour, according to Chen et al. (2019a, b), was used to identify DDC. The DDC gene from Harmonia axyridis was shown to be constitutively expressed in all embryonic stages after analysis. To investigate the melanin synthesis route of the elytra in Harmonia axyridis, they used the RNAi approach. After injecting third instar larvae with double-stranded RNA of HaDDC (dsHaDDC) at a rate of 300 ng/individual, the transcript levels of HaDDC were drastically reduced. In third instar larvae, silencing HaDDC had no substantial impact on pupation or eclosion or on mortality. We also proved that all H. adults are. When the HaDDC gene was silenced in the third larval stage, the axyridis (forms succinea, spectabilis, and conspicua) had an aberrant phenotype that manifested as decreased elytra melanin. However, melanin was still seen in other adult body parts, such as the pronotum and head. These findings show, for the first time, that the primary factor contributing to the melanisation of Harmonia axyridis’ elytra is dopamine-derived melanin. Additionally, they demonstrate that silencing of HaDDC in the third larval stages drastically decreased female egg laying and egg hatching, supporting the idea that DDC controls fecundity. As a result, DDC is probably pleiotropic with regard to its function in fecundity processes and the formation of melanin. The framework on which future research on the mechanism of pigment generation and patterning in different types of insect colours may be undertaken is strengthened by these findings, which provide unique insights into melanin formation in holometabolous insects. The multi-coloured Asian ladybird beetle Harmonia axyridis’s wings are patterned in part by the gene pannier (pnr), which codes for a transcription factor of the GATA family (Niimi and Ando 2021). In a different study, Kawakami et al. (2021) noted that Cheilomenes sexmaculata’s elytral colour polymorphism had a geographical variation from 33° N to 36°N in Japan over a 100-year period beginning in 1900. The findings showed that the genetic makeup of the mitochondrial COI gene (mtCOI) was related to different morph types according to latitude; in particular, the haplotype group of Cheilomenes sexmaculata that was more prevalent at lower latitudes tended to have

8.5 Sexual Dimorphism

239

light morph types. The distributional expansion into higher latitudes since 1900 may have been driven by dark morph types in the higher latitude genetic group rather than lower latitude light morph types, according to the haplotype of Cheilomenes sexmaculata that is prevalent in higher latitudes.

8.5

Sexual Dimorphism

Insects, in particular their natural enemies, use a variety of defence mechanisms to avoid being eaten in the wild. They have developed highly complex strategies for mimicking everything from other predators to the environment’s foliage in an effort to hinder predators’ ability to keep a clear record of the prey images they need to hunt. To provide a far wider variety of forms, cryptic insect species frequently evolve additional imitation techniques like polymorphs and masquerade. Sexual dimorphism allows for even more variation in form. Sexual size dimorphism (SSD), shape, features, colour, and parasite load differences between males and females of the same species are all considered to be a part of sexual dimorphism. Most insect species exhibit sexual dimorphism, with female-biased SSD predominating (e.g. Coleoptera, Odonata, Orthoptera, Phasmatodea). Sex-specific selective forces may be the cause of the prevalent in arthropods size and shape dimorphisms between males and females. Due to their often restricted movement, large insect females must rely on camouflage and defensive behaviours to survive. Many species have significant life histories and behavioural differences between the sexes. Furthermore, selection for features that make men more attractive to females or simply provide male’s better access to females is a popular explanation for sexual dimorphism. Sexual size dimorphism (SSD), which refers to the relationship between the size disparities between males and females in a natural population, often follows patterns of female-bias (females larger than males) or male-bias (males larger than females) (Fairbairn et al. 2007).

8.5.1

Mantodean

When aggressiveness and cannibalism are interwoven with courting, sexual size dimorphism becomes very intriguing. The majority of the research examining the connection between SSD and cannibalistic behaviours has been conducted on spiders and praying mantises, with a substantial female skew in both the SSD and cannibalistic female mantids populations (Barry et al. 2008; Prokop and Václav 2008). There has been less research done on animals with male-biased SSD and cases of males eating females (Schütz and Taborsky 2005; Aisenberg et al. 2011). The body size and camouflage abilities of predatory praying mantises are sexually dimorphic, with the males being somewhat smaller and more flight-capable and following the females utilising smell and visual cues. The bulk of these insects follow similar patterns. The majority of these sexually dimorphic mantis species continue to have sexes that are physically identical and that live in the same

240

8

Polymorphisms in Insect Predators

ecological niche, which is an interesting observation. The orchid mantises, on the other hand, are members of a lineage of highly camouflaged mantises that includes two species that are very conspicuous, appear to have abandoned camouflage coloration, and display significant sexual size dimorphism (SSD) as well as sex-dependent cryptic traits. Praying mantises blend in with their surroundings and disguise the contours of their bodies to ward off predators, making them impressive instances of disruptive camouflage. Most mantids are green, white, and brown in colour (Svenson et al. 2016). The general degree and direction of SSD, however, remain very consistent in well-researched cannibalistic systems, allowing scientists to categories particular species as having primarily male-biased or female-biased SSD. Extreme size variation in both sexes in cannibalistic creatures is uncommon and poorly understood. Both sexes in these animals may be at danger of cannibalism since either the male or female in each mating pair may be much larger than the other. It may be possible to get new knowledge about the evolution and use of SSD across taxa by investigating the effects of size on aggressive behaviour, cannibalism, and reproductive behaviour in systems with severe size variation in both sexes (Lietzenmayer et al. 2022).

8.5.2

Orthoptera

The Orthopteran subfamily pterochrozinae (Orthoptera: Tettigoniidae) contains frequent and intricate examples of polymorphism and cryptic masquerade that frequently present taxonomic challenges since the incredibly diverse and convincing forms are difficult to differentiate from one another. This subfamily contains the genus Mimetica, which is a perfect system for observing polymorphic variation of leaf morphologies due to its accuracy and precision in replicating leaf appearance, as well as its great diversity in shape and significant colour variation. In an experiment, Mikles and Song (2021) used digicam to determine whether the katydid species incisa, crenulata, viridifolia, mortuifolia, tuberata, and simoni within Mimetica display a distinct number of cryptically polymorphic forms versus continuous variation in order to clarify the species relationships within Mimetica as well as to develop a deeper understanding of the function and importance of polymorphs in insect populations and ecological phenomena. All species’ sexually dimorphic forms were strongly supported by TPS analysis, and no clear form differences were found in any incisa specimens, pointing to a more useful polymorph. In contrast, simoni and viridifolia specimens showed the most pronounced form differences, supporting the idea that this genus may contain different polymorphs but necessitating more in-depth morphometric investigations.

8.5.3

Odonate

Consider another possibility for sexual differences in colouring between species of the damselfly family Coenagrionidae, according to Sherratt and Forbes (2001).

8.5 Sexual Dimorphism

241

(Odonata: Zygoptera). Males of many of these species have distinct patterns and more vivid colours than females. However, they are non-territorial and avoid displays; in fact, male mate rivalry frequently resembles a scramble. They contend that even if females do exhibit some degree of mate preference, it is unlikely to be determined by colour or pattern. Instead, they contend that sex-related warning coloration is the primary way in which sexual dimorphism has developed in this group. First, we contend that it is almost certain that interactions between men will result in a minor cost to both parties. Then, we present some proof that guys are capable of interpreting colour as a cue to their sexual orientation. We demonstrate using a straightforward model that sexual dimorphism will easily emerge if these criteria hold. The model also demonstrates that males should have evolved significantly brighter coloration than females if females were selected to escape excessive pestering by males, as is sometimes hypothesised. Crowley and Johansson (2002) used an odonate-like parameter set to determine patterns of life history and behaviour to be anticipated in an odonate population in order to demonstrate this notion. The male-biased sex ratio was produced by the default parameter magnitudes, which led to males having smaller bodies and developing more quickly than females. There were significant effects on life-history characteristics depending on whether population growth was density dependent or density independent, as well as whether development time was set or variable. The model produced the following five broad hypotheses for odonate systems: 1. Male and female activity levels, growth rates, and mortality rates should differ more across sexes in animals with set developmental timeframes than they do in species with flexible life cycles. 2. In animals with set development durations, populations at high latitudes or high altitudes ought to be more active, emerge and reproduce at smaller sizes, and have a more skewed sex ratio towards men than populations at low latitudes and low altitudes. 3. For species with variable development durations, increased predation rates should result in higher activity levels and shorter development times in populations where mortality is activity-dependent and density-dependent. 4. For species with variable development times, activity levels should decline and development durations should lengthen in populations where density dependency is heavily mediated by mortality. 5. Males should be larger than females at emergence, and territorial species should have a more female-biased sex ratio at emergence than non-territorial species. Focused tests are obviously required because the actual evidence that already exists is typically scant and ambiguous regarding these expectations.

8.5.4

Anthocoridae

While Orius takaii, a Japanese Ryukyuan species only known from Okinawa Island, is similar in minuscule size and shape of paramere, this Ryukyuan species has a

242

8

Polymorphisms in Insect Predators

bigger body and slender antennal segment II and does not display sexual dimorphism in colours and size (Yasunaga 2000). Orius insidiosus females secrete both a non-volatile trail pheromone and a volatile sex pheromone. The benefit of females responding to the trail pheromone may be that this signal functions as a cue suggesting the chance of locating nearby prey, in contrast to how Orius insidiosus males most likely use it to discover suitable mates (Aldrich et al. 2007). The metathoracic smell gland secretion is clearly sexually dimorphic in Orius insidiosus and Orius sauteri (Uehara et al. 2019). The female abdomen is larger and has two copulatory tubes. When seen ventrally in males, the tip of the abdomen is slightly bent to the left. Both sexes exhibit a number of setae on the dorsal side. Orius nigromaritus displayed remarkable sexual dimorphism, including very small size and a shiny black body in the male, a bulbous antennal segment II, dark femora with the exception of a pale extreme apex of the metafemur, and generally slender paramere missing denticule (Yasunaga et al. 2019). The sexual dimorphism of Blaptostethus pallescens was obvious. Females possessed wider abdomens and larger bodies than males, as well as ventral copulatory tubes. In males, the abdomen was thin and had a little twist to the left. Male and female average lifespans were 52 and 40 days, respectively, according to Jose and Subramanian (2020).

8.5.5

Reduviids

In the Reduviidae, sexual dimorphism is common and can take a variety of forms, from subtle variations in body size, eye and ocellus size, or wing type to drastic dimorphism that involves wing loss or reduction in the female, which is occasionally accompanied by changes to the head, thorax, and legs. Ectrichodiinae, the millipede assassin bugs that are close cousins of Tribelocephalinae, frequently exhibit sexual dimorphism. The abundance and amount of sexual dimorphism in the Tribelocephalinae have not been well-documented because the bulk of species, mainly the male, were described from only one sex. A single apterous female specimen with small eyes served as the basis for the description of the tribelocephaline Enigmocephala deinorhyncha, which combines several uncommon or distinctive morphological features, including the very wide labium, apically bifurcating third visible labial segment, short and curved legs, and almost completely fused abdominal sternites (Rédei 2007). Given the resemblance in labium shape and the assumption that the male may be macropterous, this Borneo species could very well resemble members of the African genus Afrodecius. However, the male of this species is still unknown. The predatory adult Harpactor angulosus (Harpactorinae) showed sex dimorphism, with males being smaller than females (Pereira et al. 2009). Zelus adults show sexual dimorphism, with larger females (approximately 14 mm long and about 11 mm) than males (El-Tom 1965; Greenop 2020; Lahbib et al. 2022).

8.8 Future Directions

8.6

243

Mimic Polymorphism

Longicorn beetles (Braconidse) resemble Hymenopterous models so convincingly that seasoned collectors have documented their inability to tell them apart in flight. In tropical America, a black fossorial wasp has a propensity of fiercely flapping its wings while moving around on the ground. Two additional, distantly related insects, a reduviid (Hemiptera) and a Tettigoniid grasshopper, who share the wasp’s entirely black appearance, imitate this trait by engaging in a behaviour that is entirely unrelated to that of their relatives (Carpenter 1946).

8.7

General Reasons for Polymorphism

1. The presence of chemicals in different insect body parts demonstrates that longer chain, polymethyl-branched compounds (C35–C39, branched group) are primarily present on the antennae while shorter chain, unbranched compounds (C21– C31, linear group) predominate on the body, head, and wings. On the cuticle, a number of other substances, including as 1,3-pentacosadiene and oxygenated aliphatic molecules, are present in trace levels (Arsene et al. 2002). 2. Diet and seasons cause the morphism. 3. Another element that affects the polymorphism of many insects is population density. 4. High levels of a particular morph’s production are favoured by the long-term stability of the habitat’s heterogeneity and mate availability. 5. Flightlessness is more common in species that inhabit particular habitats (low altitudes) and is encouraged by both phenotypic and reproductive success. 6. When a certain morph predominates, the relationship between body size and fertility is positive. 7. Predators with long or short wings or wingless morphs can quickly adapt to varied settings, which is seen as ecological success.

8.8

Future Directions

1. Research on how predatory insect reproduction and flying ability are related is necessary. 2. Studies on body size, development, survival, and fertility in relation to predators with and without wings could be conducted. 3. Different predatory insects’ habitat stability and dispersal abilities could be studied. 4. In addition, more field research on this polymorphism has to be supported in order to better understand the laboratory discovery. 5. Although there is a paucity of solid experimental data, juvenile hormone (JH) and ecdysteroid (ES) govern the wing polymorphism in insects. Therefore, we recommend conducting numerous trials along this line.

244

8

Polymorphisms in Insect Predators

References Abram PK, Guerra-Grenier E, Després-Einspenner ML, Ito S, Wakamatsu K, Boivin G, Brodeur J (2015) An insect with selective control of egg coloration. Curr Biol 25:2007–2011. https://doi. org/10.1016/j.cub.2015.06.01 Ahmad M, Mofleh MM, Makhlouf H (2020) First report of polymorphism of Hippodamia variegata (Goeze) (Coleoptera: Coccinellidae) in the coastal area of Syria (Latakia Governorate). SJAR 7(3):449–453 Aisenberg A, Costa FG, González M (2011) Male sexual cannibalism in a sand-dwelling wolf spider with sex role reversal. Biol J Linn Soc 103:68–75. https://doi.org/10.1111/j.1095-8312. 2011.01631.x Aldrich JR, Oliver JE, Shifflet T, Smith CL, Dively GP (2007) Semiochemical investigations of the insidious flower bug Orius insidiosus Say. J Chem Ecol 33:1477–1493 Ambrose DP, Livingstone D (1987) Biology of Acanthaspis siva Distant, a polymorphic assassin bug (Insecta, Heteroptera, Reduviidae). Mitteilungen aus dem museum für Naturkunde in Berlin. Fortschr Zool 63(2):321–330 Ambrose DP, Livingstone D (1988) Polymorphism in Rhinocoris marginatus Fabricius (Insecta, Heteroptera, Reduviidae). Mitteilungen aus dem museum für Naturkunde in Berlin. Fortschr Zool 64(2):343–348 Ambrose DP, Livingstone D (1991) Biology of Neohaematorrhophus therasii (Hemiptera: Reduviidae). Colemania Insect Biosyst Ecol 1(1):5–9 Andrés JA, Sánchez-Guillén RA, Cordero Rivera A (2000) Molecular evidence for selection on female color polymorphism in the damselfly Ischnura graellsii. Evolution 54:2156–2161 Arsene C, Schulz S, Van Loon JJ (2002) Chemical polymorphism of the cuticular lipids of the cabbage white Pieris rapae. J Chem Ecol 28(12):2627–2631 Aspöck U, Aspöck H, Gruppe A (2020) Anchored between heaven and earth–a new flightless brown lacewing from Peru (Neuroptera, Hemerobiidae). Deut Entomol Z 67:141 Banham WMT (1990) Non-random mating in the polymorphic damselfly Ischnura elegans. PhD thesis, University of Manchester Barry KL, Holwell GI, Herberstein ME (2008) Female praying mantids use sexual cannibalism as a foraging strategy to increase fecundity. Behav Ecol 19:710–715. https://doi.org/10.1093/ beheco/arm156 Bestete LR, Torres JB, Henry CS (2019) Yellow mutant of the Neotropical green lacewing Chrysoperla externa: trait inheritance and predator performance. Entomol Exp Appl 167 (7):646–654 Brugnera R, Campos LA, Grazia J (2019) Morphology of immature stages, chromatic polymorphism of adults and natural history of Oplomus catena (Heteroptera: Pentatomidae: Asopinae). Neotrop Entomol 48(6):1046–1057 Busch AK, Wham BE, Tooker JF (2021) Life history, biology, and distribution of Pterostichus melanarius (Coleoptera: Carabidae) in North America. Environ Entomol 50(6):1257–1266 Canard M, Grimal (1988) Insect photoperiodism: various ways of regulating univoltinism in lacewings (Planipennia: Chrysopidae). Experientia 44:523–525 Carpenter GH (1946) Mimetic polymorphism. Nature 158(4008):277–279 Carpintero DL, Dellapé PM (2012) Neotropical scolopini (Heteroptera: Anthocoridae): new taxa, diagnostic characters and a key to the genera of the tribe. Acta Entomol Musei Nationalis Pragae 52(1):49–66 Chen X, Xiao D, Du X, Guo X, Zhang F, Desneux N, Wang S (2019a) The role of the dopamine melanin pathway in the ontogeny of elytral melanization in Harmonia axyridis. Front Physiol 10:1066 Chen X, Xiao D, Xiaoyan D, Zhang F, Zang L, Wang S (2019b) Impact of polymorphism and abiotic conditions on prey consumption by Harmonia axyridis. Entomol Gen 253:251–258 Cordero A (1990) The inheritance of female polymorphism in the damselfly Ischnura graellsii (Rambur) (Odonata: Coenagrionidae). Heredity 64:341–346

References

245

Cornelis M, Diez F, del Carmen CM (2016) Description of nymphs and additional information on Nabis ashworthi (Hemiptera: Heteroptera: Nabidae) from Patagonia, Argentina. AEMNP 56(1): 61–70 Cornelis M, Diez F, Coscaron MDC (2021) A new subgenus of Nabis Latreille (Heteroptera: Nabidae) from South America and a key to the subgenera worldwide, with notes about comparative morphology of the genitalia. An Acad Bras Cienc 93:4 Crowley PH, Johansson F (2002) Sexual dimorphism in Odonata: age, size, and sex ratio at emergence. Oikos 96(2):364–378 Duelli P (1992) Body coloration and color change in green lacewings (Neuroptera: Chrysopidae). In: Canard M, Aspock H, Mansell MW (eds) Current research in neuropterology. Sacco, Toulouse, pp 119–123 Duelli P, Johnson JB, Waldburger M, Henry CS (2014) A new look at adaptive body coloration and color change in “common green lacewings” of the genus Chrysoperla (Neuroptera: Chrysopidae). Ann Entomol Soc Am 107(2):382–388 El-Tom HA (1965) The biology of Zelus renardii Kolenati and Zelus socius Uhler (Hemiptera: Reduviidae). Ph.D. Thesis, University of Arizona, Tucson, AZ, USA Fairbairn DJ, Blanckenhorn WU, Székely T (eds) (2007) Sex, size and gender roles: evolutionary studies of sexual size dimorphism. Oxford University Press, Oxford Fincke OM, Jödicke R, Paulson DR, Schultz TD (2005) The evolution and frequency of female color morphs in Holarctic Odonata: why are male-like females typically the minority? Int J Odonatol 8(2):183–212 Ford EB (1940) Polymorphism and taxonomy. In: Huxley J (ed) The new systematics. Clarendon Press, Oxford, pp 493–513 Gamberale-Stille G, Johansen AI, Tullberg BS (2010) Change in protective coloration in the striated shieldbug Graphosoma lineatum (Heteroptera: Pentatomidae): predator avoidance and generalization among different life stages. Evol Ecol 24(2):423–432 Gao F, Zhao P, Webb M, Cai W (2015) A taxonomic review of the polymorphic assassin bug genus Rihirbus Stal (Hemiptera: Reduviidae: Harpactorinae). Zootaxa 3963(4):502–516 George PE (2000) Polymorphic adaptation in biology and life-table studies of Rhynocoris marginatus (fabricius) on Earias vittella (fabricius). J Biol Control 14(2):35–39 George PJ, Ambrose DP (2001) Polymorphic adaptive insecticidal resistance in Rhynocoris marginatus (Fabr.) (Het., Reduviidae) a non-target biocontrol agent. J Appl Entomol 125(4): 207–210 Greenop A (2020) Using species traits to understand the mechanisms driving pollination and pest control ecosystem services. Ph.D. Thesis, Lancaster University, Lancaster, UK Gutschick VP (1999) Biotic and abiotic consequences of differences in leaf structure. New Phytol 143:3–18 Hlaváček A, Daňková K, Benda D, Bogusch P, Hadrava J (2022) Batesian-Müllerian mimicry ring around the oriental hornet (Vespa orientalis). J Hymenopt Res 92:211–228 Howard RW, Stanley-Samuelson DW, Akre RD (1990) Biosynthesis and chemical mimicry of cuticular hydrocarbons from the obligate predator, Microdon albicomatus Novak (Diptera: Syrphidae) and its ant prey, Myrmica incompleta Provancher (Hymenoptera: Formicidae). J Kansas Entomol Soc 63:437–443 Howarth B, Clee C, Edmunds M (2000) The mimicry between British Syrphidae (Diptera) and aculeate Hymenoptera. Br J Entomol Nat Hist 13(1):1–39 Jose A, Subramanian M (2020) Biology of anthocorid predator, Blaptostethus pallescens Poppius (Heteroptera: Anthocoridae). Entomon 45(2):149–152 Joshi S, Poorten NVD, Sumanapala A, Nielsen E, Patel J, Nielsen B, Sherif M (2020) New records of polymorphism in Asian libellulid dragonflies (Insecta: Odonata). Int J Odonatol 23(4): 337–356 Jourdan H, Barbier R, Bernard J, Ferran A (1995) Antennal sensilla and sexual dimorphism of the adult ladybird beetle Semiadalia undecimnotata Schn. (Coleoptera: Coccinellidae). Int J Insect Morphol Embryol 24(3):307–322

246

8

Polymorphisms in Insect Predators

Kawakami Y, Yamazaki K, Ohashi K (2018) Effects of temperature on the expression of elytral colour polymorphism in the ladybird beetle, Menochilus sexmaculatus (Coleoptera: Coccinellidae). J Asia Pac Entomol 21(2):663–666 Kawakami Y, Yamazaki K, Ohashi K (2019) Intergenerational fluctuations in colour morph frequencies may maintain elytral polymorphisms in the ladybird beetle Cheilomenes sexmaculata (Coleoptera: Coccinellidae). Biol J Linn Soc 128(3):725–734 Kawakami Y, Yamazaki K, Ohashi K, Nakahama N (2021) Mitochondrial genetic structure reflects the geographical variation of elytral polymorphism frequency in Cheilomenes sexmaculata (Coleoptera: Coccinellidae). J Asia Pac Entomol 24(4):1270–1274 Krivokhatsky VA (2019) On variability of two Arabian myrmeleontoid lacewings (Neuroptera: Myrmeleontidae, Ascalaphidae). Entomol Rev 99(1):78–81 Lahbib N, Picciotti U, Sefa V, Boukhris-Bouhachem S, Porcelli F, Garganese F (2022) Zelus renardii Roaming in Southern Italy. Insects 13(2):158 Lattin JD (1989) Bionomics of the Nabidae. Annu Rev Entomol 34(1):383–400 Lattin JD (1999) Bionomics of the Anthocoridae. Annu Rev Entomol 44(1):207–231 Lietzenmayer LB, Goldstein LM, Pasche JM, Taylor LA (2022) Extreme natural size variation in both sexes of a sexually cannibalistic mantidfly. R Soc Open Sci 9(8):220544 Lin X, Lavine LC (2018) Endocrine regulation of a dispersal polymorphism in winged insects: a short review. Curr Opin Insect Sci 25:20–24 Mayer E (1969) Animal species and evolution. The Belknop Press of Hardwar University Press, Cambridge, pp 150–158 Mclver JD, Stonedahl G (1993) Myrmecomorphy: morphological and behavioral mimicry of ants. Annu Rev Entomol 38(1):351–377 Mikles B, Song H (2021) Examining tegmina polymorphism in leaf-masquerading katydids (Orthoptera: Tettigoniidae). Instars 6:2 Murakami Y, Hasegawa E, Watanabe S (2019) Effects of color morph on aggregation formation for hibernation in an extremely color polymorphic ladybug, Harmonia axyridis (Pallas) (Coleoptera: Coccinellidae). Entomol Ornithol Herpetol 8(219):2161–0983 Niimi T, Ando T (2021) Evo-devo of wing colour patterns in beetles. Curr Opin Genet Dev 69:97– 102 Okuda K, Chen Z (2021) Notes on the little-known assassin bug, Pygolampis breviptera ren (Heteroptera: Reduviidae: Stenopodainae). Jpn J Syst Entomol 27(1):141–144 Paleari LM (2013) Developmental biology, polymorphism and ecological aspects of Stiretrus decemguttatus (Hemiptera, Pentatomidae), an important predator of cassidine beetles. Rev Bras Entomol 57:75–83 Pereira AI, Zanuncio JC, Gil-Santana HR, Ramalho FS, Leite GL, Serrão JE (2009) Harpactor angulosus (Reduviidae: Harpactorinae), a predator of neotropical saturniids, Hylesia spp. in Brazil. Entomol News 120(2):206–212 Petchidurai G, Chitra R, Sahayaraj K (2019) Polymorphism of Rhynocoris marginatus (Fab.) (Heteroptera: Ruduviidae) on the biology, biological control potential. J Biopest 12(1):114–125 Prokop P, Václav R (2008) Seasonal aspects of sexual cannibalism in the praying mantis (Mantis religiosa). J Ethol 26:213–218 Putyatina TS, Gilev AV, Grinkov VG, Makarov AV (2022) Variation in the color pattern of the narrow-headed ant Formica exsecta (Hymenoptera: Formicidae) in European Russia. EJE 119 (1):58–68 Rédei D (2007) A new genus of tribelocephaline assassin bugs from Borneo (Hemiptera: Heteropteran: Reduviidae). Zootaxa 1465:47–53 Sahayaraj K (1991) Bioecology, ecophysiology and ethology of chosen predatory Hemipterans and their Potential in biological control (Insecta: Heteroptera: Reduviidae). Ph.D. Thesis, Madurai Kamaraj University, p 303 Sánchez-Guillén R, Van Gossum H, Cordero Rivera A (2005) Hybridization and the inheritance of female colour polymorphism in two ischnurid damselflies (Odonata: Coenagrionidae). Biol J Linn Soc 85(4):471–481

References

247

Saulich AK, Musolin DL (2009) Seasonal development and ecology of anthocorids (Heteroptera, Anthocoridae). Entomol Rev 89(5):501–528 Schütz D, Taborsky M (2005) Mate choice and sexual conflict in the size dimorphic water spider Argyroneta aquatica (Araneae, Argyronetidae). J Arachnol 33:767–775. https://doi.org/10. 1636/S03-56.1 Sherratt TN, Forbes MR (2001) Sexual differences in coloration of Coenagrionid damselflies (Odonata): a case of intraspecific aposematism? Anim Behav 62(4):653–660 Slater JA (1975) On the biology and zoogeography of Australian Lygaeidae (Hemiptera: Heteroptera) with special reference to the southwest fauna. Aust J Entomol 14(1):47–64 Sloggett JJ, Honěk A (2012) Genetic studies. In: Hodek I, Honek A, Van Emden HF (eds) Ecology and behaviour of the ladybird beetles (Coccinellidae). Wiley, Hoboken Stride GO (1956) On the mimetic association between certain species of Phonoctonus (Hemiptera, Reduviidae) and the Pyrrhocoridae. J Entomol Soc South Afr 19(1):12–28 Svenson GJ, Brannoch SK, Rodrigues HM, O’Hanlon JC, Wieland F (2016) Selection for predation, not female fecundity, explains sexual size dimorphism in the orchid mantises. Sci Rep 6(1): 1–9 Tauber MJ, Tauber CA (1971) An autosomal recessive (Neuroptera: Chrysopidae) mutantin a neuropteran. Can Entomol 103:906–907 Tauber CA, Tauber MJ, Nechols JR (1976) Yellow body, a sex-linked recessive mutant in Chrysopa. J Hered 67:119–120 Uehara T, Maeda T, Shimoda M, Fujiwara-Tsujii N, Yasui H (2019) Identification and characterization of the pheromones in the minute pirate bug Orius sauteri (Heteroptera: Anthocoridae). J Chem Ecol 45:811–817 Vennison SJ, Ambrose DP (1991) Biology and behavior of Ectomocoris xavierei Vennison & Ambrose 1990 (Heteroptera Reduviidae Piratinae). Trop Zool 4(2):251–258 Viana GG, Albuquerque GS (2009) Polymorphism in the pattern of integument markings of Ceraeochrysa caligata (Neuroptera: Chrysopidae) larvae and adults and redescription of the larval instars. Fortschr Zool 26:166–174 Wagner E (1952) Die europäischen Arten der Gattung Orius Wff. (Hem. Het. Anthocoridae). Notulae Entomol 32:22–59 Waldbauer GP (1970) Mimicry of hymenopteran antennae by Syrphidae. Psyche 77(1):45–49 Wygodzinsky PW (1966) A monograph of the Emesinae (Reduviidae, Hemiptera). Bull Am Mus Nat Hist 133:1–614 Yasunaga T (2000) A new species of the genus Orius Wolff (Heteroptera: Anthocoridae) found on Okinawa Island, Japan. Appl Entomol Zool 35:9–12 Yasunaga T, Yamada K, Duwal RK (2019) Three new species of the flower bug genus Orius (Hemiptera: Heteroptera: Anthocoridae) from Nepal. AEMNP 59(2):391–401

9

Influence of Ecological/Climatic Change

Contents 9.1 9.2

9.3

9.4

9.5 9.6

9.7 9.8 9.9

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Bioefficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.1 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.2 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.3 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.4 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.5 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.2.6 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.1 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.2 Damsel Bugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.3 Aleyrodidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.4 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.5 Miridae Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.6 Geocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.7 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.8 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.3.9 Thrips . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Macromolecular and Antioxidant Responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.4.1 Macromolecular Profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.4.2 Antioxidant Responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Temperature Tolerance Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.5.1 Cuticular Permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Cold Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.1 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.2 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.3 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.4 Geocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.5 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.6.6 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Wind Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Distribution at Field Level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Elevated Atmospheric CO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_9

250 250 251 252 254 256 257 259 261 262 264 265 266 267 268 271 273 274 275 275 276 277 279 279 280 280 282 282 283 283 285 287 288

249

250

9

Influence of Ecological/Climatic Change

9.10 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291

9.1

Introduction

Numerous ecosystem changes will result from global climate change. Large changes in species’ phenology and distribution are anticipated, and some creatures have already shown these changes. Climate change may have an impact on an individual’s performance, particularly for ectothermic creatures whose physiological processing rates are temperature-dependent. By changing animal interactions, such as interactions between predators and prey, climate change will also have an impact on ecosystems. Predation affects ecosystems in a number of significant and far-reaching ways, such as through the biocontrol of pest species, community composition, and the modulation of ecosystem functions like primary production or nutrient cycling. Therefore, it is crucial to comprehend how predator–prey interactions may change as a result of climate change. Understanding insect seasonal adaptations and extending the shelf life and shipping of predators employed in biological control programmes depend on it. Through a number of channels, climate change may have an impact on predator–prey relationships. 1. Climate change has the potential to directly impact species by altering their behaviour or geographic distribution. 2. The effects of climate change on consumptive (killing and departing) and/or non-consumptive (feeding patterns, morphology, or development rates) predator effects may alter predator–prey interactions. 3. Changing wind speeds and the impact of rising temperatures and CO2 on predator populations at different agricultural. 4. The effects of changing temperatures, humidity, photoperiod, and CO2 levels on the physiology, feeding, growth, and fecundity of predators. 5. The impact of climate change on species interactions may have an impact on communities and ecosystems. All of these factors are compiled here as biocontrol potential, development, survival, and fecundity (number of prey consumed/predator/day or per stage). The first portion focuses on predatory behaviour and the effectiveness of their biological regulation, while the second segment discusses biological features.

9.2

Bioefficiency

Abiotic variables have an impact on all living things. Poikilothermic creatures, including insects, are substantially more affected than other animals. Insect metabolism is directly influenced and accelerated by temperature. This will then have an impact on the rate of development, motility, predation, oviposition, and

9.2 Bioefficiency

251

morphogenesis. Each species has its own temperature maxima and/or requirements, and development varies along a temperature curve. In the wild, elements including plant phenology, light condition, temperature, humidity, wind velocity, and their interactions all have an impact on predation rates. Both the predator and the prey must be engaged in some form of activity for predation to take place; the predator must be hunting, and the prey must be vulnerable. Each activity times depend on the temperature (Logan et al. 2006). In this model, we assume that there is a predation-prone overlap between the microhabitat temperature intervals Tlo (estimated activity temperature range), Tm (microhabitat temperature), and Thi (temperature–humidity index). The lower and higher thresholds stand in for Tlo and Thi, or, in this case, the lower threshold and the lowest temperature at which prey are active. The maximum temperature at which the predator is active is represented by this. Of course, the prey and predator thresholds can be switched around. These presumptions can be changed to take into account temperature ranges where the predator and prey are active and accidental encounters occur (Logan et al. 2006). Given their propensity for feeding, insects’ tendency to move more when the temperature rises. Khan (2010) found that increased activity can enhance encounter rates with predators and minimise handling times (Khan 2010; Sentis et al. 2012). As the temperature rises, so does the predation rate. Due to the interaction of two or more creatures that are each separately impacted by the temperature increase, this change is a little more difficult to understand. As previously said, both predators and prey respond to an increase in temperature by being more active, so the predation rate may rise, fall, or remain the same depending on how the temperature affects the predator and prey in different ways (Kruse et al. 2008). Predator foraging ability typically grows more quickly than prey’s capacity to flee, increasing the rate of predation (Korenko et al. 2010). Models and laboratory experiments have been used to study how temperature affects the strength of the interactions between ladybeetles and aphids as predators and their victims (Sentis et al. 2012). The study’s findings show that predator searching declines in hotter weather (i.e. above the top threshold), and that predation rate is non-linear, dependent first on searching rate and then on handling time, and will decline as temperatures get closer to their thermal limits.

9.2.1

Dermaptera

Functional response studies are frequently used to assess the viability of predators as biocontrol agents. Temperature impacts trophic interactions, including linkages between insect pests and predators. We looked into how temperature affected the way an insect predator feeding on Plutella xylostella (Lepidoptera: Plutellidae) larvae, Euborellia annulipes (Dermaptera: Anisolabididae), responded functionally. Seven prey densities and three thermal conditions were used to evaluate the predation rate, type of functional response, attack rate (a′), handling time (Th), and maximal predation rate (T/Th) of the predator. Temperature affected the type of functional response that Euborellia annulipes had to Plutella xylostella, with type III

252

9

Influence of Ecological/Climatic Change

under lower temperatures (18 and 25 °C) and type II at 32 °C. As the thermal condition rose, we noticed rising a′ values at 25 and 32 °C, falling Th values, and the maximum T/Th. According to our research, Euborellia annulipes could be successful at controlling Plutella xylostella in a variety of thermal environments; however, its predation behaviour varies with temperature (da Silva Nunes et al. 2020). This study demonstrates how, after feeding on Plutella xylostella larvae (Lepidoptera: Plutellidae), Euborellia annulipes (Dermaptera: Anisolabididae) displayed a functional response that changed from Type III at low temperature (i.e. 18 °C) to Type II at higher temperature (i.e. 25 and 32 °C) (da Silva Nunes et al. 2020). When environmental changes occur, this type of Holling “disc” curve varies, indicating a shift in the biological control effectiveness of the predatory insects.

9.2.2

Miridae

On vegetable crops, species of the genus Macrolophus (Hemiptera: Miridae) are frequently utilised as biological control agents against pest insects like aphids and whiteflies. The primary biological pest management agent for aphids on tomato crops in central Greece is Macrolophus pygmaeus, which is also frequently found on eggplant and pepper plants. This predator can grow well at relatively low temperatures and can grow on vegetables with or without prey. The light phase had a major impact on the predation rate of Macrolophus pygmaeus, which was significantly greater in the dark. Temperature (20, 25, and 30 °C), life stage (each nymphal stage, male or female), and light condition (light or dark) all showed a significant three-way interaction. This was primarily due to the larger instars and adults being predated upon at much higher rates under the dark phase than the light phase [light or dark—8:16, 12:12, and 16:8 (L:D) h]. For pepper plants, a predation rate that was considerably higher in the dark than in the light was found in the fifth instar at 20 °C, in females at 25 °C, and in the third, fourth, and fifth instars as well as males and females at 30 °C. This was seen at a photoperiod of 8:16 h (L:D). Significant changes were seen for females at 20 °C, fifth instars and females at 25 °C, and third, fourth, and fifth instars plus males and females at 30 °C under the 12:12 h (L:D) photoperiod. Finally, for males and females at 25 °C, fourth and fifth instars, and males and females at 30 °C, predation rates were considerably greater in the dark phase than in the light phase at the photoperiod of 16:8 h (L:D) (Perdikis et al. 2004). The results of this interaction’s analysis show that: 1. In the dark phase, the predation rate on pepper plants was much higher than on eggplant, especially at 30 °C; and 2. Predation increased significantly more from 25 to 30 °C in the dark than in the light under photoperiods of 8:16 h (L:D), 12:12 h (L:D), and 16:8 h (L:D) (L:D). Under controlled laboratory conditions, the thermal needs of six dicyphine species—Dicyphus bolivari (two distinct strains), Dicyphus eckerleini, Dicyphus errans, Dicyphus flavoviridis, Nesidiocoris tenuis, and Macrolophus pygmaeus—

9.2 Bioefficiency

253

were determined. The developmental times at six temperatures (15–40 °C) were recorded using two experimental methods: one static and one dynamic. The latter involved figuring out the low- and high-temperature thresholds for movement. We divided the sample into two groups based on the outcomes of both methods: The species Dicyphus errans, Dicyphus errans, and Dicyphus flavoviridis were most active at low temperatures, while Nesidiocoris tenuis, Macrolophus pygmaeus, and Dicyphus bolivari performed best at high temperatures. Only two species, Dicyphus bolivari and Nesidiocoris tenuis, were able to mature at the constant temperature of 35 °C. Only two species of Dicyphus were still able to walk when it was below zero degrees—Dicyphus eckerleini and Dicyphus errans. The species that were more susceptible to higher temperatures were also more susceptible to lower temperatures. The studied species’ greater sizes appear to be more adapted to colder temperatures, whereas the smaller sizes appear to be better adapted to warmer temperatures. All species have males and females with different tolerances for heat and cold. In general, males were better acclimated to higher temperatures than females were to lower ones (Ingegno et al. 2021). Sparkes (2012) observed how predation, foraging behaviour, and longevity of Dicyphus hesperus (Heteroptera: Miridae) were affected by modest and large amplitude temperature fluctuations. According to the findings, temperature variations do have an impact on foraging behaviour and adult longevity, but the extent to which they have an impact depends on the kind of host plant. The findings of this study imply that temperature fluctuations should be taken into account in future studies since constant temperature models might not be accurate representations of low-amplitude fluctuations in nature or high-amplitude fluctuations that may come from climate change (Sparkes 2012). The experiment resulted in the death of 45 Dicyphus hesperus, which were not included in the analysis. The total amount of prey consumed was consistent across all temperature regimes. All time frames have an equal chance of experiencing predation in the low-amplitude variation. Both at 2100–0500 h and at 1310–2100 h, the likelihood of predation occurring in the high-amplitude fluctuation is equal. Nevertheless, differences between 0500–1300 and 1300–2100 and 2100–0500 and 0500–1300 h were noted (Sparkes 2012). The impact of temperature on two mirids, Cyrtorhinus lividipennis and Tytthus chinensis (Miridae), which are significant natural predators of planthoppers and leafhoppers in Asian rice fields, was recently examined by Bai et al. (2022). Tytthus chinensis and Cyrtorhinus lividipennis were studied by Bai et al. (2022) in order to better understand their thermal tolerance, fitness, predation potential, and transcriptome response in hot conditions. T. chinensis was more resistant of heat than Cyrtorhinus lividipennis and Nilaparvata lugens, its prey. In comparison to Cyrtorhinus lividipennis, Tytthus chinensis not only exhibited superior growth, survival, reproduction, and predation abilities, but it also shown stronger competitiveness when the two species co-existed in high-temperature environments. We sequenced their transcriptomes at various temperatures in order to comprehend the underlying mechanisms. The high co-regulation of heat shock protein (HSP) genes after heat treatment led to their identification and analysis. Comparing Tytthus chinensis to Cyrtorhinus lividipennis, quantitative polymerase chain reaction data

254

9

Influence of Ecological/Climatic Change

revealed that Tytthus chinensis promotes HSPs expression more quickly and powerfully over a wider temperature range in response to heat stress. Together, our findings shed light on Tytthus chinensis’ potential as a biological control agent under conditions of future global warming and revealed information about the thermal adaptation of mirid species (Bai et al. 2022).

9.2.3

Pentatomidae

A pentatomid predator named Eocanthecona furcellata’s growth and food consumption were observed in a laboratory setting at temperatures of 15, 20, 15, 30, and 35 ° C. The growth period was hastened by higher temperatures and the maximum temperature (20 °C), which was also accompanied by an increase in food consumption, shortening the time until adult maturity. Bad opposing consequences, a lower minimum temperature (15 °C), and halted egg development. The efficiency of capturing prey improved as temperature rose. Although females ingested more prey than did males, men killed a greater quantity of prey than did females (Rani 1992). Perillus bioculatus, however, appears to be primarily affected by the climate during hibernation (Péricart 2010) in order to survive outside of its native area. A few true bugs from warm climates have adapted to life in our nation (Seat et al. 2019, 2020), and recent information on the spread of Perillus bioculatus in Serbia supports the theory that the species is expanding due to climate change. Rabitsch (2008) had provided an explanation of this idea using the example of the real insect fauna of Central Europe, which is changing owing to habitat and climate. Functional response tests were used to assess the predatory prowess of Podisus maculiventris and Podisus nigrispinus on Spodoptera exigua caterpillars at three different temperatures (18, 23, and 27 °C). Prey density and predation capacity were related in both species, with higher predation rates as temperature and prey availability rose. The type II and type III functional responses, according to the results, suit the data for Podisus nigrispinus at 18 and 23 and 27 °C, respectively, the best. At 18 °C and higher temperatures, the data for Podisus maculiventris, however, fit type II and type III more accurately. The handling time decreased with rising temperature in both pentatomids. Predation rates were higher for Podisus nigrispinus than Podisus maculiventris at higher temperatures (Clercq 2001). The study also shows that as the temperature rose from 18 to 27 °C, Podisus maculiventris and Podisus nigrispinus (Hemiptera: Pentatomidae), predators of Spodoptera exigua (Lepidoptera: Noctuidae), reduced their handling time, increased their consumption rate, and changed their functional response type from Type II to Type III (Clercq 2001). In a laboratory setting, the effects of temperature on Eocanthecona furcellata, a predator of Spodoptera litura Fab. (Lepidoptera: Noctuidae), were examined at six constant temperatures from 20 to 35 °C and three gradient temperatures (20, 23, 26, 29, 32, and 35 °C). The findings demonstrated that the temperature range of 20–35 °C significantly increased the pace of development of the Eocanthecona furcellata. The pre-oviposition duration, oviposition period, and adult longevity all

9.2 Bioefficiency

255

fell significantly as temperatures rose. At 20 °C, the females had the longest longevity (36.72 days), whereas the males had the shortest (35.43 days). At 35 °C, males lived the shortest (13.07 days), and females for the shortest (14.27 days). Eocanthecona furcellata had the highest fecundity and hatchability at 29 °C, which were 271.47% and 99.82%, respectively. The intrinsic rate of natural increase (rm) and net reproduction rate (R0) of E. furcellata were both impacted by the temperature range (20–35 °C). The linear reduction in the mean generation time (T ) at 29 °C, the maximum values of R0 and rm and the lowest doubling times (DT) were, respectively, 77.88, 0.1223, and 5.6676 days (Peng et al. 2022). The results of this experiment, “Study on the Predatory Functional Response of Eocanthecona furcellata to S. litura”, revealed that the predation efficiency a/Th of the third instar nymphs of Eocanthecona furcellata against the 23rd instar larvae of Spodoptera litura reached its highest temperature at 35 °C, reaching 54.2863, followed by 49.0828 at 32 °C. The third instar nymphs of Eocanthecona furcellata had significant predation abilities at high temperatures, while their slow growth and feeble activity revealed little control at low temperatures, as indicated by the lowest temperature of 5.0650 at 20 °C. The findings indicated that temperature had a significant impact on the growth, survival, fertility, life-table parameters, and predation of Eocanthecona furcellata, with 29 °C being the ideal temperature for population growth (Peng et al. 2022). A predaceous pentatomid with the ability to manage a variety of insect pests is Arma chinensis. This study examined how Arma chinensis developed and survived at different stages at seven continuous temperatures (between 18 and 35 °C) while being fed yellow mealworms (Tenebrio molitor). Temperatures between 18 and 33 ° C (or 30 °C for eggs and nymphs in their first instar) had an inverse relationship with development times (measured in days) for the juvenile stage, the entire nymphal stage, and egg-to-adult development. The survival rate of Arma chinensis was found to be lowest at 18 °C (6.7%) and highest at temperatures between 21 and 24 °C (80–93.3%). The thermal constants for the egg, the complete nymph stage, and the egg-to-adult development were estimated to be 85.47, 334.9, and 423.8° days, respectively. The low-temperature thresholds for these stages were 14.3, 12.8, and 12.8 °C, respectively. The Taylor model provided the greatest fit for the egg data among the three non-linear models that were investigated, followed by the Briére1 model for the first instar nymph stage and the Lactin1 model for all subsequent instar stages, the entire nymphal stage, and overall development. Eggs, the entire nymphal stage, and egg-to-adult development had upper temperature thresholds predicted using the Lactin1 model of 38.57, 38.9, and 40.0 °C, respectively. The ideal temperature was determined to be 33.5 °C for the entire egg-to-adult period. The outcomes of this research can be applied to the mass breeding of this pest adversary and the creation of phenology models that depict its seasonal development (Xia et al. 2022).

256

9.2.4

9

Influence of Ecological/Climatic Change

Anthocoridae

Holling I and Flinn’s models were used to simulate the relationships of the predatory ability, prey density, and environmental temperature on the pepper Capsicum annuum in order to estimate the parameters of functional response and searching efficiency. This analysis focused on the predatory ability of female adults of Orius sauteri on female adults of the thrips Echinothrips americanus at different temperatures. The findings demonstrated that at 16, 19, 22, 25, 28, 31, and 34 °C, Orius sauteri consumed more prey as thrips density increased, whereas searching effectiveness dropped. Although Holling II gave a superior fit, both Holling II and Holling HI models simulated the functional responses at various temperatures. According to the Holling II model, Orius sauteri had the highest instantaneous attacking rate at 22 °C (1.132) and the quickest handling time on one prey at 34 ° C (0.007 day). According to the Holling III model, the lowest optimal searching density for Orius sauteri was at 19 °C (10.31 individuals), and the greatest optimal temperature for Orius sauteri to consume the most E. americanus was at 34 °C (52.67 individuals/day). Under Flinn’s model, which included two variables (temperature and prey density), the mean instantaneous attacking rate of Orius sauteri was 4.314 at any temperature. According to Flinn’s model, Orius sauteri has the greatest chance of consuming Echinothrips americanus with a prey density of 45 thrips at 34 °C. According to the findings, Orius sauteri was a powerful natural opponent for battling Echinothrips americanus (Zhu et al. 2015). In the Mashhad region of Iran, the predatory insect Anthocoris minki is a significant predator of the ash psyllid Psyllopsis repens (Psyllidae) and is widely distributed on ash trees (Fraxinus excelsior). Awareness a predator’s feeding behaviour requires an understanding of functional response parameters, handling time, and searching effectiveness. Different factors have an impact on the functional response’s kind and the rates of its characteristics. The functional reactions of Anthocoris minki to densities of Psyllopsis repens at various temperatures (15, 24, and 30 °C), prey instar (second and fourth instar stages), predator developmental stage (fourth nymphal stage and adult) and predator sex (male and female) were examined in this work. On the effectiveness of searching, the impacts of Anthocoris minki densities and interference were assessed. The nature of the functional response and its parameters were estimated using logistic regression and non-linear leastsquare regression, respectively. Nicholson’s model and linear regression were used to calculate the interference coefficient and per capita searching efficiency. Except for the trial involving adult Anthocoris minki and second-instar Psyllopsis repens nymphs of prey, which had a type III functional response, the results of all tests revealed a type II functional response. At 30 °C, both the highest attack rate and the fastest handling time were observed. There was no discernible variation in the per capita searching efficiency as densities rose. The laboratory findings indicate that Anthocoris minki may be able to control Psyllopsis repens on ash trees, but field testing is still required (Hassanzadeh-Avval et al. 2019).

9.2 Bioefficiency

9.2.5

257

Reduviidae

The temperatures have not changed the general reduviid predatory pattern, which involves arousal, approach, rostral probing (catching), injecting toxic saliva, paralysing, sucking (handling), and post-predatory activities. The measures utilised to assess the degree of biological control efficiency and bioefficacy of any natural enemies were approaching time (or attack rate) and handling time. Rhynocoris marginatus’s bioefficacy on Dysdercus cingulatus shows that depending on temperature, both predatory behaviour and bioefficacy change. For instance, third, fourth, and fifth nymphal instars of Rhynocoris marginatus as well as adults ingested 3.3, 3.2, 2.4, and 3.9 Dysdercus cingulatus, respectively, at room temperature. The predatory rate decreased while the reduviids were exposed to various temperature regimes (Fig. 9.1a), with the exception of Rhynocoris marginatus, a fifth instar, at 25 and 30 °C. When compared to the control category, it was statistically insignificant. Rhynocoris marginatus (third, fourth, and fifth nymphal instars) weight gain rose at 30 °C and at 25 °C in the fourth instar, despite the predatory rate being lower than the control. The timeframes for approaching and handling changed steadily from 10 to 35, respectively. At 35 °C, it was then significantly lowered. These predators favoured the abdomen part of Spodoptera litura and the ventral side of Dysdercus cingulatus (Fig. 9.1b).

Fig. 9.1 At varied constant temperatures (°C), the predatory rate of Rhynocoris marginatus life stages on Dysdercus cingulatus (a) and Spodoptera litura (b)

258

9

Influence of Ecological/Climatic Change

No. of Prey / day

15 10

Adult 5th Instar

5

4th Instar 3rd Instar

0 10

15

25 20 30 Temperature levels (OC)

35

RT

14

No. of Prey / day

12 10

5th Instar

8

Adult

6

4th Instar

4

3rd Instar

2 0 10

15

20 25 30 35 Temperature levels (°C)

RT

Fig. 9.2 Predatory rate of Rhynocoris fucipes on Dysdercus cingulatus (a) and Spodoptera litura (b) at different temperature levels (°C)

Rhynocoris fuscipes adults and fourth nymphal stages were given to Dysdercus cingulatus, respectively, and handling time was gradually increased up to 25 and 30 °C (Fig. 9.2a). It was evident from the figure that Rhynocoris fuscipes nymphal stages and adults took the longest to approach Spodoptera litura nymphal stages and adults at 30 and 35 °C, respectively (Fig. 9.2b). The handling of time showed a similar pattern. No matter what stage of life Rhynocoris fuscipes was in, it only ingested the most Dysdercus cingulatus at 30 °C. The largest amount of Dysdercus cingulatus was ingested by Rhynocoris marginatus life stages at 30 °C, with the exception of the fourth nymphal stages and particularly the fourth instar. On Spodoptera litura, adults made up the majority of all other life stages. Zelus renardii’s biological characteristics were examined in a lab setting at constant temperatures of 20, 25, 30, and 35 °C. The findings indicate that 25 °C produced the highest nymphal survival rate and the longest-living adult Zelus renardii (Ali and Watson 1978).

9.2 Bioefficiency

9.2.6

259

Coleoptera

The coccinellid Harmonia axyridis has both light-coloured aulica and dark-coloured nigra phenotypes. The latter group consists of the subgroups conspicua, spectabilis, and aulica, which are distinguished by melanic elytral patterns of various sizes and shapes. Variations are typically caused by seasonal and geographic factors. Coccinellids were fed a meal that included eggs from Ephestia kuehniella as well as the aphids Aphis fabae and Myzus persicae. At 10 and 30 °C, Harmonia axyridis larvae of the aulica phenotype consumed food at a considerably greater rate. Except at 20 °C, where there was no significant difference, the relative consumption rate of adults of the aulica phenotype was substantially higher than that of adults of the nigra phenotype. Both the relative consumption rate of nigra and aulica phenotypic larvae significantly rose between 10 and 20 °C and dramatically decreased between 25 and 30 °C. Although there were no significant differences between 25 and 30 °C, the relative consumption rate of adults with the aulica phenotype increased significantly between 10 and 30 °C. Between 10 and 25 °C and between 25 and 30 °C, adults’ relative consumption rates of the nigra phenotype both considerably rose and reduced. While the maximum tolerance limit for aulica phenotype larvae was higher than that for nigra phenotype larvae, the minimal tolerance limit for relative consumption rate by nigra phenotype larvae was lower than that of aulica phenotype. Both larvae and adults of the aulica phenotype had greater thermal amplitudes for relative consumption rates. The thermal optimum for the relative consumption rate of nigra and aulica larvae was the same, whereas the nigra phenotype’s thermal optimum for adults was 3.7 °C lower (Soares et al. 2003). Predators’ foraging behaviours are influenced by environmental factors, such as air temperature. The maximal prey consumption and hunting behaviour of Coccinella septempunctata at all developmental stages were observed in the current study in relation to the changing air temperature values. Instantaneous attack rate (a′) and prey handling time (Th), two crucial components of functional response, were found to be strongly correlated with temperature fluctuations. Up to a certain point, where it levels out and decreases to the point of “0” predation, the prey intake ratio per unit of time increased with increasing temperature. The foraging activities take place in a temperature range of 10–40 °C. But the highest predation rates were recorded between 20 and 23 °C and 23 and 25 °C. Regardless of the available prey densities, the lower temperature at which foraging behaviours entirely halted was between 10 and 12 °C. A high scale of co-relation parabolic (curvilinear) connection (r2 = 0.86–0.99 = 0.01) (Th) was discovered (Khan and Khan 2010). When tested in field enclosure trials, the carabid beetles Anchomenus dorsalis and Poecilus cupreus were found to decrease populations of oilseed rape insect pests emerging within winter rape crops. In comparison to the chamber’s actual temperature T1, Amara ovata, Harpalus distinguendus, and Poecilus cupreus killed a substantial amount of pollen beetle larvae at T2 (the temperatures rose by 3 °C) and T3 (temperatures increased by 5 °C). At T2, the killing of Anchomenus dorsalis noticeably more larvae than it did at T1, but Harpalus affinis did not exhibit any discernible variations. Amara ovata, Harpalus distinguendus, and Anchomenus

260

9

Influence of Ecological/Climatic Change

dorsalis all consumed considerably more biomass at T2 and T3 than they did at T1. Harpalus affinis and Poecilus cupreus did not significantly consume more biomass at different temperatures. Poecilus cupreus, one of the five carabids investigated, had the highest values for both the quantity of dead larvae and biomass consumption (Frank and Bramböck 2016). Planococcus citri and Ferrisia dasyrilii are two of the several species of mealybugs (Hemiptera: Pseudococcidae) that are preyed upon by the predatory Tenuisvalvae notata. While the effects of temperature change on insect growth and reproduction are extensively understood, the effects of any potential interactions between temperature and prey for predatory insects are not. Tenuisvalvae notata thermal needs and predation rates with both prey species were assessed at various constant temperatures. Under controlled temperatures between 18 and 38 °C, Tenuisvalvae notata larvae matured into adults. The maximum temperature for pupation was 33 and 34 °C for egg hatching, respectively. Adults reared at 32 °C failed to lay eggs and only lived for 1 week or less. Lower temperature thresholds or thermal constants for development from egg to adult were unaffected by prey species. Additionally, prey had no effect on a number of reproductive features, but temperature and prey interactions had an impact on changes in developmental times and oviposition rate with ageing. Temperature had an effect on the predatory rate of Tenuisvalvae notata, and adults generally ingested more nymphs of Planococcus citri and Ferrisia dasyrilii. These results show that Tenuisvalvae notata is well adapted to tropical and subtropical temperatures, and they suggest that it may be effective for the biological control of some native and non-native mealybugs (Ferreira et al. 2020). With a global distribution, Spodoptera litura (Lepidoptera: Noctuidae) is a significant pest of many economically significant crops. The primary management technique for it involves the use of insecticides, which has therefore resulted in insecticide resistance and control failures. This study used Spodoptera litura eggs as the prey and examined Harmonia axyridis (Coleoptera: Coccinellidae) functional response at larval and adult stages at varied temperatures ranging between 15 and 35 °C. Based on the results of the logistic model, the linear parameters of different predatory stages of Harmonia axyridis at different temperatures were considerably negative, which denotes a type II functional response. Across all predatory phases, the theoretical maximum number (T/Th) of eggs that might be devoured grew as temperature rose. According to the random predator equation, when the temperature rose, the coefficients for attack rate and handling time changed. When the target prey is being controlled biologically, often at warmer temperatures, the fourth instar and adult stages are the best possibilities. The females had the fastest handling times (0.54 h) and highest attack rates (0.189 h), whereas early instars showed lower values for these characteristics. These results demonstrate that adult predators and fourth instar predators are both effective egg consumers and potential suppressors of Spodoptera litura field populations (says Islam et al. 2020). Very recently Islam et al. (2022) also investigated impact of various temperature (15, 20, 25, and 30 °C), on predation and growth of Harmonia axyridis feeding on Acyrthosiphon pisum were studied. An examination of the predator’s two-sex, age-stage life table was

9.3 Biology

261

performed. To evaluate population growth and predation factors, computer forecasts based on bootstrap percentile confidence intervals were used. The development and predation of Harmonia axyridis varied with temperature. At 15 °C, development moved significantly more slowly than at higher temperatures. At 15 °C, the pre-adult stage’s mean duration was 41.25 days; however, at 20, 25, and 30 °C, it was 28.67, 18.35, and 13.23 days, respectively. At 15 °C, the intrinsic rate of rise was 0.0805 d1, while at 20, 25, and 30 °C, it rose to 0.1009, 0.1324, and 0.1813 d1. At 15 °C, the mean generation time (T ) was 71.96 days; at 20, 25, and 30 °C, it was 54.68, 42.64, and 29.96 days, respectively. With 743.68 eggs per female, 15 °C had the maximum fecundity, whereas 30 °C had the highest intrinsic rate of growth (r) and finite rate of increase. At 15, 20, 25, and 30 °C, respectively, the net predation rates (C0) were 4445.28, 4299.30, 3602.18, and 2624.20 aphids. The projected population and predation were inversely related to temperature. These findings can be used to model how the population of Harmonia axyridis will react to climate change and to modify IPM tactics. However, despite having better prey killing/handling skills at higher temperatures, predators Scymnus levaillanti, Adalia bipunctata, and Cycloneda sanguinea (Coleoptera: Coccinellidae) feeding on Aphis gossypii and Myzus persicae (Hemiptera: Aphididae) did not alter the type of This shows variable temperature impacts on prey–predator pairs, likely due to differences among species in their behaviour and susceptibility to temperature during hunting (Walker et al. 2020; Davidson et al. 2021).

9.3

Biology

In many insects, photoperiodic regulation of developmental time occurs. All components of development time tend to speed faster as temperature rises towards a higher threshold. A decrease in egg development, emergence time, and nymphal development results in an increase in stage progression and a decrease in adult longevity. Reproduction times lengthen as a result of the shorter nymphal (young one) stage, but population doubling times shortens (Logan et al. 2006). Development stops, reproduction stops, and if exposure lasts long enough, the insect will eventually perish once the temperature exceeds an upper threshold. Early studies suggested a number of potential causes, including: insufficient oxygen intake; acid waste buildup as a result of metabolic processes; and increased water loss due to structural changes in the lipids in the cuticle. According to recent studies, denatured proteins, changed enzyme and membrane structure, and extremely high temperatures can kill cells. Additionally, a concentration of denatured proteins causes the formation of heat shock proteins, which prevents the synthesis of new proteins (Gullan and Cranston 2005). By extrapolating the regression line and noting where it crossed the temperature axis, the developmental zero (DZ) for each instar was discovered (i.e. DZ = -ai/bi). The reciprocal of the slope (bi) of the fitted regression line, i.e., DD = 1/24 bi, was used to calculate the degree-days (DD) above the DZ necessary to

262

9

Influence of Ecological/Climatic Change

complete each stage of development. The formulas for determining the mistakes are related to the DZ parameters. The following factors relating to fecundity were noted: • The number of days between emergence and the first oviposition is the pre-oviposition period. • Oviposition period is the sum of the days between each oviposition. • The number of days between the last oviposition and mortality is the postoviposition period. • Daily oviposition is the quantity of eggs laid by each female each day. • Total fecundity is the sum of the eggs deposited by each female; • The proportion of hatched eggs equals the per cent egg viability. One of the most significant abiotic elements influencing insect survival, development rate, abundance, behaviour, and fitness is ambient temperature. Each insect species does, in fact, have a preferred temperature range as well as lower and upper growth thresholds. Insects’ behaviour, sex ratio, and population dispersal can all be impacted by low temperatures, but fertility, hatching, and survival of these animals can be negatively impacted by hot temperatures. This section examines how various environmental factors affect the biological characteristics of predators that belong to distinct insect orders such as dermapter, nabidae.

9.3.1

Dermaptera

Temperature controls a predator’s embryonic development in addition to monitoring its eating behaviour. Intentionally, the Forficula auricularia fertilised egg’s shape, ultrastructure, and tolerance to low and high temperatures were examined both before and after water absorption. The water absorption-induced volume expansion had no effect on the envelopes’ ultrastructure. The chorion was discovered to consist of five different layers and a vitelline envelope beneath which the embryo secreted a serosal cuticle. The egg’s supercooling point was 27 °C before hydration; after hydration, it was 22 °C. The egg’s heat resistance was unaffected by hydration. It survived a temperature increase of 36 °C without obvious harm. When subjected to temperature fluctuations, eggs’ morphological and physiological alterations brought on by water absorption did not affect the viability of the eggs (Chauvin et al. 1991). Temperature had a substantial impact on both the total number of eggs a Euborellia annulipes female laid and the number of eggs she laid each day over the course of her life. Egg production was about 1.5 times higher at 30 °C compared to 25 °C. The observations of other researchers who have also observed that the fertility of Euborellia annulipes fluctuates with temperature are lower than those from our study. The lifetime fecundity of Euborellia annulipes was estimated to be 186 and 91.4 eggs, respectively, by Neiswander (1944) and Koppenhöfer (1995). According to Melamed Madjar (1971), Euborellia annulipes oviposited much more eggs at high temperatures (24–34 °C) than at low temperatures (10–20 °C), with most eggs

9.3 Biology

263

being laid between 24 and 30 °C. The Euborellia annulipes produced 90 eggs on average per female. Only a small number of females—up to 250—deposited more eggs than the bulk (85%), which normally lay 0–150 eggs (Melamed Madjar 1971). The total number of eggs laid by Euborellia cincticollis in the same year ranged from 47.2 at 26.72 °C to 83.7 eggs at 22.22 °C, according to Knabke and Grigarick (1971). The earwig Euborellia annulipes, however, showed higher fecundity when compared to other species in the same genus, producing an average of 321 eggs over the course of its lifespan (Situmorang and Gabriel 1988). Females reared at 30 °C had considerably more eggs per clutch and clutches per female per day than those grown at 25 °C. The interval between clutches of Euborellia annulipes was noticeably greater when the predator was elevated at 25 °C. Neiswander (1944) and Koppenhöfer (1995) observed an average of 32.0 eggs per clutch at 25 °C and 35.1 eggs per clutch at 30 °C from a Euborellia annulipes nest. However, Bharadwaj (1966) and Rankin et al. (1995) found mean numbers of eggs per clutch of 52.7 and 62.8, respectively, which is different from what we found in our study. Neiswander (1944) claimed that the egg production of Euborellia annulipes peaked earlier in the oviposition stage and tended to drop in succeeding clutches. Female Euborellia annulipes oviposited several clutches frequently at intervals of 20 days at 28 °C, according to Rankin et al. (1995), which is comparable with the results of our experiment at 25 °C. Female Euborellia annulipes clutch sizes did not differ appreciably between 25 and 30 °C. According to Neiswander, Euborellia annulipes can lay anywhere between 2 and 7 clutches, with an average of 3.9 clutches per female (1944). Similar research by Bharadwaj (1966) revealed that this species’ females give birth to four clutches during the course of their lifetime. Knabke and Grigarick (1971) found that Euborellia annulipes adults only produce one or two clutches in the lab, but Euborellia cincticollis adults generate one to eight clutches. The biology of the striped earwig Labidura riparia was studied on the third instar fall armyworm Spodoptera frugiperda at three constant temperatures and humidity levels. Striped earwigs swallowed more larvae at 30 °C than they did at 25 or 35 °C. Humidity has no effect on feeding rate. Rogers’ random predator equation outlined how prey density affects eating at 30 °C and 80% RH. In comparison to handling rates of 22.1 and 20 preys per day, male and female search rates were 1127 and 4355 cm2 each day. Both sexes consumed more prey than the other. The striped earwig looks to be an excellent biological control agent for arthropod pests in row crops (Kharboutli and Mack 1993a, b). Euborellia annulipes has recently produced an average of 2.6 and 4.5 clutches per female (Koppenhöfer 1995; Rankin et al. 1995). These values, which were lower than those found in our study, were obtained by other researchers. This shows that factors besides temperature, examples include the predator’s geographic origin, rearing conditions, and diet, may affect some reproductive parameters of this organism, particularly the number of eggs per clutch and the number of clutches per female. Although there are no research on how nutrition affects Euborellia annulipes reproduction. According to Smith and Jones (1998) and Lemos et al. (2003), diet affects the number of eggs laid per clutch by the dermapteran predator Doru taeniatum.

264

9

Influence of Ecological/Climatic Change

There is little information known on the ecology, and biology of dermapteran predators. In order to use Euborellia annulipes in biological control programmes against the cotton boll weevil Anthonomus grandis, it is imperative to comprehend the effects of temperature on the reproductive features of mature females of this predator (Curculionidae). The goal of the study was to test the reproductive capability of female Euborellia annulipes using artificial meals in a lab setting at 25 and 30 ° C. Around 25 and 30 °C, fertility started to diminish around the 84th and 74th days of adulthood, respectively, and stopped with the death of the females in both cases. At 25 and 30 °C, respectively, Euborellia annulipes females produced an average of 206.2 and 306.0 eggs, and they lived an average of 198.4 and 149.1 days. A model that included an exponential function for the following decline in egg laying at older ages with a linear function for the increase in fecundity at young ages was used to characterise the link between age-specific fecundity and age. The formula f(x) = x exp.(-x) was used to describe this relationship. In the model, x is the age in days (age class), f(x) is the daily age-specific fecundity rate (eggs/female/day), and are constants. The model was weighted by the proportion of females contributing to the means, generating the curves, and then matched to the data using the non-linear least square technique. The parameters were predicted to be at 25 °C (1.599 and 0.356) and at 30 °C, respectively (3.226 and 0.415). At 25 and 30 °C, the model correctly predicted 78.8 and 74.8% of the variation, respectively. At both temperatures, the age-specific fecundity pattern was consistent (Nunes et al. 2020). Furthermore, Nunes et al. (2022) revealed that temperature, a critical mediator of pest population size, may have an impact on the dynamics of the predator–prey interaction. This investigation aimed to assess how female Euborellia annulipes feeding preferences were influenced by the temperature and developmental stage of Plutella xylostella individuals. Larvae and pupae were evaluated along with three temperatures—18, 25, and 32 °C—and two developmental stages—with or without choice—relative to prey capture. Female ring-legged earwigs displayed a preference for feeding larvae over pupae regardless of the temperature. Only for larvae did the temperature and prey selection conditions have an impact on the quantity of prey consumed (not for pupae). In both prey capture scenarios, 18 °C showed the lowest larval consumption. Euborellia annulipes is particularly effective at suppressing Plutella xylostella under a variety of thermal circumstances and its predation behaviour varies with temperature from 25 to 32 °C, according to da Silva et al. (2020).

9.3.2

Damsel Bugs

According to past studies, Nabis americoferus had a mean egg development time of 10 days at 24 °C, which was the same temperature as the rearing and life history (says Braman and Yeargan 1988). Depending on the climate, Nabis capsiformis nymph development takes 3–4 weeks. Soon after hatching, they become active and start eating. The nymphs have a smaller, wingless appearance than the adults. At 26 to 28 °C, male Nabis capsiformis nymph development lasts an average of 18 days

9.3 Biology

265

compared to 22.4 days for females (Hormchan et al. 1976). In Florida, the majority of adults or late instars spend the winter in weeds, ground cover, or winter vegetables like potatoes and maize. The life histories of the adults show that it is a multivoltine insect. Adult Nabis capsiformis can survive for 14.9 (males) or 21.6 (females) days under laboratory conditions (26–28 °C, 60–70% RH, and a 15:9 L:D photoperiod), with a generation period of roughly 51.6 days. Adult Nabis spp. live for over 60 days, according to the report of Rebolledo et al. (2005). The predator Nabis consimilis life cycle was identified and contrasted across the year’s seasons. In a laboratory setting, the studies were carried out in the summer (26.10.5 °C; HR 51.40.9) and the winter (20.41.0 °C; HR 60.32, 7). In comparison to winter, summer had higher levels of development, oviposition, post-oviposition, and oviposition capacity. For the I, II, III, IV, and V instars, the incubation period, viability, and length were 7.9 days, 41.7%, 4.6 days, 3.61, 7 days, 3.2 days, 2.9 days, and 4.1 days, respectively, in the summer, and 13 days, 48.2%, 6 days, 5.2 days, 4.8 days, 5.3 days, and 6.4 days, respectively, in the winter. The largest death rates are seen in the summer in the II and III instars and in the winter in the IV and V instars. In comparison to winter, summer had a higher overall death rate in the nymphal stages. In both seasons, there was a similar sex ratio (Salcedo et al. 2020). At 16, 18, or 38 °C, Tenuisvalvae notata eggs did not hatch. Only eggs incubated at 33 or 34 °C hatched (100% when females were fed Ferrisia dasyrilii or), regardless of whether the temperature was kept constant at 33, 34, 35, or 36 °C. At 34 °C, larval survival was decreased to roughly 40%. At 35 °C, no larvae developed into adults. Additionally, adults raised at 33, 34, 35, and 36 °C that were collected at 25 °C failed to lay eggs and perished within a week. Tenuisvalvae notata juveniles, on the other hand, matured into adults and multiplied when raised at 20, 22, 25, 28, or 32 °C while being fed various mealybug species. Additionally, all juvenile stages demonstrated accelerated growth (1/D) in response to rising temperatures. Mean egg development times ranged from 10.1 days at 20 °C to 4.2 days at 32 °C when fed with Ferrisia dasyrilii (Ferreira et al. 2020).

9.3.3

Aleyrodidae

In areas where Bemisia tabaci (Coleoptera: Coccinellidae) is abundant, the vital predator Axinoscymnus cardilobus (Homopteran: Aleyrodidae) can be found. Arthropod development has been discovered to be significantly influenced by temperature among other things. Seven temperature regimes with constant value (14, 17, 20, 23, 26, 29, and 32 °C) were used to study the effect of temperature on the emergence of Axinoscymnus cardilobus. The findings revealed that increased temperature had a substantial impact on the length of the egg, larval, and pupal stages. From 14 to 26 °C, the pace of development gradually grew; but, from 26 to 32 °C, it slowed down. At temperatures between 20 and 26 °C, the survival rates of several insect stages were steady, but at extreme temperatures between 32 and 14 °C, a significant decline was visible. The female’s ovipositional period shortened when temperatures rose from 17 to 32 °C. At 23 °C, the female’s greatest fecundity (225.7

266

9

Influence of Ecological/Climatic Change

eggs per female) was noted. Based on the findings of experiments conducted at temperatures between 14 and 32 °C, life tables for Axinoscymnus cardilobus were created. At 23 °C, the highest values of the reproductive rate (R0), the intrinsic capacity for increase (rm), and the finite rate of increase were 70.7, 0.059, and 1.062, respectively. The maximum and lowest values found at temperatures of 17 and 32 ° C, respectively, were 112.7 and 38.7, while the temperature increased from 17 to 32 °C, the mean generation time (T ) decreased. These findings provide important information about the introduction of Axinoscymnus cardilobus into new settings with a variety of temperature regimes (Huang et al. 2008).

9.3.4

Anthocoridae

Orius insidiosus and Orius majusculus developmental durations (Fig. 9.3) varied significantly across photoperiods at 18 °C, but no discernible trend could be seen (Van den Meiracker 1994). During the 12L:12D treatment, the temperature was 17 ° C as opposed to 18 °C during experiment 1. A 4- to 5-day delay in growth was most likely the outcome of this inadvertent 1 °C temperature decrease. Although the differences in developmental timing across the other photoperiods were substantially less, (much smaller) variations in temperature may have also been a factor. Males in both species developed more quickly than females. Ruberson et al. (1991) observed no direct association between photoperiod and developmental time, but found that nymphs of Orius insidiosus developed more quickly at L10:D14 than at longer day lengths (at 20 °C). Askari and Stern (1972) discovered that Orius tristicolor development occurred more quickly at L12:D12 than at L16:D8 (at 25.5 °C). Other Heteroptera, such as Palomena angulosa (Hori 1986, 1987, 1988) and Eysarcoris

Fig. 9.3 Developmental time of non-diapausing Orius insidiosus and Orius majusculus at 18 °C and different photoperiods (from oviposition to adult eclosion)

9.3 Biology

267

lewisi have similarly shown an accelerated effect of short-day lengths on developmental time (Hori and Inamura 1991). This might be a result of a compromise between building up metabolic reserves and reaching the diapausing stage before the commencement of unfavourable conditions (for reproduction or hibernation). Numerous Heteroptera, such as Dolycoris baccarum (Conradi-Larsen and Somme 1973) and Carbula humerigera, which experiences nymphal diapause, have been shown to develop more slowly under short day lengths (Kiritani 1985). In Pyrrhocoris apterus, nymphal development was dramatically slowed down at photoperiods around the critical one for inducing diapause (Saunders 1983). Elatophilus nigricornis, a Hemiptera Anthocoridae predator of Matsucoccus feytaudi, lays 3–139 eggs (average 57), at 13–20 °C, 16 h of light, 15–48 eggs (average 29), at 25 °C, 16 h of light, with four couples, and 2–11 eggs (average 57) at 25 °C, 16 h of light (Fabre et al. 2000).

9.3.5

Miridae Biology

The development and survival of Macrolophus pygmaeus (Hemiptera: Miridae) nymphs were studied on a range of host plants, with and without a range of insect food, and on bee pollen and pollen from Ecballium elaterium (Cucurbitaceae) in different combination. It was also investigated how temperature affected the growth and demise of Macrolophus pygmaeus nymphs. Depending on the experiment, experiments were carried out in temperature cabinets maintained at consistent temperatures, 16L:8D h photoperiod, and 65.5% RH. The research shows that Macrolophus pygmaeus can effectively complete its development on tomato, eggplant, cucumber, pepper, and green beans in the absence of insect prey. The shortest times of nymphal development on eggplant were observed for Myzus persicae, Macrosiphum euphorbiae, Aphis gossypii, and Tetranychus urticae, trailed by Trialeurodes vaporariorum. Although Macrolophus pygmaeus nymph mortality was somewhat higher when prey was absent from different host plants than when it was present, this mortality was not thought to be a barrier to the establishment of the predator. Even in the absence of prey, Macrolophus pygmaeus developed on tomato under a range of temperatures from 15 to 30 °C, with 30 °C being the ideal temperature for development. Separate offerings of bee pollen and pollen from Ecballium elaterium were sufficient for the effective development and survival of predator nymphs. When bee pollen was consumed along with other food sources like eggplant leaves and M. persicae, it significantly aided in the development and survival of the nymphs (Perdikis and Lykouressis 2000). The predatory mirid Macrolophus praeclarus is widespread throughout the Americas and is known to prey on a variety of horticultural pest species. The developmental time from egg to adult on tomato was examined using Ephestia kuehniella at five different temperatures: 20, 25, 30, 33, and 35 °C. At 20 °C, the duration took the longest (56.3 days), and at 33 °C, the shortest (22.7 days). The number of nymphs that were able to develop into adults decreased as the temperature climbed, with the highest proportion occurring at 20 °C (78.0%) and the lowest

268

9

Influence of Ecological/Climatic Change

percentage occurring at 35 °C (0%). The lowest and highest developmental thresholds, according to estimations, are 11.2 and 35.3 °C, respectively. The maximum rate of development took place at 31.7 °C, with a thermal constant of 454.0 degree-days. Ephestia kuehniella eggs were most commonly preyed upon at 30 °C. In Y-tube olfactory choice tests, Macrolophus praeclarus preferred the odours of tomatoes, sweet peppers, and eggplant more frequently than the no plant control treatment. Macrolophus praeclarus, a zoophytophagous mirid, did not cause necrotic rings on tomato plants as did Nesidiocoris tenuis, another zoophytophagous mirid. Macrolophus praeclarus phytophagy boosted defensive responses in tomato plants by activating the jasmonic acid metabolic pathway (Pérez-Hedo et al. 2021). A significant biocontrol agent of several significant arthropod pests, such as the South American tomato pinworm, Tuta absoluta, is Nesidiocoris tenuis (Hemiptera: Miridae). An experiment with a full factorial design was conducted to examine the impact of ten distinct kinds and 15, 20, 25, 30, and 35 °C constant temperatures on the plant damage and growth of Nesidiocoris tenuis. The number of adults that moulted as a result of temperature (15 °C–0.529, 20 °C–1.147, 25 °C–1.730, 30 °C– 1.714, and 35 °C–1.79) increased with temperature increase (Siscaro et al. 2019). Nesidiocoris tenuis nymphs’ developmental periods were examined at temperatures of 15, 20, 25, 30, and 35 °C, with 76.9% relative humidity and a 12:12 (L:D) h photoperiod (Sanyo, MLR-350). Nymphs’ whole developmental period lasted an average of 49.4 (0.25) days at 15 °C and 10.50 (0.07) days at 30 °C. Temperature has no discernible impact on the sex ratio. Nymphs’ overall survival rates varied from 94 to 100% and were independent of temperature. Temperatures between 15 and 30 °C had little to no impact on instar survival, whereas 35 °C considerably lowered it. The longest female lifespan was 49 days, and the mean lifespan was 27.12 (1.58 days). The number of nymphs that hatched per female, or fecundity, varied from 8 to 142. The average number of total offspring produced by each female, which was 65.0, increased significantly throughout the course of the week (5.89). However, there were no discernible variations between the first 3 weeks. The amount of nymphs that hatched during the first 3 weeks, which made up 81.9% of the total, varied greatly from week to week. Compared to weeks 5 and 6, week 4 was notably different. On week 7, there was just one female laying eggs (Marquereau et al. 2022).

9.3.6

Geocoridae

According to Cohen (1982), Lygus hesperus (Hemiptera: Miridae), the predator, was less tolerant of heat than its victim, Geocoris punctipes (Hemiptera: Lygaeidae), and the prey was more heat-resistant than the predator at preventing water loss. No specific research has been done on the role of relative humidity in the development or hatching of big-eyed bugs, although research on other predatory heteropteran insect species supports a value of 40–60% RH as the optimal environment (Javahery 1994). In Australian agricultural crops, big-eyed bugs called Geocoris spp. (Hemiptera: Geocoridae) are a frequent predator. The effects of nutrition, temperature, and

9.3 Biology

269

Fig. 9.4 Geocoris atricolor, Geocoris pallens, and Geocoris punctipes’ developmental days to temperature

photoperiod on the growth and survival of Geocoris lubra from egg to adult were examined by Mansfield et al. in 2007. Geocoris lubra nymphal survival was extremely low when raised on live aphids (Aphis gossypii), although it marginally increased when fed Helicoverpa armigera eggs. When compared to 25 °C, development was accelerated and nymphal survival significantly increased at 27 °C. Further research at 27 °C revealed photoperiod affected immature Geocoris lubra development time but not survival. At 10L:14D, development took noticeably longer. The photoperiod had no effect on the fecundity of first-generation Geocoris lubra, but egg viability was higher at 12L:12D (Mansfield et al. 2007). It was concluded that short days (light and dark periods in hours: 10:14) slow nymphal development, although photoperiod has no bearing on nymph survival (Mansfield et al. 2007). Different stable temperatures have an impact on the growth of eggs and nymphs, fecundity, and fertility in Geocoris atricolor, Geocoris pallens, and Geocoris punctipes. At 37.8 °C, all species’ eggs and nymphs were unable to finish their development. With rising temperatures, the egg incubation period grew shorter for Geocoris pallens and grew longer for Geocoris punctipes at each temperature examined. With rising temperatures, the length of the nymphal stages likewise got shorter, although at 23.9 °C, Geocoris punctipes developed the fastest and Geocoris pallens the slowest. At the other temperatures under investigation, nymphal stages lasted for the same amount of time after eggs (Geocoris pallens, Geocoris atricolor, and Geocoris punctipes). In Geocoris atricolor and Geocoris punctipes, oviposition and fertility peaked around 32.2 °C. At 35.0 °C, Geocoris pallens oviposition reached its peak. In contrast to Geocoris punctipes, females of Geocoris atricolor and G. pallens oviposited at 35.0 °C. Only female Geocoris punctipes oviposited at 23.9 °C (Dunbar and Bacon 1972). Total duration of development is in favour of 23.9 °C (Fig. 9.4). When grown on Ephestia kuehniella eggs at 20, 24, 26, 30, 33, or 36 °C, Geocoris varius and Geocoris proteus growths and reproduction. The thermal

270

9

Influence of Ecological/Climatic Change

constants (K ) and developmental thresholds (T0) of Geocoris varius eggs and nymphs were lower, at 13.3 °C and 151.1 degree-days and 13.4 °C and 433.0 degree-days, respectively. The thermal constants of Geocoris varius were 16.1 °C, 98.3 degree-days, and 16.9 °C, 226.9 degree-days, respectively. Geocoris varius eggs hatched far less frequently at 33 °C than they did at 30 °C, and none did so at 36 °C. It was lowest at 20 °C and did not significantly drop at 36 °C. The survival rate of Geocoris varius increased with temperature up to 30 °C, while in Geocoris proteus it was lowest at 20 °C. Therefore, it seems that 24–30 °C and 26–33 °C, respectively, are the best rearing temperatures for juvenile stages of both species. It might be possible to improve the efficiency of their mass production by controlling the rearing temperature in the aforementioned ranges. This would also standardise the nymphs’ developmental stages and prevent cannibalism while mass-raising them (Oida and Kadono 2012). Authors conduced that nymphs’ survival and growth rates are favourably connected with temperature increases up to a certain threshold (Oida and Kadono 2012). In 2014, Calixto and colleagues assessed the thermal requirements of the species by rearing Geocoris punctipes in climate chambers at alternating (21/11 °C, 24/18 ° C, 27/21 °C, and 30/26 °C) and constant (16.8, 21.5, 24.5, and 28.3 °C), RH 70 10%, and a 14-h photophase. The levels of survival and growth in Geocoris punctipes were the same when they were grown in both constant and varying temperatures. Five instars were observed for all temperature ranges. The duration of the egg stage and each instar, as well as that of overall larval growth, were longer and larval survival was poorer when reared at 16.8 °C, 21/11 °C, 21.5 °C, and 24/18 °C compared to 24.5 °C, 27/21 °C, 28.3 °C, and 30/26 °C. For growth and survival, Geocoris punctipes favours a temperature range of 24.5 to 30 °C. Note that 13.5 °C is its lower development threshold temperature, and 295.9 degree-days is its thermal constant. Sex ratios did not significantly differ from the male:female ratio of 1:1 in any temperature regime. Indicating that this predator will function well in crops where this pest is present, the temperature regimes at which the prey Tuta absoluta and the predator Geocoris punctipes are active are pretty comparable (Calixto et al. 2014). Bueno et al. (2016) evaluated the everyday and overall fecundity, lifespan, and life-table parameters of Geocoris punctipes under stable temperatures (16.8, 21.5, 24.5, and 28.3 °C) and proportionally varied (day/night) temperatures (21/11 °C, 24/18 °C, and 27/21 °C). Pairs of adult predators maintained at the aforementioned temperatures were descended from nymphs exposed to the same temperature regimes when they were 24 h old in Petri plates containing Anagasta kuehniella eggs and an oviposition substrate. Tests were conducted in climatic rooms with a range of temperature regimes, a RH of 70%, and a photoperiod of 14L:10D. When temperatures were low (16.8 °C, 21/11 °C, 21.5 °C, 24/18 °C) or high (24.5 °C, 27/21 °C, 28.3 °C, 30/26 °C), there were noticeable differences between treatments. Temperature also had a significant impact on reproduction, longevity, and life-table characteristics. Average temperatures between 24.5 and 30 °C were the most conducive to Geocoris punctipes’ highest reproduction rates and fastest population

9.3 Biology

271

increase, but temperature changes within any of the temperature regimes had no negative effects on either of these processes (Bueno et al. 2016).

9.3.7

Reduviidae

According to a research by Ali and Watson (1978), the ideal temperature range for Zelus renardii nymph development is between 25 and 30 °C. Additionally, these authors noted that extremely high and low temperatures greatly shorten the time needed to complete each larval stage, increase nymphal mortality, and cause many adults to become sterile and pass away shortly after emerging at 20 and 35 °C, respectively. The evolution of this species may also be influenced by other factors, such photoperiod (Ali and Watson 1978). By raising individuals at a range of consistent temperatures (22.5–35 °C), the impact of temperature on the pace of development of Pristhesancus plagipennis was ascertained. From 22.5 to 30 °C, development rates varied linearly. For egg development (13.1 °C), nymphal development (15.5 °C), and egg to adult development (15.4 °C), estimates of lower developmental thresholds were found. The nymphal and egg development rates were highest at 30 and 32.5 °C, respectively. For development from egg to adult, 845.7 degree-days were calculated to be necessary. The highest nymphal survival occurred between 25 and 30 °C (James 1992). Rhynocoris marginatus and Rhynocoris fuscipes lifespan, reproduction, and length of nymphal development were examined in the lab at six constant temperatures (10–35 °C). At 10, 15, and 35 °C, eggs were not fully formed and suffered 100% and 75% mortality, respectively. Similar to other predators, these two did not achieve adulthood, especially at temperatures between 10–15 °C and 35 °C. The overall nymphal development period of Rhynocoris marginatus reached its maximum temperature at 20 °C (87.26 days), declined at 25 °C (49.8 days), and further decreased at 30 °C (48.06 days). The longest lifetime was seen at 20 °C when freshly moulted adults of Rhynocoris marginatus were exposed to various temperature regimes (7.66 and 17.26 days for male and female respectively). With a 95.5% hatchability rate, the Rhynocoris marginatus egg hatched after 7.58 days. In Rhynocoris marginatus adults kept at various temperatures, the incubation period grew longer regardless of the temperature levels. Although Rhynocoris fuscipes had the longest longevity (35.25 and 32.5 days for male and female, respectively) at 25 ° C, it has been found that both male and female predators have the shortest and longest adult lifespans, which are 28.66 and 69.75 days at 20 °C (Sahayaraj and Sujatha 2012). The two reduviid were most fertile at 28 °C room temperature (Fig. 9.5). The lower (LTT) and upper (UTT) threshold temperatures (To) for Rhynocoris marginatus and Rhynocoris fuscipes vary and are not stable even within the life stages. Rhynocoris marginatus and Rhynocoris fuscipes have first instar and second instar lower threshold temperatures (To) that are extremely high. However, (To) requires a higher temperature for the second and fourth instars of Rhynocoris marginatus and Rhynocoris fuscipes, respectively, in order to maintain higher

272

9

Influence of Ecological/Climatic Change

Fig. 9.5 Influence of temperature on the fecundity (eggs/female) of Rhynocoris marginatus and Rhynocoris fuscipe maintained from 20 to 35 °C Table 9.1 Lower and upper threshold temperatures (To) for Rhynocoris marginatus and Rhynocoris fuscipes

Life stages Egg I instar Nymph II instar Nymph III instar Nymph IV instar Nymph V instar Nymph I—Adult

Lower threshold temperatures (To) Rhynocoris Rhynocoris marginatus fuscipes 19.51 20.53 21.16 19.41

Higher threshold temperatures (To) Rhynocoris Rhynocoris marginatus fuscipes 24.5 30.11 23.16 20.10

16.92

22.48

35.92

21.49

12.09

15.58

33.09

25.7

18.05

14.63

31.05

31.63

18.40

12.07

35.27

30.38

24.31

25.01

32.31

34.28

Nymphal lengths were found to be 7.1, 0.99, 8.3, 0, 10.1, 1, 2, and 1.16 days for I, II, III, IV, and V instars Rhynocoris marginatus, respectively. In contrast, Rhynocoris marginatus adult male and female longevity in Kharif season (July to December) was 81.3±6.3, and 116.9±9.7 days, respectively. Nymphal lengths for the I, II, III, IV, and V instars during the Rabi season (October to March) were 8.3, 11.2, 13.7, 17.9, and 22.3 days, respectively, with male and female lifespan of 90.3 and 119.9 days, respectively. Egg incubation took less time during the Kharif (9.0 days) than it did during the Rabi season (9.4 days), and the number of eggs that hatched during the Kharif (279.8 eggs and 88.4%, respectively) was higher than it was during the Rabi season (240.4 eggs and 77.3%). During the Kharif and Rabi seasons, the oviposition periods were 56.4 and 61.6 days, respectively. When compared to Rabi season, the total developmental period for the predatory bug Rhynocoris marginatus reared on the rice moth Corcyr cephalonica was shorter during the Kharif season, and there was also a higher fecundity. This indicates that the Kharif season is ideal for raising reduviid predators like Rhynocoris marginatus in natural settings (Bhoyar et al. 2021)

9.3 Biology

273

threshold temperatures (Table 9.1). Low temperatures typically cause insects to develop more slowly and spend longer in each stage of their life cycles. This may increase predation based on stage.

9.3.8

Coleoptera

Elevated temperatures have variable effects on the life histories, behaviours, and physiological characteristics of natural enemies like ladybird beetles, which in turn have varying effects on biological control. Ladybird beetles are crucially essential natural enemies that help control a variety of crop pests, like mites and aphids. Axinoscymnus cardilobus (Coleoptera: Coccinellidae), reared on Bemisia tabaci (Homoptera: Aleyrodidae) (Huang et al. 2008), Hippodamia variegata, Sitotroga crealella (Asghari et al. 2011), Cryptolaemus montrouzieri (Solangi et al. 2013), Propylea quatuordecimpunctata (Papanikolaou et al. 2013), and others have all been studied in the laboratory to see how elevated temperatures affect their growth, development, and survival in both the immature and adult stages (Ardakani et al. 2020). Impacts vary depending on the species, the experimental temperature range, and other life cycle factors including fecundity (Huang et al. 2008, Papanikolaou et al. 2013; Ardakani et al. 2020). The contribution of ladybugs to biological control is another important (temperature-dependent) factor. Exochomus nigripennis larvae on Gossyparia spuria (Ardakani et al. 2020), larvae and adults of Cheilomenes sexmaculata on Megoura japonica (Wang et al. 2013), or Micraspis discolour larvae on Brevicoryne brassicae are a few examples of the effects of temperature on the predation rates, stages of development, and prey items targeted by several species of ladybug (Hong et al. 2013). In cotton-growing regions of northwest China, two of the principal hemipteran pests’ natural enemies are Hippodamia variegata and Propylaea quatuordecimpunctata (Coleoptera: Coccinellidae). Temperature had an impact on the survival of adults of Helicoverpa variegata and Propylaea quatuordecimpunctata. Hippodamia variegata has the highest survival rate across all temperatures. In comparison to other temperatures, Propylaea quatuordecimpunctata survival at 38 °C dramatically decreased on the third day. Therefore, Propylaea quatuordecimpunctata was adversely affected by high temperatures more so than Hippodamia variegata. Temperature had an equivalent impact on the lifespan of Hippodamia variegata and Propylaea quatuordecimpunctata. Initially, only a temperature increase to 35 °C reduced longevity in Hippodamia variegata, but additional temperature increases similarly reduced longevity in Propylaea quatuordecimpunctata, which decreased from 10.96 days at 32 °C to 7.72 days at 35 °C to 3.83 days at 38 °C, respectively. Hippodamia variegata and Propylaea quatuordecimpunctata, two species of ladybird beetles, both saw an equal impact from the increased temperature on age-specific and overall fertility. Hippodamia variegata oviposition rates decreased as temperatures increased from 32 over 35 to 38 °C, going from 450.8 to 332.2 to 107.4 eggs per female. Peak oviposition shifted to days 6 and 4 at 35 and 38 °C and

274

9

Influence of Ecological/Climatic Change

day 6 at 32 °C. Total oviposition was substantially lower for Propylaea quatuordecimpunctata than for Hippodamia variegata at all temperatures. Propylaea quatuordecimpunctata oviposition rates decreased from 19.19 at 32 °C to 8.27 at 35 °C to 0 at 38 °C, respectively. Peak oviposition was seen for all temperatures on day 5 (Yang et al. 2022). With increased prey density, Hippodamia variegata and Propylaea quatuordecimpunctata had a Type II functional response, consuming a steadily decreasing proportion of Aphis gossypii (50, 100, 150, 200, 250, 300, or 350 individuals of fourth instar aphids; 350 individual). The maximum daily predation (1000 aphids) and attack rate (a = 1.12) of Hippodamia variegata occurred at 35 °C. Daily predation of Propylaea quatuordecimpunctata did not differ between temperature regimes (32, 35, or 38 °C); however, a value gradually decreased with increasing temperatures (Yang et al. 2022). Despite the fact that functional response curves are unchanged by temperature changes, predation rates may change. For instance, Harmonia axyridis’s rates of predation on Acyrthosiphon pisum nymphs caused temperatures to rise by over 15–35 °C (Yasir et al. 2021). The intake of Harmonia dimidiata by Myzus persicae decreased with temperature (24–32 °C) (Khan et al. 2016). Although Propylaea quatuordecimpunctata predation of Aphis gossypii reduced as temperature rose from 32 to 35 °C, variegata predation of Aphis gossypii was highest in this study at 35 °C. Consuming more A. gossypii regularly is Hippodamia variegata than Propylaea quatuordecimpunctata, demonstrating how strongly species-dependent temperature-mediated variations in predation rates are. Although both predator and prey act within complex, dynamic food webs where a variety of consumptive and non-consumptive effects (such as intra-specific competition and intraguild predation) determine the final outcomes (Wang et al. 2017, 2021; Yang et al. 2017), it can be challenging to translate these observations of specific predator–prey couplets made in the lab to real-world situations (Murdoch 1973). Numerous of the aforementioned impacts are further modified by field-, farm, or landscape-scale variability (Nedved et al. 2013). Therefore, semi-field trials, manipulative experiments, and observational assays in “real-world” situations are crucial to precisely predict how temperature would affect ladybird beetle fitness or biological control.

9.3.9

Thrips

Thrips are present in a variety of climatic conditions. As a result, their growth, reproduction, and other life traits varied depending on the environment. Ananthakrishnan (1993) initially assessed Thirps Bionomics. This point was also emphasised by the author. Japan and Formosa are home to the beneficial native thrips Scolothrips takahashii. In fields, it is typical of bean, citrus, pear, and tea, and is regarded as a significant spider mite predator. Intially, Yamasaki et al. (1983) looked at the lower temperature threshold for Scolothrips takahashii. development. According to a 2003 study by Gotoh and his colleagues, Scolothrips takahashii females deposited their eggs at 37.5 or 40 °C. The thermal constant was calculated to

9.4 Macromolecular and Antioxidant Responses

275

be 204.1 degree-days, and the lower threshold temperature for oviposition development was 11.7 °C. According to these findings, between seven and ten generations could reach maturity under field conditions in Ibaraki, central Japan, at the most. Although oviposition could be induced without mating, unmarried females only conceived male offspring. Females averaged 90.5 eggs per egg lay over a mean oviposition duration of 17.8 days at 25 °C. One of the most important life-history metrics is the intrinsic rate of natural increase (rm), which is low at minimum temperatures ranging from 20 °C (0.113) to 25 °C (0.195) and high at 30 °C (0.246). Comprehensive information on the distribution of Scolothrips longicornis is lacking in the publications. However, Scolothrips longicornis is a naturally beneficial thrip in the Mediterranean and Middle East. The growth, reproduction, and survival of the predatory thrips Scolothrips longicornis, which fed on the two-spotted spider mite, Tetranychus urticae, were examined under five photoperiods in a laboratory setting at 60% RH and 25 °C. Practically all Scolothrips longicornis immature stages evolved the quickest when the days were long (18:6 and 24:0 L:D). The adult durations of both sexes decreased when light duration was increased from 6 to 24 h. For both men and women, shorter lives accompanied longer days. Under a 12:12 L:D photoperiod, female Scolothrips longicornis had the longest oviposition time, the highest net reproduction rate (R0 = 15.37), the intrinsic rate of natural growth, and the finite rate of ascent. The life table factors showed a notable variance with different photoperiods. The current study’s findings demonstrated that a 12:12 L:D photoperiod is ideal for Scolothrips longicornis development and reproduction when fed Tetranychus urticae, and that it would also be the best photoperiod for mas rearing for augmentative biological control systems to maximise production (Pakyari and McNeill 2020). Pakyari et al. (2011) previously observed that temperature has an impact on the sex ratio of Scolothrips longicornis progeny growing at various temperatures with ratio values (females/ males). Therefore, even at 37 °C, where the male population rose, 75% of the offspring were female, perhaps indicating less than optimum conditions. Temperature significantly affected the persistence of Scolothrips longicornis adult thrips on longicornis that can live for 38 days at 15 °C but only 4 days at 37 °C. Similar to lifespan, the duration of both preoviposition time, during oviposition, and postoviposition period was noticeably lowered as the temperature rose. Thus, at the lowest temperature it took 6 days for the thrips to start the egg-laying process, which continued for 24 days, whereas oviposition occurs at the greatest temperature was began immediately after female emergence but only continued for a few days.

9.4

Macromolecular and Antioxidant Responses

9.4.1

Macromolecular Profile

In physiological ecology, the evolution of metabolic rate-temperature reaction standards is crucial (Terblanche et al. 2009). Furthermore, according to the metabolic cold adaptation theory, an increased metabolism brought on by an adaptation to

276

9

Influence of Ecological/Climatic Change

cold climes may benefit insects by enabling them to complete growth, development, and reproduction at relatively low temperatures throughout the breeding season (Terblanche et al. 2009). Entomologists have generally noticed that reducing the growth temperature led to increased pigment formation, polysaccharide synthesis, and levels of unsaturated fatty acids in cellular lipids. It is abundantly obvious from the findings of Sahayaraj and Sujath (2012) that the total carbohydrate content is low at 10 °C (22.2 g/mg). As the temperature rises, it steadily gets stronger until it reaches its maximum at 35 °C. Another study looked at the biochemical makeup of eggs in response to temperature. For the two reduviid species, the greatest content of total carbohydrates was discovered at 35 °C. In addition, the intestinal protein profile of Rhynocoris fuscipes, a reduviid species, gradually increased the amount of egg protein up to 30 °C. The total amount of carbohydrates was noticed in the biochemical composition of eggs in proportion to temperature; the reduviid species’ greatest content was noted at 35 °C. Additionally, the reduviid species’ egg protein levels gradually increased up to 30 °C. Both Orius majusculus and Orius aculeifer underwent 7 or 16 days of acclimation at 10, 15, or 20 °C, respectively. The effects of temperature exposure on food intake, basal metabolic rate, and body composition were next examined in both species. Our findings demonstrated that Orius majusculus and Orius aculeifer are more tolerant of famine when exposed to low temperatures. Body composition studies showed that both species have built up greater lipid stores while being exposed to colder temperatures, which, at least in part, explains why they are more tolerant to famine after being exposed to the cold. In contrast, heat acclimatisation had no effect on consumption or basal metabolic rate. According to our research, predatory arthropods exposed to cold weather develop stronger physiological systems and are better able to withstand environmental stresses like low temperatures and a lack of prey (Jensen et al. 2018).

9.4.2

Antioxidant Responses

Antioxidant defences can be activated by high temperatures to eliminate free radicals and insulate the insect from thermal stress (Marcelo and Tania 2002; Yao et al. 2007). These species-specific antioxidant defences are controlled by a collection of temperature-dependent enzymes. P. japonica starts to become more active at 39 °C for the enzymes superoxide dismutase (SOD), catalase (CAT), and glutathione S transferase (GSTs), while the POD enzyme only starts to become more active at 41 ° C. The SOD, CAT, and GST activities in larvae of Mythimna separata were significantly elevated by thermal stress, which is defined as temperatures of 30, 35, 40, and 45 °C (Ali et al. 2017). Chen et al. (2018a, b) found that when exposed to high temperatures (40, 42, and 44 °C), female adults of Ophraella communa had a survival rate that was considerably higher than that of male adults. Adult female likewise demonstrated higher antioxidant enzyme activity than adult men (Chen et al. 2018a, b). The high survival rate of the species may be directly tied to communa females experiencing heat stress. The research indicated that four

9.5 Temperature Tolerance Factors

277

antioxidant enzymes behaved differently to rising temperatures. Such chemicals undoubtedly play a substantial role in the thermal tolerance of Hippodamia variegata and Propylaea quatuordecimpunctata as SOD and CAT enzyme activity increased at higher temperatures. Hippodamia variegata’s POD enzyme activity increased at 35 ° C, whereas Propylaea quatuordecimpunctata’s POD enzyme activity dropped. Meanwhile, differences between species were discovered in (temperature-related) T-AOC activity. Hippodamia variegata may have more antioxidant defences (and a lower risk of oxidative damage) than Propylaea quatuordecimpunctata as evidenced by its increased T-AOC activity at high temperatures. These physiological patterns serve as the primary drivers of Hippodamia variegata spatiotemporal distribution and may help to explain the variances in life history features mentioned above. In adults of the Hippodamia variegata held under the three temperatures for 24 h, SOD activity increased at higher temperatures. Similar to this, SOD activity gradually rose with temperature in adults of Propylaea quatuordecimpunctata. Adults of the CAT-producing species Hippodamia variegata and Propylaea quatuordecimpunctata likewise steadily rose with temperature. Hippodamia variegata POD activity increased at 35 °C but subsequently fell by 82.6% at 38 ° C. At higher temperatures, POD activity for Propylaea quatuordecimpunctata gradually diminished. At increasing temperatures, GST activity consistently decreased for both species of ladybug. The temperatures used during treatment had a substantial impact on the T-AOC activity levels in the two adult ladybugs. Hippodamia variegate T-AOC activity rose with temperature, but Propylaea quatuordecimpunctata T-AOC activity only started to fall at 38 °C. Lastly, both species of ladybug were significantly impacted by heat stress in terms of protein content (Yang et al. 2022).

9.5

Temperature Tolerance Factors

In addition to being more sensitive to higher temperatures, the Miridae species Dicyphus bolivari (two distinct strains), Dicyphus eckerleini, Dicyphus errans, Dicyphus flavoviridis, Nesidiocoris tenuis, and Macrolophus pygmaeus were also more sensitive to lower temperatures. The studied species’ greater sizes appear to be more adapted to colder temperatures, whereas the smaller sizes appear to be better adapted to warmer temperatures. All species have males and females with different tolerances for heat and cold. In general, males were better acclimated to higher temperatures than females were to lower ones (Ingegno et al. 2021). Based on morphological characteristics and biochemical testing, 13 and 12 bacteria were found in Rhynocoris fuscipes and Rhynocoris marginatus, respectively. The bacteria that were found in Rhynocoris marginatus’s stomach included Bacillus cereus, B. magaterium, Enterobacter aerogenes, Micrococcus luteus, Corynbacterium kutcherii, Corynbacterium xerosis, Bacillus subtilis, Escheritia coli, Pseudomonas aeruginosa, Staphylococcus aureus, and Micrococcus variance. In the fore and hindgut of both predators, protease activity peaked at 25 °C, but amylase and invertase activity peaked at 35 °C. The foregut and hindgut of

278

9

Influence of Ecological/Climatic Change

Rhynocoris marginatus displayed their highest esterase activity at 20 °C. However, in Rhynocoris fuscipes, esterase activity was equal and at its peak at both room temperature and 30 °C. Except for B. subtilis and Lactobacillus delbrucki, all the bacterial species were positively linked with temperatures, with values of 0.67, 0.50, 0.84, and 0.42 for Bacillus cereus, Micrococcus variance, Staphylococcus aureus, and Corynbacterium xerosis, respectively. At 10 °C, it was more prevalent (29%) At all temperatures, the populations of Staphylococcus aureus and B. cereus were generally comparable in both reduviid species. These two bacteria were thought to be the temperature-tolerance-causing autochthonous bacteria of reduviid predators (Sahayaraj and Sujatha 2011). The effects of global warming are now increasingly apparent, and because natural enemies typically have low thermal tolerance, this could have a severe impact on pest biological management. By choosing between different surrounding temperatures, a number of insect species control their body temperature both physiologically and behaviourally. In turn, interspecific connections, even those that are pathogenic, may have an impact on behavioural thermoregulation. For instance, exposing a host to temperatures close to its thermal limits may change how they forage, how they survive and reproduce, and how they interact with their pathogens. By inducing changes in behaviour and physiology, pathogens can affect how the host regulates its body temperature. For instance, behavioural fever, or the wilful exposure to a body temperature higher than normal, is thought to constitute a component of an insect’s immunological response, a minimum of in some species. Parallel to this, exposure to a pathogen extreme temperature (referred to as ET from now on) may weaken insect immune defences, increasing infection susceptibility. This paradigm makes it seem like there is little knowledge about how ET affects insects and the pathogens that infect them, especially when it comes to how fungus pathogens affect thermal behaviour (Porras et al. 2021). In terrestrial food webs, pathogens can change a variety of physiological or behavioural characteristics of the host with cascading effects on all trophic levels. It has nearly always been studied how fungal infections affect trophic level-specific thermal tolerances and behavioural responses to extreme temperatures (ET). The hosts’ ability to tolerate heat is likewise impacted by these alterations. We looked into how the fungus Beauveria bassiana affected the behaviour and upper and lower heat tolerance of the beetle predator Hippodamia convergens and the herbivorous bug Acyrthosiphon pisum. Porras et al. (2021) studied differences in thermal tolerance limits, thermal boldness (voluntary exposure to ET), and energetic cost (ATP) posed by each reaction comparing healthy insects and insects infected with two fungal loads (thermal tolerance and boldness). A fungal infection reduced the beetles’ CTMax and CTMin values. Fungus infection altered the inclination or boldness of predator beetles and aphids to cross either warm or cold ET zones (ETZ). Adenosine triphosphate (ATP) levels increased in both insect species in response to pathogen infection, and the highest ATP levels were found in individuals who crossed the cold ETZ. Fungus decreased the thermal tolerance range, and thermal boldness behaviours to cross ET were stopped. As environmental temperatures rise, different trophic levels of a food web’s constituents will react to

9.6 Cold Storage

279

thermal stress in different ways, which might have an effect on predator–prey relationships, food web structures, and species distributions (Porras et al. 2021).

9.5.1

Cuticular Permeability

The cuticular permeability of medium-sized female nymphs varied from 8.5 JLg water lost per cm2 ‘h’ mmHg to 17.0 JLg water lost per cm2 ‘h’ mmHg. When the temperature was high, the striped earwig, Labidura riparia, and the lesser cornstalk borer both lost more water. The smaller cornstalk borer lost more water in a 24-h period than the striped earwig Labidura riparia, indicating that the effect of temperature on water loss varied by species. The slopes of the regression line of the weighted water loss at 35 and 40 °C were >2 SE larger for females of the striped earwig, Labidura riparia, relative to third instars of the lesser cornstalk borer (Kharboutli and Mack 1993a, b).

9.6

Cold Storage

Low storage temperatures are ideal for preserving entomophagous insects. Since natural insect enemies are commonly used in biological control programs, studies on cold storage have been ongoing for more than 90 years. Farmers can store enough entomophagous insects for field release under optimal weather conditions by having the ability to preserve raised biocontrol agents at low temperatures for a predetermined period of time. This also makes them available to them during times of high demand. Cold-stored natural enemies can be concurrently released in the fields during the height of a pest outbreak. Cold storage also aids in maintaining a healthy pool of natural adversaries when not needed and in lowering laboratory activity by extending their survival and delaying eclosion. The trait of cold storage resistance can be significantly influenced by a number of biotic and abiotic factors that are encountered before to, during, and after cold exposure. These factors ultimately have an impact on entomophagous insects’ morphology, behaviour, and physiology as well as their growth, longevity, fertility, parasitism, sex ratio, and other fitness indices. For a cold storage project to be implemented effectively, it is essential to comprehend these varied causes of storage and post-storage repercussions. Therefore, the potential for cold storage protocols to improve mass rearing and commercial production of bioagents is investigated in order to obtain access to the international methods, inventions, methods, devices, and wisdom involved in the process of cold storage of entomophagous insects (Rathee and Ram 2018).

280

9.6.1

9

Influence of Ecological/Climatic Change

Reduviids

Low temperature shortens the time that the eggs of both of the analysed predators spend incubating. Similar to this, Allaeocranum biannulipes, another reduviid predator, saw a shortening of the incubation period at higher temperatures (Tawfikm and Awadallaht 1983). Additionally, reduviid eggs need a particular quantity of moisture to incubate; prolonged dryness slows the development of the eggs (Vennison and Ambrose 1990). In Rhynocoris marginatus and Rhynocoris fuscipes, the impact of cold storage on egg hatching was examined at two distinct temperatures, namely 14 and 20 °C, during a range of storage times, including 1, 2, 4, 8, and 12 days. At 14 °C, neither reduviid hatched as expected. While Rhynocoris marginatus eggs could be maintained for up to 4 days with minimal nymphal mortality and only minor nymphal abnormalities, Rhynocoris fuscipes eggs were not acceptable for storage at 20 °C. However, the hatching of Rhynocoris marginatus eggs and nymphal survival were significantly impacted by storage for 8 days or longer. Rhynocoris fuscipes hatched at a rate of 99% in the control, compared to 90% for Rhynocoris marginatus. The incubation duration in both reduviids was prolonged by cold storage at 20 °C (Sahayaraj and Paulraj 2001). At several sites throughout Egypt, Coranus afn’cana was found on a range of domestic and untamed plants. Using the larval stage of Anagasta kuehnklla (Lepidoptera: Pyralidae) as a prey, the effects of cold storage temperatures (5, 10, and 15 ° C) and 70% relative humidity were investigated by El-Sebaey (2007). Male and female nymphal stages were stored for an average of 5.1, 6.43, 7.3, 8.9, 10.01, 10.31, and 9.79 weeks each at 15 °C. They exceeded storage times at 10 and 5 °C. Storage temperatures also had an impact on aspects of reproductive biology, age-specific survivability, fecundity, egg hatchability, and dietary intake (El-Sebaey 2007).

9.6.2

Pentatomidae

Temperature affected how much heat was needed for development. For example, at 17.5 and 35 °C, the eggs of Biprorulus btbax (Hemiptera: Pentatomidae) hatched after 16.2 and 2.4 days, respectively. However, Biprorulus btbax nymphs did not grow past the second stadium. All stadia of the Biprorulus btbax successfully developed between 20 and 32.5 °C. At 200 °C, it took 79.2 days to go from oviposition to adulthood, whereas only 23.9 days did so at 32.5 °C. The fifth stadium was the longest at all tested temperatures as Biprorulus btbax nymphs evolved from the first through fifth stages, with subsequent stages generally taking longer to develop. For nymphs, the greatest pace of development took place at 32.5 °C, while for Biprorulus btbax eggs, it happened at 35 °C. Except for the fourth stadium (37%), estimates of the number of degree-days (DD) and developmental zeros (DZ) required to complete each stage of development ranged from moderate to high (63–92%). The DZ for development from egg to adult was found to be 14.3 °C. However, it varied between 1.3 and 17.5 °C for each phase. From the first to the fourth stadium, the DZ increased while it was being constructed.

9.6 Cold Storage

281

But it fell short in the fifth. At 15 °C, embryonic development occurred but eggs did not hatch, which is consistent with the stage’s anticipated DZ (15.3 °C). Complete development is theoretically possible at 17.5 °C, according to the computed DZ, but there was a considerable amount of fatalities at this temperature. However, in other laboratory investigations, a few individuals have successfully finished nymphal development at this temperature. Biprorulus btbax has an upper developmental threshold that is thought to be between 35 and 37.5 °C. If field temperatures dropped below the lower developmental threshold or climbed over the upper developmental threshold, developmental rates might be anticipated to be a little quicker or slower than those seen under constant temperatures (James 1990). Southern New South Wales fields frequently see springtime field temperatures below 14.3 °C and summertime field temperatures above 35 °C. For the majority of this period, temperatures hardly rise beyond the DZ, which hinders considerable nymphal growth. Winter conditions (mean daily maxima of about 15 °C, minimum of about 5 °C; unpublished data) are not harsh enough to kill mid- to late-stadium nymphs, who can survive until the end of winter. However, the onset of cold weather can occasionally cause small numbers of late-season nymphs to get thermally “stuck”. Eggs required 50.4 degree-days (DD) to grow, and various nymphal stadiums needed between 50 and 122.8 DD; the fifth stadium required the most DD, followed by the second stadium. While nymphal development required 405 DD, the development from oviposition to adult emergence took 455.4 DD. For temperature, considering the rates of development of additional pentatomids, there are comparatively limited data available. According to Simmons and Yeargan (1988), Acroster numhilare has an approximate development threshold of 15 °C, with nymphal development rates peaking at 27 °C. The predaceous pentatomids Oechalia schellenbergii and Cermatulus nasalis have a developmental threshold for nymphal development that is only slightly above 15 °C, according to data from Awan (1988). At 35 °C, Oechalia schellenbergii and Cermatulus nasalis effectively matured from eggs to adults, and this temperature was ideal for their rate of development (Awan 1988). In habitats like those of B. bibax, Oechalia schellenbergii and Cermatulus nasalis can be found in southern Australia, and it appears that their preferred temperature range is comparable. At 25, 27.5, and 30 °C, nymphal survival was at its maximum (41–63%), while it was at its lowest (4–12%) at 20, 22.5, and 32.5 °C. There was no survival to the third stadium at 17.5 and 35 ° C, while there was 7% and 58% survival to the second stadium, respectively. Ninetyfive per cent to 100% of the cohort made it to the second stadium at 25, 27.5, and 30 ° C (James 1990). Eggs from insects fed on a diet fared better in storage at 10 °C than eggs from Podisus maculiventris fed on prey. Podisus maculiventris eggs also fared better in storage at 10 °C than eggs from insects fed on a diet. With a similar pattern of response to temperature, nutrient quality, and period of storage, nymphal survival of cold storage treatment was marginally higher than for eggs. But of the three developmental stages evaluated, adulthood provided the highest rate of survival. Adult survival was improved at 10 °C, just like it was for eggs and nymphs. The adults who were fed prey fared better than the adults who were fed diet, in contrast to

282

9

Influence of Ecological/Climatic Change

eggs and nymphs. The results show that the ability of prey-fed adults to live, breed, or lay viable eggs was not seriously influenced by 4 weeks of cold storage at 10 °C. At 4 °C, the biggest decline in survival, fecundity, and egg viability occurred due to extended storage durations. For the longer periods of storage at 10 °C, the eggs from adults who were provided a meal had the highest viability. These results demonstrate that the optimal food supply for continuous rearing is not always the greatest food source for cold storage, and that nutrient quality influences how organisms respond to cold storage at different developmental stages (Coudron et al. 2007).

9.6.3

Nabidae

Nabis pseudoferus is a key biocontrol agent for sucking pests and caterpillars. Nabis pseudoferus’s life history traits, consumption rate, and cold storage were observed. The estimated temperature, relative humidity, and photoperiod for Nabis pseudoferus on three diets—Aphis gossypii, Ephestia kuehniella eggs, and Aphis gossypii + Ephestia kuehniella eggs—were 25 °C, 60–70%, and 16:8 h, respectively (L:D). The results showed that the intrinsic rates of rise (r), depending on the feeding diets, were 0.033 ± 0.003, 0.043 ± 0.003, and 0.062 ± 0.002 day-1. Data analysis revealed that Nabis pseudoferus had a net predation rate (C0) of 1347.95 ± 276.77 for cotton aphid meals and 2259.62 ± 145.54 for diets containing both cotton aphid and Ephestia kuehniella eggs. Additionally, 72.2, 69.2, and 69.2% of the pseudoferus eggs hatched after being kept in a refrigerator at 5 °C for 7, 14, and 21 days, respectively. These discoveries might contribute to more effective breeding of Nabis pseudoferus (Ahmadi and Madadi 2021).

9.6.4

Geocoridae

Both Geocoris pallidipennis nymphs and adults could be kept in storage for 3 weeks at 5 °C without suffering deleterious effects on their survival, fertility, or susceptibility to Bemisia tabaci predation. Geocoris pallidipennis fecundities tended to decline with longer periods of cold storage, but up to 4 weeks of storage at 5 °C had little effect on their fecundity. It was possible to store Micromus angulatus adults for up to 140 days at 5 °C with honey solution as prey, exhibiting a cumulative survival rate of about 50%; however, cold storage for longer than 2 weeks was not optimal for their reproduction. M. angulatus’ aphid predation decreased as the length of cold storage increased; however, it may be able to keep food in the freezer for up to 5 weeks without it having a substantial impact. For inundative biological techniques, short-term cold storage of Geocoris pallidipennis and Micromus angulatus may be helpful (Seo et al. 2019).

9.6 Cold Storage

9.6.5

283

Neuroptera

The stage of the larva most responsive to photoperiodic cues is the third instar, which is free-living. The development of prepupal diapause is primarily controlled by photoperiod, with temperature acting as a modifying factor. Chrysopa formosa photoperiodic response curves revealed a normal long-day response across a wide range of temperatures. At 20 °C, 14.3 h was determined to be the critical day length for diapause induction, with shorter day lengths being more important at higher temperatures. At temperatures of 18 or 20 °C, the diapause reaction peaked with a short-day duration of 8L: 16D, and no diapause was elicited at any other temperature. The conclusion of diapause termination and the onset of post-diapause development were both significantly impacted by cooling temperatures. Most prepupae went into diapause after 60 days or more of cooling at 5 °C. The development of post-diapause and the onset of adulthood were coordinated by a longer duration of chilling exposure. The ability of diapausing chrysopids to be held at 5 °C for up to 300 days with good survival and rapid adult emergence seemed particularly alluring. For chrysopids that would experience diapause, the pre-diapause developmental stage was greatly prolonged. In Chrysopa formosa, pre-pupal diapause significantly improved cold tolerance. These findings point to processes that could be used to extend the shelf life of chrysopids and their long-distance shipping for use in biological control programmes (Li et al. 2018).

9.6.6

Coleoptera

The viability of cold storage has been examined for a number of coccinellids, including Coccinella septempunctata, Adalia bipunctata (Hamalainen 1977), Coccinella undecimpunctata (Abdel-Salam and Abdel-Baky 2000), and Coleomegilla maculata lengi (Gagne’ and Coderre 2001). Regarding an insect’s thermal reactions, very low or high temperatures can limit an insect’s biological and physiological activities, but the majority of biotic processes would benefit from an ideal environmental temperature. When living in cold surroundings, some ladybird species were shown to have shorter developmental times, lower survival rates, and less effective reproduction than when living in warm environments. At a number of constant temperatures (15, 20, 25, 30, 32.5, and 35 °C), the fertility of the pseudococcid predators Nephus includens and Nephus bisignatus (Coleoptera: Coccinellidae), which preyed on Planococcus citri (Hemiptera: Pseudococcidae), was investigated. Life-fecundity tables were created with extra information regarding the growth of the immature stages, and several population parameters were computed. The average total fecundities of Nephus includens were 49.2, 97.8, 162.8, 108.5, 87.4, and 31.1 eggs/female at the aforementioned temperatures, and the average longevities were 99.5, 84.7, 69.5, 61.1, 49.6, and 30.1 days, respectively. The intrinsic rates of rise (rm) were 0.014, 0.041, 0.083, 0.086, 0.077, and 0.024 females/female/day, respectively. The net reproductive rates (R0) were 8.0, 32.2, 60.7, 32.6, 20.7, and 2.6 females/female. Nephus bisignatus had average total

284

9

Influence of Ecological/Climatic Change

fecundities of 54.7, 72.1, 96.9, 56.0, and 22.8 eggs/female at 15, 20, 25, 30, and 32.5 °C, respectively, and average longevities of 116.1, 108.7, 71.8, 68.8, and 43.7 days. For females at 15, 20, 25, 30, 32.5, and 35 °C, the intrinsic rates of rise (rm) were 0.017, 0.035, 0.060, 0.051, and 0.024, respectively. The net reproductive rates (R0) for females at 15, 20, 25, 30, 32.5, and 35 °C were 13.9, 26.4, 31.3, 15.2, and 3.6 (Kontodimas et al. 2007). It was investigated how various temperatures affected certain of Nephus includens biological characteristics (Coleoptera: Coccinellidae). This species is one of Planococcus citri’s most important predators (Homoptera: Pseudococcidae). The growth time, mortality, and fecundity were evaluated at constant temperatures of 15, 20, 25, 30, and 35 °C as well as at variable temperatures of 25–35 °C (12 h at 25 °C, 12 h at 35 °C). Additionally, life tables were created for 25, 30, 35, and 25–35 °C. Compared to all other temperatures except 35 °C, the mortality was lower and the mean generation time was shorter at 30 °C. At 30 °C, the intrinsic rate of growth was highest (0.081), followed by 25–35 °C (0.076). In comparison to 30 °C, the net reproduction rate was greater at 25–35 °C. The best temperatures for mass raising of Nephus includens were found to be 30 °C and 25–35 °C based on biological data and population growth characteristics computed from the life tables. However, bulk raising of the citrus mealybug on sprouted potatoes at a temperature as high as 35 °C may result in degeneration. Consequently, 30 °C would be preferable to 25–35 °C. Coccinella septempunctata and Hippodamia convergens egg, larval, and pupal mortality. Hippodamia convergens and Coccinella septempunctata had the highest rates of egg mortality at 14 °C, reaching 85% and 49%, respectively. Only in Hippodamia convergens, though, were the variations in temperature treatments statistically significant. Temperature had a considerable impact on both predators’ overall larval and pupal mortality, with 14 °C being the highest. At 14 °C, both predators experienced much higher mortality in their first larval instar than in all subsequent instars. Only at 14 °C, when death was considerably higher in Hippodamia convergens than in Coccinella septempunctata, were there significant differences in mortality between the two predators at each temperature. Increased temperatures drastically shortened the time spent in the egg, larval, or pupal stages as well as the entire pre-imaginal development of Hippodamia convergens and Coccinella septempunctata. In its natural habitats in Asia, the predatory ladybird Cheilomenes (Menochilus) sexmaculata (Coleoptera: Coccinellidae) preys on a number of herbivorous arthropod species. Cheilomenes sexmaculata developmental time was strongly influenced by environmental temperature at the egg stage, first through fourth larval stages, prepupal stage, pupal stage, and overall developmental period. The findings demonstrated that when environmental temperature rose, each developmental stage’s duration shrank. Additionally, at warmer temperatures, the adult eclosion rate was much higher. Cheilomenes sexmaculata average copulation, pre-oviposition, and oviposition times varied dramatically with respect to environmental temperature. Environmental temperature also had a major impact on fecundity and egg hatchability (Wang et al. 2013). A colder climate might significantly lengthen the time

9.7 Wind Velocity

285

between copulation and pre-oviposition. In general, larger rates of sperm or spermatophores transmission in insects may be associated with a longer copulation period. The length of the photoperiod also had a substantial impact on Cheilomenes sexmaculata developmental periods during the egg stage, first through fourth larval stages, prepupal and pupal stages, and overall developmental period. The length of the photoperiod had a big impact on how long copulation, pre-oviposition, and oviposition lasted. According to observations, photoperiod also had a considerable impact on fecundity and egg hatchability. Cheilomenes sexmaculata longevity was considerably influenced by the photoperiod and gender both singly and in combination (Wang et al. 2013). Harmonia axyridis (Coccinellidae) is a significant biological pesticide in Asia’s original distribution area. To spend the winter, the “chill intolerant” ladybird beetle either migrates (always across considerable distances) to physically sheltered areas, or it acclimates to the cold (Pervez and Omkar 2006). The ability of Harmonia axyridis to endure cold temperatures was the subject of multiple studies (Berkvens et al. 2010; however, there is little information on how cold storage affects the ladybird’s fitness). According to a recent study, after the first four storage periods, female adult survival rates decreased as storage times lengthened, and Ha-artificial diet had a significantly higher survival rate than Ha-aphid. However, storage length had a different effect on Ha-aphid survival than it did on that of the Ha-artificial diet (significant diet storage duration interaction). For example, at storage periods of 15 and 30, almost 80.0% of Ha-aphid and roughly 95.5% of Ha-artificial diet survived; during 45- and 60-day storage intervals, these survival rates rose to 92.7% and 68.0%, respectively. At 60 days, they fell to roughly 50% and 68.0%. Male adult survival rates decreased as storage times increased, and the artificial food fared marginally better in terms of survival than the aphid diet. There was no clear correlation between diet and storage time, although 45 days of storage drastically reduced adult survival for both feeding methods. In particular, the survival rates of Ha-aphid and Ha-artificial diet were 92.0% and 95.2%, respectively, after 15 and 30 days of storage; however, after 45 and 60 days of storage, the survival rates declined to extremely low levels (Ha-aphid: 48.5% and 52.8%, respectively; Ha-artificial diet: 207 75.6% and 73.7%, respectively) (Sun et al. 2019).

9.7

Wind Velocity

Even though wind is a component of the environment that is almost invariably present, little work has been done to assemble our knowledge of how wind affects interspecific interactions. In order to summarise our existing knowledge of the effects of wind, we looked through the literature with a focus on interactions between predators and prey. We identified three main ways that wind affects interactions between predators and prey: physical disruption, movement, and identification of other species. We discovered that wind can have a variety of consequences that either amplify or reduce the impact of predators on their prey.

286

9

Influence of Ecological/Climatic Change

These impacts, meanwhile, are context-dependent, and predicting how a slowing wind speed may affect how different species interact will depend on the precise characteristics of the predator, prey, and ecosystem in which they coexist (Cherry and Barton 2017). A predator–prey interaction model with wind effects and anti-predator behaviour has been constructed in this study. Prey is thought to logistically grow in the absence of predators. Additionally, it is presumable that the prey engages in anti-predator behaviour (group defence) in opposition to the predator. The effects of wind direction have been taken into consideration when analysing our suggested model. The density of predators is thought to have decreased as a result of the wind impact. Following a study of the boundedness of each system solution, all potential equilibrium points are identified. The local stability of our proposed system has been investigated near these equilibrium locations. The Hopf bifurcation of the system has been investigated in relation to predator inhibition, wind impacts, and antipredator behaviour (b1). By stabilising our suggested system, the effects of wind direction and prey–anti-predator behaviour are diminished (says Panja 2022). A predator–prey model including wind as a component of the predation function is created and evaluated by Barman et al. in 2022. Two alternative types of modified functional responses have been developed to represent the overall dynamics of the linked system since wind can either slow down or speed up how quickly predators consume their prey. It has been demonstrated that the systems’ solutions’ positivity and boundedness support their well-posedness. The stability of equilibrium points and their state have been investigated under the proper parametric constraints. Through Hopf bifurcation, it has been shown that wind can sometimes change how stable the coexistence equilibrium point is. On the other hand, it is impossible to assume that the wind’s strength will remain consistent throughout time. We employ computer simulation to analyse the relevant systems after adjusting the functional responses to account for regularly varying wind flow. Our research revealed a strong association between a system’s functional response and the impact of wind on its dynamics. The article concludes with a thorough analysis of our entire investigation, with all the analytical findings supported statistically. In cotton, Gossypium hirsutum, field research was carried out at the University of Arizona in the United States from 1997 to 1999. Dislodgment was correlated with wind speed (m s-1), rainfall, and predator densities, while rates of predation were correlated with densities of Geocoris spp., Orius tristicolor, Chrysoperla carnea, and Lygus (Naranjo and Ellsworth 2005). At the field level, there was a positive association between the populations of Hippodamia convergens adults, Coccinella septempunctata larvae, and Coccinellidae adults (Elliott et al. 2006). The populations of Hippodamia convergens larvae, Coccinella septempunctata adults, Coccinellidae larvae, Nabidae nymphs and adults, Chrysopidae young ones and adults, Carabidae, and Staphylinidae, on the other hand, showed negative connection (Elliott et al. 2006). Wind speed had a highly significant favourable impact on the population of Chrysoperla carnea in cabbage in the 2008–2009 seasons, but only a marginally significant positive impact in the following season (El-Fakharany 2010). In cotton

9.8 Distribution at Field Level

287

fields, Coccinella undecimpunctata populations both in 2006 and 2007 observations displayed a negative connection with wind speed (Khodier et al. 2011). In cowpea environments, a different species, Coccinella septempunctata, likewise revealed a negative association between its populations and wind speed (Kumar and Kumar 2015). On soybean and Glycine max plants, three predatory insect species, including Chrysoperla carnea, Coccinella undecimpunctata, and Orius spp., were prevalent (variety Giza 111). According to Khattab et al., there populations have not been affected by wind speed (2019). In Mulberry Field, there was no predatory population that positively or naturally correlated with wing velocity (Das et al. 2021). The majority of studies indicate that wind velocity plays no part in controlling the population of predators.

9.8

Distribution at Field Level

According to field investigations, Cheilomenes sexmaculata’s developmental period lasted noticeably longer in colder years than it did in warmer ones (Zhang et al. 1980). In a Dutch apple orchard, observational data collected over a 13-year period were examined to determine the phenology and abundance of the first brood of the common earwig, Forficula auricularia. According to the findings, nymphs in their fourth instar appeared between 12 June and 10 July, and they achieved adulthood between 8 and 30 July. The observed number of earwigs in shelter traps and the total number of day-degrees above a thermal threshold of 6 °C since 1 January were shown to be strongly correlated. The most earwigs could be caught in the trees after the first brood’s nymphs reached the fourth instar, which typically happens between 600 and 750 day-degrees (Helsen et al. 1998). Most provinces in southern China, as well as the middle and lower portions of the Yangtze River, are home to Cheilomenes sexmaculata (Pang et al. 2004). According to Mouly et al. (2018), it was possible to assess the influence of abiotic conditions on the prevalence of reduviid, Isyndus heroes by looking at their abundance in an organic mango orchard. The initial flowering phase (January) and vegetative phase were when reduviid populations peaked (September–December). According to the correlation matrix, the population of Isyndus heroes and relative humidity were significantly associated, although the highest and lowest temperatures were significantly correlated. Further, the significant variables were regressed, and the maximum temperature (R2 = 0.62) with a single meteorological factor had the highest coefficient of determination. But multiple regression analyses showed that the maximum and minimum temperatures might account for up to 49% of the variability (Mouly et al. 2018). Distribution of Rhynocoris marginatus and Rhynocoris fuscipes throughout India was recorded. The northern section of India is extremely chilly from the end of the Southeast Monsoon to December, while the southern part is significantly warm. Consequently, there is a lot of variance in both pest damage and predatory effectiveness across the geographic range of these species, which probably reflects

288

9

Influence of Ecological/Climatic Change

heterogeneity in their biological performance under local conditions, especially their responses to temperature.

9.9

Elevated Atmospheric CO2

Since the Industrial Revolution, there has been a sharp increase in the global atmospheric CO2 concentration, which is thought to be a major contributor to climate change. It has been found that herbivore insects are most affected by the changes brought about by elevated CO2 in their host plants. The intensity and regularity of insect outbreaks on crops may subsequently be affected by these changes. Elevated CO2 modifies the phenotype of plants and has an effect on plant quality by changing how carbon and nitrogen resources are distributed among primary and secondary metabolites in plant tissue. Reduced nitrogen in plant tissue affects the generation of plant nutrients and secondary metabolites, and this has cascading effects up the food chain that can change the multi-trophic interactions in the agroecosystem. In general, it has been discovered that increased atmospheric CO2 has detrimental effects on plant quality, which in turn has detrimental consequences on insect herbivore performance. But when pests make up for poorer food quality by eating more plant tissue, crop loss may rise. Rarely have the impacts of increased CO2 on higher trophic levels been studied. However, a flow-through effect from the plant to higher consumers (predators and parasitoids) occurred to alter the third trophic level. Elevated CO2 had minimal direct impact on natural enemies of insect herbivores. Basically, increased CO2 may change the size and make-up of food insects for predators and/or interfere with parasitoid development synchronisation. Therefore, the complicated reactions of natural enemies to high CO2 were governed by the inconsistent responses of the insect preys. Oechalia schellenbergii (Heteroptera: Pentatomidae) that failed to reach adulthood took longer to develop on the elevated CO2-prey diet, which consisted of Helicoverpa armigera (Lepidoptera: Noctuidae) feeding on pea (Pisum sativum) foliage. However, those that were successful in eclosing were unaffected by the CO2 treatment in terms of development time, adult weight. When given Helicoverpa armigera of comparable size from various CO2 treatments, the Oechalia schellenbergii did not exhibit compensatory feeding. Apparently because these Helicoverpa armigera were smaller and simpler to control, the Oechalia schellenbergii functioned best when fed Helicoverpa armigera larvae from the elevated-CO2 condition. Overall, the findings suggest that increased CO2 may help generalist predators by increasing prey vulnerability, which would raise the risk of predation for pest species (Coll and Hughes 2008). Two generations of Chinese lacewings, Chrysopa sinica (Neuroptera: Chrysopidae), raised on cotton, and Gossypium hirsutum, growing in high CO2 environments, were studied for their growth and predatory abilities (double ambient vs. ambient). Higher ambient CO2 concentrations reduced the length of larval development, pupae survival rate, and weight of adult female Chrysopa sinica,

9.9 Elevated Atmospheric CO2

289

but had no discernible effects on the survival rate of the individual larval stages, adult female fecundity, rate of egg hatching, or adult lifetime. In situations with increased CO2, Chrysopa sinica larvae’s capacity to prey on Aphis gossypii in the third instar and during their whole larval life was much less effective. With various CO2 treatments, the quantity of aphids consumed by the Chrysopa sinica population of the first generation did not change much; however, the population of the second generation ingested noticeably less aphids with elevated CO2. We hypothesise that due to Chrysopa sinica diminished capacity to prey on Aphis gossypii in an environment with elevated CO2 concentrations, A. gossypii may become a more problematic pest (Gao et al. 2010). In cotton and wheat fields, predatory ladybirds like Harmonia axyridis and Propylaea japonica (Coleoptera: Coccinellidae) are common. Their primary food sources include lepidopteran pest eggs, immature larvae, spider mites, cotton and wheat aphids, and spider mites. Green lacewing Chrysopa sinica belongs to the predatory family Chrysopidae (Neuroptera). In China, cotton aphids are also mostly parasitised by Lysiphlebia japonica (Hymenoptera: Aphidiinae), which has the ability to significantly reduce cotton aphid populations in the early summer. When consuming cotton aphids, higher CO2 lengthened the development of Propylaea japonica but shortened the larval and pupal stages of Harmonia axyridis, Chrysopa sinica, and Lysiphlebia japonica (Gao et al. 2008). Additionally, increased CO2 decreased Chrysopa sinica predation while increasing Harmonia axyridis predation of aphids (Gao et al. 2010). Elevated CO2 prolonged the embryonic time but decreased mean relative growth rates and aphid consumption rates in Harmonia axyridis, a predator of wheat aphids. In open-top chambers, the effects of high CO2 (550 and 750 L/L vs. ambient CO2) on Harmonia axyridis, a predator in the third trophic level, were investigated. While the abundance of the parasitised aphid (Sitobion avenae) by A. picipes exhibited a considerable increase in 550 (12.5%) and 750 (19.6%) L/L CO2, respectively, compared to ambient CO2, the effect of higher CO2 on the growth and development of Harmonia axyridis was modest or non-existent. Additionally, under 750 L/L CO2 compared to ambient CO2, there was a substantial reduction (10%) in A. picipes’ emergence rate. Both the predator and the parasitoid significantly reduced aphid numbers, notably for A. picipes in environments with high CO2. Additionally, compared to those fed ambient CO2-grown wheat plants, Harmonia axyridis and A. picipes preferred to prey on/parasitise more Sitobion avenae aphids than those afflicted on the 550 (9.1% and 16.9%) and 750 (23% and 25.7%) L/L CO2-grown wheat plants. These preliminary findings suggest that increased CO2 significantly alters the predator/predilection parasites for preying on wheat aphids. Elevated CO2 can improve A. picipes biocontrol effectiveness against Sitobion avenae, but it also negatively affects the growth and development of A. picipes-parasitic Sitobion avenae (Jun Chen et al. 2007). The leaf-miner Dialectica scalariella (Lepidoptera: Gracillariidae) may take longer to grow in high CO2, exposing their larva longer to search by parasitoids or predators (Smith and Jones 1998), or attack rates from natural enemies may increase because larger mine areas draw more foes. Due to their apparent superior ability to

290

9

Influence of Ecological/Climatic Change

subdue the smaller bollworm larvae, predatory bugs appeared to be more effective under higher CO2 levels. The coccinellid predator Leis axyridis (Coccinellidae), which preys on the herbivorous aphid Aphis gossypii (Hemiptera: Aphididae), also ingested more prey when CO2 levels were higher (Chen et al. 2005). However, Chen et al. (2007) discovered that high CO2 had little of an effect on Harmonia axyridis, a coccinellid pest predator, on another aphid pest called Sitobion avenae (Hemiptera: Aphididae) (Coccinellidae). In a dual-choice experiment, plants raised under each treatment are then offered to gravid hoverfly females. Aphids raised under ambient or increased CO2 settings are additionally fed directly to emerging Episyrphus balteatus larvae, which are then regularly measured and weighed until pupation. Samples of the odours released by the plant–aphid interaction are taken. On plants cultivated in environments with ambient (450 ppm) CO2, more eggs are laid. However, there is no discernible difference between the two groups of predatory larvae raised at various CO2 concentrations, suggesting that the CO2 concentration has no impact on the calibre of the aphids they consume. The amount of aphid alarm pheromone on the plant– aphid relationship raised by the elevated CO2 condition (800 ppm) is lower, despite the fact that plant volatiles do not differ between ambient and elevated CO2-treated plants. This shows that altered semiochemical emissions caused by increased CO2 concentrations affect aphid predator oviposition behaviour (Boullis et al. 2018). The aphid predator Episyrphus balteatus fed either specialised or generalist aphids raised on either of two host plants under laboratory circumstances to examine the effects of elevated atmospheric CO2 levels on plants. Elevated CO2 enhanced the sinigrin content of the host plant and the specialist aphid in the sinigrin-containing host plant (black mustard), but it decreased the already extremely low levels in the generalist aphid. Elevated CO2 lengthened the time it took for a predator to grow while reducing fertility. Consequently, with rising atmospheric CO2, personal fitness marginally declined. Sinigrin severely slowed the predator’s development and reduced its fertility. Fitness suffered dramatically as a result. Plant and prey alone, as well as interactions between host plant prey type and CO2-prey type, had a substantial impact on the consumption rate. It will be easier to comprehend the nature and direction of their effects on natural enemies as a component of the tritrophic system with further research on the effects of climate change parameters (e.g. greenhouse gases such as CO2, ozone (O3), nitrogen dioxide (NO2), etc.) separately and jointly under controlled environmental conditions (Sadeghi Namaghi et al. 2018). On the tritrophic interactions of cowpea (Vigna unguiculata subsp. unguiculata), legume aphid Aphis craccivora, and coccinellid predator Menochilus sexmaculatus, there are direct and indirect impacts of temperature and elevated CO2 (eCO2). With an increase in temperature from 20 to 35 °C, there was a considerable reduction in development time (DT), an increase in reproductive rate (RR), and an increase in eCO2 over ambient CO2 (aCO2). The summation degree day (DD) for both the nymphal (232.9) and adult (247.1) phases at eCO2 and aCO2 across six temperatures showed a reduction in the mean degree day, DD requirement of both the nymphal (75.8) and adult stages (157.2) at eCO2 over aCO2. The steady increase in “rm” and

References

291

“R0” with temperature followed a non-linear trend, reaching maximum values at 27 ° C with shorter “T” throughout temperatures of 20–35 °C at eCO2, demonstrating a considerable difference in growth and development at the second trophic level. Menochilus sexmaculatus on Aphis craccivora showed decreased grub duration (23%) with increased predation capacity (19%) at eCO2 over ambient, indicating the incidence of Aphis craccivora is anticipated to be higher with enhanced predation in the future climate change scenario (Rao et al. 2018).

9.10

Conclusions

Low- and high-temperature tolerances/thresholds could be used in a way that would be helpful for BIPM mass production. Temperature tolerance has been provided by autochthonous microbes of the predators. Temperature tolerance has been modulated by macromolecules, hydrolytic enzymes, free radicals, and altered the permeability of exocuticles. Wind velocity has a negligible impact on controlling the predator population in the field. Cold tolerance paved the path for additional pest control.

References Abdel-Salam AH, Abdel-Baky NF (2000) Possible storage of Coccinella undecimpunctata (Col., Coccinellidae) under low temperature and its effect on some biological characteristics. J Appl Entomol 124:169–176 Ahmadi Z, Madadi H (2021) Life table, consumption rate and cold storage of Nabis pseudoferus (Hemiptera: Nabidae), the predator of agricultural crop pests. J Entomol Soc Iran 41(1):73–100 Ali AA, Watson TF (1978) Effect of temperature on development and survival of Zelus renardii. Environ Entomol 7(6):889–890 Ali A, Rashid MA, Huang QY, Wong C, Lei CL (2017) Response of antioxidant enzymes in Mythimna separate (Lepidoptera: Noctuidae) exposed to thermal stress. Bull Entomol Res 107: 382–390 Ananthakrishnan TN (1993) Bionomics of thrips. Annu Rev Entomol 38(1):71–92 Ardakani HR, Samih MA, Ravan S, Mokhtari A (2020) Effect of temperature on the development and predatory potential of Exochomus nigripennis (Erichson) (Col.: Coccinellidae) fed on Gossyparia spuria (Modeer) (Hem.: Eriococcidae). Int J Trop Insect Sci 40:723–728 Asghari F, Samie MA, Mahdian K, Basirat M, Izadi H (2011) Effects of temperature on some biological characteristics of Hippodamia variegata (Goeze) reared on common pistachio psylla Agonoscena pistaciae Burckhardt and Lauterer and on Angoumois grain moth, Sitotroga crealella Olivier, under laboratory conditions. Iran J Plant Prot Sci 42:137–149 Askari A, Stern VM (1972) Biology and feeding habits of Orius tristicolor (Hemiptera: Anthocoridae). Ann Entomol Soc Am 65(1):96–100 Awan MS (1988) Development and mating behaviour of Oechalia schellenbergii (GuérinMéneville) and Cermatulus nasalis (Westwood) (Hemiptera: Pentatomidae). Aust J Entomol 27(3):183–187 Bai Y, Quais M, Zhou W, Zhu ZR (2022) Consequences of elevated temperature on the biology, predation, and competitiveness of two mirid predators in the rice ecosystem. J Pest Sci 95(2): 901–916 Barman D, Roy J, Alam S (2022) Impact of wind in the dynamics of prey–predator interactions. Math Comput Simul 191:49–81

292

9

Influence of Ecological/Climatic Change

Berkvens N, Bale JS, Berkvens D, Tirry L, De Clercq P (2010) Cold tolerance of the harlequin ladybird Harmonia axyridis in Europe. J Insect Physiol 56:438–444 Bharadwaj RK (1966) Observations on the bionomics of Euborellia annulipes (Dermaptera: Labiduridae). Ann Entomol Soc Am 59(3):441–450 Bhoyar ES, Wadaskar RM, Dadhe CG, Sormare SG (2021) Influence of season on biology of predatory reduviid bug, Rhynocoris marginatus. J Entomol Zool Stud 9(6):259–262 Boullis A, Francis F, Verheggen F (2018) Aphid–hoverfly interactions under elevated CO2 concentrations: oviposition and larval development. Physiol Entomol 43(3):245–250 Braman SK, Yeargan KV (1988) Comparison of developmental and reproductive rates of Nabis americoferus, N. roseipennis, and N. rufusculus (Hemiptera: Nabidae). Ann Entomol Soc Am 81:923–930 Bueno VHP, Calixto AM, Montes FC, van Lenteren JC (2016) Reproduction and population parameters of the nearctic predator Geocoris punctipes at constant and varying temperature regimes. J Appl Entomol 140(5):323–333 Calixto AM, Bueno VH, Montes FC, Van Lenteren JC (2014) Development and thermal requirements of the nearctic predator Geocoris punctipes (Hemiptera: Geocoridae) reared at constant and alternating temperatures and fed on Anagasta kuehniella (Lepidoptera: Pyralidae) eggs. Eur J Entomol 111(4):521 Chauvin G, Hamon C, Vancassel M, Vannier G (1991) The eggs of Forficula auricularia L. (Dermaptera, Forficulidae): ultrastructure and resistance to low and high temperatures. Can J Zool 69(11):2873–2878 Chen FJ, Ge F, Parajulee MN (2005) Impact of elevated CO2 on tri-trophic interaction of Gossypium hirsutum, Aphis gossypii, and Leis axyridis. Environ Entomol 34:37–46 Chen FJ, Gang W, Megha N, Parajulee FG (2007) Impact of elevated CO2 on the third trophic level: a predator Harmonia axyridis and a parasitoid Aphidius picipes. Biocontrol Sci 17:313–324 Chen H, Zheng X, Luo M, Solangi GS, Wan F, Zhou Z (2018a) Effect of short-term hightemperature exposure on the life history parameters of Ophraella communa. Sci Rep 8:13969 Chen HS, Solangi GS, Guo JY, Wan FH, Zhou ZS (2018b) Antioxidant responses of ragweed leaf beetle Ophraella communa (Coleoptera: Chrysomelidae) exposed to thermal stress. Front Physiol 9:808–816 Cherry MJ, Barton BT (2017) Effects of wind on predator-prey interactions. Food Webs 13:92–97 Clercq D (2001) Functional response of the predators Podisus maculiventris (Say) and Podisus nigrispinus (Dallas)(Het., Pentatomidae) to the beet armyworm, Spodoptera exigua (Hübner) (Lep., Noctuidae): effect of temperature. J Appl Entomol 125(3):131–134 Cohen AC (1982) Water and temperature relations of two hemipteran members of a predator-prey complex. Environ Entomol 11(3):715–719 Coll M, Hughes L (2008) Effects of elevated CO2 on an insect omnivore: a test for nutritional effects mediated by host plants and prey. Agric Ecosyst Environ 123(4):271–279 Conradi-Larsen EM, Somme L (1973) Notes on the biology of Dolycoris baccarum L. (Het., Pentatomidae). Nor Entomol Tidsskr 20:245–247 Coudron TA, Ellersieck MR, Shelby KS (2007) Influence of diet on long-term cold storage of the predator Podisus maculiventris (Say)(Heteroptera: Pentatomidae). Biol Control 42(2):186–195 Das A, Mondal B, Chatterjee H (2021) Population dynamics of major insect pests and their predatory fauna in mulberry. Indian J Entomol 83(3):310–314 Davidson AT, Hamman EA, McCoy MW, Vonesh JR (2021) Asymmetrical effects of temperature on stage-structured predator–prey interactions. Funct Ecol 35:1041–1054 Dunbar DM, Bacon OG (1972) Influence of temperature on development and reproduction of Geocoris atricolor, G. pallens, and G. punctipes (Heteroptera: Lygaeidae) from California. Environ Entomol 1(5):596–599 El-Fakharany SK (2010) Population fluctuation of thrips, whitefly and associated predators in cabbage plantations as influenced by weather factors and toxic compounds. J Plant Prot Pathol 1(11):885–897

References

293

Elliott NC, Tao FL, Fuentes-Granados R, Giles KL, Elliott DT, Greenstone MH, Royer TA (2006) D-vac sampling for predatory arthropods in winter wheat. Biol Control 38(3):325–330 El-Sebaey EI (2007) Effect of cold storage on some biological characteristics of Coranis africana El-sebaey (heteroptefta: reduviidae). Egypt J Agric Res 85(2):475–487 Fabre JP, Menassieu P, Foing JJ, Chalon A (2000) Biology and ecology of Elatophilus nigricornis Zetterstedt (Hemiptera Anthocoridae) predator of Matsucoccus feytaudi Ducasse (Homoptera Matsucoccidae) in the South-East of France. Ann For Sci 57(8):777–792 Ferreira LF, Silva Torres CS, Venette RC, Torres JB (2020) Temperature and prey assessment on the performance of the mealybug predator Tenuisvalvae notata (Coleoptera: Coccinellidae). Aust J Entomol 59(1):178–188 Frank T, Bramböck M (2016) Predatory beetles feed more pest beetles at rising temperature. BMC Ecol 16(1):1–7 Gagne I, Coderre D (2001) Cold storage of Coleomegilla maculata larvae. Biocontrol Sci Tech 11: 361–369 Gao F, Zhu S, Sun Y, Du L, Parajulee MN, Kang L, Ge F (2008) Interactive effects of elevated CO2 and cotton cultuivar on tri-trophic interaction of Gossypium hirsutum, Ahpis gossyppii, and Propylaea japonica. Environ Entomol 37:29–37 Gao F, Chen F, Ge F (2010) Elevated CO2 lessens predation of Chrysopa sinica on Aphis gossypii. Entomol Exp Appl 135(2):135–140 Gullan PJ, Cranston PS (2005) The insects: an outline of entomology, 3rd edn. Blackwell, Malden Hamalainen M (1977) Storing dormant Coccinella septempunctata and Adalia bipunctata (Col., Coccinellidae) adults in the laboratory. Ann Agric Fenn 16:184–187 Hassanzadeh-Avval M, Sadeghi-Namaghi H, Fekrat L (2019) Factors influencing functional response, handling time and searching efficiency of Anthocoris minki Dohrn (Hem.: Anthocoridae) as predator of Psyllopsis repens Loginova (Hem.: Psyllidae). Phytoparasitica 47(3):341–350 Helsen H, Vaal F, Blommers L (1998) Phenology of the common earwig Forficula auricularia L. (Dermaptera: Forficulidae) in an apple orchard. Int J Pest Manag 44(2):75–79 Hong BM, Binh TTT, Hang VTT (2013) Effect of temperature on the life cycle and predatory capacity of ladybird beetle Micraspis discolor Fabricius (Coleoptera: Coccinellidae). Acad J Biol 35:37–42 Hori K (1986) Effect of photoperiod on nymphal growth of Palomena angulosa Motschulsky (Hemiptera: Peutatomidae). Appl Entomol Zool 21:597–605 Hori K (1987) Effects of stationary and changing photoperiods on nymphal growth of Palomena angulosa Motschulsky (Hemiptera: Pentatomidae). Appl Entomol Zool 22:528–532 Hori K (1988) Effects of stationary photoperiod on nymphal growth, feeding and digestive physiology of Palomena angulosa Motschulsky (Hemiptera: Pentatomidae). Appl Entomol Zool 23:401–406 Hori K, Inamura R (1991) Effect of stationary photoperiod on reproductive diapause, nymphal growth, feeding and digestive physiology of Eysarcoris lewisi Distant (Heteroptera: Pentatomidae). Appl Entomol Zool 26:493–499 Hormchan P, Schuster MF, Hepner LW (1976) Biology of Tropiconabis capsiformis. Ann Entomol Soc Am 69:1016–1018 Huang Z, Ren S, Musa PD (2008) Effects of temperature on development, survival, longevity, and fecundity of the Bemisia tabaci Gennadius (Homoptera: Aleyrodidae) predator, Axinoscymnus cardilobus (Coleoptera: Coccinellidae). Biol Control 46(2):209–215 Ingegno BL, Messelink GJ, Leman A, Sacco D, Tavella L (2021) Development and thermal activity thresholds of European mirid predatory bugs. Biol Control 152:104423 Islam Y, Shah FM, Shah MA, Musa Khan M, Rasheed MA, Ur Rehman S, Zhou X (2020) Temperature-dependent functional response of Harmonia axyridis (Coleoptera: Coccinellidae) on the eggs of Spodoptera litura (Lepidoptera: Noctuidae) in laboratory. Insects 11(9):583 Islam Y, Güncan A, Fan Y, Zhou X, Naeem A, Shah FM (2022) Age-stage, two-sex life table and predation parameters of Harmonia axyridis Pallas (Coleoptera: Coccinellidae), reared on

294

9

Influence of Ecological/Climatic Change

Acyrthosiphon pisum (Harris)(Hemiptera: Aphididae), at four different temperatures. Crop Prot 159:106029 James DG (1990) Development and survivorship of Biprorulus bibax (Hemiptera: Pentatomidae) under a range of constant temperatures. Environ Entomol 19(4):874–877 James DG (1992) Effect of temperature on development and survival of Pristhesancus plagipennis [Hem.: Reduviidae]. Entomophaga 37(2):259–264 Javahery M (1994) Development of eggs in some true bugs (Hemiptera–Heteroptera). Part I. Pentatomoidea. Can Entomol 126(2):401–433 Jensen K, Michaelsen JV, Larsen MT, Kristensen TN, Holmstrup M, Overgaard J (2018) Increased lipid accumulation but not reduced metabolism explains improved starvation tolerance in coldacclimated arthropod predators. Sci Nat 105(11):1–9 Jun Chen F, Wu G, Parajulee MN, Ge F (2007) Impact of elevated CO2 on the third trophic level: a predator Harmonia axyridis and a parasitoid Aphidius picipes. Biocontrol Sci Tech 17(3): 313–324 Khan MR (2010) The relationship between temperature and the functional response of Coccinella septempunctata (L.) (Coleoptera:Coccinellidae). Pak J Zool 42:461–466 Khan MR, Khan MR (2010) The relationship between temperature and the functional response of Coccinella septempunctata (L.) (Coleoptera: Coccinellidae). Pak J Zool 42(4):461–466 Khan J, Haq EU, Mehmood T, Blouch A, Fateh J (2016) Effect of temperature on the biology and predatory potential, of Harmonia dimidiata (Fab.) (Coleoptera: Coccinellidae) feeding on Myzus persicae (Sulzer) (Hemiptera: Aphididae) aphid. Int J Environ Agric Biotechnol 1: 342–349 Kharboutli MS, Mack TP (1993a) Effect of temperature, humidity, and prey density on feeding rate of the striped earwig (Dermaptera: Labiduridae). Environ Entomol 22(5):1134–1139 Kharboutli MS, Mack TP (1993b) Tolerance of the striped earwig (Dermaptera: Labiduridae) to hot and dry conditions. Environ Entomol 22(3):663–668 Khattab MA, Hegazy FH, Eissa GM, Khalafalla EM, Mesbah II (2019) Population fluctuation of some piercing-sucking insects and its relation to associated predators and the prevailing weather factors in soybean fields at Kafr el-Sheikh governorate. Egypt J Agric Res 97(1):147–158 Khodier IA, El-Khodary AS, Yousef AE, Mansour MR (2011) Detecting the predatory insects associated with the cotton aphid, Aphis gossypii glov. with special references to the efficiency of the predator Coccinella undecimpunctata L. AT Kafr EL-Sheikh Region. J Plant Prot Pathol 11: 1015–1022 Kiritani Y (1985) Effect of stationary and changing photoperiods on nymphal development in Carbula humerigera (Heteroptera: Pentatomidae). Appl Entomol Zool 20(3):257–263 Knabke J, Grigarick A (1971) Biology of the African earwig, Euborellia cincticollis (Gerstaecker) in California and comparative notes on Euborellia annulipes (Lucas). Hilgardia 41(7):157–194 Kontodimas DC, Milonas PG, Stathas GJ, Economou LP, Kavallieratos NG (2007) Life table parameters of the pseudococcid predators Nephus includens and Nephus bisignatus (Coleoptera: Coccinelidae). Eur J Entomol 104(3):407 Koppenhöfer AM (1995) Bionomics of the earwig species Euborellia annulipes in Western Kenya (Dermaptera: Carcinophoridae). Entomology Gen:81–85 Korenko S, Pekar S, Honek A (2010) Predation activity of two winter-active spiders (Araneae: Anyphaenidae, Philodromidae). J Therm Biol 35:112–116 Kruse PD, Toft S, Sunderland KD (2008) Temperature and prey capture: opposite relationships in two predator taxa. Ecol Entomol 33:305–312 Kumar A, Kumar A (2015) Effect of abiotic and biotic factors on incidence of pests and predator in cowpea [Vigna unguiculata (L.) walp.]. Legume Res 38(1):121–125 Lemos WP, Ramalho FDS, Zanuncio JC (2003) Age-dependent fecundity and life-fertility tables for Euborellia annulipes (Lucas)(Dermaptera: Anisolabididae) a cotton boll weevil predator in laboratory studies with an artificial diet. Environ Entomol 32(3):592–601

References

295

Li Y, Wang M, Gao F, Zhang H, Chen H, Wang M, Zhang L (2018) Exploiting diapause and cold tolerance to enhance the use of the green lacewing Chrysopa formosa for biological control. Biol Control 127:116–126 Logan JD, Wolesensky W, Joern A (2006) Temperature-dependent phenology and predation in arthropod systems. Ecol Model 196:471–482 Mansfield S, Scholz B, Armitage S, Johnson ML (2007) Effects of diet, temperature and photoperiod on development and survival of the bigeyed bug, Geocoris lubra. Biocontrol 52(1):63–74 Marcelo HL, Tania ZS (2002) Animal response to drastic changes in oxygen availability and physiological oxidative stress. Comp Biochem Phys C 13:537–556 Marquereau L, Cottineau JS, Fontaine O, Chiroleu F, Reynaud B, Delatte H (2022) Life history parameters and predation capacities of Nesidiocoris volucer: a new biological control agent for tomato crop. Bull Entomol Res 112:584. https://doi.org/10.1017/S0007485321001164 Melamed Madjar V (1971) Bionomics and ecology of the earwig anisolabis (Euborellia) annulipes Luc.(Labiduridae Dermaptera) in Israel 1. Zeitschrift für Angew Entomol 69(14):170–176 Mouly R, Shivananda TN, Verghese A (2018) Effect of temperature on minor invertebrate predator reduviid Isyndus heros (Fab.) (Hemiptera: Reduviidae). Curr Sci 115(5):983–986 Murdoch WW (1973) The functional response of aphid predators. J Appl Ecol 10:335–342 Naranjo SE, Ellsworth PC (2005) Mortality dynamics and population regulation in Bemisia tabaci. Entomol Exp Appl 116(2):93–108 Nedved O, Fois X, Ungerova D, Kalushkov P (2013) Alien vs. predator—the native lacewing Chrysoperla carnea is the superior intraguild predator in trials against the invasive ladybird Harmonia axyridis. Bull Insectol 66:73–78 Neiswander CR (1944) Ring-legged earwig, Euborellia annulipes (Lucas). Ohio Exp Agric Station Bull 648:15 Nunes GDS, Paulo HHD, Dias WSR, Bortoli SAD (2020) Feeding preference of Euborellia annulipes to Plutella xylostella: effects of temperature and prey development stage. Pesqui Agropecu Trop 50. https://doi.org/10.1590/1983-40632020v5057778 Nunes GDS, Dantas TAV, Figueiredo WRS, Oliveira RD, Batista JDL, Vacari A, Bortoli SAD (2022) The life history of Euborellia annulipes (Lucas) (Dermaptera: Anisolabididae) fed onlarvae and pupae of Plutella xylostella (L.) (Lepidoptera: Plutellidae). Turk J Zool 46 (2):175–185 Oida H, Kadono F (2012) Development of Geocoris varius and G. proteus (Hemiptera: Geocoridae) provided with Ephestia kuehniella (Lepidoptera: Pyralidae) eggs. Appl Entomol Zool 47(4):365–372 Pakyari H, McNeill MR (2020) Effects of photoperiod on development and demographic parameters of the predatory thrips Scolothrips longicornis fed on Tetranychus urticae. Bull Entomol Res 110(5):620–629 Pakyari H, Fathipour Y, Enkegaard A (2011) Effect of temperature on life table parameters of predatory thrips Scolothrips longicornis (Thysanoptera: Thripidae) fed on twospotted spider mites (Acari: Tetranychidae). J Econ Entomol 104(3):799–805 Pang H, Ren S, Zeng T, Pang X (2004) The species diversity and application of coccinellids in China. Guangdong Science and Technology Press, Guangzhou Panja P (2022) Impacts of wind and anti-predator behaviour on predator-prey dynamics: a modelling study. Int J Comput Sci Math 15(4):396–407 Papanikolaou NE, Milonas PG, Kontodimas DC, Demiris N, Matsinos YG (2013) Temperaturedependent development, survival, longevity, and fecundity of Propylea quatuordecimpunctata (Coleoptera: Coccinellidae). Ann Entomol Soc Am 106:228–234 Peng Z, Zhang X, Gu X, Meng F, Zhang L, He Y, Xian J (2022) Influence of temperature on the development, reproduction, and predation of the predatory stink bug, Eocanthecona furcellata (Wolff) (Heteroptera: Pentatomidae). SSRN: https://ssrn.com/abstract=4078780, https://doi. org/10.2139/ssrn.4078780

296

9

Influence of Ecological/Climatic Change

Perdikis D, Lykouressis D (2000) Effects of various items, host plants, and temperatures on the development and survival of Macrolophus pygmaeus Rambur (Hemiptera: Miridae). Biol Control 17(1):55–60 Perdikis DC, Lykouressis DP, Economou LP (2004) Influence of light-dark phase, host plant, temperature, and their interactions on the predation rate in an insect predator. Environ Entomol 33(5):1137–1144 Pérez-Hedo M, Gallego C, Roda A, Kostyk B, Triana M, Alférez F, Urbaneja A (2021) Biological traits of the predatory mirid Macrolophus praeclarus, a candidate biocontrol agent for the Neotropical region. Bull Entomol Res 111(4):429–437 Péricart J (2010) Hémiptères pentatomoidea euro-mediterranéens (Bd. 3) Podopinae et Asopinae. Faune de France 93:290 Pervez A, Omkar GAK (2006) Ecology and biological control application of multicoloured Asian ladybird, Harmonia axyridis: a review. Biocontrol Sci Tech 16:111–128 Porras MF, Agudelo-Cantero GA, Santiago-Martínez MG, Navas CA, Loeschcke V, Sørensen JG, Rajotte EG (2021) Fungal infections lead to shifts in thermal tolerance and voluntary exposure to extreme temperatures in both prey and predator insects. Sci Rep 11(1):1–9 Rabitsch W (2008) Alien true bugs of Europe (Insecta: Hemiptera: Heteroptera). Zootaxa 1827:1– 44 Rani PU (1992) Temperature-induced effects on predation and growth of Eocanthecona furcellata (Wolff)(Pentatomidae: Heteroptera). BioControl 6(2):72–76 Rankin SM, Palmer JO, Larocque L, Risser AL (1995) Life history characteristics of ringlegged earwig (Dermaptera: Labiduridae): emphasis on ovarian development. Ann Entomol Soc Am 88(6):887–893 Rao MS, Shaila O, Sreelakshmi P, Vennila S, Vanaja M, Rao AS, Reddy KS (2018) Tritrophic interactions of cowpea [Vigna unguiculata subsp unguiculata (L.)], aphids [Aphis craccivora (Koch)] and Coccinellids [Menochilus sexmaculatus (Fab.)] under eCO2 and eTemp. J Asia Pac Entomol 21(2):531–537 Rathee M, Ram P (2018) Retracted article: impact of cold storage on the performance of entomophagous insects: an overview. Phytoparasitica 46(4):421–449 Rebolledo R, Villegas G, Klein C, Aguilera A (2005) Fluctuación poblacional, capacidad depredadora y longevidad de Nabis punctipennis Blanchard (Hemiptera: Nabidae). Agric Tec (Chile) 65:442–446 Ruberson JR, Bush L, Kring TJ (1991) Photoperiodic effect on diapause induction and development in the predator Orius insidiosus (Heteroptera: Anthocoridae). Environ Entomol 20:786–789 Sadeghi Namaghi H, Amiri Jami A, Shoor M (2018) Aphid suitability for the predatory hoverfly Episyrphus balteatus altered with elevating atmospheric CO2 and sinigrin. Entomol Sci 21(1): 12–21 Sahayaraj K, Paulraj MG (2001) Effect of cold storage on egg hatching in two reduviid predators Rhynocoris marginatus Fab. and R. fuscipes Fab. (Hemiptera Reduviidae). Belg J Entomol 3: 201–207 Sahayaraj K, Sujatha S (2011) Temperature-dependent biology & physiology reduviids. Nova Science, Hauppauge, p 146 Sahayaraj K and Sujatha S (2012) Temperature-dependent biology and physiology of reduviids. Nova Science Publishers, New York, pp. 146 Salcedo C, Iannacone J, Alvariño L (2020) Ciclo biológico de Nabis consimilis Reuter, 1912 (Heteroptera: Nabidae) bajo condiciones de laboratorio. Revista Electrónica de Divulgación de Metodologías emergentes en el desarrollo de las STEM 2(2):3–21 Saunders DS (1983) A diapause induction-termination asymmetry in the photoperiodic responses of the linden bug, Pyrrhocoris apterus and an effect of near-critical photoperiods on development. J Insect Physiol 29(5):399–405 Seat J, Vujić M, Nadaždin B (2019) New faunal data on true bugs (Heteroptera) in Serbia. Acta Entomol Serbica 24(1):95–99

References

297

Seat J, Nadaždin B, Sciban M (2020) Acrosternum heegeri (Heteroptera: Pentatomidae) in Serbia. Acta Entomol Serbica 25(1):77–81 Sentis A, Hemptinne JL, Brodeur J (2012) Using functional response modelling to investigate the effect of temperature on predator feeding rate and energetic efficiency. Oecologia, Online first, 24 Jan 2012. https://doi.org/10.1007/s00442-012-2255-6 Seo M, Kim JH, Seo BY, Park B, Lee SG, Kim KH, Cho JR (2019) Effect of cold storage on quality of Geocoris pallidipennis and Micromus angulatus. Korean J Appl Entomol 58(2):151–157 da Silva Nunes G, Truzi CC, Cardoso CP, Vieira NF, Ramalho DG, de Souza JM, De Bortoli SA (2020) Temperature-dependent functional response of Euborellia annulipes (Dermaptera: Anisolabididae) preying on Plutella xylostella (Lepidoptera: Plutellidae) larvae. J Therm Biol 93:1–5. https://doi.org/10.1016/j.jtherbio.2020.102686 Simmons AM, Yeargan KV (1988) Development and survivorship of the green stink bug, Acrosternum hilare (Hemiptera: Pentatomidae) on soybean. Environ Entomol 17(3):527–532 Siscaro G, Lo Pumo C, Tropea Garzia G, Tortorici S, Gugliuzzo A, Ricupero M, Zappalà L (2019) Temperature and tomato variety influence the development and the plant damage induced by the zoophytophagous mirid bug Nesidiocoris tenuis. Pest Manag Sci 92(3):1049–1056 Situmorang J, Gabriel BP (1988) Biology of two species of predatory earwigs Nala lividipes (Dufour) (Dermaptera: Labiduridae) and Euborellia (Euborellia) annulata (Fabricius) (Dermaptera: Carcinophoridae). Philipp Entomol 7(3):215–238 Smith PHD, Jones TH (1998) Effects of elevated CO2 on the chrysanthemum leafminer, Chromatomyia syngenesiae: a greenhouse study. Glob Change Biol Bioenergy 4:287–291 Soares AO, Coderre D, Schanderl H (2003) Effect of temperature and intraspecific allometry on predation by two phenotypes of Harmonia axyridis Pallas (Coleoptera: Coccinellidae). Environ Entomol 32(5):939–944 Solangi GS, Karamaouna F, Kontodimas D, Milonas P, Lohar MK, Abro GH, Mahmood R (2013) Effect of high temperatures on survival and longevity of the predator Cryptolaemus montrouzieri Mulsant. Phytoparasitica 41:213–219 Sparkes J (2012) Adaptive foraging behaviour and impact on longevity of an omnivorous insect in response to the extreme temperature fluctuation events associated with climate change. Electronic Theses and Dissertations, University of Windsor, 4843, p 110. https://scholar.uwindsor. ca/etd/4843 Sun YX, Hao YN, Liu CZ, Wang SS (2019) Artificial diet is fruitful pre-storage nutrition for longterm cold storage of laboratory-reared Harmonia axyridis (Pallas) adults. Biol Control 139: 104075 Tawfikm FS, Awadallaht (1983) Effect of temperature and relative humidity on various stages of the predator, Allaeocranum biannulipes (Mortr. et. sign.) (Hemiptera: reduviidae). Bull Entomol Soc Egypt 64:239–250 Terblanche JS, Clusella-Trullas S, Deere JA (2009) Directional evolution of the slope of the metabolic rate-temperature relationship is correlated with climate. Physiol Biochem Zool 82: 495–503. https://doi.org/10.1086/605361 Van den Meiracker RAF (1994) Induction and termination of diapause in Orius predatory bugs. Entomol Exp Appl 73(2):127–137 Vennison SJ, Ambrose DP (1990) Biology and behaviour of the Assassin bug, Ectomocoris vishnu Distant (Hemiptera: Reduviidae: Piratinae) from South India. J Soil Biology Ecol 10(2):89–102 Walker R, Wilder SM, González AL (2020) Temperature dependency of predation: increased killing rates and prey mass consumption by predators with warming. Ecol Evol 10:9696–9706 Wang S, Tan XL, Guo XJ, Zhang F (2013) Effect of temperature and photoperiod on the development, reproduction, and predation of the predatory ladybird Cheilomenes sexmaculata (Coleoptera: Coccinellidae). J Econ Entomol 106(6):2621–2629 Wang B, Liu Y, Wang GR (2017) Chemosensory genes in the antennal transcriptome of two syrphid species, Episyrphus balteatus and Eupeodes corollae (Diptera: Syrphidae). BMC Genomics 18:586

298

9

Influence of Ecological/Climatic Change

Wang GF, Dong ZK, Li LL, Lu ZB, Li C, Men XY (2021) Intraguild interference of spiders and ladybirds in a cotton field: evidence from their active time and aphid control efficiency. Entomol Res 51:533–539 Xia PL, Wang B, Xie XF, Feng Y, Huang Y (2022) Effect of temperature on survival and immature development of Arma chinensis. J Asia Pac Entomol 25(3):101927 Yamasaki Y, Yoshioka K, Takeuchi F (1983) Bionomics and predation of Scolothrips sp. Proc Assoc Pl Prot Shikoku 18:83–86 Yang F, Wang Q, Wang DM, Xu B, Xu JX, Lu YH, Harwood JD (2017) Intraguild predation among three common Coccinellids (Coleoptera: Coccinellidae) in China: detection using DNA-based gut-content analysis. Environ Entomol 46:1–10 Yang Q, Liu J, Wyckhuys KA, Yang Y, Lu Y (2022) Impact of heat stress on the predatory ladybugs Hippodamia variegata and Propylaea quatuordecimpunctata. Insects 13(3):306 Yao DU, Chun-Sen MA, Zhao QH, Gang MA (2007) Effects of heat stress on physiological and biochemical mechanisms of insects: a literature review. Acta Ecol Sin 27:1565–1572 Yasir I, Farhan MS, Xu RB, Muhammad R, Miao YB, Li XH, Zhou XM (2021) Functional response of Harmonia axyridis preying on Acyrthosiphon pisum nymphs: the effect of temperature. Sci Rep 11:13565 Zhang S, Jiang Y, Shen R (1980) The brief report of the research about Menochilus sexmaculata (Coleoptera: Coccinellidae). Nat Enemy 4:13Ð16 Zhu L, Ge Z, Gong Y, Shi B, Wang S, Wei S (2015) Effects of temperature on predation of the thrips Echinothrips americanus (Thysanoptera: Thripidae) by the predatory bug Orius sauteri (Heteroptera: Anthocoridae). Acta Photonica Sin 42(2):229–236

Predation Ethology of Various Orders

10

Contents 10.1 10.2 10.3

10.4 10.5 10.6 10.7 10.8 10.9

10.10

10.11

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Characteristics of Good Predator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.3.1 Lepidoptera as a Prey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.3.2 Collembolans as a Prey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.3.3 Millipedes as Prey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.3.4 Beetle and Weevil as Prey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.3.5 Group Feeding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Coleopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Isopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Comparison Between Coleoptera and Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Orthoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hemiptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.1 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.2 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.3 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.4 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.5 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.6 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.7 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9.8 Diptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Expression of Bt Toxins Through Their Preys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.10.1 In Cotton-Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.10.2 In Rice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.10.3 Geocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.10.4 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.10.5 Thysanura . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Intraguild Predation (IGP) or Competition Dominates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_10

300 301 302 303 303 304 304 304 305 306 307 308 309 309 309 315 316 318 321 321 324 324 325 326 327 328 328 329 332

299

300

10 Predation Ethology of Various Orders

Best Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.12.1 Miridae and Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.12.2 Coleoptera and Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.12.3 Coleoptera and Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.12.4 Coleoptera and Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.12.5 Coleoptera, Hemiptera, and Diptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.12.6 Among Hemipterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.13 Role of Molecular and Other Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.14 Anti-Predatory Acts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.15 Chemical Ecology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.15.1 Field Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.16 Future Recommendation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.12

10.1

333 333 334 334 335 335 336 337 339 341 345 347 347

Introduction

A natural enemy’s (parasitoid or predator’s) foraging behaviour refers to the process through which it looks for oviposition sites and food supplies for its survival, growth, and reproductive success. In Hymenoptera, predatory behaviour in the strictest sense—where the predator attacks and eats its own prey—is proportionally less significant. This behaviour is limited to adult feeding strategies like host feeding, which can be fatal in a few ichneumonids and dryinids, and direct predation by adult tenthredinids hunting other insects for self-nutrition. In Hymenoptera, adult wasps operate as provisioning predators and prey primarily on larvae, eating on them. Undoubtedly, provisioning predation has developed from the parasitoid style of life, and one of the essential aspects of this change is the host’s relocation. Only the aculeate Hymenoptera exhibit host relocation, a set of behaviours involving the migration of the host (or prey) from the location where it was initially discovered and captured to a hidden site. In general, increasing the biological management of herbivorous pests requires a better understanding of the foraging behaviour of natural enemies. Definitions Predation has been described as a competitive biological relationship in which a consumer (predator) feeds on a target organism (prey), which is always alive when the predator first attacks it. Typically, the prey dies as a result of this interaction. Predator often kills its victim shortly after attacking it, and over the course of its career, the predator will murder multiple prey individuals, frequently swallowing them in their entirety.

10.2

10.2

Characteristics of Good Predator

301

Characteristics of Good Predator

1. The significance of vision: The existence of compound eyes shows that predatory and parasitic insects place a high value on vision. The eyesight of these same arthropods has, however, been examined far less frequently, despite the fact that olfaction is very well understood for many taxa. Using information from recent studies and publications, Lim and Ben-Yakir (2020) evaluated and summarised the visual sensory systems of parasitoids and predatory arthropods. It is well known that many predatory insects, including ladybird beetles, utilise visual cues together with chromatic sensitivity and geometric shape perception during foraging. Due to their weak eyes, hymenopteran parasitoids rely more on olfactory signals than other senses. While certain species can even distinguish between shape and size, the majority of them have trichromatic vision. Arthropods that are predators and parasites need a combination of cues rather than just eyesight to find their prey or hosts (Lim and Ben-Yakir 2020). Among the visual cues are: • Size and location of the prey or their colony; • Form and colour of the plants; • Size and density of the plant patch; • Form and movement of the plants; • Shape and position of the preys. • The aroma of plants. • The smell of the plants’ related preys. 2. Mimicry: A predator or parasite mimics a signal from a different species in order to take advantage of the signal’s recipient. In some of the most astounding instances, a predator species copies the sophisticated sexual signals of its victim (aggressive mimicry). 3. The prey’s defensive secretions: Olfactory stimulation from the prey’s defensive secretions seems to improve the response and aid in the continuation of feeding (Haridass and Ananthakrishnan 1980). 4. Movement/mobility: Reduviids’ predatory behaviour is influenced by moving visual and olfactory inputs (Parker 1965). 5. Prey (hosts): Consume and/or kill as many of them as necessary in order to satisfy their hunger. 6. Group feeding: Most hemipteran predators adopt this strategy because it increases per capita food intake while decreasing the cost of capture. Group hunting followed by food sharing boosts participants’ fitness (Alain et al. 2013). 7. Organs/body parts: The Harpactorinae, which feed on caterpillars, paralyse and kill their prey by stabbing it with their stylets and injecting it with saliva two to three times in quick succession. The mandibular and maxillary stylets of Reduviidae have intricate barbs, teeth, and tubercles that show a wide range of alterations and evolutionary advancement (Cobben 1978). Antennal and labial tip sensilla (Sensilla trichodea, Multilobed sensilla, and Sensilla basiconica) were responsible for the predation behaviour of Pentatomidae. The salivary system of reduviids is also very complex (Haridass 1978), and the anterior lobes of the main

302

10 Predation Ethology of Various Orders

glands are concerned with the secretion of neurotoxic substances involved in the paralysis and death of the prey (Haridass 1978). 8. Organs-Raptorial forelegs: Raptorial forelegs are structures that are used to catch prey and have undergone numerous evolutionary cycles in a variety of insect lineages. Currently, predatory insects like mantid lacewings (Neuroptera), mantises (Mantodea), water scorpions (Hemiptera), and dance flies have raptorial forelegs (Diptera). 9. Scheduling of time for different activities: Scheduling of an animal’s time and effort for foraging involves a series of fundamentally economic decisions, judgments that depend on the quantity of investment capital available and the expected return from each of the distinct behavioural paths. Predators spend a significant amount of time capturing, paralysing, and sucking prey. A predator seeks to capture, paralyse, suckle, and kill a larger quantity of prey in the shortest amount of time possible.

10.3

Hymenoptera

Hymenoptera predatory behaviour In Hymenoptera, predation primarily occurs for larval feeding and, less frequently, for adult self-nutrition. Few reports of direct predation by larvae are known in many families of otherwise typical parasitoids. Most occurrences of predation for larval nutrition are carried out by adult wasps acting as provisioning predators (Godfray 1994). Only social vespids, which malaxate their prey before giving it to the larvae (Spradbery 1973), and some solitary eumenine and apoid wasps, which bite or pierce the prey to suck up fluids, occasionally also partially consuming prey tissue (Cowan 1991; O’Neill 2001), appear to be capable of adult self-nutrition derived from provisioning predation. Tenthredinid sawflies and a few apocritan parasitoid wasps that prey directly on their hosts are known to engage in a very small amount of adult predatory feeding that is unrelated to larval provisioning. Adolescent predatory feeding. Although several adult sawflies in the Tenthredinidae family have been seen feeding on other insects, basal hymenopteran lineages contain phytophagous larvae (Iwata 1976). The direct hunting and feeding of several holometabolous insects by at least 63 species in 27 genera has been documented. These holometabolous insects include adult flies (Bibionidae, Calliphoridae, Empididae, Fanniidae, Scatophagidae, Stratiomyidae, Syrphidae, Tachinidae, and Tipulidae) and adult sawflies (Iwata 1976). Another notable instance of adult predation is documented in one Ichneumonidae species, which harbours a variety of moth larvae (Leius 1961). Adult females must first consume some host tissue, typically haemolymph, in order to oviposit. If not for the fact that extremely small hosts might be completely devoured by the females during the feeding process, this could be merely considered an example of host feeding (Jervis and Kidd 1986). Another family where adult females are known to prey on their auchenorrhynchous hosts is the Dryinidae (Clausen 1962; Waloff 1974). Three species’ captive females have reportedly been observed feeding partially on the hosts they are living on before either discarding them or parasitising them (Waloff 1974).

10.3

Hymenoptera

303

Fig. 10.1 Young hemipteran and lepidopteran species are preyed upon by hymenopteran predators. A carabronid wasp was eating an aphid pest (a), a yellow spider was feeding a caterpillar (b), and a carabronid wasp was carrying a caterpillar (c)

Typically, they ate the day’s first host to be seized. These instances of adult predatory eating are supported by either anecdotal field observations or by actions taken during experiments. Although it appears that Ichneumonidae and Dryinidae behave similarly in nature, it must be kept in mind that laboratory conditions may have influenced the outcomes seen. Unfortunately, there is not enough evidence on Hymenoptera to assess the significance and effects of this kind of predation in the field. Predatory behaviour varies between species as well as between prey and prey (Fig. 10.1). The quality of the food, the level of competition, the distribution of food resources through time and place, the risk of predation or parasitisation, and the distribution of food resources are all factors that affect ant foraging behaviour (Csata and Dussutour 2019).

10.3.1 Lepidoptera as a Prey Oecophylla smaragdina (Hymenoptera: Formicidae) on Pteroma pendula (Lepidoptera: Psychidae) had three key stages in its life cycle. The gaster was raised at a 90° angle during stage 1 when the ant detected a prey (a larva or pupa of Pteroma pendula) from a distance. The ant then approached the victim and bit it (Stage 2) before spraying formic acid onto the Pteroma pendula (Stage 3; this is distinctively recognisable, with the elevated gaster arching forwards). As a result, the prey is neutralised, hoisted (Stage 4), and then carried to the nest (Stage 5) (Pierre and Idris 2013).

10.3.2 Collembolans as a Prey Pyramica benten, a short-mandibulate variety, ambushes its victim without going near it in their hunting activity, capturing it close to the mouthparts (collembolans of Entomobryidae). The likelihood of success in this covert hunting is increased by attracting the target by body smearing with organic material. Pyramica and strumigenys are closely related, although strumigenys are long-mandibulate forms.

304

10 Predation Ethology of Various Orders

On Collembola, both genera are species of specialised predators. These dacetines and the coleopterans are contrasted for their hunting strategies and species diversity because Coleoptera is another group that has evolved specialist predators on Collembola (Masuko 2009).

10.3.3 Millipedes as Prey Based on field observations of Thaumatomyrmex ants (Hymenoptera: Formicidae) feeding on millipedes of the family Polyxenidae and a field and laboratory study of Thaumatomyrmex contumax. Before feeding, Thaumatomyrmex grabs its prey and removes its body-covering setae. Thaumatomyrmex exhibits this specific behaviour in at least two species, which are members of two different species-groupings (Brandão et al. 1991).

10.3.4 Beetle and Weevil as Prey When feeding on Myopopone castanea (Hymenoptera: Formicidae) ants, Oryctes rhinoceros (Coleoptera: Scarabaeidae) larvae displayed colour changes in the cuticle. It starts to turn brownish before eventually turning black. Ants kill their prey by biting and stinging them. The haemolymph liquid is next consumed and sucked from the larvae until only the cuticle is left. The ant’s bites tore the prey’s body to pieces. Even first and second instar larvae, which in the case of the prey larvae just left the mandible, can be completely consumed by ants. The cuticle was becoming darker and wrinkled as the haemolymph had been sucked out by the ants, and the third instar is the largest. Amblyopone family ants often feed on and drain haemolymph from their victim. For their brood or colony, these ants typically bring the parts of their prey back to the nest. However, Myopopone castanea, which preys on Oryctes rhinoceros, operates differently. Myopopone castanea ants bring their larvae to the body after the prey has died, and they have haemolymph from the animal together (Tobing and Kuswardani 2018). Near the anthesized spikelets in the Serting oil palm plantation, a big-headed ant (Pheidole megacephala, Formicidae) feasts on an adult pollinating the weevil Elaeidobius kamerunicus (Muhammad Luqman et al. 2018).

10.3.5 Group Feeding The ants (Azteca cf. lanuginosa) ambush their prey by simultaneously attacking insects that descend on the upper surface of the leaf while they are positioned in sitand-wait groups along the leaf margin. Beetles, bees, and butterflies, most of which are two times longer than ants, can be captured by the workers (Morais 1994).

10.4

10.4

Coleopterans

305

Coleopterans

Pharoscymnus numidicus has a greater seeking ability as the larvae get older, as measured by mobility and survival rates. The predator’s searching behaviour is haphazard; it does not turn towards its prey, and physical circumstances like light might affect the predator’s movement’s direction. When there is a natural food shortage, the predator’s probability of survival is increased if it can eat other foods. Its cannibalistic habit is crucial to its survival in this regard. The eating capacity is influenced by the larvae’s instar, sex, history of food consumption, air temperature, and host insect population density (Kehat 1968). Every insect has a different predation method. Cuese and distance were the two strategies used by the diving beetle to catch the victim. Two instars of the predaceous diving beetle larva, Dytiscus verticalis, were studied to determine the reactive distance and cues employed to find prey. Regardless of the sort of prey-cue applied, the earlier instar larvae showed shorter response distances than the later instar larvae. The impact of distance varied depending on the cue’s type and the distance to the beetle larvae. These beetle larvae also appeared to use chemical and tactile cues in addition to visual signals to find and capture prey. The complete attack sequence from the beetle larvae seems to require chemical cues (Formanowicz 1987). In order to investigate the prey detection and predatory behaviour and strategies of 12 species of Carabidae and one species of Staphylinidae, Wheater (1989) used video and orientation techniques. The species that were looked at included Pterostichus madidus, Pterostichus melanarius, Pterostichus niger, Abax parallelepipedus, and Staphylinus olens. Other species included Cicindela campestris, Cychrus caraboides, Carabus problematicus, Carabus violaceus, Calosoma maderae, Nebria complanata, and Scarites abbreviates. All of the investigated species were seen to react to prey when they came into contact with it (either through tactile or gustatory reception) while moving about. Orientation chamber tests revealed that some animals would turn towards prey even in the absence of interaction. Some creatures used vision to find their way (Carabus campestris, Carabus maderae, Scarites abbreviates, and Abax parallelepipedus). When tested with moving prey, this happens most frequently. In some, orienting to prey was place while olfactory signals were present (Pterostichus madidus, Pterostichus melanarius, Pterostichus niger and Abax parallelepipedus). The Pterostichini family included all of the species that used this technique of prey detection, and it appears that the receptors involved are located on the antennae. Some of these species were discovered to react to slug mucus even though some of them did not orient towards prey in the absence of touch cues (Carabus caraboides, Carabus problematicus, Carabus violaceus and Scarites abbreviates). The terminal extremities of the palps are likely where the receptors for this are located (Wheater 1989). When a feeding beetle first comes across a prey item in a patch, it often accelerates its search behaviour for a predetermined “give-up” time period. Finding a prey initiates a three-dimensional search activity in certain species that hunt in two dimensions. Once the prey is found, animals usually transition to a specific

306

10 Predation Ethology of Various Orders

prey-catching behaviour. In this stage of feeding, numerous physical and behavioural adaptations are in operation, particularly in specialised species. Unique arrays of adaptations involving sight, behaviour, and anatomy in both adults and larvae have been discovered via careful study of prey capture in numerous European species that hunt springtails. The majorities of carabid adults kills and dismember prey with the aid of their well-developed mandibles. Specialist species that prey on snails appear to immobilise their victim by biting, blocking the slugs’ defence response of producing mucus. The development of terrestrial wildlife! Predators of carabids, including mammals, birds, amphibians, and reptiles, have likely been an important driver in the evolution of these creatures. Morphological, physiological, and behavioural defensive mechanisms make up the extensive arsenal of antipredator mechanisms. For instance, cryptic or warning colours, mimicry, a small body form, dorso-ventral flattening, large eyes, and long legs are morphological characteristics of arboreal carabids. Beetles that are inactive rest in protected areas, such as under stones, in soil crevices, or on the undersides of leaves; night activity is also suggested to be a predator defence. Beetles that have been attacked flee and hide, dive into water, exhibit catalepsy, spew crop contents or digestive fluid, and then bite the assailant. Stridulation is another common and powerful deterrent. Many carabid species have noticeable elytral patches that may serve to divert attacks away from the important anterior body sections. Carabids have been found to exhibit Batesian mimicry, and tiger beetles have been found to exhibit Mullerian mimicry. Large species also have fused elytra and toughened cuticles that act as structural defences against predators. The most effective defence is the secretion of substances from carabids’ pygidial glands, which are always present (Lovei and Sunderland 1996). Agonum dorsale, Prerostichus melanarius, and Harpalus rufipes, three carabid predators, did not distinguish between healthy and ill larvae of the cabbage moth Mamestra brussicae (Lepidoptera: Noctuidae) as prey (Vasconcelos et al. 1996). Exochomus flaviventris (Coleoptera: Coccinellidae) is a polyphagous predator that is indigenous to sub-Saharan Africa. In Central Africa, it is the most active predator of the cassava mealybug Phenacoccus manihoti (Homoptera: Pseudococcidae), a major pest of cassava Manihot esculenta (Euphorbiaceae). However, cassava–mealybug and mealybug–predator bitrophic systems have been studied extensively both in the laboratory and in the field. Le Rü and Makaya (2001) studied the interaction of cassava odor, mealybug, plant–mealybug complex, and plant–mealybug–natural enemy complex mealybug predator Exochomus flaviventris olfactory responses. The two species of rove beetles behaved in quite different ways. Tachyporus hypnorum was primarily found on the plants, but Drusilla canaliculata was almost solely found on the ground. Tachyporus hypnorum caused more aphids to fall to the ground, which led to a larger rise in the development of the winged morph (Balog et al. 2013).

10.5

Isopterans

Ants and termites interact in a variety of ways, including cohabitation, mutualism, competition for breeding sites, and—perhaps most significantly—as predators and prey. Even if geographical hurdles, physical barriers (termite mounds and sheetings),

10.6

Dermaptera

307

or partitioning of microhabitats prevent them from ever meeting in nature, termites and their developmental stages are likely to constitute adequate prey for the majority of ant species. Because of the presence of a good amount of lipids, proteins, minerals, carbohydrates, and micronutrients can be found in termites. On instance, canopy ants like Polyrhachis ypsilon forage and nest in the forest canopy whereas wood/soil-eating termites like Dicuspiditermes nemorosus search for substrates rich in organic materials underground (Tuma et al. 2020). Ant species that are specialised predators of other arthropods are prone to engage in opportunistic predation on termites. If given the chance, several ant species may feed on exposed termites, such as when they come across termite individuals while foraging. Therefore, several ant species with different eating patterns are probably going to occasionally feed on termites (Holldobler and Wilson 1990). The generalist ant species Pheidole spp. and Camponotus spp. are the most well-known and frequently mentioned termite-eating ant species (Holldobler and Wilson 1990). In addition to opportunistic predators, termite-eating ants are another type of predator. These are primarily members of the genus Dorylus and the subfamilies Ponerinae and Myrmicinae (Culliney and Grace 2000).

10.6

Dermaptera

Since 1886, reports of earwigs eating insect pests have been made. Predaceous earwigs (Proreus simulans and Proreus melanocephalus) have been discovered in sugarcane (Saccharum officinarum) and paddy stubble leaf sheaths and borer holes in India in the past. More species have been seen in this study to tunnel into sugarcane borers and attack them in the field, eating on both adults and nymphs. To assess their potential in the control of pests, additional research on the biology and population fluctuations of earwigs is currently being conducted (Ramamurthi and Solayappan 1980). Under controlled laboratory conditions, the red-legged earth mite, Halotydeus destructor, was eaten by the European earwig, Forficula auricularia. At the two higher densities of 50 and 75 mites per dish, there was no difference between males and females, despite the fact that female earwigs consumed more mites at the lowest density of 25 mites per dish. Vermiculite was added to the dish to replicate a variety of habitats; however, this had a negative effect on how quickly food was consumed (Weiss and McDonald 1998). Labidura riparia (Labiduridae), Euborellia annulipes (Anisolabididae), Doru taeniatum and Doru lineare (Forficulidae), Euborellia annulipes (Anisolabididae), and the European earwig Forficula auricularia (Forficulidae) have all been noted as significant predators of stem borers and leaf. In America’s corn fields, the predatory earwig Doru luteipes (Forficulidae) is frequently observed. For example, the autumn armyworm Spodoptera frugiperda (Lepidoptera: Noctuidae), sugarcane borer Diatraea saccharalis (Lepidoptera: Crambidae), maize earworm Helicoverpa zea (Lepidoptera: Noctuidae), Ascia monuste orseis (Lepidoptera: Pieridae), and aphids are all controlled by (Naranjo-Guevara et al. 2017).

308

10 Predation Ethology of Various Orders

At Brooke’s Point, Palawan, Philippines, adults of Proreus simulans, a significant predator of Ostrinia furnacalis larvae, were gathered from flooded, rain-fed lowland rice. They were discovered both inside the feeding chambers of Pelopidas mathias, which are made out of a collection of leaves connected by silk, and inside leaves folded by Cnaphalocrocis medinalis. The larvae were completely digested, eaten in pieces, or harmed by the earwig’s forceps. Prior to Brooke’s Point, there was no sugarcane growing close to the rice fields, contrary to the traditional theory that this predator lives in upland areas, particularly in sugarcane fields from whence it migrates to maize fields to pupate (Barrion and Litsinger 1985). The earwig Forficula auricularia was regarded by fruit growers as both a natural enemy of aphids and psyllids and a nuisance in orchards (Jana et al. 2021).

10.7

Comparison Between Coleoptera and Dermaptera

Earwigs and coccinellid larvae attacked colonies the most out of all predators that were seen doing so. Even though earwigs performed more visits and assaults than coccinellid larvae, the total attack time was longer for coccinellid larvae because each attack lasted longer. Since most attacks for both taxa were brief and some were quite lengthy, mean attack durations were longer than median attack durations. Ten assaults were made in total by all other predators. Although ants frequently interacted with colonies, they hardly ever appeared to transport aphids, and tending behaviours could not be distinguished. Since we released them into the study area, earwigs were seen throughout the entire experiment. The other major predators were not spotted until 3 July (chrysopid larvae) or 4 July, and coccinellid larvae were not seen until that day (syrphid larvae). Except for one individual spotted at 13:30, all earwig sightings took place between 19:00 and 7:00, and 97% of sightings took place between 21:00 and 4:30. This diel periodicity coincided with sunrise and sunset (20:6 min throughout our investigation) (5:1 min). Coccinellid larvae were visible both at night and during the day. In film that included all 24 h of the day, ants were most prevalent at night (22:00–5:00; 55% of appearances per hourly 10-s sample period). The majority of ant sightings occurred during the day between 6: 00 and 9:00 in the morning (23%) and 18:00 to 21:00 in the evening (16%), with only 6% of sightings occurring between 10:00 and 17:00. There weren’t enough visits to other taxa to effectively describe their diel periodicity. When earwig activity on trunks and branches was compared, it was found that earwigs visited trunks substantially more frequently per hour of video, but that there were equivalent amounts of visits, attacks, and attack times per hour at both sites. On trunks rather than branches, ants were more prevalent. Ants were primarily small and exhibited a strong negative connection with earwig–aphid average attack duration, earwig–aphid attacks per earwig visit, and earwig–aphid attack time per earwig visit time. Ants only attacked coccinellid larvae eight times. Coccinellid larvae altered their direction and speed after coming into contact with an ant in five of the interactions. The coccinellid larvae remained put, lifted their abdomens, and the ant moved on in the other three encounters. Additionally, two coccinellid adults on

10.9

Hemiptera

309

trunks and one chrysopid adult on a branch were agitated by ants. These predators never touched the ant; they always took off. Non-ant antagonistic contacts were rather few. Never have earwigs been seen provoking other predator species. In five antagonistic intra-species contacts, earwigs made a hasty exit after making contact. A coccinellid larva made contact with a syrphid larva twice before separating. A syrphid larva once collided with a coccinellid larva, causing the latter to fall from the tree. Additionally, we saw one coccinellid larva consume a smaller one and another drop to the ground after coming into contact with another coccinellid larva (Orpet et al. 2019).

10.8

Orthoptera

Neither the adults nor the larvae of Grylloblatta campodeiformis (Orthoptera, Grylloblattodea) can survive without eating live or recently deceased animals. The antennae locate prey, which is then caught by the mandibles. Five different types of sensilla are documented from the antennae and palpi, all of which are likely used to identify prey (Pritchard and Scholefield 1978). The male Kobonga oxleyi cicada mating song attracted the female Spotted Predatory Katydid, Chlorobalius leucoviridis (Orthoptera: Tettigoniidae), who preyed on the cicada as it approached (Marshall and Hill 2009).

10.9

Hemiptera

In a laboratory setting, it was demonstrated that the wasps Coleomegilla maculata (Coccinellidae), Podisus maculiventris (Hemiptera: Pentatomidae), and Astata occidentalis (Hymenoptera: Sphecidae) attack the eggs of Halyomorpha halys (Morrison et al. 2016). The first instar nymphs of the Halyomorpha halys can be preyed upon by hemipterans like nabids and reduviids, and this is how nabids and Podisus maculiventris are able to prey upon the second instar nymphs. When tested on all developmental instars, including the adult stage, Euthyrhynchus floridanus (Hemiptera: Pentatomidae), it has shown some potential for biological control (Bulgarini et al. 2021).

10.9.1 Reduviids Sphedanolestes variabilis predated in the same manner as other non-tibial pad harpactorine reduviids: arousal, approach, capture, probing, piercing, and sucking post-predatory cleaning (Ambrose et al. 2009). The location, catch, immobilisation, and consumption of prey can all be broadly classified as stimuli-response-mediated sequences of discrete events that make up Rhynocoris marginatus feeding behaviour (Haridass et al. 1988) (Fig. 10.2). In addition, feeding of Zelus reindeer and another harpactorinae feeding was also recorded (Fig. 10.2).

310

10 Predation Ethology of Various Orders

Fig. 10.2 Zelus reindeer feeding on leafhopper Asian bug (contribution by Ahmet Gunaor), Rhynocoris marginatus feeding on cotton mealybugs, Harpactorinae feeding on a Chalcid wasp (Brachymeria sp.)

(a) Prey capturing: Rhynocoris marginatus is always stimulated by visual cues, and a moving prey causes the starving and roving predator to become more active. Once the prey is located, the predators lift the body off the ground and move slowly and steadily towards it while the antennae are pointed forward. The rostrum is fully extended forward to roughly 90° and the fore legs are lifted off the ground, as if to pin down the prey, when the prey is within striking distance, well before the antennae could make any contact with the prey body. Similar reactions are induced when an inactive object, such as a dead caterpillar, is carried in front of a hungry predator. These insects’ nymphal stages also exhibit comparable behaviours. (b) Prey immobilisation: As soon as the predator gets close to the prey, it tries to hold it down with its forelegs’ tibial pads and pokes the prey with its long rostrum before injecting its poisonous saliva. For salivary injection, the thorax or segments behind it are always chosen. The tibial pads are less effective in holding down hairy caterpillar–prey; therefore, two to three rostrum-equipped probes are needed to choose the best location before inserting stylets and injecting saliva. When larger caterpillars are met, the predator does not attempt to firmly grasp the prey with its forelegs, instead simply jabs with its fully extended rostrum and injects poisonous saliva. When attacked in this way, caterpillars react furiously and attempt to flee, but the predator is pursuing them and keeps jabbing them. Larger caterpillars almost always manage to avoid such attacks. When a predator attacks, little and medium-sized caterpillars also lash out forcefully, twisting and rolling their bodies and spitting out body fluids and saliva. In these situations, the predator lets go of its grip but continues to closely pursue the writhing victim until it is finally immobilised and killed within 30 to 50 s with its stylets still in place. The predator, after pinning down and injecting the saliva, totally lifts the prey off the ground when the prey body size is small or when, as in the laboratory, maggots and grubs were supplied as meal. Unable to escape, the prey struggles dangling at the rostral tip until death. The predator originally displayed alarm responses when grasshoppers were provided as prey by rising its body off the ground, lifting its forelegs, and

10.9

Hemiptera

311

extending its rostrum and antennae forward. Along with ongoing visual stimulation. The predator approaches carefully, pins the prey down with its forelegs, and then attempts to insert stylets and inject saliva between the inter-segmental membranes of the thorax or abdomen. However, 7 out of 10 times, the grasshoppers responded strongly when the predator touched or prodded them; they either fled or jumped into hiding spots. The grasshoppers have been observed kicking the predator with their rear legs on multiple times. Rhynooris marginatus could only occasionally successfully capture grasshoppers as prey if they were little nymphs or small-sized adults (Oxya nitidula, Cyrtacanthacris ranacea, and Euprepocnemis alacris alacris). Rhynooris marginatus was unable to capture larger grasshoppers with strong hind legs. (c) Prey transport: Rhynocoris marginatus frequently transports the immobilised and slain prey to a remote location prior to the start of feeding. Whenever there is ground-level predation, the predator drags the dead prey with the rostrum as it walks backward with the stylets still in place. Typically, secluded areas, cracks in the earth and under large stones are chosen. The deceased prey is retained at the rostral tip and quickly transported over short distances when predation on plants occurs. When the prey is small, it is also held dangling at the rostral tip, and feeding could start right away after the prey has been killed. When nymphs attack their prey in groups, the dead prey is not dragged around. The predators take their time cleaning their antennae of the bodily fluids and vomited salivary secretions emitted by the prey during attack before hauling the prey for meal. This grooming is done with the antennal comb, which is located on the inner top borders of the fore tibia. After cleaning, the predator re-inserts the stylet into the body of the prey so that it can be transported and then fed. (d) Prey consumption: This is the final phase of feeding behaviour, and it may start right away after killing the prey or after moving it to a remote location, depending on the size of the prey and the predation habitat. If the prey is small or borne at the rostral tip, the feeding location is never altered; if the prey is larger, the feeding site is occasionally altered. In each of these situations, the predator’s feeding activity doesn’t end until its abdomen has fully distended. Adults who are completely satisfied do not feed for the following 20–24 h. The post-predatory act was afterwards incorporated as a further act. The predator began cleaning its tibiae, antennae, and rostrum after sucking the victim at all feasible locations. Depending on the prey encountered, III, IV, and V nymphal instars and adults’ post-predatory actions differed in length. The predator’s last act of predation consisted of dragging the host’s empty case off and cleaning its antennae and rostrum with its fore tibial pads (Rajan et al. 2017). Similar to other general predators, reduviids’ antennae, eyes, and tibial comb play a crucial part in feeding. Through certain experiments, Claver and Ambrose (2001) were able to demonstrate it. The eye-blinded, tibial comb-coated, and antennectomised predators all showed a delayed arousal response. The approach reaction was not considerably altered in the blinded or tibial comb-coated reduviid predators, but it was in the predators whose complete antennae had been removed.

312

10 Predation Ethology of Various Orders

Fig. 10.3 Rihirbus trochantericus (a) Cydoncoris gilvus (b), Panthus bimaculatus (c), Euagoras plagiatus (d), and Sphedanolestes signatus (e) predating on Helopeltis in cashew

Predators with blindness and tibial comb coating as well as predators with their complete antennae and pedicel removed all had a major impact on their ability to capture. Antennectomy, eye blindness, and tibial comb coating had little effect on the prey’s ability to be paralysed. As a result of sequentially removing antennae in segment-by-segment fashion, a progressive decrease in sucking duration and the number of sucking sites was found. Predators with tibial comb coating and blindness showed reductions in sucking length and number of sucking locations (Claver and Ambrose 2001). Additionally, it was discovered that the assassin insect Sycanus dichotomus ate an adult Elaeidobius kamerunicus pollinating weevil. In Malaysian oil palm fields, this assassin bug targets bagworms and nettle caterpillars. It is a generalist predator, a key bagworm predator, and a tool for Metisa plana biological control (Muhammad Luqman et al. 2018). Compared to Coranus vitellinus, Coranus spiniscutis nymphs started feeding after eclosion in just 1 h. It was discovered that nymphs were quite active and chose relatively small-sized prey (Ambrose and Livingstone 1985). In

10.9

Hemiptera

313

Fig. 10.4 Rhynocoris fuscipes phase contrast photographs of terminal segment (a), stylet (b), magnified tip region of mandibular stylet (c), different view of mandibular tip (d, e, and f). Br brushlike, F furrow, H hook, PDF partially digested food material, 1R first row, 2R second row, 3R third row, 4R fourth row, SC salivary canal, TO terminal opening

India, it was discovered that harpactorine reduviids (Reduviidae: Harpactorinae) were feeding on the sucking pest Helopeltis antonii in cashew. This technique of predation is known as “pin and jab”. Endochus albomaculatus, Epidaus bicolor, Euagoras plagiatus, Irantha armipes, Panthous bimaculatus, and Sphedanolestes signatus are just a few of the six species of harpactorine reduviids that have been observed feeding on Helopeltis antonii (Srikumar et al. 2014) (Fig. 10.3). Reduced stylet structure plays an important function in predation: Kumar and Sahayaraj described the stylet structure (2012). They claim that Rhynocoris marginatus has a rostrum with three segments, the middle segment being longer than the base and terminal segments. Three different forms of fine sensilla, long spines (LS), medium spines (MS), and short spines, are present on the terminal rostral segment (SS). Trichobothria is also present at the terminal segment’s tip (T). Stylet The stylet is made up of four long, hair-like structures that are bundled together and have a sharp point at the tip. Adductor muscles are used to attach the base of the stylet to the head (AM). Each pair of maxillary stylets (6260 m) inside and mandible stylets (5770 m) outside can be found in the Rhynocoris marginatus stylet bundle. Similar to the maxillary stylet, the mandibular stylet is joined by the adductor muscles (AM). The maxillary stylet has a spear-like shape and a sharp tip (Fig. 10.4a), after which it has a narrow tube. There is a furrow (F) on both sides of the junction where the tube meets the tip. It has a connection to the SC, or central salivary canal (Fig. 10.4b). The right side includes 23 barbs in the middle and 28 sharply curled brush-like barbs (Br) on the lateral sides that point away from the head (Fig. 10.4c). The lateral border of the stylet has sharp, brush-like barbs (Br).

314

10 Predation Ethology of Various Orders

Table 10.1 Predatory behavioural tools to demarcate the subfamilies Acanthaspidinae (A), Harpactorinae (H), and Piratinae (P) Tools Arousal Approach

Acanthaspidinae A Fast (1–4 s) Fast (1–20 s)

Capturing

Pouncing with legs (3–30 s) Slowly (10 s–Z min)

Harpactorinae Slow (5–6 s) Slow (or) waiting (20 s–1 h) Pinning with rostrum (10 min-d-S h) Quickly (2–60 s)

Firm predator grip released but hold the prey with legs (21 min–3 h) More (12–20)

Lifting the prey with the rostrum (or) holding the prey with legs (20 min–1 h) Less (8–15)

Piratinae Instantaneous Aggressive (1–4 s) Pouncing with legs (2–15 s) Slowly (ZO s-3 min) Predator grip released and sucked at its ease (1–6 h) More (2Q–30)

In early instars

In all instars

Absent

In all instars

In early instars

In all instars

Paralysing Sucking

No. of sucking sites Nymphal congregational feeding Nymphal cannibalism

Source: Ambrose (1987)

Barbs have distinct barbs laterally and centrally, and are short (42.2 mm) at the tip and long (73.7 mm) at the distal end. The mandibular stylet is highly pointed and sharp, and it opens distinctly with an opening (TO) (Fig. 10.4d) followed by a salivary canal (SC) (Fig. 10.4e) that joins with the digestive canal of the salivary apparatus. The barbs are triangular plates with dimensions of left = 37.5 m, right = 58.4 m, and base (62.5 m). Reduviids eat food that hasn’t fully digested through the terminal opening (Fig. 10.4f). The gap between the two plates measured 37.4 0.3 m. The tip’s dorsal aspect reveals four rows of small barbs. These tiny barbs are 20.8 mm at the base and 16.6 mm on the right. Barbs are spaced apart by 29.12 0.10 m. The distance between the first and second row was measured to be 62.4 0.1 m, while the distance between the second, third, and fourth rows was 24.9 0.1 m. These four rows of barbs (1R, 2R, 3R, and 4R) are all oriented towards the head. The primary salivary duct is reached via the SC. However, the adductor muscles in the head are what attach the distal end of the maxillary stylet to the head (AM). Both mandibular and maxillary stylets can be used for feeding. Various feeding acts of reduviid were different in various genuses as specified below in Table 10.1. Specialised Structures for Predation Phymatinae species of ambush bugs have highly modified sub-chelate or chelate fore legs, and thread-legged bugs (Emesinae) lure their ant prey with unique abdomen glands. Thread-legged bugs have raptorial fore legs that mimic those of preying mantises. Fossula spongiosa, an adhesive

10.9

Hemiptera

315

structure on the tibiae that aids in prey capture and is common in Reduviidae and other Cimicomorpha, is another distinctive feature. The largest Reduviidae subfamily, Harpactorinae, demonstrates a unique predation technique: some species are known to exploit sticky materials to increase predation success, a phenomenon known as “sticky trap predation”. Depending on where the sticky substance comes from, there are two different sorts of this predation method. The resin bug tribes Ectinoderini, Apiomerini, and Diaspidiini are notable for gathering plant resins, which are then applied to the legs and bodies of the insects to help them catch prey. The tribe Harpactorini’s species of the huge (>60 spp.) genus Zelus, however, use an endogenous source of sticky chemicals to grab their prey. On the fore tibiae of Zelus leucogrammus and Zelus luridus, respectively, a specific epidermal gland is present. They came to the conclusion that these glands are in charge of producing the viscous coating that covers these assassin bugs’ fore tibiae. Zelus species with sticky glands have specialised setae that look like the trichomes on sundew plant leaves (e.g. Drosera). The fore tibiae of Zelus contain large numbers of what are known as sundew hairs or sundew setae. They may have a role in keeping the secreted viscous substances in place. Currently, Reduviidae outside of the genus Zelus lack sundew setae. In addition to sticky glands, we also saw typical class 3 cutaneous glands, which include saccules that resemble ovoids or balloons. The current investigation is not focused on this sort of gland because it appears to be common in the Harpactorinae and Reduviidae (Zhang 2012).

10.9.2 Miridae The following components of each predator’s behavioural routines were identified after an hour of observation: moving, searching, resting, and coming into contact with prey, preying, feeding on leaves, and cleaning. Then, predation, encounter, and acceptance rates were computed (Martínez et al. 2022). In the lab, predatory behaviour was seen in the phyline mirid Rhinacloa forticornis. Thrips orientalis and hawaiiensis nymphs and adults were exposed to jasmine blossoms and bean pods that were infected with larval and adult stages of the pest. The larvae of only two thrips species—Thrips orientalis and Thrips hawaiiensis—were successfully attacked. The way Rhinacloa forticornis handled Thrips orientalis and Thrips hawaiiensis, particularly the mirid’s use of its fore tarsi to position and manipulate prey and its probing in different body regions, as well as the feeding process, which involved the ingestion and egestion of fluids and thus suggested a potential role for extra-oral digestion of Thrips orientalis and Thrips hawaiiensis tissues, were suggestive of similar (Culliney 2014). Although predation occurs naturally when there is a lack of suitable plant food, studies of this plant bug eating seemingly preferentially on vegetable matter when there is accessible prey suggested that phytophagy is a significant, possibly dominating, feeding mode for this plant bug. Nesidiocoris tenuis is used primarily on tomato, but also on eggplant and sweet pepper, to control whiteflies (Hemiptera: Aleyrodidae), leafminers (Lepidoptera: Gelechiidae), thrips (Thysanoptera: Thripidae), and spider mites (Arachnida: Tetranychidae) (Marquereau et al. 2022). The majority of the United States and

316

10 Predation Ethology of Various Orders

Canada are home to Deraeocoris nebulosus, which is widespread in the eastern states. It was maybe the first mirid in North America to have predaceous traits attributed to it. One of the most significant of the real bugs that consumed scale insects, it was noted as a predator of females and potentially eggs of “canker-moth” [spring cankerworm, Paleacrita vernata, or fall cankerworm, Alsophila pometaria]. It primarily consumes the European red mite, Panonychus ulmi, as well as the woolly apple aphid, Eriosoma lanigerum, the clover aphid, Nearctaphis bakeri, the eyespotted bud moth, Spilonota ocellana, and crawlers of the terrapin scale, Lecanium nigrofasciatum. It also consumes the larvae of the codling moth (Wheeler et al. 1975). The predatory bug Cyrtorhinus lividipennis is a significant natural opponent of rice planthopper eggs and young nymphs in Asian rice systems. This species can persist when delphacid prey is scarce by moving to other prey, such as conspecifics. Cyrtorhinus lividipennis is known to target Lepidoptera eggs and larvae, including those of the rice leaf folder Cnaphalocrocis medinalis, rice stem borer Chilo suppressalis, and pink rice borer Sesamia inferens, even though it is widely recognised as a prominent predator of delphacid eggs and nymphs. In rice fields, Cyrtorhinus lividipennis can prey on N. lugens eggs at rates as high as 70%. Individual Cyrtorhinus lividipennis nymphs and adults can eat 7.5 and 10.2 N. lugens daily, according to laboratory experiments (Zhu et al. 2014). Manly’s preference index indicated that when offered with Tuta absoluta, Engytatus varians strongly preferred Neoleucinodes elegantalis, although no such preference was apparent for Macrolophus basicornis. On Tuta absoluta, Macrolophus basicornis spent more time seeking for both prey and needed less time to find fresh eggs than Engytatus varians did. Macrolophus basicornis had a higher attack rate and was more effective as a predator against both pests. Prey finding, resting, and feeding (including sucking of leaflets) took up more of both mirids’ time than the other activities combined. However, compared to Macrolophus basicornis, Engytatus varians spent more time eating on plant tissue. These findings show that both mirids are effective biocontrol agents for Neoleucinodes elegantalis and Tuta absoluta (Martínez et al. 2022). Our findings lead us to the conclusion that both mirids could be taken into account in pest management plans for tomato crops. In a lab setting, two predators were observed attacking the eggs of the two lepidopteran pests Neoleucinodes elegantalis (Lepidoptera: Crambidae) and Tuta absoluta (Lepidoptera: Gelechiidae). These predators were Macrolophus basicornis and Engytatus varians (Miridae). Manly’s preference index was used to measure the two mirids’ predation rates over a period of 24 h (Martínez et al. 2022).

10.9.3 Nabidae Nabis spp. can attack lepidopteran larvae successfully despite their defensive behaviours by continuously probing with an extended foreleg until the prey is accustomed to its presence, according to Arnold (1971), who also summarised the feeding behaviours of Nabis. Once alerted to nearby prey, the predators Nabicula

10.9

Hemiptera

317

subcoleoptrata, Nabis americoferus, and Sinea diadema must find and attack the prey object (says Freund 1998). If their target gained weight or ate food from adjacent hosts, nabids attack them. For instance, Nabis roseipennis that did not attack prey lost weight over the course of 24 h, whereas those that did acquired weight from lima leaves, including Epilachna varivestis (Coleoptera: Coccinellidae) and Anthonomus grandis (Coleoptera: Curculionidae) that attacked prey. During the late summer and early fall, the Pacific damsel bug (PDB), Nabis kinbergii, is frequently observed on lucerne, Medicago sativa. Aphids (bluegreen lucerne aphid, Acyrthosiphon kondoii, Acyrthosiphon pisum (Harris), mirids (Australian crop mirid, Sidnia kinbergi, potato mirid, Calocoris norvegicus), and other small insects are among the prey this polyphagous predator consumes (Siddique and Chapman 1987). Nabis encountered and consumed more Geocoris while grazing for these herbivores. Surprisingly, the frequency of intraguild predation was not highly correlated with Geocoris or Nabis densities or the farming method, indicating that prey community structure is more significant than predator community structure. The detritus-feeding fly Scaptomyza pallida was a food source for Geocoris, and we discovered evidence of this with increasing predator evenness. In light of the increased risk of intraguild predation, this would be compatible with Geocoris changing to more feeding on the ground, where S. pallida would be quite prevalent. Findings imply that while detritivores may facilitate a shift to safer foraging on the ground, herbivorous prey may increase intraguild predation of Geocoris in the foliage. This gives more proof that depending on the prey species and predator behaviour, prey richness and variety can either increase or decrease predator– predator interference. Two new exotic stink bugs to California, the bagrada bug and the brown marmorated stink bug (BMSB), Halyomorpha halys and Bagrada hilaris, are now found widespread over the state. An investigation was done to find local natural adversaries of their eggs. The identification of arthropods attacking these eggs and their relative impact were made easier with the help of a field camera. Eggs of the BMSB, which oviposit naturally onto plant foliage, were less likely to be eaten than eggs of the bagrada bug, which oviposit onto and into soil. The Asian cockroach Blatta orientalis (Orthoptera), the spider Trachelas spp., the carabid Laemostenus complanatus (Coleoptera), the European earwig Forficula auricularia, and the ringlegged earwig Euborellia annulipes (Dermaptera), were the main predators of BMSB eggs (Araneae). Solenopsis xyloni and Monomorium ergatogyna, two ant species (Formicidae), and a species of Collembola were the two most prevalent and significant predators of bagrada bugs (Entomobryidae). To examine the relative importance of BMSB egg predators, a Predation Index was created. Finally, two startling, unexpected findings are reported: Collembola appears to prey on the sentinel eggs of the bagrada insect through a synergistic relationship with the native ant M. ergatogyna and the spider Trachelas spp.

318

10 Predation Ethology of Various Orders

10.9.4 Pentatomidae The majority of the predatory Pentatomidae stinkbugs, according to De Clercq (2000), are Asopinae. Caterpillars that are both open and dormant and are concealed in silken webs are both attacked by Picromerus bidens (Family Pentatomidae). The bug normally approached the larval mass, stalked a larva with its proboscis, walked a few centimetres away, and sucked with its proboscis extending onwards when feeding on exposed caterpillars. The bug returns to the group of larvae and chooses a different caterpillar after finishing the butterfly larva. The bug first searched the web surface for an entrance before puncturing the encasing silk and poking its proboscis within the web for a brief period of time until it pierced a larva (Konvička et al. 2005). According to De Clercq (2000), Asopinae is where the majority of the predatory Pentatomidae stinkbugs reside. Sahayaraj and Fernandez (2021) noticed Antilochus coquebertii (Heteroptera: Pyrrhocoridae) feeding on Dysdercus koenigii. The normal pattern of adult A. coquebertii’s predatory behaviour included arousal, approach, capture, rostral probing, paralysis, feeding, and post-feeding act. In a feeding action known as a “case and pounce”, Antilochus coquebertii will actively pursue their prey, Dysdercus koenigii, pounce over them, and grip them with their legs. Tested predators are always present at feeding times. The total durations of the male and female Antilochus coquebertii sequential activities on Dysdercus koenigii adults are mentioned below. The average data of a predator versus three distinct preys shows that male predators considerably took less time than female predators to arouse, approach, capture, explore, and paralyse their prey. Between male and female predators, post-predatory behaviour was inconsequential. Compared to a female predator, the Antilochus coquebertii female took a considerably shorter time to feed (FT) three adults of Dysdercus koenigii. The following is a description of the feeding steps: Arousal The predator agitatedly observes the prey’s movements, straightens its legs, and frequently vibrates its antennae. The pyrrhocorids were stimulated by the moving prey, which caused the Antilochus coquebertii to awaken from akinesis and reach a high level of excitement. When a predator is aroused, it travels slowly or swiftly (depending on how quickly the prey is moving) in the direction of the prey while pointing its rostrum and antennae forward. When the prey was bigger than the predator, it would frequently repeat the approach. Capture The predator closes up on the prey and either pounces over it (Fig. 10.5a– d) or modifies its rate of movement to that of the prey (Fig. 10.5c, d), capturing it for later immobilisation by rostral probing. Then, grasp the prey firmly by both its front and back legs. Nymphs, however, did not leap over the prey; rather, they caught it with either their forelegs or simply their forelegs (Fig. 10.5a, b). Rostral Probing Following the successful capture of the prey, the predator first probes at the base of the antennae (3.5 times), followed by the junction between the

10.9

Hemiptera

319

Fig. 10.5 Feeding behaviour of Antilochus coqueberti nymphs (a, b) and adult (c–f) against Dysdercus koenigii. Antilochus coquebertii fifth instar nymph feeding on fifth instar (a) and newly emerging adult red cotton bug (b), adult predator pounces over the prey (c, d) and feeding in frons region (e) or ventral side of abdominal portion (f) of the prey

head and thorax regions, the frons region, the region between the forewing and the scutellum, the connexivum, the cervical membrane, or other softer areas (1.0 times), which results in the paralysis of the prey. The majority of antennal bases or frons, though (about 60%), were chosen for rostral probing. By injecting the prey’s salivary secretion through the inter-segmental membranes at the base of the antennae, the frons area, the neck, the connexivum, the cervical membrane, or the thoracic regions, predators can paralyse their victims. The predator relinquished its hold on the prey after paralysing it and pulled back its rostrum to signal paralysis. The absence of the prey’s antennae clicking action indicated that the victim was completely paralysed. Feeding The predator rotates the prey and, while facing the upper ventral side (Fig. 10.5d), holds it between its fore and hind legs (Fig. 10.5e), sucking out the body’s contents by inserting its rostrum in various locations, such as the leg joints, inter-segmental regions, and the frons region. However, adults favoured the head and ventral side of the abdomen, whereas nymphs largely favoured the head areas. Adults also drank the pre-digested bodily fluid of the victim (Fig. 10.5e, f). Less females (5–7 min) than males had piercing and sucking sites (7–10 min). On rare occasions, the same site was chosen again, but this time the stylet was withdrawn and the needle put into a different location. The antennae are always facing horizontally while being fed.

320

10 Predation Ethology of Various Orders

Post-Predatory Acts After sucking the victim at all feasible locations, the predator drops the carcass and begins to clean its tibiae, antennae, and rostrum. Results show that compared to male predators, female predators spent substantially less time to approach and handle Dysdercus koenigii adults. However, during a feeding, males barely outnumbered females in terms of the amount of prey they consumed (3 h). However, females had a higher capacity for predatory behaviour than males.

10.9.4.1 Nymphal Foraging Behaviour The Picromerus bidens nymphs stayed near the edge of the provincialis larval Euphydryas aurinia web during predation and only looked at the surface of the older nest webbing. When they came upon a potential meal (the closest caterpillar), they attacked the larva by piercing it with their stylet. They then elevated the prey with their beak and moved away, hiding between leaves about 20 to 30 cm from the nest, where they subsequently ate the food. Only after the nest was in the light and the larvae were still inside the silk web did a second instar nymph approach the larval mass. Picromerus bidens adults were visible on a Euphydryas aurinia provincialis larval web, but no instances of caterpillar predation were ever documented. We only captured one adult predator feeding on a Euphydryas aurinia provincialis larva on the surface of the larval web (Cianferoni and Dioli 2019). 10.9.4.2 Adult Feeding Behaviour When adults of Eoeanthecona (=Cantheconidia) jureellata were given to Spodoptera litura larvae in their fifth instar, they almost instantly stretched their antennae straight out in front of their heads in the direction of their prey. The predator then positions itself directly towards the prey, pivoting if necessary and moving forward. Although the antennae frequently make contact with the larva, it is not necessary for the prey to be accepted. Rostrum is stretched close to the victim, which is then quickly pounced upon and punctured. No specific body area appeared to be preferred as the site of puncture when a soft-bodied larva is given. The larger larvae that this predator targeted were noted to stop squirming and become immobile quickly after the body wall was penetrated. Three to 10 min after being punctured by the stylets, the prey’s struggle came to an end. Because the bugs instantly begin sucking after the stylet is inserted, it is impossible to distinguish between poisonous and feeding effects with such delayed reactions. During the subsequent 2–5 h of feeding, the feeding place is rarely changed. The maxillary stylets are continuously extended, retracted, and moved about inside the prey, as is typical for pentatomids, whereas the mandibular stylets are punctured only shallowly, securing the prey at the point of puncture with their re-curved hooks. The antennae are pointed horizontally while feeding. The bug frequently carries the prey, impaled in the stylets, to a vertical surface, remaining there while feeding until the prey turns sac-like (Rani and Havukkala 1993).

10.9

Hemiptera

321

10.9.5 Anthocoridae In greenhouse and field crop environments, different species of Orius are significant predators on Thysanoptera, mites, and the eggs of Lepidoptera. The Thysanoptera biocontrol agent Montandoniola maraquesi has been widely used on figs and olives. The distribution of immature pear psylla Psylla pyricola and Anthocoris nemoralis’ searching behaviour showed a strong association. The bottom leaf surface, which was home to 74% of the immature pear psylla Psylla pyricola, was the area that the Anthocoris nemoralis spent 71% of its time searching. The top leaf surface received 29% of the predator’s search time, compared to 26% for psylla immature distribution. Nearly half of Anthocoris nemoralis’s search time (48%) was spent along the leaf midribs, where 64% of the psylla immatures were discovered. Regarding certain leaves, Anthocoris nemoralis searching behaviour was not random. The majority of the time was spent searching the lower leaf periphery, upper and lower surfaces, and midrib of the leaf (Brunner and Burts 1975). When looking for aphids on sycamore leaves, Anthocoris nemorum and Anthocoris confusus were observed acting similarly (Dixon and Russel 1972). The majority of their time was spent perusing the principal veins and the leaf margins by these predators. Under laboratory settings, Orius insidiosus seeking patterns on soybean trifoliolates exposed to different Sericothrips variabilis densities were detected. The predator’s searching rate and eating period were examined in relation to prey density. With an increase in prey density, predator seeking speed and feeding duration per soybean thrips decreased. The upper midrib of the soybean trifoliolate was where the majority of predator–prey interactions took place. In 45% and 56% of all interactions with adult male and female Orius insidiosus, respectively, adult soybean thrips were captured. Running was discovered to be the main means of escape for soybean thrips from Orius insidiosus attacks (Isenhour and Yeargan 1981).

10.9.6 Neuroptera Larvae of coniopterygids are predatory. A wide range of slow-moving arthropod prey, such as mites, scale insects, insect eggs, coccids, aphids, and phylloxerans, have been documented as being consumed by adults and larvae. Few generalist predators known to prey on antlions have been the subject of studies. Recent research by Segev et al. (2020) demonstrates that antlions (Neuroptera) are a group of predator insects that sit and wait for prey to approach them. Some species are also experts at building conical pit-traps in the ground to capture prey. When gripping prey, larger setae in predatory insects are a more cost-effective alternative to fangs or spines. On the inner border of the mandibles of several larval Nymphidae and Myrmeleontidae, stiff, expanded setae are present (Neuroptera). Dilaridae can successfully feed on the eggs and larvae of weak, fragile insects. The creation of funnel-shaped pitfall traps in sandy ground, where the antlion larvae (Neuroptera: Myrmeleontidae) wait for prey, is a well-known predatory strategy. Reticulitermes flavipes (Rhinotermitidae), Prenolepis imparis (Formicidae), and Alphitobius

322

10 Predation Ethology of Various Orders

diaperinus (Tenebrionidae) were given to it, each measuring 5.7 mm, 4.22 mm, and 6.13 mm. The behavioural catalogue of Myrmeleon mobilis revealed the following 12 distinct predatory behaviours: (Napolitano 1998). (a) Attack. Sand is expelled from the pit by rapidly flicking the head back and forth while shutting the mandibles. (b) Holding. The mandibles are tightly clamped around the prey. (c) Submergence. The larva moves down and back into the substrate while holding the prey until neither the complete larva nor at least some of the prey is visible. (d) Emergence. The larva moves up and forward while holding its prey until the entire prey and at least a portion of the larva’s head and mandibles are visible. (e) Beating of prey. The larva rapidly flicks its head up and down while holding its prey (4–5 beats each bout), frequently drumming the prey on the substrate. (f) Feeding. Fluids are taken from the prey while at least one mandible tip is entered, frequently alternating with mandibular probing and manipulation of the prey. (g) Pit sweeping. The head is flicked quickly back and forth, depositing dirt on the dorsal surface as it moves laterally. (h) Head roll, no. 8. Sediment gathers in the pit’s middle as the head is raised and swept in a circular motion along the pit wall. (i) Clearing the prey. Prey is placed on the dorsal head surface using the mandibles, after which it is quickly ejected back by the head. (j) Grooming. One mandible’s tip is pushed along the groove on the inner edge of the mandible opposite it. (k) Quiescence. Larva is immobile for at least 7 s without a meal. (l) Jaw Set 12. The larva fully opens its mandibles as it digs beneath the sand. The tips of the mandibles, antennae, and eyes are still visible. All prey kinds often displayed a fundamental sequence of actions in the following order: attack, holding, submersion, emergence, and feeding. Following the end of feeding, maintenance behaviour (such as head rolling, pit clearing, grooming, and prey clearing) usually follows before jaw setting. Prey beating activity made up the majority of the behavioural difference: 90% of beetle prey trials ended in prey beating, compared to 20% of ant trials and 10% of termite trials. The termite’s (2.00 SE) and the ant’s (8.90 SE) mean frequency of preybeating bouts was significantly different from the beetle’s (42.40 SE), but the latter two were not significantly different (Napolitano 1998). Ambushing a Group By increasing per capita food intake and lowering the cost of capture, group hunting followed by food sharing increases members’ fitness. In actuality, preys are captured more successfully and with less danger of harm in addition to avoiding predation. This has been observed in social spiders, some ant species, fish, birds, cetaceans, otters, carnivorans, and chimpanzees, among other vertebrate species. Adalia bipunctata, an aphidophagous ladybird, first instar larvae, Euthyrhynchus floridanus, a predatory pentatomid (Heteroptera) nymph and adults (Ables 1975),

10.9

Hemiptera

323

Fig. 10.6 Group feeding behaviour of Rhynocoris marginatus nymphs on caterpillars with (a) and without hairs (b), Antilochus coquebertii third instar nymphs feeding on a beetle (c)

and some harpactorine assassin bug nymphs (Heteroptera; Reduviidae) were all observed to engage in group hunting and prey sharing (Sahayaraj 1991; Jackson et al. 2010). In search groups of ten or more individuals of various origins (not necessarily siblings), Agriosphorus dohrni nymphs (Harpactorinae) devote brief amounts of time to active foraging; tiny preys are consumed selfishly, while large preys are shared (Inoue 1983). It is a unique property shared by numerous reduviids. An extensive study was carried out using Zelus annulosus to discover this fact (Alain et al. 2013). They made the following observations: The first instar nymphs in the field stayed together under the leaf where they hatched until the first moult and only seldom climbed up onto the upper side of that leaf where they all sat and waited, ambushing in a group with their forelegs skyward. The nymphs in their second and third instars spent a lot more time ambushing in groups on the upper side of their leaves (Alain et al. 2013). However, in a lab setting, nymphs typically elevate and stretch their forelegs when ambushing. The nymphs warily approached prey that had been located at a distance by visual, chemical, or vibration cues, elevating their forelegs higher than usual. Once they were far enough away, they rammed their forelegs into the preys that were clinging to the sundew setae. The nymphs then extended their rostrums towards the preys and bit them in the inter-segmental zone, rendering them unconscious. The Zelus annulosus hoisted the preys by extending their rostrum, and then rotated them with their forelegs while spitting saliva and sucking out digested materials, turning the preys’ corresponding sections black in the process. The Zelus annulosus nymphs can spin and bite their preys up to ten times in succession before throwing away the empty preys (Alain et al. 2013). As a result, it is likely that communal feeding by Zelus annulosus nymphs helps siblings decrease their risk of starvation, shorten their exposure to predators, synchronise their developmental stages to prevent cannibalism while allowing fair mutual prey sharing to continue, and ultimately increase their chances of growing up. The whole happens through a self-sustaining process governed by straightforward rules, such as clutch laying causing group living and early instar ambushes (non-kins are tolerated), sharing a prey or not depending on its size and the successful nymph’s level of hunger, and last instars attracting siblings to share large preys (Alain et al. 2013). Group feeding is a common phenomenon of heteropteran nymphs (Fig. 10.6) and adults.

324

10 Predation Ethology of Various Orders

10.9.7 Syrphids Omkar and Mishra (2016) summarises the following statement regarding Syrphid predatory behaviour in 2016. He asserts that both visual and smell signals are used by syrphids to determine the location of their prey sources. Instead of pollen, whose quantity had no effect, Episyrphus balteatus exhibits improved behavioural reactions to smaller yellow flowers with higher nectar concentrations? Syrphids start foraging as soon as they find their prey source, and their success is controlled by plant size and aphid abundance more so than by temperature. The optimal forging theory, which contends that there is an ideal window for oviposition and that syrphids oviposit in controlled numbers at an ideal density in an aphid patch in order to successfully assure the survival of their young, is also claimed to apply to syrphids. While later instars of Episyrphus balteatus show no reaction to aphid–plant complex smells, the initial instars do demonstrate aphid olfactory cue and not honeydew-guided short distance prey seeking. On the other side, these cues encourage larvae to eat. Syrphid species vary in how effectively they consume their prey; however, Episyrphus balteatus larvae are often regarded as the most prevalent and effective aphidophagous predators, devouring up to 100 species of aphids. However, the predation rate varies depending on the prey and instar stage. For syrphids feeding on the same species of aphids, it was discovered that there are species-specific changes in predation capacity relative to body size. This went against their presumption that, regardless of species, there would be a robust link between larval size and predation potential. The findings imply that it is impossible to categorise syrphid species only based on body size in order to translate estimations of predation capability from the lab to the field. Syrphids exhibit a type II functional response to varied aphid concentrations, as demonstrated by laboratory experiments, which show that prey intake increases curvilinearly with increasing prey density. However, other research suggests that while certain syrphids demonstrate a type II functional response in their first instar, type III response is more frequently seen in later instars. Additionally, these functional responses alter in response to the surface characteristics of the host plant. The size and hunger of the syrphid larvae as well as the level of prey content depletion determine the length of time spent handling prey and the quantity of prey devoured.

10.9.8 Diptera The sit-and-wait tactic is used by many predatory insects, including antlions, dragonflies, mantids, water bugs, and several types of flies (Cloarec 1990). For instance, the term “sit-and-wait behaviour” describes a strategy in which a predator choose a specific location to perch, waits for a potential prey to approach, and then ambushes it. Predators defend their territories more effectively, are less likely to be themselves preyed upon, and expend less energy while perched (Tomé et al. 2011). Predatory sit-and-wait species choose their habitats based on biotic and abiotic parameters such temperature, food availability, competition, and predation danger

10.10

Expression of Bt Toxins Through Their Preys

325

(Scharf and Ovadia 2006). Particularly predatory flies are dance flies (Diptera: Hybotidae), both as adults and larvae. The spike-armed hind legs of Chvalaea yolkamini are used to attack and handle prey, and its short proboscis pierces and sucks on inter-segmental areas because they are less sclerotised. The feeding period lasted between 1.27 to 15.24 min. Furthermore, Chvalaea yolkamini had better results attacking large prey; however, it appears that the predator’s size is the upper limit for prey size. When ambushing huge Brachycera as opposed to relatively small nematocerous Diptera and Hemiptera, the individuals were more successful. No effect of temperature or humidity was seen, but predation success was favoured by higher perches. Our research provides the first in-depth knowledge of a sit-andwait hybotid fly’s perching and hunting behaviour. We put up theories that may be tested by taking into account co-variables like wind and prey colour in order to better understand predatory animals’ sit-and-wait behaviour (Jaume-Schinkel et al. 2022). The feeding acts are as follows: In this study, the predator Chvalaea yolkamini waited for flying insects to pass by before performing an airstrike to capture the prey. They typically flew for small distances and short times (on average, 3 s). When a prey capture effort was unsuccessful, the animal went back to its original perch and typically did not try to catch another prey throughout the observation time. Only three times, during a single observation period (maximum time = 20 min), were two or three attempts seen, generally on slower and larger prey, but none of these attempts were successful. A total of 283 flights were observed, with flying times ranging from 0.3 s to 27.9 s. Total flying duration was 12.1 min, and each person spent an average of 10.56 s in the air. Thus, 98.61% of the time the dancing flies under study were perched, whereas just 1.29% of their time was spent searching. Hunting behaviour is straightforward. Predators remained perched until a prey item flew by; at that point, they began to fly towards the prey and attempted to capture it with their rear legs. Usually, if an attack fails, they quickly retreat to their perching location and leave without being pursued. Similar to how they rapidly returned to their perches after a successful flight and began to feed, pulling their prey into their proboscis with their hind legs. The average feeding time for tiny prey items was 2.65 min (1.27 min, 4.42 maximum), and the average feeding time for large items was 6.77 min (1.37 min, 15.24 max). While feasting, the predator rotated the victim slightly with the aid of its hind legs. It’s possible that the soft tissue there made insertion of the fly’s proboscis easier because that’s where it generally starts. The mesopleural region was the fly’s next target, followed by the abdomen. The proboscis was seen moving as the animal was feeding. Following a meal, predators kicked the carcass away with their back legs lowered. The predator then immediately groomed its face and rear legs and typically remained perched for the remainder of the observation period.

10.10 Expression of Bt Toxins Through Their Preys Average concentrations of Cry1Ac toxin were 0.23 and 0.25 g Cry1Ac g1 of fresh tissue in cotton leaves from cages housing nymphs from 2 July to the first week of August and adult during reproductive peak from the first week of August to

326

10 Predation Ethology of Various Orders

30 September, respectively. Through beet armyworm (BAW) larvae (0.18 and 0.19 g Cry1Ac g1 of fresh weight), about 81% and 76% of the toxin originally expressed in Bt cotton leaves were exposed to predator nymphs and adults. Despite the amount of Cry1Ac toxin found in plants and prey, which exposed big-eyed bug nymphs and adults both directly and indirectly, no toxin was found in the bodies of adult predators.

10.10.1 In Cotton-Coleoptera In laboratory experiments, newly hatched Helicoverpa armigera larvae fed on transgenic Bt cotton (Event 93–4) suffered higher rates of predation by the coccinellids, Propylaea japonica and Coccinella septempunctata, than non-Bt-fed Helicoverpa armigera larvae. This suggests that Propylaea japonica (Coleoptera: Coccinellidae) does not avoid prey that (Cui and Xia 1999). Variable and erratic effects of Bt-fed prey on predators can be seen. When Chrysopa sinica larvae are fed Bemisia tabaci nymphs raised on transgenic Bt cotton NuCOTN 33B as opposed to non-Bt cotton cultivar Simian 3, their development is slowed down and their survivability is reduced. When Propylaea japonica was fed Bemisia tabaci nymphs raised on transgenic Bt cotton, there were no differences in pre-imaginal growth, larval survival, or adult body mass. The fact that both Bt-Bemisia-fed predators had reduced adult survival rates, however, may mean that Bemisia tabaci is not the best food for either Chrysopa sinica or Propylaea japonica (Guo et al. 2004). When the predator Propylaea japonica is raised on the ideal prey species Aphis gossypii, a different reaction can be obtained. Along with abiotic, chemical, and biological factors, the structure of the plant (canaby) and the distribution of various natural enemies at different sites in the field also influence the interaction of GM crops and animals associated with them. Abundance of canopy- and ground-dwelling predators was monitored in three pairs of commercial Bt and non-Bt cotton fields by various methods at three seasons by Torres and Ruberson (2005). Result data from all three seasons were pooled for each sampling method, results revels that two natural enemies were favoring Bt (Nabis spp. and spiders) and four favoring non-Bt cotton [Hippodamia convergens and lady beetle eggs and Geocoris uliginosus ]. Analyses of predator community dynamics using principal response curves showed that the abundance of ground-dwelling predators was not affected by cotton type, whereas the abundance of canopy predators varied across seasons, with no particular trend for either cotton type. The tender tips of transgenic GK-12 and NuCOTN 33B (both Bt cotton) showed an average of 49.2 ng Cry1Ab/g plant fresh mass (FM) and 94.2 ng Cry1Ac/g plant FM, respectively. When raised on GK-12 and NuCOTN 33B cotton, respectively, Aphis gossypii adults and nymphs had 6.0 ng/g FM Bt toxin and 4.0 ng/g FM Bt toxin concentrations. Simian 3 (the control cotton cultivar) tender tips and Aphis gossypii raised on this cultivar both lacked Bt toxin. The predator Propylaea japonica acquired the Bt toxin from its prey, which was raised on Bt cotton plants. When the adult fed on Bt plant-associated prey for longer periods of time

10.10

Expression of Bt Toxins Through Their Preys

327

(5 to 20 days), the quantity of accumulated Bt toxin rose from 13.5 ng/g predator FM to 34.0 ng/g predator FM when fed cotton aphids raised by GK12 and from 9.0 to 24.0 ng/g when fed cotton aphids raised by NuCOTN 33B. Bt toxin was passed down to offspring in the case of NuCOTN 33B host plants; freshly born, unfed P. japonica larvae from parents fed NuCOTN 33B-reared cotton aphids contained 15.0 ng/g predator FM of Bt toxin. Additionally, the non-Bt control samples of adult Propylaea japonica were shown to contain trace quantities of the toxin (Zhang et al. 2006a). When Bt cotton plants were employed as a host plant for the prey aphids, survival, mating habits, and fecundity of Propylaea japonica were not significantly impacted. Contrarily, even while traditional ladybird beetles developed more slowly than either Bt GK or Bt Nu ladybird beetles, their larval stage was much shorter (6.6 days) than that of Bt GK ladybird beetles but was otherwise identical to that of Bt Nu ladybird beetles (6.5 days). Compared to female Bt Nu or traditional ladybird beetles, female Bt GK ladybird beetles were heavier (Zhang et al. 2006a).

10.10.2 In Rice By using an enzyme immunosorbent assay approach, it was possible to measure the levels of Cry1Ab toxin expressed in KMD1 and KMD2 rice plants as well as the amount of toxin conveyed to Nilaparvata lugens eating these plants. In the lab, the development characteristics of Propylaea japonica raised on KMD1- or KMD2-fed Nilaparvata lugens were evaluated. The findings demonstrated that from the booting to grain filling stage, the concentration of Cry1Ab in rice leaves and stems dramatically increased, and then reduced as the plants grew older. Nymphs and adults of Nilaparvata lugens seen feeding on Bt rice plants had Cry1Ab. Propylaea japonica that had eaten KMD1- or KMD2-fed N. lugens nymphs as larvae did not vary substantially from Propylaea japonica that had eaten nymphs that had been fed XS11 in terms of development time, pupation, adult eclosion, pupal weight, or maleadult motor activity. The non-target insect N. lugens and its predator Propylaea japonica are exposed to Cry1Ab toxin from transgenic cry1Ab rice; however, the toxin had no effect on the predator’s development through tritrophic interactions, according to our findings (Bai et al. 2006). When compared to Bt GK or Bt Nu ladybird beetles, male conventional ladybird beetles had a lower body mass in their second instar, but by the time they reached adulthood, this difference had vanished. Instead, there was a non-significant trend for male conventional ladybird beetles to have more mass than either Bt GK or Bt Nu ladybird beetles. Even though female Bt Nu and Bt GK ladybird beetles engaged in significantly more matings and were 2 days younger than female conventional ladybird beetles when they first mated, the proportions of mated female Bt Nu and Bt GK ladybird beetles that oviposited were lower than those of mated female conventional ladybird beetles. Compared to regular ladybird beetles, Bt GK ladybird beetles tended to have more deformities. Both conventional and Bt Nu ladybird beetles did not exhibit this tendency (Zhang et al. 2006a).

328

10 Predation Ethology of Various Orders

10.10.3 Geocoridae The omnivorous big-eyed insect Geocoris punctipes (Geocoridae), a significant predator in cotton Welds, may come into contact with the Bt toxin by animals fed Bt cotton or through plant feeding if it is continuously expressed in plants during the growing season. Both types of prey were used to feed the predator: Helicoverpa zea eggs (CEW) and Spodoptera exigua larvae (BAW), which carried the Bt Cry1Ac toxin to the higher trophic level. In 2010, Bahar et al. (2010) explored whether the transgenic Bt cotton affected a generalist predator’s foraging behaviour on Helicoverpa armigera eggs or neonate larvae in the presence of the aphid Aphis gossypii (green lacewing larvae: Mallada signata Neuroptera: Chrysopidae) (Homoptera: Aphididae). Aphid nymphs, “H. armigera” (eggs or larvae), or both, attacked Bt and conventional cotton plants. Two 4-day-old “Mallada signata” larvae were introduced onto each plant during lacewing treatments. A laptop computer was used to record observations of lacewing activity (resting, searching, or feeding) and position (upper side of leaf, lower side of leaf, stem, petiole, square, flower, or boll) across three times (morning, noon, and afternoon) each day for three consecutive days. The lower surface of leaves accounted for the majority of green lacewing larvae’s time in all cases (26%– 60%), followed by fruit buds (cotton squares) (20%–22%). However, in green house cubicles, resting was the most frequent activity (68%) followed by resting (27%) and feeding (5%). In growth cabinet studies, searching was the most common activity (53%) followed by resting (40%) and feeding (7%). However, these analyses found no discernible change in activity between transgenic Bt and regular cotton. Overall, Bt cotton plants had little to no impact on how Mallada signata foraged (Bahar et al. 2010).

10.10.4 Neuroptera In maize, Chrysoperla carnea is a significant predatory insect. The performance of three prey herbivores (Rhopalosiphum padi, Tetranychus urticae, and Spodoptera littoralis) on transgenic Bt and non-transgenic maize plants, the intake of the Cry1Ab toxin by the three herbivores, and the effects on C. carnea when fed each of the prey species were all examined in order to evaluate the ecological effects of Bt maize, expressing the C For Rhopalosiphum padi and Tetranychus urticae, the intrinsic rate of natural increase (rm) was used as a performance indicator. Herbivores raised on Bt or non-transgenic plants showed no differences in this metric, according to the research. In contrast, Spodoptera littoralis larvae fed Bt maize showed a higher death rate and a slower pace of development than those fed control maize plants. An immunological assay was used to gauge the different herbivores’ consumption of the Cry1Ab toxin (ELISA). Tetranychus urticae and Spodoptera littoralis had the highest levels of the Cry1Ab toxin, while Rhopalosiphum padi had just minimal amounts. The survival, growth, or weight of Chrysoperla carnea were unaffected by feeding it Tetranychus urticae, which was shown to possess the Cry1Ab toxin, or

10.10

Expression of Bt Toxins Through Their Preys

329

R. padi, which does not consume the toxin. When predators were fed S. littoralis larvae raised on Bt maize, on the other hand, a significant rise in mortality and a holdup in development were seen. The detrimental effects seen on Chrysoperla carnea when fed Spodoptera littoralis could be attributed to a combination interaction between poor prey quality and the Cry1Ab toxin. These results are reviewed in relation to the ecological dangers of Bt maize on Chrysoperla carnea (Dutton et al. 2002).

10.10.5 Thysanura The effects of Bt maize expressing the Cry1Ab toxin on the thrips Frankliniella tenuicornis (Thysanoptera: Thripidae) were researched as part of a risk assessment method, and possible dangers for predators feeding on thrips on Bt maize were evaluated. By comparing life-table parameters between Frankliniella tenuicornis grown on Bt and non-Bt maize, the impacts of Bt maize on this species were evaluated. To assess the potential exposure of predators when feeding on thrips, the Cry1Ab toxin content in various stages of Frankliniella tenuicornis fed on Bt maize and the persistence of the toxin in adults were determined. Additionally, Chrysoperla carnea (Neuroptera: Chrysopidae) was utilised as a model predator to examine how prey and predator behaviour may affect how much the Bt toxin is exposed to a natural adversary. According to results from life-table parameter analysis, Frankliniella tenuicornis was unaffected when it was raised on Bt maize. This suggests that there is little chance of predators being impacted by prey quality. Bt content was highest in thrips larvae and adults and lowest in the pre-pupal and pupal phases, which do not feed. In adult Frankliniella tenuicornis, the Cry1Ab toxin had a brief half-life and decreased by 97% in the first 24 h. Young Chrysoperla carnea larvae had varying degrees of success in predation on the various thrips stages, showing that the prey stage can also affect how much Bt toxin is exposed to predators. When the present understanding of the vulnerability of the main thrips predators is combined, our results indicate that there is no risk to predators when consuming thrips in or near Bt maize fields since there is minimal chance for prey quality-mediated effects, the thrips have a low toxin concentration, and the thrips have a brief persistence (Obrist et al. 2005). The beet armyworm, Spodoptera exigua (Lepidoptera: Noctuidae), and the predator Podisus maculiventris (Heteroptera: Pentatomidae), as well as the two-spotted spider mite, Tetranychus urticae (Acarina: Tetranychidae), and the predatory big-eyed bug Geocoris punctipes (Heteroptera: Heter (Heteroptera: Anthocoridae). We measured the levels of the toxin Cry1Ac in cotton plants, pests, and predators, as well as the impacts of continuous feeding on Spodoptera exigua larvae fed either Bt or non-Bt cotton on the life cycle features of Podisus maculiventris. The Cry1Ac toxin may be transferred from all three herbivores to their respective predators. T. urticae showed among the herbivores 16.8 times more toxin in their bodies than that expressed in Bt cotton plant, followed by Spodoptera exigua (1.05 times) and immatures and adults of Frankliniella occidentalis (0.63 and 0.73 times,

330

10 Predation Ethology of Various Orders

respectively). Four percent, 40%, 17%, and 14% of the toxin present in the corresponding herbivorous prey were found in the predators Geocoris punctipes, Podisus maculiventris, O. insidiosus, and N. roseipennis, respectively. Regardless of the prey’s food source, Podisus maculiventris displayed similar life history traits (developmental period, survival, longevity, and fertility). Consequently, depending on the herbivore species, varying amounts of the Cry1Ac toxin are transmitted to natural enemies through non-target herbivores, yet lifelong exposure to the toxin by the predator Podisus maculiventris through its prey had no effect on the predator’s life history. Predatory heteropterans are not harmed by feeding on Cry1Ac-contaminated non-target herbivores, according to the results presented here and those previously published. As a result, growing Bt cotton may offer a chance to preserve these predators in cotton ecosystems by lowering insecticide use (Torres and Ruberson 2008). As expected for a largely susceptible lepidopteran species, Zhang et al. (2006a) found that a Bt cotton variety expressing both Cry1Ab and Cry1Ac toxins and a variety expressing solely Cry1Ac toxin decreased weight gain, developmental rate, and survival for second instar Spodoptera litura. First instars of the predatory coccinellid Propylea japonica were fed on these contaminated larvae, which resulted in lower weight increase (18%) and survival of the predators compared to those reared on prey from non-Bt cotton. Contrarily, when comparing the effects of prey reared on a single or dual Bt toxins transformed variety (Cry1Ab and Cry1Ac), the predators fed prey reared on the dual-gene plants exhibited greater weight gain (14%) and larval survival (20%) than those given prey from the single-gene plant, despite the fact that S. litura larvae reared on the dual-gene plants contained 26% more toxin than larva Because prey larvae carried more toxins to the third trophic level, predator larvae developed and thrived better on them. This finding implies that the findings of these tests during their brief evaluation time may have been influenced by variables other than the poisons themselves and contaminated prey. Additionally, this same predator, Propylea japonica, showed similar growth, survival, and fecundity to predators raised on prey from non-Bt plants when fed either cotton aphid, Aphis gossypii reared on Bt cotton varieties, or planthopper, Nilaparvata lugens (Homoptera: Delphacidae), reared on C (Zhang et al. 2006b; Bai et al. 2006). Whether obtained by prey or directly from pure toxins, retention of ingested Bt toxins in non-targeted herbivores and predators seems to be quite brief. Cry1Ac levels in the predator P. maculiventris in the current study reduced to 13% and 0% within 24 h and 48 h of the predators switching from Bt-intoxicated food to BAW larvae reared on non-Bt cotton. G. punctipes consuming 16 and 32 ppm purified Cry1Ac toxin water concentrations likewise demonstrated short toxin retention (Torres et al. 2006). According to these authors, quantifiable amounts at lower and higher Cry1Ac concentrations were present for up to 24 and 48 h, respectively, and significant declines were seen 1, 12, 24, 48, and 72 h after intake. Additionally, G. punctipes faeces in the time ranges of 0–12, 12–24, and 24–48 h had substantial quantities of toxin. When the Bt maize-raised thrips F. tenuicornis were switched to non-Bt maize for 24 h, only about 3% of the initial toxin levels present on day 0 remained in the thrips’

10.10

Expression of Bt Toxins Through Their Preys

331

bodies, and only a trace of toxin was detected at day 4. Obrist et al. (2005) quantified the time of retention and excretion of Cry1Ab toxin by the thrips, and they discovered additionally, the author noted that the thrips’ dry faeces had high concentrations of the Cry1Ab toxin. These findings, along with those for Podisus maculiventris and G. punctipes, imply that the persistence of Bt toxins acquired by herbivores and transferred to predators is quite low. The fate of ingested Bt toxins in the digestive system of heteropterans is currently unknown. We can speculate that, given the large levels seen in their faeces, the toxin is either expelled as undigested food or, at the very least; a portion of it is sufficiently intact to cause an ELISA reaction. Whether the poison is broken down or not, the predatory heteropterans’ brief toxin persistence in their bodies is likely due to their pre-oral digestion and liquid feeding habits, which allow for significantly shorter food retention durations in the stomach for digestion and absorption than solid feeders. According to the findings of Dutton et al. (2002), Obrist et al. (2005, 2006), and the ones presented here, there is a chance that predatory heteropterans will have detectable levels of Bt toxin in the wild; however, this chance will depend on the quantity and types of prey consumed, as well as the prey’s ability to absorb and concentrate the toxin. In this instance, the small predatory heteropterans (Geocoris, Orius, and Nabis) are more likely to acquire detectable levels of Bt toxin from the higher Bt toxin concentration in particular herbivorous prey, such as two-spotted spider mites and western flower thrips, compared to prey with lower toxin concentrations, such as BAW larvae. On the other hand, larger predators that eat a lot of infected prey, such the Podisus maculiventris, can pick up enough poison from this low-concentration prey to reach detectable levels. The absence or minimisation of negative effects on natural enemies is crucial because the final product may lead to a synergistic outcome when predators consume more prey that are partially affected by the toxin expressed in the plant without affecting predator performance when considering the multi-pest cotton ecosystem (Bt against lepidopteran larvae plus natural enemies against non-targeted Bt pests). BAW larvae and other partially sensitive lepidopteran species show delayed growth rates in the Bt cotton ecosystem, extending the window of larval sensitivity to these predators and resulting in a favourable interaction between host plant resistance and natural enemies (Bottrell et al. 1998). According to research by Torres et al. (2006), Podisus maculiventris swallowed 10.1 more BAW larvae of the same age that were raised on Bt cotton than those that were raised on non-Bt cotton (5.7 larvae). The current research’s findings, however, make it evident that greater ingestion of intoxicated caterpillars has no negative effects on the life history of this predator. The big-eyed insect, G. punctipes, also displayed similar life history traits when fed BAW larvae housed on cotton plants that were either Bt- or non-Bt-treated in the wild (Torres and Ruberson 2006). We concentrated on the life history traits of only Podisus maculiventris despite the fact that all three of the predatory bugs investigated in this study are significant in cotton fields and that the results of this study showed their exposure to the Cry1Ac toxin through prey fed Bt cotton. This was done for two main reasons. First of all, it is a frequent predator in cotton fields like the other heteropterans investigated here; nevertheless, it is the only heteropteran predator discovered to have detectable

332

10 Predation Ethology of Various Orders

amounts of Cry1Ac toxin in feral field populations (Torres et al. 2006). Second, prior research examined the growth and reproduction of G. punctipes, which feeds on BAW larvae raised on Bt cotton (Torres and Ruberson 2006). Torres and Ruberson (2006) assessed G. punctipes’ growth and reproduction over the course of two cotton seasons while feeding it young BAW larvae trapped on Bt cotton (Cry1Ac toxin; variety DPL 458RR) and non-Bt cotton (variety DPL 5690RR). The life history of the predator was unaffected by Bt and non-Bt cotton plants, or by the toxin being ingested by prey, as was the case in the current investigation. The predators Orius majusculus and O. insidiosus both prey on thrips Anaphothrips obscurus and lepidopteran Ostrinia nubilalis larvae, respectively, that were raised on Bt maize (Zwahlen et al. 2000; Al-Deeb et al. 2001). Nabids haven’t been researched yet. However, when all of these findings are taken into account, there isn’t much proof that Bt toxins harm predatory bugs, which are a significant predator in many agroecosystems (at least Cry1Ac). Our findings add to the growing body of knowledge on the population dynamics of naturally occurring predatory taxa in Bt cotton fields by providing more detailed information on the tritrophic interactions of Bt cotton, herbivores, and predatory heteropterans. Although management of herbivorous pests not targeted by the Bt toxin still benefits from suppression given by predators and parasitoids prevalent in the crop agroecosystem, genetically engineered cotton expressing Cry1Ac or combination Cry1Ac/2Ab is particularly effective against numerous lepidopteran larvae. Even while levels of toxin in the prey can be much greater than those expressed by the genetically modified plants, the results of the earlier tests mentioned above and those of the current study come to the same conclusion: predatory bugs are unaffected when feeding on prey carrying Bt toxins. We investigated whether the Cry1Ac toxin was transferred up the food chain from the Bt insect-resistant cotton variety A26–5 to the herbivorous prey Helicoverpa armigera and then to the non-target predatory insect Arma chinensis in order to evaluate the potential ecological impact (Fallou). Cry1Ac protein was found in trace amounts in both Helicoverpa armigera and Arma chinensis, suggesting that Cry1Ac might be passed from one species to another. However, after treatment with Bt-fed or Bt-free H. armigera, there were no variations in the survival rate, development, or fecundity of Arma chinensis. Due to our discovery that Cry1Ac could not connect to receptors in the midgut, it is likely that exposure to Cry1Ac had no effect on the expression of the detoxification genes in Arma chinensis. These findings demonstrate that the Bt cotton A26–5 has minimal to no harmful effects on Arma chinensis (Ma et al. 2022).

10.11 Intraguild Predation (IGP) or Competition Dominates Competition and predation are two important animal interactions that frequently occur together in nature. When potential competitors for a common prey (the extraguild prey: EGP) instead choose to prey on one another, this phenomenon is known as intraguild predation (IGP). Studies on biological control now focus a lot

10.12

Best Examples

333

on interactions between predator guild members and how they affect prey populations. When intraguild predation (IGP) or competition predominates, these intraguild interactions may lead to the additive or synergistic predation of target pests or to antagonistic interactions that reduce predation efficiency, relieving the prey from predation pressure. Every IGP encounter involves a top predator (aggressor), also known as an intraguild predator (IG predator), and one or more predatory species that have been attacked and killed by the top predator (also known as intraguild prey, or IG prey). IGP, which is a pervasive interaction in several guilds that includes polyphagous predators and, to a lesser extent, specialists, may have an impact on the participating species’ spatial distribution, seeking behaviour, and abundance. Prey density, habitat complexity, body size, aggression, and other factors. Predation that involves two co-occurring predator species competing for the same prey results in the death and eventual ingestion of one of the predators is known as intergroup predation. The top predator is known as the intraguild predator, the prey that it consumes is known as the intraguild prey, and their shared resource is known as the extraguild prey. The composition, quantity, and dynamics of natural enemy guilds are all impacted by IGP, which has been recognised as one of the major mortality drivers that may destabilise communities. IGP is a vital survival strategy that is especially prevalent in generalist predators, yet it frequently gets in the way of these biological control agents’ efficiency. In aphid colonies, the green lacewing Chrysoperla carnea (Neuroptera: Chrysopidae) and the variegated ladybird Hippodamia variegata are two significant predators of the black bean aphid Aphis fabae (Coleoptera: Coccinellidae). These two species are among the most prevalent aphid predators in various agroecosystems, but due to field survey taxonomic constraints, their current presence is infrequently reported. The green lacewing’s larval stage is predatory, but the ladybird’s larvae and adults are both predators. As a result, both species are likely to engage in IGP interactions with other members of the aphidophagous guild as well as with one another.

10.12 Best Examples 10.12.1 Miridae and Anthocoridae Miridae, as well as the coexistence of the generalist predators Macrolophus pygmaeus (Miridae) and Orius laevigatus (Anthocoridae), and their management of two pests (aphids and thrips) in a commercial sweet pepper crop grown in a greenhouse. Both predators feed on thrips and aphids, ingest pollen and nectar from sweet pepper blossoms, and Orius laevigatus preys on Macrolophus pygmaeus within its own guild. Two predator species coexisted for 8 months in the greenhouse with low pest numbers, according to observations. The predators obviously controlled thrips and aphids more effectively while working jointly, according to a greenhouse trial. Orius laevigatus greatly improved the control of thrips, and

334

10 Predation Ethology of Various Orders

Macrolophus pygmaeus significantly improved the control of aphids. Therefore, combining Macrolophus pygmaeus and Orius laevigatus inoculative releases to control both thrips and aphids in greenhouse-grown sweet pepper appears to be a good approach. The predators may remain in a crop for a considerable amount of time, and they work in tandem to control both pests (Messelink and Janssen 2014).

10.12.2 Coleoptera and Coleoptera The Pistachio psyllid, Agonoscena pistaciae (Hemiptera: Psyllidae), a significant pest of pistachio, is preyed upon by Oenopia conglobata and Menochilus sexmaculatus (Coleoptera: Coccinellidae). To assess the possibility of intraguild predation (IGP) between these two species, identify any potential asymmetries in IGP interactions, and characterise the sensitivity of IGP to varying densities of two extraguild (EG) prey, Agonoscena pistaciae and Aphis gossypii, Ranjbar et al. (2020) conducted a laboratory experiment (Hemiptera: Aphididae). We assessed IGP rates in susceptible early life stages of the other species using female adults and fourth instar stages that had both been deprived for 12 h (eggs, first and second instar stages). After 12 h, prey consumption for IG and EG was recorded. All parameters affected IGP rates, but the biggest variance was caused by differences in IG predator species, EG prey life stages, and EG prey densities. IGP dropped parabolically as prey concentrations rose. The IGP of eggs was higher than that of first instars, which was higher than second instars, indicating a fall in palatability as the organism aged. In both predation and IGP, the larger Menochilus sexmaculatus showed greater voracity than the smaller Oenopia conglobata. In all treatments, both species consumed more Agonoscena pistaciae than Aphis gossypii, which might be due to shorter handling times or lower food value per prey (Ranjbar et al. 2020).

10.12.3 Coleoptera and Dermaptera In Petri dishes, 24 h of predation by 1st (H1) to 4th (H4) instar larvae of Harmonia axyridis (Coleoptera: Coccinellidae) and 3rd (F3) and 4th (F4) instar nymphs of Forficula auricularia (Dermaptera: Forficulidae) were observed. The densities used were 5, 10, 25, 50, and 75 (RAA). With age, both predator species became more ferocious. For H1, H2, H3, H4, F3, and F4, respectively, the satiation level, or the greatest number of eaten RAA, reached an average of 17.8, 36.3, 49.4, 79.6, 62.9, and 84.3 aphids per day. Few signs of intraguild predation were present in the combined treatments. At modest RAA densities (25 aphids), 2.9% of H1 and H2 died or vanished. When the two predators were together in the same arena, they consumed 7.9% less RAA than the total of their individual consumptions, at the non-limiting RAA 41 density of 200. However, the advantages of combining these two predators to maximise prey consumption greatly surpassed this effect, which was quite tiny. For the remaining six RAA densities, a significant part (> 90%) of the RAA was devoured by predators working together. Our findings support the use

10.12

Best Examples

335

of these two predator species in conjunction, particularly when their eldest instars are used, as promising aphidophagous possibilities in the early spring against RAA (Dib et al. 2020).

10.12.4 Coleoptera and Neuroptera Zarei et al. studied intragroup predation between the lacewing Chrysoperla carnea and the ladybird Hippodamia variegata in simple laboratory settings and more complicated microcosm environments (2020). Six combinations of predator life stages were given three initial concentrations of 50, 150, and 400 Aphis fabae third instar nymphs each, along with a control without aphids (second and third larval instars of lacewing and third and fourth instars and adult females of ladybird). After 24, 48, and 72 h, the amount of aphid density and the presence of IGP were examined. In both the simple arena (reaching 0.6 between larvae in the absence of EGP and 0.3 between lacewing larvae and ladybird females) and microcosm setting, the IGP intensity (IGP level, IL) was comparable (0.3 without EGP). According to a negative exponential relationship, increasing EGP density decreased IL in both situations. IGP was asymmetric (overall average asymmetry was 0.82 in simple arena and 0.93 in microcosm; the difference was not significant) and largely in favour of C. carnea larvae, with the exception of when Chrysoperla carnea second larvae were combined with Hippodamia variegata fourth larvae and adults. During the course of the experiment, the IGP’s direction partially changed but not other properties. The frequency of IGP interactions between aphid predators in actual environments and how they affect aphid biological control (Zarei et al. 2020).

10.12.5 Coleoptera, Hemiptera, and Diptera For the management of aphids on greenhouse crops, Adalia bipunctata (Coleoptera Coccinellidae), Macrolophus pygmaeus (Hemiptera Miridae), and Aphidoletes aphidimyza (Diptera Cecidomyiidae) are advised. Devee et al. studied the behaviour of these three aphidophagous predators and their instars in combinations of two or three species, including Adalia bipunctata (second, third, and fourth larval stage), Macrolophus pygmaeus (second and fifth nymphal instar), and Aphidoletes aphidimyza (second and third larval stage), in the absence and presence of extraguild (Devee et al. 2018). Predations on other guilds and within them were both noted. Interguild predation shows that no killing incidents between Adalia bipunctata and Macrolophus pygmaeus were documented. The third instar of Adalia bipunctata, however, ingested 20% of the third instar larvae of A. aphidimyza. Adalia bipunctata (fourth instar) ingested 15% and 25% of the second and third instar larvae of Aphidoletes aphidimyza, respectively. Fifteen percent of the larvae of Aphidoletes aphidimyza (third instar) were devoured by Macrolophus pygmaeus (fifth instar). It was discovered dead in two and one replicates, respectively, when Aphidoletes aphidimyza second instar was employed with either Adalia bipunctata

336

10 Predation Ethology of Various Orders

fourth instar or Macrolophus pygmaeus fifth instar or with Adalia bipunctata third instar and Macrolophus pygmaeus fifth instar (Devee et al. 2018). When each predator was tested separately, the results of the aphid consumption were greatly influenced by the predator’s species and instar. The Adalia bipunctata fourth instar stage had the largest consumption, followed by its third and second instar stages. The latter consumed more aphids than either Macrolophus pygmaeus fifth or Aphidoletes aphidimyza third, but there was no discernible difference between their predation rates and those of the predators’ instars. Between the observed values and the expected values, there were no discernible discrepancies. In the cases of Adalia bipunctata fourth, A. aphidimyza second, and Macrolophus pygmaeus fifth instars and Aphidoletes aphidimyza third, Aphidoletes aphidimyza third, and Macrolophus pygmaeus second instars, the observed values were much higher than the anticipated one (Devee et al. 2018).

10.12.6 Among Hemipterans IGP significantly altered the interactions between the effects of lacewings and either Zelus renardii or Nabis’s predators on the population reduction of the aphid Aphis gossypii. Generalist predators in cotton include members of the following four genera of predatory true bugs (Hemiptera): Geocoris spp. (Geocoris pollens, Geocoris punctipes, and Geocoris atricolor (Lygaeidae), Nabis spp. (Nabis alternatus and Nabis americoferus (Nabidae), and Zelus spp. (Zelus renardii). i) Nabis are described as feeding on aphids, leafhoppers, Lygus bugs, and lepidopteran caterpillars; (ii) Geocoris are described as feeding on lepidopteran caterpillars, Lygus bugs, whiteflies (mostly Trialeurodes vaporariorum and Bemisia tabaci), leafhoppers (mostly Empoasca spp.), spider mites (Tetranychus spp.), and a (iv) Orius tristicolor is said to prey on small arthropods like lepidopteran caterpillars, thrips (mainly Frankliniella occidentalis), spider mites, whiteflies, aphids, Lygus nymphs, and insect eggs. The nymphal or adult stages of other predators significantly decreased the chance of lacewing larvae surviving (mostly Zelus and Nabis, but also perhaps Geocoris). Strong connections between lacewings and Nabis and Zelus renardii were noted by Rosenheim et al. (1993), and there was some indication of a weaker contact between lacewings and Geocoris. This is the first evidence that IGP is influencing features of arthropod community organisation that go beyond the predator guild in Aphis gossypii predators (Rosenheim et al. 1993). We examined predatory interactions between adult females and first instars of Dicyphus cerastii vs. Nesidiocoris tenuis as well as Dicyphus cerastii vs. Macrolophus pygmaeus, as this species is also naturally present in horticultural crops in Portugal, to determine whether intraguild predation (IGP) can explain the declining abundance of Dicyphus cerastii. The same three species’ cannibalistic interactions were also examined. Each experiment was conducted in a Petri dish arena under controlled laboratory circumstances with or without the use of eggs of Ephestia kuehniella (Lepidoptera: Pyralidae) as a substitute prey. Conspecific and heterospecific nymphs were both preyed upon, but only when there was no

10.13

Role of Molecular and Other Techniques

337

other food available. Dicyphus cerastii and M. pygmaeus engaged in symmetrical and reciprocal intragroup predation. IGP, on the other hand, was asymmetrical, favouring Dicyphus cerastii over Nesidiocoris tenuis. These mirid species did not considerably differ in their cannibalism. According to our findings, Dicyphus cerastii is more able to devour intraguild prey than Nesidiocoris tenuis. IGP on tiny nymphs cannot account for the change in abundance between Dicyphus cerastii and Nesidiocoris tenuis (Oliveira et al. 2022).

10.13 Role of Molecular and Other Techniques In 2004 and 2005, the average percentage of coccinellids that had DNA from other coccinellids in their guts, according to non-weighted data, was 46.8% and 58.9%, respectively. In 2004, but only somewhat in 2005, the ranking of predators in terms of IGP strength was altered by using the weighted DS50 values (DNA detection success over time). Propylea quatuordecimpunctata was ranked above Coleomegilla maculata lengi, Harmonia axyridis, and Coccinella septempunctata in 2004 based on raw data. The ranking of Harmonia axyridis, Coccinella septempunctata, and P. quatuordecimpunctata over C. maculata was determined by using weighted DS50 values. Because of this, H. axyridis and Coccinella septempunctata undervalue the relative importance of IGP when using raw data. In 2005, species-specific relative IGP rates were more comparable. It was essentially unchanged when weighted DS50 values were used, with the exception that the relative strengths of IGP for Coccinella septempunctata and Propylea quatuordecimpunctata were the same. The ranking using raw was: Coleomegilla maculata lengi > Harmonia axyridis > Coccinella septempunctata > Propylea quatuordecimpunctata (Gagnon et al. 2011). Second, the findings suggest that IGP is reciprocal, with all four species of coccinellids feeding on all three of the other species. Although IGP levels were high in both years, there were differences in the relative number of intraguild prey species. Harmonia axyridis was a prominent intraguild prey species in 2004, but Propylea quatuordecimpunctata and Coccinella septempunctata took over as the predominant intraguild prey species in 2005 (Gagnon et al. 2011). When the data from the 2 years are combined, 11.8% of the intraguild predators had the DNA of two additional coccinellid species in their guts, and 1.4% of the sampled coccinellids had three intraguild prey species simultaneously in their guts. Harmonia axyridis and C. maculata consumed two intraguild prey species the most frequently (48.1% of all instances and 35.7%, respectively), whereas only Harmonia axyridis consumed three intraguild prey species (Gagnon et al. 2011). The IGP between predators and their preys has been suggested using recent past molecular techniques. This method is mostly used for varied ecological conditions, distribution patterns, etc. For instance, molecular methods (polymerase chain reaction) were used to check the guts of 177 Harmonia axyridis larvae that were fieldcollected in 2010 in England, France, Germany, Slovakia, and the Czech Republic for the presence of different insect (coccinellid, syrphid, and chrysopid) prey DNA.

338

10 Predation Ethology of Various Orders

Adalia decempunctata, Adalia bipunctata, and Episyrphus balteatus were three of the four target prey species that were found in the intestines of Harmonia axyridis at the following rates: 9.6%, 2.8%, and 2.8%, respectively. IGP was not found on Chrysoperla carnea. In England, France, Slovakia, and the Czech Republic, IGP detection of at least one target species was made, but not in Germany. These findings support the notion that Harmonia axyridis is a very versatile predator (Brown et al. 2015). Another contemporary method for tracking predator movement from one crop to the adjacent crop is isotopic analysis. Madeira et al. examined the migration patterns of Orius majusculus, Orius niger, and Nabis provencalis between alfalfa and maize (Madeira et al. 2019). According to the authors, carbon and nitrogen stable isotope analysis was used to determine whether one crop was a “donor” and/or whether another was a “receptor”. All heteropteran species can transfer between crops, according to the results, although the movement of the predators is species-specific, and the function of alfalfa and maize as “donor” or “receptor” varies. The movement is also impacted by agricultural management techniques like cutting alfalfa. N. provencalis returned to alfalfa from maize at the same time as the other heteropterans that initially colonised maize. Heteropterans travel from alfalfa to maize as a result of cutting it, but the timing of the practise also influences the movement. Cutting of the alfalfa had little effect on its recolonisation, and some of the heteropterans discovered during the alfalfa’s regrowth period appeared to have persisted within the plant (mostly Nabis provencalis), while others appeared to have originated from maize (mainly Orius majusculus and Orius niger). Orius majusculus shifted to alfalfa later than the other two predators after maize was harvested or dried. Alfalfa and maize adjacent fields or a mosaic of both crops at the farm and landscape levels could improve conservation biological control due to the heteropteran’s ability to travel between the two crops and their varied roles as “donor” or “receptor” (Madeira et al. 2019). Identification of predators with a given prey’s specialised predators is a crucial function of the molecular approach. Stronger linkages between predator and prey are required as a result of the evolution of one or more behavioural, morphological, or physiological adaptations for predation. Reduviids are primarily generalists when it comes to feeding on other arthropods; however, some groups are thought to specialise on particular prey, such as some Holoptilinae on ants, Phonoctonus (Harpactorinae) on pyrrhocorids, some Emesinae on spiders, and most likely all Ectrichodiinae on millipedes. Reduviids may also include suspected specialised termite predators in certain groupings. In the Neotropics, nautitermes species are specialised predators by Micrauchenus lineola (Harpactorinae) nymphs. As species of Phonolibes, Lophocephala, and particularly Tegea have been seen to prey on termites, specifically Nasutitermes, in the Old World, Tegeini (Harpactorinae) may all be obligate termite predators. In captivity, some genera of Cetherinae, such as Eupheno and Cethera, have consumed termites of unknown species. Two species of Acanthaspis (Reduviinae) have been recorded to feed on termites, but not primarily, while Neivacoris steini (Reduviinae) has been discovered in Brazilian termite nests (likely Cornitermes [Nasutitermitinae]). PCR was used to analyse DNA from the

10.14

Anti-Predatory Acts

339

gastrointestinal contents of 50 people using primers specific to the prey (Gordon and Weirauch 2016).

10.14 Anti-Predatory Acts “Primary defences” are anti-predatory mechanisms that reduce the possibility of physical contact between predator and prey, whereas “secondary defences” start once subjugation or contact has started. Formic acid, p-benzoquinones, and other phenolic chemicals are among the numerous toxic and irritating exudations that millipedes, caterpillars, beetles, and ants emit when they pass away (Eisner et al. 1963). Reduviid predators have a distinctive habit once the prey is rendered immobile by such repellents, spending a lot of time cleaning the body and antennae with their forelegs and wiping the body surfaces on the ground (Haridass and Ananthakrishnan 1981). Many insects have developed the ability to cryptically shape and colour their bodies to help them hide from predators. Crypsis, warning colouring, and mimicry are only a few examples of the principles covered by adaptive coloration. There are also many more, less well-known visual phenomena including masquerade, countershading, and disruptive coloration. An important contribution to understanding the behavioural relevance of animal colour patterns was made by early naturalists and bug enthusiasts. Many insects use visual, chemical, tactile, electric, and, in this case, acoustic cues to reduce the likelihood of being noticed when potentially observable by an observer. Vision, hearing, and olfaction are the main sensory modalities used by insects to hunt insects, with a bias towards vision in daytime hunters and towards hearing and olfaction in night-time hunters. For instance, crypsis, masquerade or disruptive camouflage, autotomy or falling, or simulate death are some examples of primary defences that prevent detection. For instance, morphological defences like the presence of hairs, spines, or hard exoskeletons, chemical defences like toxins, and behavioural defences like aggressive behaviour are examples of secondary defences that deter attack after identification. These secondary defences also include complex signalling defensive tactics like aposematism and imitation (Wedmann et al. 2021). Other insects, spiders, fish, frogs, lizards, birds, and mammals are some of these predators. Primary Defences The nymphs of the West African assassin bugs Paredocla and Acanthaspis spp. wrap their bodies with a “dust coat” of dirt, sand, and soil, and they also stow a “backpack” of heavier items, such empty prey corpses and plant pieces, on their abdomens. Spiders, geckos, and centipedes are three of its probable predators. According to Brandt and Mahsberg’s observation, camouflaged Paredocla and Acanthaspis spp. nymphs had a much higher survival rate than denuded bugs (2002). Acanthaspis, Alloeocranum, Edocla, Empyrocoris, Holotrichius, Leogorrus, Paredocla, Tetroxia, Reduviinae, Cetherinae, Salyavatinae, and Stenopodainae of the reduviidae were also shown to exhibit camouflaging behaviours. Many creatures, including reptiles, amphibians, mammals, birds, fish,

340

10 Predation Ethology of Various Orders

echinoderms, crustaceans, spiders, and insects, frequently experience autotomy, or appendage removal. The only order within the class Insecta that frequently sheds and regenerates lost legs is the Phasmida, despite the fact that many species within the phylum Arthropoda experience autotomy. In the lab, difficulties related to moulting account for the loss of legs 30% of the time. A sampled population in the field revealed that 40% of people had lost or regenerated legs (Maginnis 2008). Reduviidae Numerous reduviid predators have been observed using the primary defence of pretending to be dead. Adults and nymphs of Rhinocoris albopunctatus suddenly fell to the ground and pretended to be dead (Nyiira 1970). These behavioural traits were described as being typical of the Harpactorinae by Miller (1956). Nymphs and adults of Irantha armipes and Sphedanolestes signatus both displayed death feigning during predation (Srikumar et al. 2014). Coleoptera Phidippus audax (Salticidae) spiders attacked Rhinocyllus conicus adults on multiple occasions while they were moving but released them after they stopped moving. This behaviour seemed to discourage the spiders from attacking these adults. The population of Rhinocyllus conicus was not affected by the general spider population since it was too low (less than 1 per 100 plants) (Dowd and Kok 1981). The unlearned adoption of a motionless posture by a prey person in response to physical contact or being in close proximity to a predator—rather than an injury caused by the predator—is known as tonic immobility (TI) (or another antagonist). The prey’s physical vulnerability if the attack is pursued or the predator’s ability to locate or identify it through sensory means is unaffected by the prey’s position. Even after the predator has left, the prey remains in a condition of motor inhibition for a while, and during this period, they are less receptive to external stimuli (although monitoring of the environment can still occur). The prey can return to its initial physiological state after TI if there is no mortality or harm (Humphreys and Ruxton 2018). However, the predator has planned to follow the prey’s behaviour while also determining the best moment to capture it. Secondary Defences raptorial limbs present (found in mantid insects), There are two rows of at least seven strongly re-curved teeth in front of at least three weakly re-curved teeth on the inner surface of the right maxillary stylet of Deraeocoris nigritulus (Hemiptera: Miridae), all of which point away from the head. Emesinae, or thread-legged bugs, wait for their prey and use their lengthy raptorial forelegs, which are armed with reduviid tubercles and spines, to pounce on it. Furthermore, harpactorines use their long rostrum to pin and poke their victim. The welldeveloped tibial pads of the peiratines, reduviines, and some ectrichodines are used to chase, pounce on, and capture their prey. The mandibular stylet tip of Pentatomidae, Asopinae has five irregular teeth and three long, pointed hooks; the apices of the right maxilla have small teeth and a few short barbs along the edge of the food canal. They also have a distribution and abundance of receptor sensilla with nanopores (St1, Sb3). The coccinellid larvae are concealed from the prey by their

10.15

Chemical Ecology

341

peculiar shapes, stealthy movements, and maybe chemical camouflage. Pupae were assaulted by prey, but their thick coat of hair provided protection. Adults were also attacked, and they reacted by either running away or by pressing their bodies firmly against the surface of the plant.

10.15 Chemical Ecology Natural enemies use a variety of environmental cues during each of the aforementioned foraging behavioural phases to locate their host and prey (hereafter jointly referred to as herbivores). Although natural enemies take advantage of kairomones and visual cues, the former is well established to be crucial to their feeding habits. Kairomones are chemical cues that help organisms that can sense them interact with other species. During their foraging behaviour, or the location of oviposition sites and feeding resources in nature, natural enemies listen in on kairomones. Natural enemies are likely to react more strongly to kairomone combinations than to a single kairomone. Through the attraction and retention of natural enemies, kairomonebased lures are used to improve biological control tactics in an environmentally acceptable manner. This reduces insect pest numbers and agricultural loss. Arthropod herbivore damage causes a variety of plant metabolic processes, which leads to the release of herbivore-induced plant volatiles (HIPVs). The ecological roles that these HIPVs perform in mediating interactions among various trophic levels are significant. For instance, predators, parasitoids, and natural enemies of herbivores use HIPVs as clues to find their host or prey. Natural enemies have evolved to take use of information about the host or prey suitability based on HIPVs since the HIPV blend differs depending on the herbivore identity and developmental stage. The first molecules released by plants in response to herbivore damage are often green leaf volatiles (GLV), whereas MeSA is released later and is still detectable from plants hours or days later. Terpenoids, phenylpropanoids/benzenoids, and fatty acid derivatives (often referred to as “green leaf volatiles” or “GLVs”) are the main components of HIPV blends. Generally speaking, plants that have been harmed by herbivores have a temporal dynamic of volatile emission; GLVs (6-carbon aldehydes, alcohols, and acetates) are released first, followed by terpenoids. Terpenoids and GLVs have both been proven to be essential chemicals for attracting predatory wasps or hosts to host or prey-damaged plants. In contrast to terpenoids, which are present in blends released later and are more particular of the herbivore’s identity, GLVs are widely distributed throughout plant species and provide quick but generic information about host or prey location for natural enemies. So it is anticipated that generalist natural enemies will react to early HIPV blends, where GLVs are predominate, and specialists will react to later blends, where terpenoids predominate (NaranjoGuevara et al. 2017). Depending on what time of day the damage was caused, plants react to herbivory differently. For instance, night-time herbivory cause’s lima bean plants to produce more jasmonic acid than it does during the day, which alters the HIPV blend.

342

10 Predation Ethology of Various Orders

Fig. 10.7 Typical instances of chemical interactions above ground in a tritrophic system. Direct Interaction with Diln Herbivory-Induced Plant Volatiles (HiPlV), Prey, and Predator Pheromones (PyPh and PrPh, respectively), and Volatile Organic Compounds (VOCs)

Because some HIPVs require photosynthesis for their synthesis, diurnal and nocturnal plant volatile emissions differ as well. Only a few studies have evaluated species that appear to be active during the day to determine how natural enemies using nocturnal HIPVs look for prey. Several predators are night-active species, as opposed to parasitoids, which are diurnal. As a result, they must investigate night-time HIPV mixes, either because to their dial rhythm or because they are only attracted to nocturnal HIPVs (Naranjo-Guevara et al. 2017). Each trophic level’s species produce chemical cues (emitters) that other organisms can detect to mediate intraspecific and/or interspecific interactions (receivers). Direct or indirect interactions are also possible. In contrast to indirect interactions, which typically include a third species, direct interactions involve two creatures working together directly without the need for an intermediary. In order to draw natural enemies (receivers) for the purpose of suppressing herbivores (intermediary), plants (emitters) also release HIPVs and OIPVs (indirect interactions). Zoophytophagous insects are sap-sucking predators that feed on plants. As a result of their feeding, plants generate volatiles that draw in both conspecific and heterospecific predators. In Fig. 10.7, this interaction is depicted. When mediating intraspecific interactions between the emitter and its conspecifics, pheromones are highly species-specific (Krost 2008). These categories can be used to classify them: Sex pheromones, aggregation pheromones, marking pheromones, and alarm pheromones. Chrysopa pallens, a predator, uses species-specific semiochemicals, such as the identification and discrimination of odorant-binding proteins (OBPs) and chemosensory proteins (CSPs), to identify and engage with their prey (Li et al. 2015). Predaceous arthropods demonstrate hostility towards their prey after locating the target (Lorenz 1966). In predaceous arthropods, body size and opponent age are

10.15

Chemical Ecology

343

factors that influence aggression and/or fighting success. Numerous genes, including cyp6Q20, tyramine receptor, octopamine receptor, and metabotropic glutamate receptor B, have been implicated in the control of complex behavioural traits, according to earlier investigations (Alaux et al. 2009). The quinolizidine alkaloids are withheld from their leguminous hosts by the larvae of Uresiphita reversalis (Lepidoptera: Pyralidae), who principally store them in the cuticle. With the final larval moult, stored alkaloids are lost. The potato tuber moth, Phthorimaea operculella (Gelichiidae), which is ordinarily appetising to two hymenopteran predators, the Argentine ant, Iridomyrmex humilis (Formicidae), and the paper wasp, Mischocyttarus flavitarsus, was treated with extracts of late instar stages and of pupae (Vespidae). In contrast to untreated prey, Phthorimaea operculella larvae treated with alkaloid extracts of Uresiphita reversalis larval exuviae or surface extracts of entire larvae were repulsive to both predators. However, pupal exuviae extracts applied to Phthorimaea operculella were ineffective as a deterrent. Sparteine and cytisine, the genuine alkaloids, were used to treat Phthorimaea operculella larvae, and these hymenopteran predators were similarly deterred. A successful method of defence against at least two frequent predators of lepidopteran larvae appears to be the storage of modest but concentrated concentrations of plant secondary chemicals in the cuticle (Montllor et al. 1991). Salicylaldehyde, the primary component of the prey leaf beetle Phratora vitellinae’s larval secretion, proved extremely attractive to the syrphid fly predator Parasyrphus nigritarsis (Diptera: Syrphidae) (Kopf et al. 1997). According to Reddy et al. (2002), the predator Chrysoperla carnea was particularly drawn to the volatile component of larval frass from Plutella xylostella-infested cabbage plants called allyl isothiocyanate (Neuroptera: Chrysopidae). To the best of our knowledge, the role of HIPVs in the foraging behaviour of predators has only been studied in day-active predatory arthropods, such as mirids (Drukker et al. 2000; Moayeri et al. 2007), pentatomids (Dicke 1999; Weissbecker et al. 1999; van Loon et al. 2000), and anthocorids (Drukker et al. 2000). Rhynocoris marginatus easily interacts with hexane extracts of Mylabris pustulata, Spodopter litura, and Helicoverpa armigera. In comparison to Spodopter litura (40%, 12.5%, 37.5%, 12.5%, and 62.5% of II, III, IV, V, and adult, respectively) and Mylabris pustulata (40%, 37.5%, 25%, 10%, and 40% of II, III, IV, V, and adult, respectively), the response was higher for Helicoverpa armigera extract (50%, 70%, 50%, 50%, and 60% of II, III, IV, V, and adult, respectively) (Sahayaraj 2008). Similar to this, when Rhynocoris marginatus was given water fractions from Helicovrpa armigera, Spodopter litura, and Mylabris pustulata as sources of chemical cues, it chose Helicoverpa armigera (Sahayaraj and Delma 2004). However, Rhynocoris marginatus responded most strongly to Spodopter litura rather than Mylabris pustulata extract. From observations, it was abundantly evident that Rhynocoris marginatus exclusively exhibited behaviours that involved approaching and rostrum protrusion. Similar findings were made by Yasuda and Wakamura (1996) and Yasuda (1997), who demonstrated that the predatory stinkbug Eocanthecona furcellata (Heteroptera Pentatomidae) was drawn to the larval extracts of S. litura from a distance and protruded its proboscis when it was close to the source of the odour (Sahayaraj 2008).

344

10 Predation Ethology of Various Orders

According to the study of Spodopter litura extract, the kairomone extract contained di-n-octyl phthalate, bis (2-ethylhexyl) phthalate, 1,2-benzene dicarboxylic acid, diisooctyl ester, 3,3-dimethyl octane, and 4-methyl decane. Eight substances, including 3-hexyne 2, 5-diol, (E, Z)-2-hexane-1-ol, acetate, (E, Z)6,10-dimethyl-5,9-undecadien-2-one, 1,2-benzenedicarboxylic acid, diiso octyl phthalate, 3,5-dimethyl octane, and 2-methyl heptadecane, were discovered in Mylabris indica. The hexane fraction of the Helicoverpa armigera extract contained compounds such as tridecane, octacosane, 1-iododecane, octodecane, eicosane, pentacosane, heptacosane, and dotriacontane (Sahayaraj 2008). Lenin et al. 2011’s preliminary observations show that Rhynocoris kumarii responded to all of the hexane extracts of the examined insect pests, including Achaea janata, Mylabris indica, and Dysdercus cingulatus. Rhynocoris kumarii responded most to Helicoverpa armigera (4.8 min), Spodopter litura (3.2 min), and Achaea janata (2.5 min) of the lepidopterans, then Dysdercus cingulatus (2.3 min) of the hemipterans, and least to M. indica of the coleopterans (1.9 min). As a result, Rhynocoris kumarii showed the following preferences for the five insect pests’ tested body extracts: Achaea janata > Dysdercus cingulatus > Mylabris indica > Helicoverpa armigera (Lenin et al. (2011). The influence of aphid-induced plant volatiles on ladybird beetle searching behavior was also available in the literature (Ninkovic et al. 2001). In comparison to non-natal HIPVs, a predator of the gall-forming fly Eurosta solidaginis exhibits preferences towards natal HIPVs (Rhodes et al. 2012). As evidenced by the predatory bug Orius laevigatus (Hemiptera: Anthocoridae), which was only drawn to a mixture of 1:2.3 (R)-lavandulyl acetate and neryl (S)-2-methylbutanoate, the main ingredients of Frankliniella occidentalis’ (Thysanoptera: Thripidae) aggregation pheromone, other natural enemy species are drawn to (Vaello et al. 2017). The predator Astata occidentalis is drawn to the presence of methyl (E, E, Z)-2,4,6decatrienoate by stink bugs (Thyanta pallidovirens) and bark beetles (Ips spp. and Dryophthorus americanus) (Cottrell et al. 2014). Methyl (E, E, Z)-2,4,6decatrienoate is attracted to by predatory digger wasp Astata occidentalis (Hymenoptera: Sphecidae) as a host-finding kairomone. In the United States, this substance is utilised in conjunction with the recently discovered H. halys aggregation pheromone to monitor H. halys throughout the growing season (Morrison et al. 2016). The predator Aphidoletes aphidimyza (Diptera: Cecidomyiidae) was drawn to the phenyl acetaldehyde in honeydew produced by the aphid Aphis gossypii (Hemiptera: Aphididae) (Watanabe et al. 2016). The prey and prey host odour reception in two species of lacewings, Chrysoperla carnea and Chrysopa oculata, was examined in a different study by Zhu et al. (2005). Chrysoperla carnea antennae significantly increased their EAG response in response to 2-Phenylethanol, one of the volatiles released from the host plants of their prey (alfalfa and corn). Adult Chrysoperla carnea, mostly female, were attracted in large numbers to traps that were baited with this substance. Only Chrysopa oculata adults were attracted by the sex pheromone component (1R,4aS,7S,7aR)-nepetalactol of an aphid species, Acyrthosiphon pisum. The olfactory neurons of Chrysopa oculata only responded to the aphid sex pheromone

10.15

Chemical Ecology

345

component, but those of Chrysoperla carnea responded to both components, according to single sensillum recordings. As evidenced by the predatory bug Orius laevigatus (Hemiptera: Anthocoridae), which was only drawn to a mixture of 1:2.3 (R)-lavandulyl acetate and neryl (S)-2methylbutanoate, the main ingredients of Frankliniella occidentalis’ (Thysanoptera: Thripidae) aggregation pheromone, other natural enemy species are drawn to a (Vaello et al. 2017). The most plausible defence mechanisms include defensive alkaloids, pyrazines, and quinolones, which shield coccinellid eggs from predators. The alkaloids are known to be present in all developmental stages and are produced by coccinellids. Citrus plants infected with California red scale (Aonidiella aurantii) release volatile organic compounds (VOCs), like methyl salicylate and D-limonene, which cause the predator Rhyzobius lophanthae to act in certain ways (Alsabte et al. 2022).

10.15.1 Field Application Numerous volatiles have been suggested for use at the field level to control pests. For instance, (Z)-3-hexenol can be used to draw natural enemies into field plants that are plagued with herbivores (Wei et al. 2011). Methyl salicylate (MeSA) applications by dispensers in cranberry and soybean fields increased predation and parasitism rates together with a decrease in the number of herbivores on plants (Mallinger et al. 2011; Rodriguez-Saona et al. 2011). The list of synthetic herbivore-induced plant volatiles used in fields to entice natural enemies is also shown in the table below (Ayelo et al. 2021) (Table 10.2). A single compound is working well, or the combination of them working and applied at the field level is very important for field applications. Maeda et al. (2015) compared the attractiveness of a mixture of cis-3-hexen-1-ol, cis-3hexenyl acetate, and methyl salicylate (MeSA), and each of these compounds alone, to a predatory ladybird beetle, Stethorus punctum picipes, to clarify whether a mixture of several volatiles may attract more natural enemies at different spatial scales. In an olfactometer experiment, Stethorus punctum picipes females were attracted to the volatile mixture but not to single volatiles. Male Stethorus punctum picipes were not attracted to this mixture. In a small-scale open-field test using sticky cards, Stethorus punctum picipes were attracted to the volatile mixture and MeSA alone. In the vineyard trapping test, significantly more Stethorus punctum picipes were trapped on sticky cards baited with a mixture of volatiles than on singlevolatile-baited cards.

346

10 Predation Ethology of Various Orders

Table 10.2 Synthetic kairomones utilised for the attraction of various predators at field level (Ayelo et al. 2021) Kairomone Methyl salicylate

Mixture of methyl salicylate, (Z)-3-hexenol and (Z)-3-hexenyl acetate (Z)-3-hexenol and (Z)-3-hexenyl acetate

2-Phenylethanol

cis-a-Bergamotene Octanal

Predators Chrysopa nigricornis Chrysopa oculata Geocoris pallens Orius tristicolor Orius insidiosus Stethorus punctum picipes Coccinella septempunctata Stethorus punctum picipes

Order/Family Chrysopidae

Field crop Vitis vinifera

Geocoridae Anthocoridae

Vitis vinifera Vitis vinifera

Coccinellidae

Vitis vinifera

Coccinellidae

Glycine max

Coccinellidae

Vitis vinifera

Orius tristicolor Coccinella septempunctata Orius similes

Anthocoridae Coccinellidae

Anaphes iole

Mymaridae

Chrysoperla carnea Eupeodes volucris Eupeodes fumipennis Geocoris punctipes

Chrysopidae Syrphidae

Vitis vinifera Gossypium hirsutum Gossypium hirsutum Gossypium hirsutum Glycine max Apple tree

Anthocoridae

Lygaeidae

Chrysopidae

Nicotiana attenuata Gossypium hirsutum Sophora japonica Gossypium hirsutum Malus domestica Apple trees

Chrysopidae

Apple trees

Chrysopidae

Apple trees

Staphylinidae

Brassica oleracea

Miridae

Nonanal

Deraeocoris punctulatus Harmonia axyridis

3,7-Dimethyl-1,3,6-octatriene

Orius similes

Anthocoridae

Isopropanol

Chrysopa quadripunctata Chrysopa nigricornis

Chrysopidae

Squalene Iridodial Mixture of methyl salicylate and Iridodial Benzaldehyde Dimethyl disulfide

Chrysopa oculata Chrysopa nigricornis Chrysoperla plorabunda Aleochara bilineata Aleochara bipustulata

Coccinellidae

References

347

10.16 Future Recommendation • There is a dearth of accurate information about the real predatory behaviours of different predatory insects. Therefore, we advised starting the research. • To effectively utilise biological control, more research should be done on how visual, olfactory, and touch chemosensory inputs are incorporated into behavioural decisions. • More research is required to determine which volatile organic compounds (VOCs) attract different types of predators. Group feeding predators can be identified and chosen for biological control. Specialised predators for a particular prey are important to identify. • Whether the intraguild predation affects biological control favourably or unfavourably, and whether more high-throughput technology may be used to prove the IGP. • Different synthetic kairomone should be evaluated under common plantation crops either alone or in combination with natural kairomone as primary and secondary anti-predatory actions of various predators should be carried out.

References Ables JR (1975) Notes on the biology of the predacious pentatomid Euthyrhynchus floridanus (L.). J Georgia Entomol Soc 10:353–356 Alain D, Messika R, Frédéric A et al (2013) Altruism during predation in an assassin bug. Naturwissenschaften 100:913–922 Alaux C, Saurabh S, Linda H et al (2009) Honey bee aggression supports a link between gene regulation and behavioral evolution. Proc Natl Acad Sci USA 106:15400–15405 Al-Deeb MA, Wilde GE, Higgins RA (2001) No effect of bacillus thuringiensis corn and bacillus thuringiensis on the predator Orius insidiosus (Hemiptera: Anthocoridae). Environ Entomol 30:625–629 Alsabte A, Ahmed QH, Kayahan A et al (2022) Effects of volatile organic compounds (VOCs) emitted by citrus infested with Aonidiella aurantii on the predator Rhyzobius lophanthae attraction. Phytoparasitica 50:1–9 Ambrose DP (1987) Biological, behavioral and morphological tools in the biosystematics of Reduviidae (Insecta—Heteroptera—Reduviidae). Proc Anim Sci 96(5):499–508 Ambrose DP, Livingstone D (1985) Development of Coranus vitellinus distant (Hemiptera: Reduviidae: Harpactorinae). J Soil Biol Ecol 5:65–71 Ambrose DP, Sebasti RXJ, Nagarajan K et al (2009) Biology, behavior and functional response of Sphedanolestes variabilis Distant (Insecta: Hemiptera: Reduviidae: Harpactorinae), a potential predator of lepidopteran pests. Entomol Croat 13(2):33–44 Arnold JW (1971) Feeding behavior of a predaceous bug (Hemiptera: Nabidae). Can J Zool 49(1): 131–132 Ayelo PM, Pirk CW, Yusuf AA, Chailleux A, Mohamed SA, Deletre E (2021) Exploring the kairomone-based foraging behaviour of natural enemies to enhance biological control: a review. Front Ecol Evol 9:641974 Bahar MH et al (2010) Comparing the foraging behaviour of green lacewing larvae (‘ Mallada signata’) on transgenic Bt cotton and conventional cotton using ‘Helicoverpa armigera’ eggs or neonate larvae and cotton aphids (‘Aphis gossypii’) as prey. In global entomology in New

348

10 Predation Ethology of Various Orders

Zealand: New Zealand Entomological Society’s annual meeting 2010. The Entomological Society of New Zealand Bai YY, Jiang MX, Cheng JA et al (2006) Effects of Cry1Ab toxin on Propylea japonica (Thunberg) (Coleoptera: Coccinellidae) through its prey, Nilaparvata lugens Sta°l (Homoptera: Delphacidae), feeding on transgenic Bt rice. Environ Entomol 35:1130–1136. Transgenic Res (2008) 17:345–354 353 123 Balog A, Mehrparvar M, Weisser WW (2013) Polyphagous predatory rove beetles (Coleoptera: Staphylinidae) induce winged morphs in the pea aphid Acyrthosiphon pisum (Hemiptera: Aphididae). Eur J Entomol 110(1):153–157 Barrion A, Litsinger JA (1985) Proreus simulans (Dermaptera: Chelisochidae), a predator of rice leaffolder (LF) and skipper larvae. Int Rice Res Newsletter 10(1):25 Bottrell DG, Barbosa P, Gould F (1998) Manipulating natural enemies by plant variety selection and modification: a realistic strategy? Annu Rev Entomol 43:347–367 Brandão CRF, Diniz JLM, Tomotake EM (1991) Thaumatomyrmex strips millipedes for prey: a novel predatory behaviour in ants, and the first case of sympatry in the genus (hymenoptera: Formicidae). Insect Soc 38(4):335–344 Brandt M, Mahsberg D (2002) Bugs with a backpack: the function of nymphal camouflage in the west African assassin bugs Paredocla and Acanthaspis spp. Anim Behav 63(2):277–284 Brown PMJ, Ingels B, Wheatley A et al (2015) Intraguild predation by Harmonia axyridis (Coleoptera: Coccinellidae) on native insects in Europe: molecular detection from field samples. Entomol Sci 18(1):130–133 Brunner JF, Burts EC (1975) Searching behavior and growth rates of Anthocoris nemoralis, a predator of the pear psylla, Psylla pyricola. Ann Entomol Soc Am 68:311–315 Bulgarini G, Badra Z, Leonardi S et al (2021) Predatory ability of generalist predators on eggs, young nymphs, and adults of the invasive Halyomorpha halys in southern Europe. BioControl 66(3):355–366 Cianferoni MPF, Dioli AFP (2019) Predation by nymphs of Picromerus bidens (Heteroptera Pentatomidae Asopinae) on caterpillars of Euphydryas aurinia provincialis (Lepidoptera Nymphalidae) in Italy. REDIA 102:89–94 Clausen CP (1962) Entomophagous insects. Hafner, New York Claver MA, Ambrose DP (2001) Impact of antennectomy, eye blinding and tibial comb coating on the predatory behaviour of Rhynocoris kumarii Ambrose and Livingstone (Het., Reduviidae) on Spodoptera litura Fabr.(Lep., Noctuidae). J Appl Entomol 125(9–10):519–525 Cloarec A (1990) Factors influencing the choice of predatory tactics in a water bug, Diploychus indicus Venk. & Rao (Heteropoda: Belostomatidae). Anim Behav 40(2):262–271. https://doi. org/10.1016/S0003-3472(05)80921-2 Cobben RH (1978) Evolutionary trends in Heteroptera. Part II. Mouthpart structures and feeding strategies. Mededelingen Landbouwhogeschool, Wageningen, p 407 Cottrell TE, Landolt PJ, Zhang QH et al (2014) A chemical lure for stink bugs (Hemiptera: Pentatomidae) is used as a kairomone by Astata occidentalis (hymenoptera: Sphecidae). Fla Entomol 97:233–237 Cowan DP (1991) In: Ross KG, Matthews RW (eds) The solitary and presocial Vespidae. Cornell University Press, Ithaca, NY, p 33 Csata E, Dussutour A (2019) Nutrient regulation in ants (hymenoptera: Formicidae): a review. Myrmecol News 29:111–124 Cui JJ, Xia JY (1999) Effects of transgenic Bt cotton on the population dynamics of natural enemies. Acta Gossypii Sinica 11(2):84–91 Culliney TW (2014) Notes on predatory behaviour in Rhinacloa forticornis (Hemiptera: Miridae). Curr Agri Res 2(1):01. https://doi.org/10.12944/CARJ.2.1.01 Culliney TW, Grace JK (2000) Prospects for the biological control of Bterranean termites (Isoptera: Rhinotermitidae), with special reference to Coptotermes formosanus. Bull Entomol Res 90:9– 21

References

349

De Clercq P (2000) Predaceous stinkbugs (Pentatomidae: Asopinae). In: Schaefer CW, Panizzi AR (eds) Heteroptera of economic importance. CRC, Boca Raton, FL, pp 737–889 Devee A, Arvaniti K, Perdikis D (2018) Intraguild predation among three aphidophagous predators. Bull Insectol 71(1):11–19 Dib H, Siegwart M, Delattre T et al (2020) Does combining Forficula auricularia L.(Dermaptera: Forficulidae) with Harmonia axyridis Pallas (Coleoptera: Coccinellidae) enhance predation of rosy apple aphid, Dysaphis plantaginea Passerini (Hemiptera: Aphididae)? Biol Control 151:104394 Dicke M (1999) Evolution of induced indirect defence of plants. In: Tollrian R, Harvell CD (eds) The ecology and evolution of inducible defences. Princeton University Press, Princeton, NJ, pp 62–88 Dixon AFG, Russel RJ (1972) The effectiveness of Anthocoris ncmorum and A. conjusus (Hemiptera: Anthocoridae) as predators of the sycamore aphid, Drcpanosiphum platanoides. II. Searching behavior and the incidence of predation in the field. Entomol Exp Appl 15:35–50 Dowd PF, Kok LT (1981) Predators of Rhinocyllus conicus (Coleoptera: Curculionidae) in Virginia. Environ Entomol 10(1):136–138 Drukker B, Bruin J, Sabelis MW (2000) Anthocorid predators learn to associate herbivore-induced volatiles with presence or absence pf prey. Entomol Exp Appl 25:260–265 Dutton A, Klein H, Romeis J et al (2002) Uptake of Bt toxin by herbivores feeding on transgenic maize and consequences for the predator Chrysoperla carnea. Ecol Entomol 27:441–447 Ellington J, southward M, Carrillo T (1997) association among cotton arthropods. Environ Entomol 26:1004–1008 Eisner HE, Eisner T, Hurst JJ (1963) Hydrogen cyanide and benzaldehyde produced by millipedes. Chem Ind:124–125 Freund RL (1998) Prey location and enemy avoidance behavior of three predatory bugs (Heteroptera: Nabidae, Reduviidae). University of South Dakota Formanowicz DR Jr (1987) Foraging tactics of Dytiscus verticalis larvae (Coleoptera: Dytiscidae): prey detection, reactive distance and predator size. J Kans Entomol Soc 60:92–99 Gagnon AÈ, Doyon J, Heimpel GE, Brodeur J (2011) Prey DNA detection success following digestion by intraguild predators: influence of prey and predator species. Mol Ecol Res 11 (6):1022–1032 Godfray HCJ (1994) Parasitoids: behavioral and evolutionary ecology. Princeton University Press, Princeton Gordon ER, Weirauch C (2016) Efficient capture of natural history data reveals prey conservatism of cryptic termite predators. Mol Phylogenet Evol l94:65–73 Guo JY, Wan FH, Dong L (2004) Survival and development of immature Chrysopa sinica and Propylaea japonica feeding on Bemisia tabaci propagated on transgenic Bt cotton. Chin J Biol Control 20(3):164 Haridass ET (1978) Biological and ethological studies on some south Indian reduviids (Hemiptera: Reduviidae). PhD thesis, University of Madras, India Haridass ET, Ananthakrishnan TN (1980) Models for the predatory behaviour of some reduviids from southern India (Insecta—Heteroptera—Reduviidae). Proc Anim Sci 89(4):387–402 Haridass ET, Ananthakrishnan TN (1981) A functional morphology of the salivary system in some reduviidae (Insecta-Heteroptera). Proc Indian Acad Sci (Anim Sci) 90:140–160 Haridass ET, Noble MM, Balu A (1988) Predatory behaviour of Rhinocoris marginatus Fabricius (Harpactorinae—Reduviidae—Heteroptera—Insecta). Proc Anim Sci 97(1):41–48 Holldobler B, Wilson EO (1990) The ants. Harvard University Press, Berlin Humphreys RK, Ruxton GD (2018) A review of thanatosis (death feigning) as an anti-predator behavior. Behav Ecol Sociobiol 72(2):1–16 Inoue H (1983) Nymphal cannibalism in relation to oviposition behavior of adults in the assassin bug, Agriosphodrus dohrni Signoret. Res Popul Ecol 25:189–197

350

10 Predation Ethology of Various Orders

Isenhour DJ, Yeargan KV (1981) Interactive behavior of Orius insidiosus [hem.: Anthocoridae] and Sericothrips variabilis [Thys.: Thripidae]: predator searching strategies and prey escape tactics. Entomophaga 26(2):213–219 Iwata K (1976) Evolution of instinct: comparative behavior of hymenoptera. Amerind Press, New Delhi Jackson RR, Salm K, Nelson XJ (2010) Specialized prey selection behavior of two east African assassin bugs, Scipinnia repax and Nagusta sp. that prey on social jumping spiders. J Insect Sci 10:82 Jana N, Michal S, Jana O et al (2021) Forficula auricularia (Dermaptera) in orchards: monitoring seasonal activity, the effect of pesticides, and the perception of European fruit growers on its role as a predator or pest. Pest Manag Sci 77(4):1694–1704 Jaume-Schinkel S, Machado BL, Breno GM et al (2022) In natura sit-and-wait behaviour and predation success of a Neotropical dance fly (Diptera: Hybotidae). J Nat Hist 56(1–4):1–14 Jervis MA, Kidd NAC (1986) Host-feeding strategies in hymenopteran parasitoids. Biol Rev 61: 395 Kehat M (1968) The feeding behaviour of Pharoscymnus numidicus (Coccinellidae), predator of the date palm scale Parlatoria blanchardi. Entomol Exp Appl 11(1):30–42 Konvička M, Hula V, Fric Z (2005) Picromerus bidens (Heteroptera: Pentatomidae) as predator of the Checkerspot Euphydryas aurinia (Lepidoptera: Nymphalidae). Entomol Fenn 16:233–236 Kopf A, Rank NE, Roininen H et al (1997) Defensive larval secretions of leaf beetles attract a specialist predator parasyrphus nigritarsis. Ecol Entomol 22:176–183. https://doi.org/10.1046/j. 1365-2311.1997.t01-1-00061.x Krost C (2008) Chemical communication. In: Jorgensen SE, Fath BD (eds) Encyclopedia of ecology. Elsevier, Oxford, pp 557–575 Kumar SM, Sahayaraj K (2012) Gross morphology and histology of head and salivary apparatus of the predatory bug, Rhynocoris marginatus. J Insect Sci 12(1):19 Le Rü B, Makaya MJP (2001) Prey habitat location by the cassava mealybug predator Exochomus flaviventris: olfactory responses to odor of plant, mealybug, plant–mealybug complex, and plant–mealybug–natural enemy complex. J Insect Behav 14(5):557–572 Leius K (1961) Influence of various foods on fecundity and longevity of adults of Scambius buolianae (Htg.) (hymenoptera: Ichneumonidae). Can Entomol 93:1079–1084 Lenin EA, Anbarasi R, Nagarajan K et al (2011) Kairomonal ecology of assassin bug, Rhynocoris kumarii Ambrose and Livingstone (Hemiptera: Reduviidae) and chosen prey species. Insect Pest management, a current scenario. Entomology Research Unit, St. Xavier’s College, Palayamkottai, pp 326–332 Li ZQ, Zhang S, Luo JY et al (2015) Odorant-binding proteins display high affinities for behavioral attractants and repellents in the natural predator Chrysopa pallens. Comp Biochem Phys A 185:51–57 Lim UT, Ben-Yakir D (2020) Visual sensory systems of predatory and parasitic arthropods. Biocontrol Sci Tech 30(7):728–739 Lorenz KZ (1966) On Aggression. Harcourt, Brace & World, New York Lovei GL, Sunderland KD (1996) Ecology and behavior of ground beetles (Coleoptera: Carabidae). Annu Rev Entomol 41(1):231–256 Ma Y, Zhang M, Luo J et al (2022) Transgenic Bt cotton expressing Cry1Ac/1Ab does not have detrimental effects on the predator Arma chinensis through its prey Helicoverpa armigera. J Pestic Sci 96:317–326 Madeira F, di Lascio A, Costantini ML et al (2019) Intercrop movement of heteropteran predators between alfalfa and maize examined by stable isotope analysis. J Pestic Sci 92(2):757–767 Maeda T, Kishimoto H, Wright LC et al (2015) Mixture of synthetic herbivore-induced plant volatiles attracts more Stethorus punctum picipes (Casey)(Coleoptera: Coccinellidae) than a single volatile. J Insect Behav 28(2):126–137 Maginnis TL (2008) Autotomy in a stick insect (Insecta: Phasmida): predation versus molting. Fla Entomol 91(1):126–127

References

351

Mallinger RE, Hogg DB, Gratton C (2011) Methyl salicylate attracts natural enemies and reduces populations of soybean aphids (Hemiptera: Aphididae) in soybean agroecosystems. J Econ Entomol 104:115–124 Marquereau L, Cottineau JS, Fontaine O et al (2022) Life history parameters and predation capacities of Nesidiocoris volucer: a new biological control agent for tomato crop. Bull Entomol Res 112:584–592. https://doi.org/10.1017/S0007485321001164 Marshall DC, Hill KB (2009) Versatile aggressive mimicry of cicadas by an Australian predatory katydid. PloS One 4(1):e4185 Martínez LD, de los Angeles MRM, Bueno VHP et al (2022) Predation behaviour and prey preference of two neotropical mirids against two key lepidopteran pests in tomato. Int J Trop Insect Sci 42(1):815–825 Masuko K (2009) Studies on the predatory biology of oriental dacetine ants (hymenoptera: Formicidae) II. Novel prey specialization in Pyramica benten. J Nat Hist 43(13–14):825–841 Messelink GJ, Janssen A (2014) Increased control of thrips and aphids in greenhouses with two species of generalist predatory bugs involved in intraguild predation. Biol Control 79:1–7 Miller NCE (1956) The biology of the Heterop tera, 1st edn. Leornard Hill, London, p 162 Moayeri HRS, Ashouri A, Poll L et al (2007) Olfactory response of a predatory mirid to herbivore induced plant volatiles: multiple herbivory vs. single herbivory. J Appl Entomol 131(5): 326–332 Montllor CB, Bernays EA, Cornelius ML (1991) Responses of two hymenopteran predators to surface chemistry of their prey: significance for an alkaloid-sequestering caterpillar. J Chem Ecol 17(2):391–399 Morais HC (1994) Coordinated group ambush: a new predatory behavior in Azteca ants (Dolichoderinae). Insect Soc 41(3):339–342 Morrison WR, Mathews CR, Leskey TC (2016) Frequency, efficiency, and physical characteristics of predation by generalist predators of brown marmorated stink bug (Hemiptera: Pentatomidae) eggs. Biol Control 97:120–130 Muhammad Luqman HA, Dzulhelmi MN, Idris AB (2018) The potential natural predators of Elaeidobius kamerunicus Faust, 1878 (Coleoptera: Curculionidae) in Malaysia. Serangga 22:240–252 Napolitano JF (1998) Predatory behavior of a pit-making antlion, Myrmeleon mobilis (Neuroptera: Myrmeleontidae). Fla Entomol 81(4):562–566 Naranjo-Guevara N, Peñaflor MFG, Cabezas-Guerrero MF et al (2017) Nocturnal herbivoreinduced plant volatiles attract the generalist predatory earwig Doru luteipes Scudder. Sci Nat 104(9):1–11 Ninkovic V, Al Abassi S, Pettersson J (2001) The influence of aphid-induced plant volatiles on ladybird beetle searching behavior. Biol Control 21(2):191–195 Nyiira ZM (1970) The biology and behavior of Rhinocoris albopunctatus (Hemiptera: Reduviidae). Ann Entomol Soc Am 63(5):1224–1227 O’Neill KM (2001) Solitary wasps: behavior and natural history. Comstock Publishing Associates, Ithaca, NY, p 26 Obrist LB, Klein H, Dutton A et al (2005) Effects of Bt maize on Frankliniella tenuicornis and exposure of thrips predators to prey-mediated Bt toxin. Entomol Exp Appl 115:409–416 Obrist LB, Dutton A, Romeis J et al (2006) Biological activity of Cry1Ab toxin expressed by Bt maize following ingestion by herbivores arthropods and exposure of the predator Chrysoperla carnea. BioControl 51:31–48 Oliveira CM, Silva TCS, Torres JB et al (2022) Cannibalism and intraguild predation between endemic and introduced coccidophagous lady beetles (Coleoptera: Coccinellidae). Austral Entomol 61:225. https://doi.org/10.1111/aen.12589 Omkar O, Mishra G (2016) Syrphid flies (the hovering agents). In: Omkar O (ed) Ecofriendly pest management for food security. Academic Press, San Diego, pp 259–279 Orpet RJ, Crowder DW, Jones VP (2019) Woolly apple aphid generalist predator feeding behavior assessed through video observation in an apple orchard. J Insect Behav 32(2):153–163

352

10 Predation Ethology of Various Orders

Parker AH (1965) The predatory behaviour and life history of Pisilus tipuliformis Fabricius (Hemiptera: Reduviidae). Entomol Exp Appl 8(1):1–12 Pierre EM, Idris AH (2013) Studies on the predatory activities of Oecophylla smaragdina (Hymenoptera: Formicidae) on Pteroma pendula (Lepidoptera: Psychidae) in oil palm plantations in Teluk Intan, Perak (Malaysia). Asian Myrmecol 5(1):163–176 Pritchard G, Scholefield P (1978) Observations on the food, feeding behaviour, and associated sense organs of Grylloblatta campodeiformis (Grylloblattodea). Can Entomol 110(2):205–212 Rajan SJ, Suneetha N, Sathish R (2017) Biology and predatory behaviour of an assassin bug, Sycanus collaris (Fabricius) on rice meal moth, Corcyra cephalonica (Stainston) and leaf armyworm, Spodoptera litura (Fabricius). Agric Update 12(5):1181–1186 Ramamurthi BN, Solayappan AR (1980) Dermapteran predators in the biological regulation of sugarcane borers in India. Curr Sci 49(2):72–73 Rani PU, Havukkala I (1993) Predatory and mating behaviour of Eanocthecona furcellata (Wolff.) (Heteroptera: Pentatomidae) a promising natural enemy of Lepidopterous larvae. J Biol Control 7(1):9–11 Ranjbar F, Michaud JP, Jalali MA et al (2020) Intraguild predation between two lady beetle predators is more sensitive to density than species of extraguild prey. BioControl 65(6):713–725 Reddy GVP, Holopainen JK, Guerrero A (2002) Olfactory responses of Plutella xylostella natural enemies to host pheromone, larval frass, and green leaf cabbage volatiles. J Chem Ecol 28:131– 143. https://doi.org/10.1023/a:1013519003944 Rhodes BC, Blair CP, Takahashi MK et al (2012) The role of olfactory cues in the sequential radiation of a gall boring beetle, Mordellistena convicta. Ecol Entomol 37:500–507 Rodriguez-Saona C, Kaplan I, Braasch J et al (2011) Field responses of predaceous arthropods to methyl salicylate: a meta-analysis and case study in cranberries. Biol Control 59:294–303 Rosenheim JA, Wilhoit LR, Armer CA (1993) Influence of intraguild predation among generalist insect predators on the suppression of an herbivore population. Oecologia 96(3):439–449 Sahayaraj K (1991) Bioecology, ecophysiology and ethology of chosen predatory Hemipterans and their potential in biological control (Insecta: Heteroptera: Reduviidae). PhD Thesis, Madurai Kamaraj University, Madurai Sahayaraj K (2008) Approaching and rostrum protrusion behaviours of Rhynocoris marginatus on three prey chemical cues. Bull Insectol 61:233–237 Sahayaraj K, Delma JCR (2004) Chemically mediated behaviour of the predator Rhynocoris marginatus (fab.) (Heteroptera: Reduviidae) searching insect prey. Belg J Entomol 6:75–81 Sahayaraj K, Fernandez SM (2021) The predation behavior and the prey size preferences of Antilochus coquebertii (Pyrrhocoridae) against Dysdercus koenigii (Pyrrhocoridae). Int J Trop Insect Sci 41(2):1763–1769 Scharf I, Ovadia O (2006) Factors influencing site abandonment and site selection in a sit-and-wait predator: a review of pit-building antlion larvae. J Insect Behav 19(2):197–218. https://doi.org/ 10.1007/s10905-006-9017-4 Segev N, Gavish-Regev E, Berger-Tal O (2020) Sit-and-wait prey: first field observations of scorpions preying on antlions (Neuroptera). Isr J Ecol Evol 66(1–2):57–62 Siddique AB, Chapman RB (1987) Effect of prey type and quantity on the reproduction, development, and survival of Pacific damsel bug, Nabis kinbergii Reuter (Hemiptera: Nabidae). NZ J Zool 14(3):343–349 Spradbery JP (1973) Wasps: an account of the biology and natural history of solitary and social wasps. Sidwick & Jackson, London Srikumar KK, Bhat PS, Raviprasad TN et al (2014) Biology and behavior of six species of Reduviids (Hemiptera: Reduviidae: Harpactorinae) in a Cashew Ecosystem 1. J Agric Urban Entomol 30(1):65–81 Tobing MC, Kuswardani RA (2018) The potential of Myopopone castanea (hymenoptera: Formicidae) as a predator for Oryctes rhinoceros Linn. Larvae (Coleoptera: Scarabaeidae). In. J Phys Conf Ser 1116(5):052074

References

353

Tomé R, Dias MP, Chumbinho AC, Bloise C (2011) Influence of perch height and vegetation structure on the foraging behavior of little owls Athene noctua: how to achieve the same success in two distinct habitats. Ardea 99(1):17–26. https://doi.org/10.5253/078.099.0103 Torres JB, Ruberson JR (2005) Canopy- and ground-dwelling predator arthropods in Bt and non-Bt cotton fields: patterns and mechanisms. Environ Entomol 34:1242–1256 Torres JB, Ruberson JR (2006) Interactions of Bt-cotton and the omnivorous big-eyed bug Geocoris punctipes (say), a key predator in cotton fields. Biol Control 39:47–57 Torres JB, Ruberson JR (2008) Interactions of bacillus thuringiensis Cry1Ac toxin in genetically engineered cotton with predatory heteropterans. Transgenic Res 17(3):345–354 Torres JB, Ruberson JR, Adang MJ (2006) Expression of Bacillus thuringiensis Cry1Ac protein in cotton plants, acquisition by pests and predators: a tritrophic analysis. Agric For Entomol 8:191–202 Tuma J, Eggleton P, Fayle TM (2020) Ant-termite interactions: an important but under-explored ecological linkage. Biol Rev Camb Philos Soc 95(3):555–572 Vaello T, Casas JL, Pineda A et al (2017) Olfactory response of the predatory bug Orius laevigatus (Hemiptera: Anthocoridae) to the aggregation pheromone of its prey, Frankliniella occidentalis (Thysanoptera: Thripidae). Environ Entomol 46:1115–1119. https://doi.org/10.1093/ee/nvx141 van Loon JJ, de Vos EW, Dicke M (2000) Orientation behaviour of the predatory hemipteran Perillus bioculatus to plant and prey odours. Entomol Exp Appl 96(1):51–58 Vasconcelos SD, Williams T, Hails RS et al (1996) Prey selection and baculovirus dissemination by carabid predators of Lepidoptera. Ecol Entomol 21(1):98–104 Waloff N (1974) Biology and behaviour of some species of Dryinidae (hymenoptera). J Entomol 49:97–109 Watanabe H, Yano E, Higashida K et al (2016) An attractant of the aphidophagous gall midge Aphidoletes aphidimyza from honeydew of Aphis gossypii. J Chem Ecol 42:149–155. https:// doi.org/10.1007/s10886-016-0666-2 Wedmann S, Kment P, Campos LA et al (2021) Bizarre morphology in extinct Eocene bugs (Heteroptera: Pentatomidae). R Soc Open Sci 8(12):211466 Wei J, Kang L, Wei J et al (2011) Roles of (Z)-3-hexenol in plant-insect interactions. Plant Signal Behav 6:369–371 Weiss MJ, McDonald G (1998) European earwig, Forficula auricularia L.. (Dermaptera: Forficulidae), as a predator of the redlegged earth mite, Halotydeus destructor (Tucker) (Acarina: Penthaleidae). Aust J Entomol 37(2):183–185 Weissbecker B, Van Loon JJ, Dicke M (1999) Electroantennogram responses of a predator, Perillus bioculatus, and its prey, Leptinotarsa decemlineata, to plant volatiles. J Chem Ecol 25 (10):2313–2325 Wheater CP (1989) Prey detection by some predatory Coleoptera (Carabidae and Staphylinidae). J Zool 218(2):171–185 Wheeler AG Jr, Stinner BR, Henry TJ (1975) Biology and nymphal stages of Deraeocoris nebulosus (Hemiptera: Miridae), a predator of arthropod pests on ornamentals. Ann Entomol Soc Am 68(6):1063–1068 Yasuda T (1997) Chemical cues from Spodoptera litura larvae elicit prey-locating behavior by the predatory stink bug, Eocanthecona furcellata. Entomol Exp Appl 82:349–354 Yasuda T, Wakamura S (1996) Behavioral responses in prey location of the predatory stink bug, Eocanthecona furcellata, to chemical cues in the larvae of Spodoptera litura. Entomol Exp Appl 81:91–96 Zarei M, Madadi H, Zamani AA et al (2020) Intraguild predation between Chrysoperla carnea (Neuroptera: Chrysopidae) and Hippodamia variegata (Coleoptera: Coccinellidae) at various extraguild prey densities and arena complexities. Insects 11(5):288 Zhang G (2012) Taxonomy, Phylogenetics, comparative morphology and evolution of assassin bugs (Hemiptera: Reduviidae) with emphasis on Harpactorinae and Peiratinae. PhD thesis, University of California, Riverside, USA. p 436

354

10 Predation Ethology of Various Orders

Zhang GF, Wan FH, Liu WX, Guo JY (2006a) Early instar response to plant-delivered Bt-toxin in a herbivore (Spodoptera litura) and a predator (Propylea japonica). Crop Prot 25:527–533 Zhang GF, Wan FH, Lövei GL et al (2006b) Transmission of Bt toxin to the predator Propylaea japonica (Coleoptera: Coccinellidae) through its aphid prey feeding on transgenic Bt cotton. Environ Entomol 35(1):143–150 Zhu J, Obrycki JJ, Ochieng SA et al (2005) Attraction of two lacewing species to volatiles produced by host plants and aphid prey. Naturwissenschaften 92(6):277–281 Zhu P, Lu Z, Heong K et al (2014) Selection of nectar plants for use in ecological engineering to promote biological control of Rice pests by the predatory bug, Cyrtorhinus lividipennis, (Heteroptera: Miridae). PloS One 9(9):e108669. https://doi.org/10.1371/journal.pone.0108669 Zwahlen C, Nentwig W, Bigler R et al (2000) Tritrophic interactions of transgenic Bacillus thuringiensis corn, Anaphothrips obscurus (Thysanoptera: Thripidae), and the predator Orius majusculus (Heteroptera: Anthocoridae). Environ Entomol 29:846–850

Mating Behaviour and Reproductive Biology of Insect Predators

11

Contents 11.1 11.2 11.3 11.4

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Multiple Mating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dictyoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hemiptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.4.1 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.4.2 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.4.3 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.5 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.6 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.7 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.8 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.9 Chemical Ecology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.10 Future Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

11.1

355 356 356 358 358 362 364 365 366 368 370 372 373 373

Introduction

In sexually reproducing animals, fertilisation and insemination (the exchange of male sperm for female sperm) constitute the process of reproduction (the fusion of a sperm and an egg to create a diploid zygote). In many insect species, including predators and parasitoids, fertilisation and insemination take place at different times ranging from a few minutes to several years (as in many social insects). The behavioural actions leading up to insemination that guarantee successful sperm transfer by the male and uptake by the female are referred to as “mating behaviour”. In many species, post-copulatory male behaviours that have evolved in response to sperm competition are also included in this definition. We examine mating behaviour primarily in accordance with these divisions: pair formation, courtship, copulation, insemination, and the actions that immediately follow insemination, including temporary pair maintenance (Hardy et al. 2007). # The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_11

355

356

11 Mating Behaviour and Reproductive Biology of Insect Predators

The length of the mating period seems to affect egg viability and fertility. In many insects, males who have been mating for a long time produce huge ejaculates full of nutrients that the females require for reproduction and somatic upkeep. Therefore, females with larger ejaculates use these nutrients to develop more eggs, which direct natural selection in favour of mates who also produce large ejaculates.

11.2

Multiple Mating

Males who repeatedly mate usually do so to improve their fitness by raising the paternity share they receive from their multiple matings with different females. For females, on the other hand, such multiple mating assures fertilisation as well as a wider selection of genes. The origins and effects of polyandry have been thoroughly researched in recent years. The main consequences of many matings, other from energy and time loss, are 1. 2. 3. 4.

increased predation rates, high physical injury, increased likelihood of parasite/pathogen infection, and, decreased longevity.

Females who mate repeatedly have shorter lifespans because mating and reproduction require different amounts of energy, a process known as resource partitioning. The advantages of multiple mating might be both direct and indirect (genetic). Direct material advantages include: 1. improved egg laying brought on by an increase in accessory gland secretions (proteins), 2. sperm replacement, and, 3. defence against infanticidal males or avoiding rejection costs associated with harassing males (Hosken and Stockley 2003). The likelihood of genetic compatibility is raised, offspring viability and fitness are increased, sperm fertilisation efficiency is increased, and genetic diversity is increased. These are the indirect advantages.

11.3

Dictyoptera

There were three different courtship behaviours seen in Chinese praying mantis: 1. the male approached when the female was looking at him, 2. the male approached the female from behind, and, 3. the female approached the male.

11.3

Dictyoptera

357

The courtship behaviour of the male consisted of 11 distinct components for pattern 1 (the one that was most frequently observed), and they all arose at the same time 10 days following the last moult. The time to intromission decreased with age in accordance with changes in sexual behaviour in males. Males who were sexually mature experienced this drop in time to intromission more quickly, suggesting an experiential component. Cannibalism in mantises happened in conjunction with sexual behaviour when females were starving, rather than being causally related to it. In just one out of 69 studies, the female decapitated the male before the introduction. The male and female’s courtship behaviours may reduce intraspecific aggression during courtship, which lowers the risk of cannibalism. A frequent trait of most animals, but not particularly noticeable in mantids, is mating calls. Hierodula patellifera (Dictyoptera: Mantidae) virgin mantis females display a distinctive calling posture. The female folds the ventral abdomen, stretching it away from the wings and exposing its dorsal side when holding the body below a branch or leaf. The pumping motions go along with the curling. This calling position is often adopted by females 14 days following adult moult, and it is correlated with their nutritional stage. Females display the position every day until they mate once it has been initiated. The behaviour entirely disappears after mating. Males are drawn to unmarried women who are calling, but they are not drawn to mated female. The traits of the males’ posture and their responding behaviour suggest that sex pheromones are released during this female calling (Perez 2005). Gemeno and Claramunt (2006) studied and characterised the Mantis religiosa courtship behaviour, which they detailed as follows: The male quickly became entirely motionless and turned his head in the direction of the female after being placed in the cage. Following ocular fixation, the male began a very leisurely, nearly straight-line approach towards the female, devoid of any overt displays of courtship. We began 78 runs, of which 38 were rejected. The majority of these runs were in the NP treatment, and the majority of these runs were eliminated because the females moved their heads in the direction of the males or left before the second betweenprey-presentation time was complete. Females from failed runs were seen to slightly extend their front legs when they became fascinated on the males. The prey was lowered in front of the females, who remained still but instantly rotated their heads towards its direction and followed its swing with similar motions of their bodies. The females occasionally stretched their raptorial legs in the direction of the prey, but only in the P treatment were they permitted to strike and take the prey. Males occasionally moved “rapidly” (in the sense of a very sluggish approach) towards females when they moved forward or laterally in reaction to the motions of the prey, but only for the brief period of time that the female was moving. The replication would stop if the female left her platform or fell to the ground. Otherwise, the male would move forward until she stopped, at which point he would stop. Males kept their eyes fixed on the females at all times and never swung their heads to follow the movement of the prey. The prey was consumed by all P females in less than 10 min, and they always followed that up by using their mouthparts to clean their forelegs. They then reverted to their ambush position, folding their forelegs close to their bodies. Males who came very close to females eventually sprang into the air and

358

11 Mating Behaviour and Reproductive Biology of Insect Predators

landed on them. Males spun 180 degrees in the air and landed on the female inverted in 77% of flying leaps. After landing, these men quickly made a 180-degree turn. In comparison to NP, there were noticeably more matings in P. Females only stretched their forelegs and occasionally let go of the prey they were eating on when males made physical contact with them.

11.4

Hemiptera

11.4.1 Reduviidae In a lab, Sphedanolestes variabilis was observed mating as follows: Arousal, approach, riding over, elongation of the genitalia, copulation, ejection of the female’s spermatophore capsule, and cleaning after mating (Fig. 11.1). Sphedanolestes variabilis mating behaviour, which included the distinctive ridingover occurrence, demonstrated its harpactorine nature (Ambrose et al. 2009). Coranus spiniscutis engaged in polygamous and polyandrous mating. The following mating act sequence was seen in both Coranus spiniscutis species (Claver and Reegan 2010). As reported in other reduviids like Ectomocoris tibialis and Acanthespis pedestris (Ambrose and Livingstone 1978), Rhynocoris kumarii (Ambrose and Livingstone 1987a), Coranus vitellinus, and others, freshly collected Coranus spiniscutis were immediately aroused by the sight of opposite sex in 51.9 s (laboratory reared 1.4 min) (Ambrose and Livingstone 1987b). With reference to their ecomorphological characteristics, the diversity of eggs and oviposition

Fig. 11.1 Cydnocoris gilvus male (a) female (b) male riding over the female (c) and copulation act (d)

11.4

Hemiptera

359

behaviour of 14 species of reduviids belonging to five subfamilies— Acanthaspidinae, Ectrichodiinae, Harpactorinae, Piratinae, and Stenopodinae—is explained (Vennison and Ambrose 1990). Ambrose and Vennison (1990) also describe and provide illustrations for the spermatophore capsules of 21 species of reduviids that belong to four subfamilies: Acanthaspidinae (5 species), Ectrichodiinae (1 species), Harpactorinae (11 species), and Piratinae (4 species). Approach It was also noted in Sphedanolestes sp. and Coranus sp. that the aroused males approached by chasing the females in 8.2 and 14.4 min for freshly collected and laboratory-reared Coranus spiniscutis, respectively, with expanded antennae (Kumar 1993). The still female indicated her desire by extending her antennae. The females occasionally managed to flee an approaching male (Claver and Reegan 2010). After the antennal fencing, as described in several other harpactorine reduviids, the male positioned his legs on the female (Ambrose 1999). Nuptial embrace with riding over: The males rode over the ladies while extending their rostrums, pressing her head region for a short period of time before pressing the pterothorax region with the tip of their rostrals. Prior to copulation, the male moved the female slightly to one side while rostral pinning. For Coranus spiniscutis that had just been collected and those that had been raised in a lab, the riding over times were 17.1 and 27.5 min, respectively. Extension of the genitalia and successful connection: The male made several attempts to establish a connection before succeeding. This act lasted for 0.5 s in newly collected Coranus spiniscutis and 1.6 min in laboratory-raised Coranus spiniscutis, respectively (Claver and Reegan 2010). After successful attachment, the male’s distinctive prothorax rostral pinning was noticed to have relaxed. According to Vennison (1988), Coranus spiniscutis mated in the dorsoventral posture, as seen in Sphedanolestes reclinatus, Coranus soosaii, and a number of other harpactorine reduviids (Ambrose 1999). Copulation When there was a disruption during copulation, Coranus spiniscutis male displayed prothorax rostral pinning once more. The male was seen grooming the female genitalia with its hind legs during copulation and then became down just before the termination of copulation, with their tibiae brushing against one other or the ground (Ambrose 1999). For field-collected and laboratory-raised Coranus spiniscutis, the lengths of copulation were 88 min and 106 min, respectively. At the conclusion of copulation and the subsequent separation of mating partners, both sexes were shown to droop their antennae. Both the male and the female left the area of the coupling after separation. Both sex partners engaged in the post-copulatory behaviours of Coranus spiniscutis, including brushing of the genitalia, grooming of the antennae, cleaning of the legs, and wing beating. In the case of freshly collected and laboratory-reared Coranus spiniscutis, respectively, post copulatory activities lasted for 5.0 and 6.1 min. Coranus spiniscutis did not exhibit post-copulatory cannibalism of female over male, as described for some Harpactorinae and Peiratinae reduviids (Ambrose 1999). Freshly obtained C. spiniscutis expelled the spermatophore capsule following termination of copulation in 14.2 min, whereas

360

11 Mating Behaviour and Reproductive Biology of Insect Predators

laboratory-reared specimens did so in 12.5 min, demonstrating the successful completion of copulation (Ambrose 1999; Claver and Reegan 2010). Following emergence, the first successful copulation occurs at least 20 days later. After each successful copulation, 45 min later, the spermatophore capsule is released. The capsule has a small circular body, an extended collar (2.35 mm long), and a rectangle shape (1 mm long by 0.39 mm wide) (length 0.51 mm and width 0.45 mm). At the moment of ejection, the capsule is translucent white and gradually turns pale yellow. There has only ever been one instance of successful copulation in a female adult. However, making the assumption that monogamy is the norm for this species would be premature. It has been discovered that neither this species nor Acanthaspis pedestris and Acanthaspis siva exhibit seasonal fluctuations in the intensity of mating. There were only 9 insect mating partners seen during 20 observations. In a lab setting, the sequential acts of mating behaviour—arousal, approach, nuptial clasp, including riding over the copula—were shown. After a brief interval of up to 2 min, the insects move about the arena. The male stops moving and adjusts his antennae so they are facing the female. The sexual arousal appears to be significantly influenced by sight. The antennae’s outward thrust, the tibia’s juxtaposition, and the erection of the tibia show arousal, which causes the animal to assume a particular pouncing posture that is distinct from arousal during predation. Sex-deprived males have been seen to rise to mate almost instantly when they approach the female; the female remains stationary during this process. The male approaches the woman and carefully mounts or jumps on her. In contrast, females respond by making comparable motions like rostral stridulation and antennal extension. Additionally, they groom their legs and antennae while remaining still in order to surrender to the approaching males. The guy rotates 360 degrees once he is atop the female. While mounted on the female, this action was unusual; instead, tarsi and antennae were typically used to touch the female. In relation to the female, the man positions himself dorsolaterally, either on her right or left side. With the male’s three legs holding her dorsally and ventrally, the female is immobile in this position. Sometimes the male will stretch the rostrum and place it at a roughly 90-degree angle on the female’s head. With his rear legs, the male pushes the female’s abdomen’s tip towards him. In order to introduce the aedeagus, the male’s parameres are released and help immobilise the female’s genitalia. Throughout the copulation, the female stayed still. Rhynocori marginatus rides over the female before extending its genitalia and making contact. The male held the female by the abdomen with his middle and hind legs after copulation and then reverted to his former position above her. The female can now move around while the male is riding on her. The male departs the female after descending from her on top. Acts performed after copulation, such as wiping the legs and brushing the genitalia. In Sycanus galbanus, the order of the mating behaviours was arousal, approach, riding over, and copulation (Nitin et al. 2017). Figure 11.1 shows a mating Cydnocoris gilvus. Approach and Arousal When mating partners saw someone of the opposing sex, they became excited. Reduviid males that had just emerged were quickly stimulated

11.4

Hemiptera

361

Fig. 11.2 Mating behaviour of Epidaus bicolor (a), Euagoras plagiatus (b), and Panthus bimaculatus (c and d)

upon seeing females of their own species. Males approached ladies while holding out their rostrum and antennae. Once the males placed their legs over the ladies and touched them with their antennae, the approach reaction was complete.

Riding Over In the riding-over dorsoventral position, the male held the female in place with his forelegs and placed his labial tip against the female’s prothorax. This action took close to 15 to 20 min. Copulation The mating couple stayed stationary throughout copulation as males stretched their penises and formed contact with the females towards the conclusion of riding over. Copulation lasted between 25 and 30 min. Both males and females droop their antennae at the conclusion of copulation, which is followed by the separation of the mating partners. The ejection of the spermatophore capsule by females following the cessation of copulation served as proof that copulation was successfully completed. In this species, post-copulatory cannibalism of the males by the females has not been seen. The mating behaviours of Epidaus bicolor (a), Euagoras plagiatus (b), and Panthus bimaculatus (c and d) are shown in Fig. 11.2. Figure 11.3 also shows the riding and copulation behaviours of Rihirbus trochantericus. Arousal, approach, riding over, and copulation are the most recent names for the six species of reduviids’ successive mating behaviours (Srikumar et al. 2014).

362

11 Mating Behaviour and Reproductive Biology of Insect Predators

Fig. 11.3 Rihirbus trochantericus riding over (a), and copulation (b) acts

Arousal and Approach The approach of the opposing sex excited the mating partners. Upon spotting females of their own species, virgin reduviid males were instantly aroused. According to time interval, the degree of arousal varied among species (P. bimaculatus, Epidaus bicolor, Endochus albomaculatus, Euagoras plagiatus, Sphedanolestes signatus, and Irantha armipes). Virgin males stretched their rostrum and antennae to approach females. Once the males placed their legs over the ladies and touched them with their antennae, the approach reaction was complete. Riding Over While still in the dorsoventral riding over position, the male applied pressure to the female’s prothorax region with the tip of his labial. The duration of the riding over behaviour was in the order P. bimaculatus, and it was substantially longer than the other mating activities. Epidaus bicolor, Euagoras plagiatus and Endochus albomaculatus. Sphedanolestes signatus and Irantha armipes. Copulation At the conclusion of riding over, males extended their genitalia and made contact with the females. During the course of copulation, they did not move. The length of copulation varied between species (P. bimaculatus, E. bicolor. Endochus albomaculatus, E. plagiatus, I. armipes, and Sphedanolestes signatus). Both males and females droop their antennae at the conclusion of copulation, which is followed by the separation of the mating partners. The ejection of the spermatophore capsule by females following the cessation of copulation served as proof that copulation was successfully completed. Only in E. bicolor was post-copulatory cannibalism of males by females observed (Srikumar et al. 2014). Table 11.1 provides a summary of the mating habits of reduviidae subfamilies such Acanthaspidinae (A), Harpactorinae (H), and Piratinae (P).

11.4.2 Pentatomidae Antennal touch between the two sexes was a common way for Eocanthecona furcellota (Heteroptera: Pentatomidae) to start a mating process. Slowly approaching

11.4

Hemiptera

363

Table 11.1 Mating behavioural demarcate the subfamilies Acanthaspidinae (A), Harpactorinae (H), and Piratinae (P) Tools Premating period in days Post mating period prior to oviposition Acts of mating-arousal Approach Pre-copulatory cannibalism of incompatible partner Pre-copulatory riding Duration of copulation Time for ejection of spermatophore capsule

A Short (9–15) Long (12–39

H Long (10–20) Shortest (5–10)

Fast (2 s–20 min) Similar to predatory (3–40 s) Rarely present

Slow (2 s–20 min) Lethargic (10 s–10 min) Absent

Absent Short (9 min–2 h)

Present (8 s–3 h) Long (20 min–8 h)

More (34 min–8 h)

Less (7 min–4 h)

P Shortest (6–8) Longest (31–40) Fastest (2–3 s) Aggressive (1–5 s) Present Absent Very short (17–20 min) More (40 min–8 h)

Source: Ambrose (1987)

from behind the female, the male stroked her posterior abdomen with his antennae before quickly mounting. For effective copulation, it was occasionally necessary to make many mounting attempts. The copulating pair was frequently surrounded by other bugs, mainly males, who occasionally tried to touch them with their extended proboscis. At 25 °C, the copulation lasted 6 to 8 h. Males occasionally took the position of a rider on another guy. Women were polygamous. Pre-copulatory and mating behaviour were inhibited under unfavourable circumstances such starvation or temperatures below 20 °C (Rani and Havukkala 1993). The predatory stink bug Podisus nigrispinus (Hemiptera: Pentatomidae; Asopinae) exhibits a variety of mating behaviours. These behaviours are detailed, together with the vibratory signals released on man-made and natural substrate during courtship and copulation (Laumann et al. 2013). Abdominal vibration and trembling in both sexes produce vibrational signals in the Podisus nigrispinus, which play a critical role in successful copulation. One species-specific female and two male songs are produced by abdomen vibration and released during reproductive behaviour in the phytophagous stink insect, Podisus nigrispinus. These songs do not display the calling function commonly found in phytophagous stink bugs. Furthermore, the Podisus nigrispinus generates tremulatory signals that are not gender- or species-specific. Tremulatory signals, which appear on a plant as a series of easily repeated pulses, are comparable to the Pentatominae stink bug’s calling songs. These signals might indicate the presence of a partner, but in other behavioural circumstances, they might serve a different purpose, like serving as an alert or an advertisement. Both methods generate vibratory signals that plants transmit as a low-pass filter, boosting the proportion of low-frequency components (Laumann et al. 2013). Mating behaviour of Antilochus coqueberti is also similar to other heteropterans like reduviids (Fig. 11.4).

364

11 Mating Behaviour and Reproductive Biology of Insect Predators

Fig. 11.4 Mating behaviour of Antilochus coqueberti

11.4.3 Anthocoridae Horton et al. (2001) found that as male length or relative size increased, the copulation duration considerably decreased (male length divided by female length). The length of copulation was unaffected by female length. Increases in male size, warmth, and copulation length all enhanced the likelihood of conception. Large males had a higher success rate than tiny males in fertilising females under chilly conditions and for very brief copulations. The findings are in line with the theory that larger male bodies help to decrease the physical challenges that this species’ male’s face while trying to force sperm through the aedeagus or achieve intromission. Flying is not a behaviour of mating partners during copulation. Instead, they follow the larger female’s lead and migrate together. Males begin mating 3–5 days after they are born, while females begin mating 2–5 days after reaching sexual maturity. Males typically mate at least ten times weekly. After a single mating, female big-eyed bugs continued to lay eggs for up to 30 days. Females of the tested species of Anthocoris minki (Hemiptera: Anthocoridae) oviposited 1–2 days following their initial mating. No matter how long it took between mating occasions, the successfully mated females consistently shown a distinct propensity to reject any further copulation. When the same man was successively inseminated three different females with 2 days between matings, there was no discernible change in the copulation times. The time spent in the first mating compared to the second and third matings was significantly different when the three mating episodes occurred on the same day or separated by 1 day. Additionally, a gradual decrease in copulation duration has been seen among the three mating events when the gap between copulations was 1 day (15.16, 11.31, and 8.35 min, respectively) (Abdalla et al. 2018). One of this genus’ most prevalent and geographically extensive species is the predatory insect Anthocoris nemoralis. It is endemic to England, where it has undergone substantial research in both agricultural and non-agricultural environments. The bug can be found all over western and central Europe, in Finland and the Netherlands to the north, in Turkey to the east, and in the Mediterranean area to the south, including areas of North Africa. The information on Anthocoris nemoralis’s mating inclination suggests the following (Horton et al. 2000), presuming that the male is the one who looks for partners.

11.5

Syrphids

365

That females from California were easier to locate than females from England, possibly due to differences in levels of female activity or pheromone production. That males from the UK population were either more active than males from California, or (if females do produce a pheromone) that males from England more readily perceived and responded to the sex pheromone. It should be emphasised that statistics on the length of intromission seem to support our findings regarding mating propensities. As a result, in the no-choice experiments, the time to intromission for UK males was weakly correlated with being shorter than for California males, further indicating that UK males were more adept at finding partners than California males. Time to intromission for females revealed a small trend to be shorter for females from California than females from the UK, indicating that males found females from California more quickly than did males from England (Horton et al. 2000). Insect mating times were different between the two populations in a second way. In the no-choice tests, the average copulation time was 3.1 min longer when a guy from England was engaged (16.9 versus 13.8 min) (Horton et al. 2000). Studies with Anthocoris have demonstrated that the length of copulation varies greatly among species in this genus. Anthocoris antevolens (12.4 min; Shimizu 1967), Anthocoris nemorum (7.5 min; Hill 1957), and Anthocoris whitei (89 min; unpublished data). We cannot explain why males of Anthocoris nemoralis from the two populations showed substantial differences in duration since we do not know what traits of the insect or its environment impact copulation duration in these predators. The amount of sperm that each male from the two populations transfers to the female during mating would be interesting to learn about.

11.5

Syrphids

Three subfamilies, the Syrphinae, Milesiinae, and Microdontinae, make up the family Syrphidae. The majority of aphidophagous syrphids are members of the Syrphinae subfamily. While a few species are aquatic or survive in extremely damp environments, the majority of species are terrestrial. At least 25% of the terrestrial forms are predatory, primarily aphidophagous. There are more than 4700 species in the world, and the Indian subcontinent alone is home to 312 species in 71 genera. In syrphids, mating can take place both in flight and at rest. Some species’ males exhibit violent territorial behaviour towards other species’ males, while other species’ males congregate to entice females with “lekking” displays. The male and female of I. scutellaris hovered mouth to mouth in the air during courtship before mating (Joshi et al. 1999). Various species have different mating cycles. Makhmoor and Verma (1987) noted that E. balteatus repeated matings lasted 1–2 s apiece, whereas E. corollae repeated matings lasted 1–6 h. In B. serarius and P. tibialis, copulation lasts 10–13 min, whereas it takes 30 min in A. nubilipennis. I. scutellaris was found to engage in repeated mating sessions that lasted 50–65 min apiece, compared to 15–20 min for P. serratus, P. yerburiensis, and 20–30 min for D. aegrota (Joshi et al. 1999). The syrphid fly’s pre-oviposition

366

11 Mating Behaviour and Reproductive Biology of Insect Predators

time ranges from 2 to 5 days depending on the species. Native to tropical and subtropical areas of the New World, Pseudodorus clavatus is one of the most prevalent and significant predators of citrus aphids in Florida, the Caribbean, and South America. Aphis spiraecola and Toxoptera citricida are the two main aphid species that attack citrus in Florida. Pseudodorus clavatus adults displayed lek-like mating behaviour, according to Belliure and Michaud (2001). Males hovered in groups and seemed to compete for position with one another, while females sat nearby and occasionally entered the male swarm to take a partner. In flight, with the female stationary, and with the male carrying her, mating was seen. Males and females in individual cages were seen to mate repeatedly and frequently during their reproductive lives, both before and after the start of oviposition, and to occasionally stay in copula for more than an hour (up to an hour and 45 min). Merodon equestris exhibits the same in-flight mating behaviour as Pseudodorus clavatus, with the male grabbing and carrying the female (Conn 1979). Although lek formation may be a result of being caged, other syrphids have been known to engage in this type of mating behaviour (Waldbauer 1990), and we saw Pseudodorus clavatus males vying for females and trying to push other males out of copula. The multiple matings and protracted stays in copula are additional behaviours that are compatible with mateguarding and sperm competition, respectively.

11.6

Hymenoptera

The “female-calling syndrome” and the “male-aggregation syndrome” are two major groups into which Holldobler and Bartz (1985) divided ant mating tactics. The females in the former attract males by releasing a sex pheromone at calling spots close to their home nest in the latter. In the latter, swarms of males assemble, luring females in for mating. Different generations of males may frequent a particular swarm site over the course of several years. Due to the sex ratios at swarms being heavily slanted towards men, both the scramble rivalry and the interference competition among males may be fierce (e.g. Holldobler 1976). While female ants may only mate once, serial mating appears to be the norm for many species (Holldobler and Wilson 1990). Due to their capacity to physically reject males, females may also be able to decide on the kind and number of partners (Holldobler 1976). The ant Formica subpolitas mating method was described as follows by O'Neill (1994). Swarming Male Formica subpolita made visits to the plants above the swarms in between flights. In the uppermost portion of the vegetation, up to 1 m above it, and within 2 m of the ground, males flew in erratic patterns. Flying males often had their bodies angled at a 45-degree inclination to the ground, facing upwind (i.e. with venter down and head up). Males altered their position to keep facing upwind when the wind direction varied throughout the day or between days. They did not approach other flying insects while in flight. They took a few seconds to fly, then landed on the stems and leaves of the flora, walking a few centimetres while frequently abruptly changing directions.

11.6

Hymenoptera

367

Copulation There have been up to 51 matings recorded at swarms in a single day. Females stayed motionless on plants between 0.1 and 1.5 m high before mating. Males either flew to females perched on the plant directly or walked to them after landing close by. A common manoeuvre involved a male mounting a female dorsally, turning to face her, and then grabbing her by the legs. The male prodded the female with the tip of his abdomen until genital contact was made if she did not resist. The guy immediately let go of his hold on the leg as insertion took place, flipping backwards to become venter up. Between the beginning of bite and the end of copulation, there was an average delay of 35 s. A total of 62 s were spent during copulation, including 5 observations in which the female did not appear to be biting. When there were other males around, the males who were not in copula mounted the female’s thorax and probed the abdomen in the same way a male would if he were the only one with her. However, it doesn’t seem like any direct attempts were made to get rid of the copulating male. Instead, each of the up to five males just makes an attempt to mount and enter his genitalia. Attempts at homosexual mating were frequent at specific times, particularly late in the daily swarm phase. The mounting and probing technique resembled that used for copulations. Even guys in spider webs or being carried by spiders were mounted by other males. Following copulation, the male promptly dispersed after separating from the female, typically stooping into the undergrowth. It is unknown if they went back to the swarm. It’s conceivable that they did because, when individual pairs were kept in vials, I frequently saw a male mate twice with the same female. Some females stayed put and later mated once again. Others either departed the swarm area right after or did so after briefly using their mouthparts to groom their genitalia. Some left by tumbling to the ground, moving into the open, and then rising to a height of 1–4 m before taking off and flying away from the region. They may have also just taken off from their perches atop the bushes and flown away. They found it difficult to take flight, especially from perches close to the ground. Multiple Mating After mating, not all females departed from the swarm. Some moved around the swarm, moving to a different location, or flew to a nearby shrub to mate once more. I am unaware of the greatest number of matings that females have experienced because I am unaware of their histories before my observations. On Rhus plants, however, I saw eight unrestrained females having repeated matings. Before rejecting additional mating efforts, six of these females mated twice, one mated three times, and one mated four times. All of the matings took place within intervals of 10 and 22 min for the last two females. Females would abandon their perches, fly away while raising their wings, or turn the tip of their abdomens away as a male prodded them with his. Additionally, I observed nine previously mated females that were put into insect nets with males. Six of these females mated with at least one additional male before rejecting up to seven other male efforts at mating. Symmorphus allobrogus mating behaviour, which explains why the males are willing to mount and mate. The male exhibits many behaviours, such as the way he positions himself during copulation and the frequency and intensity of his antennation. The study of the male’s copulatory and post-copulatory wooing

368

11 Mating Behaviour and Reproductive Biology of Insect Predators

demonstrates how monandry is maintained. The wasp exhibits a variety of secondary sexual characteristics, and its mating behaviour is based on both phylogenetically distinct sexual mating traits and sexual selection. There are considerable differences in the length of the mating phases and the quantity of male antennation series during the pre-copulatory, copulatory, and post-copulatory phases of mounting. The conduct of both sexes in the pre-copulatory phase is nearly equally important for mating success, which is mostly dependent on the actions of the male during the remounting phase. While during copulation, male activity has little impact on its duration; nevertheless, female conduct has a critical impact, causing it to end earlier (Dar et al. 2021). It was noticed that male wasps left the nests before female wasps did. During each day of observation, the emerging males chewed the cell partitions and were seen flying out of the nests or occasionally making patrolling flights very close to the established nests. When the female wasp first emerged, flying male wasps were seen approaching her for mating. Over the emerging ladies, the flying males were seen to briefly touch their abdomens with their antennae. The next step was to climb over the attractive females in an effort to copulate. The male wasp was also seen entering the nest cavities in an effort to extract newly hatched non-melanised females using its mandibles. Foraging females were also observed during the observations patrolling the observation nests looking for empty tubes to build their nests in. The males who were patrolling were seen to briefly touch the females who were foraging. The attempted male wasp was immediately rejected by the contacting females (Amala et al. 2022). We urgently need information on whether predators can recognise and utilise host sex pheromones and over what distances both predators/parasites and conspecific males actually perceive host sex pheromones in the field.

11.7

Dermaptera

Dermaptera’s adult longevity is decreased during mating. Reproductive activity, especially in females, shortened their life span. Adult Marava arachidis (Dermaptera; Forficulidae) females lived longer than men, for example. The lifespan of unmated females was also substantially greater than that of unmated males. No appreciable differences in longevity between the sexes were seen in the case of mated individuals. Despite the lengthy adult lifespan, the highest number of mornings was 4 per couple, with an average of 2.43. As previously indicated, it appears that female reproduction had a greater impact on longevity than did male sexual activity. The duration of each mating averaged 114 min, reaching a maximum of 350 min (Patel and Habib 1978). The male forceps of the species Forficula auricularia were used as tactile stimuli for the female later in the courtship process and during early displays of courtship. Guys with the forceps removed did not mate, according to a study of those males. Female forcep length was normally distributed, whereas male forcep length had a bimodal distribution and was positively allometric. Males without an advantage in

11.7

Dermaptera

369

mating were those with longer forceps. During courting, receptive females displayed behavioural activity (Walker and Fell 2001). Tirathaba rufivena is preyed upon by Chelisoches morio (Dermaptera: Chelisochidae), a significant predator (Lepidoptera: Pyralidae). When Chelisoches morio male and female adults mate, their abdomens unite, and the process can take anywhere between 2 and 40 min. Multiple matings are possible for females. Four to six days after mating, female adults start to lay eggs, which are often laid in masses of 35.7 on average. The females lay an average of 140.17 eggs apiece in total (Zhong et al. 2016). Denticulated Echinostoma The following traits of this species were discovered through staged mating tests, including surgical manipulation of male penes: (1) Males only use one of the paired penes for a single genital coupling; (2) both penes are probably functional; (3) there are no persistent biases in the use of the penes; (4) the direction of rotation of the male abdomen during genital coupling is unrelated to the pattern of male penis use; (5) sperm are transferred directly into the spermatheca, a female sperm storage organ; and (6) releasing of spermatospore capsule. The male aggressively courted the female by directing his abdomen and forceps towards the female abdomen, concurrently rotating the abdomen 180° around the anterior-posterior axis, then retreating to make genital contact (n = 19). Regardless of when the observation began, the first copulation typically occurred within a few minutes (up to 38 min). Surgically manipulating the male penes of Echinostoma denticulatum during mating studies revealed the following traits for this species: (1) Males only use one of the paired penes for a single genital coupling; (2) both penes are probably functional; (3) there are no persistent biases in the use of the penes; (4) the direction of rotation of the male abdomen during genital coupling is unrelated to the pattern of male penis use; (5) sperm are transferred directly into the spermatheca, a female sperm storage organ; and (6) ejaculation of spermatospore capsule. The male aggressively courted the female by directing his abdomen and forceps towards the female abdomen, concurrently rotating the abdomen 180° around the anterior-posterior axis, then retreating to make genital contact. Regardless of when the observation began, the first copulation typically occurred within a few minutes (up to 38 min). Repeated mattings in the same pairings of Echinostoma denticulatum were seen in 12 of the 19 cases. Two females of Echinostoma denticulatum were later reared and dissected, and it was discovered that neither of them laid eggs nor had any sperm visible in the spermatheca. The 17 inseminated females mated up to four times, compared to these two uninseminated females, who mated two and six times, respectively. Copulation lasted 153 min. The initial copulation lasted on average longer than the second and third copulations. However, when we just looked at the instances where the male successfully inseminated the female, there were no discernible differences between them. After their first mating, the males continued to court the female frequently, but she typically retreated from the courting male and only accepted him after a specific amount of time had passed. With a few exceptions,

370

11 Mating Behaviour and Reproductive Biology of Insect Predators

copulations were repeated in situations where the female was inseminated at intervals of 8–24 h. The two uninseminated females of Echinostoma denticulatum, on the other hand, repeatedly mated with the same male, with one exception, at intervals of less than 4 h (Kamimura and Lee 2014).

11.8

Coleoptera

In-depth research was done on Propylea dissecta mating habits. Males mature before females; hence, protandry was obvious. Male and female ladybirds were predicted to reach sexual maturity at ages of 7.12 and 9.33 days, respectively. The six steps of male courtship were approach, watch, scrutinise, embrace, mount, and attempt. While an embrace could be an expression of appeasement, an examination helps to identify a match. Male attraction to females is first triggered by chemical cues. The recognition of mates seems to be aided secondarily by touch and visual signals. The time that it took for mating to complete (275.4 min) was between unpaired individuals. When previously mated individuals conceived, it was significantly shorter (176.6 min) (Ahmad 2005). The predatory coccinellid Cryptolaemus montrouzieri single female beetles laid substantially more eggs in a continual mating environment than females maintained similarly with either 2 males or 1 male. However, correlation analysis revealed a highly significant inverse relationship between the number of days and the clutch size with three mates. When compared to other treatments, both the quantity and frequency of matings decreased over time (Jayanthi et al. 2013). When there were unpaired males present, mating in M. sexmaculatus started earlier (1.5 min). But when it appeared that there were unmated females available, the maximum time to begin mating was displayed (12.9 min). In the presence of a female that had been mated three times, a male that had been mated three times, and no adult, the mating procedure took 6.7, 3.5, and 1.8 min, respectively (Shashwat 2018). Adult male Harmonia dimidiata (Coleoptera: Coccinellidae) started wooing by approaching an adult female in order to execute mating. The male ladybird stopped for 1.6 s and hovered 1.8 cm away to observe the female. Other species of ladybirds have been observed to exhibit similar male behaviour (Omkar and Pervez 2005). The male gave the female a brief examination (1.0 s) before giving her a bear hug, as described by Omkar and Pervez (2005) for Propylea dissecta. The male joined the female’s latero-posterior side with his body. After that, he turned around and walked towards the front of the female. The male discharged a yellowish sticky substance from the tibia-femoral joint of his hind legs practically in front of the female’s mouth when he touched her head with his aedeagus as he circled the female from her head end (that is, from the head and pronotum) and advanced towards the posterior end. Reflex blood is released by Harmonia species as a first line of defence against intruders. The males mounted the females after the females licked the reflex blood. The male then sought to mate by mounting the female from the back and making genital contact. After protracted mating, the male would circle the female’s elytra to

11.8

Coleoptera

371

end the procedure. Males who had previously mated and those who hadn’t both exhibited this behaviour. Pervez and Jahan (2020) noted that with each subsequent mating, the amount of reflex blood secreted by the male reduced. When their spouse was not interested in copulating, the females likewise exhibited odd behaviour. They would touch their mate’s belly with their antennae and then flee. It can be a strategy to get him to mate. In Harmonia dimidiata, aposematic fluid is given as a wedding gift during mating. When agitated, ladybirds expel defensive chemicals. Larvae bleed from the dorsal glands, while adults bleed from the tibia-femoral joints. It can stick to a predator’s legs, antennae, and mouthparts because it is sticky and quickly coagulates when exposed to air. This protective chemical in Harmonia dimidiata also acts as a lure or nuptial gift to encourage or start mating (Pervez and Jahan 2020). The authors came to the following conclusions about Harmonia dimidiata mating: I mating in H. dimidiata is very long, with (v) the duration of mating of virgins being the longest, after which it decreases with each subsequent mating. (1) Males provide a courtship gift in the form of reflex blood, which is readily accepted by the female. (2) The amount offered decreases with each subsequent mating. Age or social status of men and women may have an impact on fecundity function. Although age in mating systems has been theoretically and practically examined in other insects to better understand its evolutionary significance, predatory ladybirds have only received a scant amount of research. Adalia bipunctata, a two-spotted ladybird, exhibits protogyny as determined by age-related mating incidence, yet Propylea dissecta, another ladybird, has protandry. While females of all ages (newly emerging to 30 days old) mated with the older males, Propylea dissecta males began mating at 3 days of age (10 to 30 days-old). Though reluctant, older males forcefully mated younger Propylea dissecta females (5 days old pooled). The younger males palpated the older females (10 to 30 days old) before mating with them if they became motionless. Only a small percentage (0.29% of males and 0.29% of females) of younger Propylea dissecta adults mated either with similar or older individuals of the opposite sex. Older adults (10 to 30 days-old) displayed 100% mating with each other. The information on younger (1- to 5-day-old) individuals’ reproductive characteristics and mating habits. After mating, the females did not expel any spermatophore or jelly-like substance. With an increase in male age, there was a sigmoidal increase in mate-seeking behaviour. Male Propylea dissecta were 0% unwilling to mate at 2 days or younger, but were more 81.11% to 83.33% willing at 10 days or older. Propylea dissecta females showed a similar sigmoidal increase in their desire to mate, with 51.11% to 52.22% of the older females (10 to 30 days-old) being open to mating (Pervez and Richmond 2004). Its fertility is greatly influenced by the length and frequency of mating as well as the size and age of the spouse. When individuals of the Propylea dissecta species were treated to five consecutive matings, the length of the mating period decreased significantly. With an increase in matings, fecundity and percent egg viability both considerably rose (Ahmad 2005). It was discovered that mating length was dependent on the ages of the mating partners, and that older individuals mated for longer than younger ones. The causes

372

11 Mating Behaviour and Reproductive Biology of Insect Predators

of the prolonged mating are unknown. However, it may be brought on by the elder male’s stronger mate-seeking drive or by multiple or prolonged sperm transfers as a result of a protracted mate-deprivation period. Adalia bipunctata, an aphidophagous two-spot ladybird, has multiple ejaculations, meaning that sperm transfer happens twice or three times during a single mating. This prolongs the mating process. Ejaculations that are extended or repeated appear to improve progeny production (Majerus 1994). Propylea japonica, a ladybird, was shown to have a relatively lengthy mating duration when compared to other ladybirds like Harmonia axyridis, Coccinella septempunctata, and Cheilomenes sexmaculata, which all mate only once (Obata and Johki 1991).

11.9

Chemical Ecology

Because mating signals are frequently obvious to predators and parasites, viability selection is anticipated to counter the sexual selection that would otherwise favour their evolution. Many secondary sexual characteristics can be an attempt to strike a balance between appeal and evasion of detection. The majority of instances of this signal exploitation originate from species that use audio or olfactory mating signals, and there are just a small number of instances of visual signal exploitation that can be supported. The severity of sexual selection on male or parental features can be affected by the fact that males are typically the signalling sex, making them more vulnerable to predators or parasitoids that identify prey or hosts by sexual signals. When mediating intraspecific interactions between the emitter and its conspecifics, pheromones are highly species-specific. Sex pheromones, aggregation pheromones, marking pheromones, and alarm pheromones are just a few of the subcategories that they fall under. According to Wicker-Thomas (2007), sex pheromones, along with visual, tactile, auditory, and other cues, are recognised for many dipteran species and are crucial in courtship behaviour. Recently, pheromones for several dipteran species have been discovered. This survey includes a wide range of species from all the researched families. With a particular emphasis on the sex pheromones involved, the review analyses various courtship behaviours in Diptera. Pheromones are volatile substances that have a long range of action and are found in the Nematocera suborder. They originate from terpenes or short-chain alkanes containing acetoxy groups (Cecidomyidae) (Psychodidae). Pheromones in the Cyclorrhapha can be non-volatile, unsaturated or methyl-branched hydrocarbons that act by contact or volatile, produced from alkanes or terpenes (Tephritidae) (the other subgenera). For mating, identifying a suitable host or prey, or finding a shelter, insects employ a variety of signals. These cues can be olfactory, gustatory, aural, visual, tactile, or visual-tactile. While pheromones from males of the predator Nabis pseudoferus (Hemiptera: Nabidae) attracted conspecific females and increased their responses to HIPVs from wheat plants infested with Rhopalosiphum padi (Hemiptera: Aphididae), pheromones from males of the mirid predator Nesidiocoris tenuis (Hemiptera: Miridae) attracted (Cabello et al. 2017). In agricultural plants,

References

373

kairomone-based lures are used, and natural enemies may locate herbivores based on background odours. The (E)-2-octenyl acetate, a rather frequent component of the volatile semiochemicals produced by a number of heteropteran species, attracted male bugs to adult female Geocoris punctipes (Marques et al. 2000). Recent discoveries demonstrate that ladybirds produce sexual pheromones. The volatile mixture of the sex pheromone Harmonia axyridis was found to have a large amount of the compound (-)-b-caryophyllene (Fassotte et al. 2014). Previous studies discovered a considerable electrophysiological difference between the male and female antennae of Harmonia axyridis in response to (-)-b-caryophyllene, demonstrating that the males were more sensitive (Verheggen et al. 2007).

11.10 Future Recommendations Very limited works are available for mating behaviour of various predators; hence, we suggest carrying out as much as work to know the exact mating acts and chemicals responsible for the same.

References Abdalla ET, Yanik E, Achiri DT (2018) Reproductive performance and mating behavior of the predatory bug, AnthocorisminkiDohrn, 1860 (Hemiptera: Anthocoridae). Turk Entomol Derg 42 (2):117–125 Ahmad P (2005) Mating behavior of an aphidophagous ladybird beetle, Propyleadissecta (Mulsant). Insect Sci 12(1):37–44 Amala U, Kandan A, Timalapur MS (2022) Nesting and predatory behaviour of potter wasp, Rhynchiumbrunneumbrunneum (Eumeninae: Vespidae: hymenoptera) in an urban farm landscape. Biocontrol Sci Tech 32:794. https://doi.org/10.1080/09583157.2022.2047612 Ambrose DP (1987) Biological, behavioural and morphological tools in the biosystematics of Reduviidae (Insecta, Heteroptera, Reduviidae). Proc Ani Sci 96:499–508 Ambrose DP (1999) Assassin bugs. Science Publishers, Enfield, NH, p 337 Ambrose DP, Livingstone D (1978) On the mating behaviour of Ectomocoris fibialis Distant and Acanthaspis pedestris Stal two reduviids of the scrub jungles of Palghat gap. J Madras Univ 41:69–79 Ambrose DP, Livingstone D (1987a) Mating behavior and the impact of mating on oviposition pattern and hatchability in Rhinocoris kumarii. Ecol Environ 5(1):156–161 Ambrose DP, Livingstone D (1987b) Mating behaviour of Coranus vitellinus (Insecta: Heteroptera: Reduviidae). In: Aspects of behaviour. New College, Madras, pp 22–26 Ambrose DP, Vennison SJ (1990) Diversity of spermatophore capsules in reduviids (Insecta, Heteroptera, Reduviidae) of South India. Mitteilungenausdem museum fürNaturkunde in Berlin. Zoologisches Museum und Institutfür Spezielle Zoologie (Berlin) 66(2):309–317 Ambrose DP, Sebasti Rajan XJ, Nagarajan K et al (2009) Biology, behaviour and functional response of Sphedanolestes variabilis (Insecta: Hemiptera: Reduviidae: Harpactorinae), a potential predator of lepidopteran pests. Entomol Croat 13(2):33–44 Belliure B, Michaud JP (2001) Biology and behavior of Pseudodorusclavatus (Diptera: Syrphidae), an important predator of citrus aphids. Ann Entomol Soc Am 94(1):91–96 Cabello T, Rodriguez-Manzaneque M, Gallego JR (2017) Can the pheromones of predators modulate responses to herbivore-induced plant volatiles? Ann Appl Biol 170:369–378

374

11 Mating Behaviour and Reproductive Biology of Insect Predators

Claver MA, Reegan AD (2010) Biology and mating behaviour of Coranusspiniscutis Reuter (Hemiptera: Reduviidae), a key predator of rice gandhi bug Leptocorisa varicornis Fabricius. JBiopest 3(2):437 Conn DLT (1979) Morphological and behavioural differences in populations of Merodon equestris (F.) (Dipt.,Syrphidae). Entomol Mon Mag 114:1364–1367 Dar SA, Sayed S, El-Sharnouby M et al (2021) Mating behaviour and behavioural ecology of a predatory wasp, Symmorphusallobrogus (de Saussure) (hymenoptera: Eumeninae). Saudi J Biol Sci 28(7):3720–3726 Fassotte B, Fischer C, Durieux D et al (2014) First evidence of a volatile sex pheromone in lady beetles. PloS One 9:e115011 Gemeno C, Claramunt J (2006) Sexual approach in the praying mantid Mantis religiosa (L.). J Insect Behav 19(6):731–740 Hardy I, Ode PJ, Siva-Jothy M (2007) Mating behaviour. In: Insects as natural enemies. Springer, Dordrecht, pp 219–260 Hill AR (1957) The biology of Anthocoris nemorum (L.) in Scotland (Hemiptera: Anthocoridae). Trans R Entomol Soc Lond 109:379–394 Holldobler B (1976) The behavioral ecology of mating in harvester ants (hymenoptera: Formicidae: Pogonomyrmex). Behav Ecol Sociobiol 1:405–423 Holldobler B, Bartz SH (1985) The sociobiology of reproduction in ants. In: Holldobler B, Lindauer M (eds) Experimental behavioral ecology and sociobiology. Sinauer Associates, Sunderland, MA, pp 237–257 Holldobler B, Wilson EO (1990) The ants. Harvard University Press, Cambridge, MA Horton DR, Hinojosa T, Lewis TM (2000) Mating preference, mating propensity, and reproductive traits in Anthocoris nemoralis (Heteroptera: Anthocoridae): a comparison of California and United Kingdom populations. Ann Entomol Soc Am 93(3):663–672 Horton DR, Lewis TM, Hinojosa T (2001) Copulation duration and probability of insemination in Anthocoriswhitei (Hemiptera: Anthocoridae) as a function of male body size. Can Entomol 133 (1):109–117 Hosken DJ, Stockley P (2003) Benefits of polyandry: a life history perspective. In: Evolutionary biology. Springer, Boston, MA, pp 173–194 Jayanthi PK, Sangeetha P, Verghese A (2013) Influence of polyandry on clutch size of the predatory coccinellid, Cryptolaemus montrouzieri (Coleoptera: Coccinellidae). Fla Entomol 96:1073– 1076 Joshi S, Ballal CR, Rao NS (1999) Species complex, population density and dominance structure of aphidophagoussyrphids in cowpea ecosystem. Entomon 24(3):203–214 Kamimura Y, Lee CY (2014) Mating and genital coupling in the primitive earwig species Echinosomade nticulatum (Pygidicranidae): implications for genital evolution in dermapteran phylogeny. Arthropod Syst Phylo 72(1):11–21 Kumar SP (1993) Biology and behavior of chosen assassin bugs, (Insecta: Heteroptera: Reduviidae). PhD thesis, Madurai Kamaraj University, Madurai, India Laumann RA, Kavčič A, Moraes MC et al (2013) Reproductive behaviour and vibratory communication of the neotropical predatory stink bug Podisusnigrispinus. Physiol Entomol 38 (1):71–80 Majerus MEN (1994) Entomologist’s monthly magazine. Female promiscuity maintains high fertility in ladybirds (Coleoptera: Coccinellidae). Entomol Mon Mag 130:205–209 Makhmoor HD, Verma AK (1987) Bionomics of major aphido phagoussyrphids occurring in mid-hill regions of Himachal Pradesh. J Biol Control 1(1):23–31 Marques FDA, McElfresh JS, Millar JG (2000) Female-produced sex pheromone of the predatory bug Geocoris punctipes. J Chem Ecol 26(12):2843–2855 Nitin KS, Bhat PS, Raviprasad TN et al (2017) Biology, behaviour and predatory efficiency of Sycanusgalbanus distant. Hemiptera: Reduviidae: Harpactorinae recorded in cashew plantations. J Entomol Zool Stud 5:524–530

References

375

Obata S, Johki Y (1991) Proceedings of 4th meeting of the IOBC WG ecology of Aphidophaga. Comparative study on copulatory behaviour in four species of aphidophagous ladybirds, pp 207–212 Omkar O, Pervez A (2005) Mating behavior of an aphidophagous ladybird beetle, Propyleadissecta (Mulsant). Insect Sci 12:37–44 O'Neill KM (1994) The male mating strategy of the ant Formica subpolita Mayr (hymenoptera: Formicidae): swarming, mating, and predation risk. Psyche 101(1–2):93–108 Patel PN, Habib EM (1978) Biological and behavioral studies of an ovoviviparous earwig, Maravaarachidis (Yersin, 1860) (Dermaptera; Forficulidae). Rev Biol Trop 26(2):385–389 Perez B (2005) Calling behaviour in the female praying mantis, Hierodula patellifera. Physiol Entomol 30(1):42–47 Pervez A, Jahan M (2020) Mating behaviour of the predaceous ladybird, Harmonia dimidiata. Eur J Environ Sci 10(1):5–8 Pervez A, Richmond AS (2004) The influence of age on reproductive performance of the predatory ladybird beetle, Propylea dissecta. J Insect Sci 4(1):22 Rani PU, Havukkala I (1993) Predatory and mating behaviour of Eocantheconafurcellata (Wolff.) (Heteroptera: Pentatomidae) a promising natural enemy of Lepidopterous larvae. J Biol Control 7(1):9–11 Shashwat S (2018) Dynamics of oviposition in Aphidophagous ladybird beetles. PhD thesis, University of Lucknow Shimizu JT (1967) A biology of Anthocorisantevolens white, a predator of pear psylla (Hemiptera: Anthocoridae). MS thesis. University of California Berkeley Srikumar KK, Bhat PS, Raviprasad TN et al (2014) Biology and behavior of six species of Reduviids (Hemiptera: Reduviidae: Harpactorinae) in a cashew Ecosystem1. J Agric Urban Entomol 30(1):65–81 Vennison SJ (1988) Bioecology and ethology of assassin bugs (Insecta: Heteroptera: Reduviidae). (Ph.D. thesis). Madurai Kamaraj University, Madurai, Madurai Vennison SJ, Ambrose DP (1990) Diversity of eggs and ovipositional behaviour in Reduviids (Insecta, Heteroptera, Reduviidae) of South India. Mitteilungenausdem Museum fürNaturkunde in Berlin Zoologisches Museum und Institutfür Spezielle Zoologie, Berlin 66(2):319–331 Verheggen FJ, Fagel Q, Heuskin S et al (2007) Electrophysiological and behavioral responses of the multicolored Asian lady beetle, Harmoniaaxyridis Pallas, to sesquiterpenesemio chemicals. J Chem Ecol 33:2148–2155 Waldbauer GP (1990) Hilltopping by males of Eupeodesvolucris (Diptera: Syrphidae). Gt Lakes Entomol 23:175–176 Walker KA, Fell RD (2001) Courtship roles of male and female European earwigs, Forficula auricularia L. (Dermaptera: Forficulidae), and sexual use of forceps. J Insect Behav 14(1):1–17 Wicker-Thomas C (2007) Pheromonal communication involved in courtship behavior in Diptera. J Insect Physiol 53(11):1089–1100 Zhong B, Lv C, Qin W (2016) Preliminary study on biology and feeding capacity of Cheli soches morio (Fabricius) (Dermaptera:Chelisochidae) on Tirathabarufivena (Walker). SpringerPlus 5: 1944

Offense and Defence Mechanism of Insect Predators

12

Contents 12.1 12.2 12.3

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Behavioural Defence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Structural Defences of Plants–Herbivory–Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.3.1 Mouthparts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.3.2 Claw . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.3.3 Plant Trichomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.3.4 Epicuticular Waxes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.4 Chemical Defence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.4.1 Cardiotonic Steroids (CTS) as a Chemical Defence . . . . . . . . . . . . . . . . . . . . . . . . . 12.5 Cryptic Colouration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.5.1 Aposematism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.5.2 Green-Brown Polymorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.6 Masquerade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7 Parental Care . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7.1 Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7.2 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7.3 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7.4 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.7.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.8 Hibernation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.9 Camouflaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.9.1 Mantids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.9.2 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.9.3 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.10 Recommendations for Future Works . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

378 378 379 380 382 382 383 384 388 390 390 391 392 392 392 393 395 395 396 396 398 399 400 402 402 402

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_12

377

378

12.1

12

Offense and Defence Mechanism of Insect Predators

Introduction

In contrast to predatory aggression, offence and defensive behaviours are two types of aggression that are displayed by one member of a species against another. Offense is approaching and attacking while the attacker employs successful attack strategies. Behavioural adjustments, structural specialisations, and biochemical clarifications all play a role in defence. In actuality, the patterns of wounding/killing sustained during combat may be used to identify between the offensive and defensive animals. Furthermore, if the encounter occurs outside or in a big arena, the aggressive animal will approach its opponent confidently or will chase after them. When an opportunity for escape arises, the defensive animal’s locomotion is characterised by escape attempts and flight locomotion, the act of a member of one species—the predator— killing and devouring a member of a different species—the prey. Predators catch their prey by foraging, which can involve hunting and other forms of hunting as well as simply waiting for prey to ambush them. While some predators use camouflage, others use aggressive imitation to pose as a non-threatening creature.

12.2

Behavioural Defence

Predation plays a crucial role in the survival of wild animals, affecting fitness-related factors like nutrition, reproduction, and frequently, ultimately, mortality. Primary and secondary defences are two different categories of anti-predation activities. Secondary defence behaviours act after the commencement of a capture effort and lower the likelihood that it will be successful, whereas primary defence activities diminish the likelihood that a predator will initiate a capture attempt. The second type of defensive behaviour, known as anti-collision behaviour, lessens physical harm brought on by collisions with other items or a predator attack (Yamawaki 2017). Lamon, Topoff, and others examined the anti-predation behaviours of three ant species of the genus Camponotus in field and laboratory investigations (1981). They claim that contact with the army ant Neivamyrmex nigrescens caused colonies of Camponotus festinatus to remove their nests and brood. Camponotus oereatus and Camponotus vicinus colonies recruited members of the main caste to protect their nests. Tactile interactions between Camponotus nestmates were necessary for the start of an evacuation or an active recruiting campaign. A high level of opponent specification by Camponotus is shown from the absence of evacuation or recruitment following other disturbances that did not entail army ant interaction. When worker morphology was compared, it became clear that the degree of worker polymorphism and each species’ defence mechanism for their nests were closely related. These results imply that caste polymorphism and colony defence systems have both evolved as a result of interspecific ant predation (Lamon and Topoff 1981). Numerous predatory insects locate their prey visually and fixate on it by moving their body or head in its direction. In order to perform fixation, they modify the

12.3

Structural Defences of Plants–Herbivory–Predators

379

amplitude and direction of their turning in accordance with the location of the prey in their field of vision. Young Coleomegilla maculata lengi coccinellids typically exhibit escape reflexes (dropping, escaping, withdrawing), and they do not survive when captured by Chrysoperla rufilabris lacewings. Older larvae may actively protect themselves by squirming or biting, even after being punctured, in addition to escape behaviours. Compared to the first (35%) and second (19%) instars, the proportion of physical encounters that led to coccinellid death was much lower for older instars (2%). The ratio of Chrysoperla rufilabris to Coleomegilla maculata lengi size affected the capture effectiveness of C. rufilabris. The most frequent protective strategy used by Coleomegilla maculata lengi larvae was to drop off the leaves, and this helped to lower overall mortality of the first instars to a level similar to that seen for older stages. Chrysoperla rufilabris, on the other hand, were able to stop Coleomegilla maculata lengi from falling by impaling and holding their prey in the air or by falling to the ground alongside the ladybirds and continuing their attack (Lucas et al. 1997). Another form of behaviour seen in the predatory insects is counterattack. In Stigmaeopsis species, counterattacks have been examined, but rarely in other web-nesting species. We looked into how the predatory thrip Schizotetranychus brevisetosus, which lives in Quercus glauca, responds to attacks (Fagaceae). A miteinfested piece of leaf was used to introduce Scolothrips takahashii nymphs, which immediately entered the nest but were chased out by females using their stylets. In contrast, in a different experiment, we were unable to distinguish between treatments—10 adult females per leaf or none per leaf—in terms of the quantity of eggs ingested after 48 h. This was likely due to the Scolothrips takahashii constantly returning to the nest. A significant impact of web presence on predation was not also found. This study demonstrates that Schizotetranychus brevisetosus counterattacks occur both inside and outside the nest, and the latter specifically is the first account of such behaviour in Acaridae, even though the defensive impact is imperceptible (Ito and Ioku 2022).

12.3

Structural Defences of Plants–Herbivory–Predators

There hasn’t been much discussion of the notion that huge size could serve an antipredator role in insects, while small size helps insect defence. However, insects’ use of great size as an evolved anti-predator defence is really minimal. However, we think that big body size may be under positive directional selection since it can significantly increase insect survival against both invertebrate and vertebrate predators in some situations. Despite having large spines to protect their dorsal surfaces, Curinus coeruleus (Chilocorinae) larvae had the smallest mandibles, moved the slowest, and were the least effective in interspecific larval conflict. In choice experiments using dead larvae, Olla v-nigrum (Coccinellinae) and Harmonia axyridis (Coccinellinae) larvae appeared to find third instar Curinus coeruleus large spines less appetising as food, but they were ineffective in Petri dish arenas as a defence against these species. The

380

12

Offense and Defence Mechanism of Insect Predators

Olla v-nigrum larvae had the smoothest dorsal surface and intermediate-sized mandibles, but they moved the fastest, which helped them survive intraguild battles. The largest mandibles were found in Harmonia axyridis larvae, which were intermediate in terms of dorsal spines and movement speed. The only species among the coccinellids to successfully struggle against Chrysoperla rufilabris larvae of a similar age was this species, which was the most effective intraguild combatant. Chrysoperla rufilabris impaled coccinellid larvae at a reasonably safe distance thanks to its speed, agility, and long mandibles. The spines of Curinus coeruleus larvae prevented Chrysoperla rufilabris from attacking them laterally, but they did not offer long-lasting defence against repeated assaults. Although the importance of dorsal spines as protective structures was not completely ruled out, success in these encounters appeared to be mostly a consequence of offensive weapons (mandible size and shape) and speed of movement. The highest rates of larval cannibalism were found in Chrysoperla rufilabris, and they largely corresponded to each species’ levels of violence during interspecific conflict Michaud and Grant (2003). Maltose and sucrose, two of the most stimulating sugars in Pterostichus oblongopunctatus, were the ones that caused the antenna sugar-sensitive neuron innervating the contact chemosensilla to be activated. Rotting wood can be found among Pterostichus oblongopunctatus’s hibernation locations. The sugars that make up cellulose and hemicelluloses are cellobiose, arabinose, xylose, mannose, rhamnose, and galactose. During enzymatic wood decomposition, brown-rot fungus releases them. None of them activated the sugar-sensitive neuron in the antenna. Pterostichus oblongopunctatus does not include them in its search for hibernation places, which includes decaying wood (Merivee et al. 2008).

12.3.1 Mouthparts Many predatory insects are actively hunting. Many ants (Formicidae), ground beetles (Carabidae), and tiger beetles (Cicindelidae) run over the ground vigorously in search of prey. Numerous predatory insects that walk or crawl, such as ladybird beetles (Coccinellidae), lacewing larvae (Chrysopidae), syrphid fly larvae (Syrphidae), and predatory bugs (Hemiptera), eat eggs and young larvae of more mobile species as well as sedentary insects like aphids and scale insects. Other flyers that prey on insects are nimble and actively grab flying insects. Large eyes and powerful wings are characteristics of dragonflies and damselflies (Odonata), which prey mostly on mosquitoes and other small flying insects. Insects at least the size of themselves, including bees and wasps, are frequently attacked and subdued by robber flies (Asilidae), who also seize their victims on the flight. They have powerful legs and piercing mouthparts for sucking up the secretions of their victim. Other insects, like a large number of social wasps (Vespidae), fly and catch insects from the ground or foliage (Weseloh and Hare 2009). Therefore, it is important to carefully study and emphasise the anatomy and specialisations of predatory insects’ mouthparts and legs.

12.3

Structural Defences of Plants–Herbivory–Predators

381

When the mouthparts’ shape varies, the feeding mechanism also does. The paired mandibles, paired maxillae, and unpaired labium are the three appendages that make up an insect’s mouthparts. Additionally, the labrum and the hypopharynx are head structures. While the ectognathous mouthparts of Insecta articulate externally on the head capsule, the no insect lineages of Hexapoda have entognathous mouthparts that are hidden inside the head. Particularly in winged insects, distinctive mouthpart adaptations evolved in relation to a variety of food sources, leading to specialisation in eating and improved functional performance. Only four predatory pentatomid mouthparts from the family Pentatomidae were examined. Podisus maculiventris, Eocanthecona furcellata, Canthecona furcellata, Perillus bioculatus, the mouthparts of Podisus bidens, Podisus lewisi, and Canthecona bhoutanica are similar to those of other true bugs in that they consist of a pair of separated mandibular and interlocked maxillary stylets that are enclosed in the groove of the labium, a long, four-segmented labium, and a tapered labrum (Lm). In front of at least three weakly re-curved teeth on the inner surface of the right maxillary stylet are two rows of at least seven highly re-curved teeth. The mouthparts (rostrum) of the insect are placed against the sternum in a parallel position to the body when it is sleeping or not feeding (Wang et al. 2020). Most heteropterans were also found to have a similar structure. All heteropterans employ their stylets and digestive enzymes to feed at the same time. They consume food by puncturing and severing tissues with their stylets and then liquefying it into nutrient-rich slurry by pumping digestive enzymes through the salivary canal. The alimentary canal is used to consume the food slurry, which is then delivered into there to be further broken down and absorbed. The reduviid rostrum is typically three-segmented, short, and thick. It is often curved and arched, and the labium is kept in a longitudinal groove (the friction groove) in the middle of the front thoracic and abdominal plates when the animal is at rest. Reduviid species’ mouthpart morphological details have only occasionally been revealed. Prior research mainly focused on the terminal labial sensilla, gross mandibular and maxillary morphology, interlocking mechanisms of the mandible and maxilla, and gross labium and labrum morphology. To ascertain how much variation in mouthparts occurs among various species, particularly those with very narrow feeding specialisation; more in-depth information on fine structure is required. Miridae are a family of historically rapacious predators that comprise both generalist and prey-specific specialists, such as mites, lace bugs (Tingidae), and thrips (Thysanoptera). Odonata adults seek their prey while flying, whereas the larvae are aquatic predators. The labium of Odonata larvae differs from that of adults in the most striking ways. A prehensile labial mask made from the larval labium is employed to catch potentially swift-moving species up to the predator’s own size. However, the mandibles and maxillae are hardly different from those of adults. The most distinguishing feature of odonatan larvae, and one that makes them somewhat unusual among insects, is the transformation of the labium into a prehensile mask. In varying degrees of detail, structural elements of the Odonata larval mouthparts and the operational mechanism of this catching device have been studied. The

382

12

Offense and Defence Mechanism of Insect Predators

labium’s functionality revealed that the branchial chamber, an internal rectal organ used by the respiratory system of Anisoptera, is the primary source of the increase in haemolymph pressure responsible for the labium’s elongation (Büsse et al. 2017). Coccinellids have five different pairs of mandibles, two pairs of maxillae, an unpaired labium, and an unpaired hypopharynx as their mouthparts. According to their length, morphology, and distribution, two types of sensilla chaetica, four types of sensilla basiconica, two types of sensilla styloconica, two types of sensilla placodea, and one type of cuticular pore were identified. Seven different forms of sensilla were found in total, and 14 different types of sensilla can be determined based on their lengths and distribution.

12.3.2 Claw The prey is fixed between the femur (upper arm) and tibia (lower arm) of the insect leg, and the claw is fully retractable. These predatory claws have a wide variety of morphologies. Some species have straight, spine-covered claws, while others have curved smooth claws. We have created a physical model to study the mechanics of the femur-tibia type of predatory insect claws, ultimately attempting to explain why some insect species have curved claws rather than straight ones (Petie and Muller 2007). The key outcomes include 1. Curved claws are more effective at driving prey away from the pivot point than straight claws when compared, which eliminates the requirement for frictiongenerating structures. 2. There is a place in the curved claw model where the force acting on the prey is exactly zero. This is due to the femur and tibias natural opposing and lining forces. The prey is perfectly clamped and not being forced out of the claw at this point. Straight claws do not have this characteristic. 3. The prey cannot be positioned in the curled claw more than a certain maximum distance from the pivoting point. Because moment arms are close to zero at this maximum location, the resulting force on the prey reaches high values. 4. The resulting force is directed towards the pivoting point between the zero position and the maximum position, stabilising prey fixation.

12.3.3 Plant Trichomes Plant trichomes and secretions can interfere with the search movement of small natural enemies and reduce their foraging efficiency. Green lacewings move slower on tobacco, which is well defended with trichomes, than on cotton, which has relatively few trichomes. Similarly, they movement of coccinellid beetles searching on tobacco is inhibited by the plants’ hairy trichomes and exudates from tobacco trichomes inhibit the movement of Trichogramma wasps. Coccinellid beetles,

12.3

Structural Defences of Plants–Herbivory–Predators

383

preying on aphid pests of the plant Mentzelia pumila, were incapacitated by the plants’ trichomes (Lambert 2007). According to Lambert (2007), the tomato trichomes proved to be lethal to the nymphs. Of 30 nymphs, 27% died within 2 days of being placed on the plant. Furthermore, 47% died by the fourth day on the plant. For nymphs on plastic plants, only one nymph died by the second day (6.7%) and two (13.3%) by the fourth day. Many of the nymphs that died on the tomato plants were entrapped in the long non-glandular trichomes on the stem of the plant and did not meet the shorter glandular trichomes. This entrapment was caused by the mechanical defensive features of the no glandular trichomes, which trap insects, rather than the “sticky” chemical features of the glandular trichomes. However, several of the nymphs that were still alive were affected by the exudates from the glandular trichomes. Of the nymphs that were still alive, 31% were stuck to the plant. Six of the remaining nymphs repeatedly fell off the plants during the observation time due to the excessive gumming on their legs from the glandular exudates. When the live nymphs were viewed under a dissecting scope after the trials were completed, all were found to have gummed legs from the trichome exudates. The electron micrographs also showed that the legs of the nymphs on tomato plants (after four days) had considerable gumming compared to the non-gummed legs of nymphs never placed on tomato plants. The tarsal claw, which predators use to hold on to the plant, was completely encased, rendering it unusable. The same gumming was observed on nymphal leg hairs. These leg hairs were mostly or completely covered over with the trichome exudates (Lambert (2007), Plant trichomes also entrap pollen from the environment. This pollen increases the abundance of predatory arthropods, subsequently decreasing herbivore abundance (Van Wyk et al. 2019). Because alternate resources such as pollen and insect carrion are very common in the environment, accumulating those resources may be an important and largely overlooked way in which plants attract natural enemies. Moreover, glandular residues can be induced by damage to leaves, increasing visitation by natural enemies (Karban et al. 2019).

12.3.4 Epicuticular Waxes In order to feed, phytophagous insects and their natural adversaries typically adhere to plants. Usually, this entails adhering to the lipophilic layer of substances, or “epicuticular waxes”, that coats all main plant surfaces. Epicuticular waxes, which appear as crystalline waxy “blooms”, can protect plants from herbivory by preventing phytophagous insects from adhering to the plants. Contrarily, epicuticular waxes blooming may lessen parasitoid and predatory attachment, thereby liberating populations of phytophagous insects from the control of their natural adversaries. Therefore, the overall impact of epicuticular wax flowers on herbivory should differ depending on the system. Epicuticular waxes often reduce the attachment pressures that insects can produce on plants when they have been measured. Predators and some herbivores favour plants with blooms of epicuticular wax when they are foraging. Some of these appear to have physiological or behavioural

384

12

Offense and Defence Mechanism of Insect Predators

modifications that either enhance attachment to EW blooms or compensate for a reduction in attachment to surfaces of plants bearing epicuticular waxes blooms. It is unknown how epicuticular wax blossoms interfere with insect adhesion or how insects can overcome this difficulty. In this review, certain hypotheses are presented. Their clarification may aid in understanding how organisms cling to plants, with implications for improving biological management of some insect pests (Eigenbrode 2004). There are three criteria that are frequently used to identify enemy-free territory, regardless of whether the plant has trichomes, glandular trichomes, epicuticular wax, or any other exudation by the plants (Wheeler and Krimmel 2015): 1. Prey gain from predator exclusion (prey fitness is higher within enemy-free space than it is outside enemy-free space when predators are present), 2. Predators are crucial (they reduce prey fitness outside enemy-free space), and. 3. Enemy-free space comes at a price (lower prey fitness in the absence of predators).

12.4

Chemical Defence

Few chemicals are frequently generalist signals in predator–prey interactions that the prey emits to communicate information to a number of predator species with different sensory capacities. As a result, it is more likely that anti-predator substances will be comparable across species and developmental stages. There are several different signals that can be used to identify the presence of predators, including chemical cues. Predators produce chemical signals to communicate, to protect themselves from desiccation, or possibly as a by-product of their activities. Chemical communication is crucial for a variety of biological activities in insects, such as courtship behaviour, kin or species recognition, prey–predator or plant–insect interactions, and chemical communication during courtship. Additionally, predator repellent substances may combine visual and aural cues to enhance long-term memory retention of unwanted prey or partners. In their nymphal and adult phases, true bugs have a wide range of defence compounds, including terpenes, phenolics, pregnanes, cardiac glycosides, alkanes, aliphatic esters, and aldehydes. In Heteroptera, for instance, (E)-2-Alkenals are essentially ubiquitous and are known to serve as intra-specific signals like aggregation and sex pheromones. The dorsal abdominal gland secretions of the nymphs of Riptortus pedestris (Heteroptera: Alydidae), Thasus acutangulus (Heteroptera: Coreidae), and Euschistus biformis (Heteroptera: Pentatomidae) adults, as well as the metathoracic scent gland secretions of these insects, were identified as (E)-2-Hexenal, Tenodera aridifolia sinensis, a Chinese praying mantid (Mantodea: Mantidae), moved away from the location of interaction when exposed to (E)-2-hexenal, (E)-2-octenal, and (E)-2-octenyl acetate, but not to 4-OHE or (E)-2-hexenyl acetate. Additionally, the likelihood of prey escape and rejection by predators was well predicted by the proportion of iridoid glycosides present in each individual caterpillar. Higher levels of iridoid glycosides in caterpillars increased their likelihood of

12.4

Chemical Defence

385

being rejected and escaping from predators (Dyer and Deane Bowers 1996). Similar investigations were conducted in the same year by Traugott and Stamp (1996) with the hemipteran predator Podisus maculiventris and its victim, Manduca sexta, which was fed tomatine and chlorogenic acid. Both inexperienced and tomatineexperienced predators discovered Manduca sexta, a tomatine-fed prey, more quickly. Podisus maculiventris and predators exposed to chlorogenic acid were more motivated to start looking for prey. However, seasoned Podisus maculiventris stinkbugs were less inclined to attack both caterpillars treated with tomatine and caterpillars fed with chlorogenic acid compared to novice predators. These findings show that allelochemical-fed prey was simpler for Podisus maculiventris stinkbugs to find, but allelochemical-containing prey Manduca sexta frequently prevented predation by skilled Podisus maculiventris stinkbugs (Traugott and Stamp 1996). Utetheisa ornatrix eggs are naturally preyed upon by the larval form of the green lacewing (Ceraeochrysa cubana), a moth that secretes pyrrolizidine alkaloids from its larval food plant (Crotalaria spp.). Usually, utetheisa eggs contain the alkaloid. Experimentally created alkaloid-free Utetheisa eggs are pierced by the larva’s razorsharp tubular jaws and sucked out. Conversely, alkaloid-rich eggs are not accepted. The larva conducts an examination procedure before attacking a Utetheisa egg cluster, which typically contains 20 eggs. If any of the two to three eggs it prods and/or pierces contain alkaloid, it passes “negative judgment” on the other eggs in the cluster and flees. The larva’s generalisation makes sense given how little the alkaloid content of the eggs within clusters varies. However, there is a significant difference in egg alkaloid content between clusters; therefore, it is reasonable to assume a large range in the palatability of clusters in nature. Therefore, it must be adaptive for the larva to evaluate each cluster for acceptability, just as it must be adaptive for Utetheisa to lay her eggs in huge clusters and distribute alkaloid equally among the eggs in a cluster (Eisner et al. 2000). Methyl salicylate (MeSA) has been shown in laboratory research to be appealing to the predatory bug Anthocoris nemoralis (Anthocoridae) (Drukker et al. 2000). However, more recently, it was discovered that MeSA, (Z)-3-hexenyl acetate, Chrysopa nigricornis (Chrysopidae), Geocoris pallens (Gecoridae), Stethorus punctum picipes (Coccinellidae), and species of Syrphidae (James) were also present (2003a, b). Brachinus sclopeta, Anchomenus dorsalis, and Chlaenius velutinus are three species with chemical defences and aposematic colours, and Pseudophonus rufipes, Steropus melas italicus, Amara anthobia, Poecilus cupreus, and Calathus fuscipes are five species without these features (Bonacci et al. 2006). Due to the prey’s movement pattern, Ocypus olens preferred to attack prey without chemical defences and with aposematic colour patterns (Bonacci et al. 2006). By having potent predator-deterrent toxins and night-time roosting, lubber grasshoppers Romalea microptera and Taeniopoda eques are able to avoid most vertebrate predation. Results demonstrate a correlation between chemical protection and large size in insects, with chemically defended species often being larger than non-defended grasshoppers. Insects may favour large size if vertebrate predation is no longer a dominant selective force (Whitman and Vincent 2008).

386

12

Offense and Defence Mechanism of Insect Predators

Cuticular hydrocarbons, also known as long-chain hydrocarbons, make up the majority of the cuticular chemical signatures in insects (CHCs). They play a crucial role in chemical communication and aid in reducing desiccation. In fact, they mediate interactions between prey and predator, plant and insect, courtship, dominance, species recognition, chemical mimicry, and fertility cues. Lucas et al. looked at the reactions of the termites Reticulitermes grassei and Reticulitermes flavipes to the presence of the cuticular chemicals made by a predator, the ant species Lasius niger (2018). The findings demonstrate that while the invasive Reticulitermes flavipes was unaffected by the predator odour, the native Reticulitermes grassei did become less aggressive. Both species’ mortality rates and caste ratios were unaffected (Lucas et al. 2018). The quantity of plant volatile compounds is directly related to the number of preys (VOCs). The amount of honeydew on all plants and the release of the aphid alarm pheromone (E)-b-farnesene both rose as aphid density increased. The amount of honeydew and (E)-b-farnesene were unaffected by plant resistance; hence, differences in predator consumption rates on the various plants were probably unaffected by these two herbivores’ cues. Plant resistance and plant resistance-byaphid density had no effect on the amount of overall VOCs or the content (blend of compounds) of the VOCs released by plants with greater aphid densities. The predator community was more diversified and plentiful on low-resistance plants, which increased prey intake (Kersch-Becker et al. (2017). In respect to chemicals, Kersch-Becker et al. (2017) proposed four different types of mechanisms in plant–herbivore–predator interactions. Here is a quick description of them. Plant resistance influences the variety and abundance of predators, which in turn indirectly controls herbivore populations (mechanism 1). Because plant defences can hinder feeding and have a negative influence on predator fitness and performance, more resistant plants typically house fewer predators. Prey suppression is frequently strengthened, for instance, by increasing predator variety. Because different herbivores are more or less attractive to predators (based on size or quality), plant resistance lowers rates of predator intake (mechanism 2). By affecting the predator’s capacity to locate prey, plant features can also reduce the pace at which predators consume prey (mechanism 3). By lowering allure cues like herbivore chemical cues (such alarm pheromones and excreta) and herbivore-induced plant volatiles, plant resistance lowers predator consumption rates (mechanism 4). In Plutella xylostella larvae that had their glucosinolate sulfatase genes silenced, toxic isothiocyanates accumulated throughout the body, hindering larval development and adult reproduction. Chrysoperla carnea, a predatory lacewing, however, rapidly metabolised ingested isothiocyanates through a general conjugation pathway, with no adverse effects on survival, reproduction, or even prey selection. These findings show how plant defences and detoxification have a significant impact on herbivore fitness but may only have a minor impact on a third trophic level (Sun et al. 2019). By securing cardenolides from Digitalis purpurea but not from Adonis vernalis and their natural adversary Chrysoperla carnea, Lygaeus equestris and Horvathiolus superbus were given protection (Pokharel et al. 2020). Findings also show that interactions between predators and their prey are very context-specific,

12.4

Chemical Defence

387

and that the choice of host plant can influence how well it protects against different predator types based on structural variations within the same class of chemical compounds (Pokharel et al. 2020). The Phyllotreta armoraciae horseradish flea beetle has a two-part chemical defence that consists of glucosinolates that are sequestered and an insect myrosinase that can change the non-toxic glucosinolates into deterring isothiocyanates. Here, we demonstrate that while insect myrosinase activity varied up to 43-fold throughout ontogenetic stages, the amounts of sequestered glucosinolates only varied by a factor of 2 across beetle ontogeny. In particular, myrosinase activity was 43.4 times higher in Phyllotreta armoraciae larvae than in pupae, despite the fact that glucosinolate levels were 1.5 times lower in the larvae. Only larvae emitted significant amounts of harmful isothiocyanates in response to attacks from the generalist predator Harmonia axyridis, which is consistent with the different levels of myrosinase activity in larvae and pupae. While pupae of Phyllotreta armoraciae were destroyed, larvae repelled the predator and survived one attack. The accumulation of glucosinolates in larvae and its subsequent interaction with Harmonia axyridis were affected by feeding Phyllotreta armoraciae larvae on plants that varied in their glucosinolate content and plant myrosinase activity. Compared to larvae with high glucosinolate levels, those with low levels of sequestered glucosinolates were far more vulnerable to predators. Our findings show that certain ontogenetic stages of Phyllotreta armoraciae are protected from predation by sequestered plant defence chemicals. The significant impact of plant defensive chemicals on sequestration suggests that predators were crucial in the development of this specialised herbivore’s host usage. Further research is needed on the various flea beetle life stages’ dietary preferences and defence mechanisms against ecologically relevant predator populations (Sporer et al. 2020). Compared to myb28myb29, the body weight of Chrysoperla carnea larvae-fed wild-type (glucosinolates)-reared aphids was dramatically reduced (transgenic) without aliphatic glucosinolates-reared Brevicoryne brassicae (Hemiptera: Aphididae); a statistical difference in predator weights appeared as early as five days after hatching and intensified during the remainder of our experimentation (17 days post-hatching). The male and female insects displayed the same pattern. After 20 days of feeding on aphids raised in the wild, the mortality rate of Chrysoperla carnea larvae reached about 70%; however, only 15% of Chrysoperla carnea fed on aphids raised in the myb28myb29 strain perished over the same time period. At the conclusion of the experiment, Chrysoperla carnea larvae fed with wild-type-reared aphids had much worse pupation success (less than 30%) than counterparts fed with myb28myb29-reared aphids (about 80%). (25 days post-hatching). Additionally, compared to lacewings fed with aphids raised on wild-type Arabidopsis thaliana, Chrysoperla carnea showed a shorter pupal stage length and a higher emergence success from myb28myb29reared aphids (Sun et al. 2021). Both the phytophagous Euschistus heros and its predator Podisus nigrispinus include the chemicals (E)-2-hexenal, hexenoic acid, (E)-2-decenal, tridecane, tetradecane, and pentadecane, though in varying levels in males and females. The scent glands of stink bugs (Pentatomidae) create volatile chemicals with foul scents

388

12

Offense and Defence Mechanism of Insect Predators

that serve as alarm and defence signals against natural enemies. The protective role of these chemicals is supported by the repellent and irritating effects of the smell gland discharge of Podisus nigrispinus and Euschistus heros on other species as well as between sexes of the same species. Pentatomidae scent gland compounds’ chemical identification may have an impact on insect behaviour and have negative impacts on other insects (Lima et al. 2021). Table 12.1 represents the plants and its chemicals which are responsible for prey–predator interaction.

12.4.1 Cardiotonic Steroids (CTS) as a Chemical Defence CTS are a broad class of substances generated from triterpenoids that are typically found in plants but can also be found in animals. They have the transmembrane protein Na+, K+ ATPase as their physiological target (NKA). Cardenolides and bufadienolides, the two primary families of CTS chemicals, have different steroid backbone and lactone group structures. The majority of cardenolides are made by plants and have a steroid backbone structure with a five-membered lactone group and a sugar moiety connected to the first carbon ring’s C-3 position. Every continent has prey organisms that contain CTS and the diversity and concentration of cardenolides varies among different prey species and individuals. Cardenolides and bufadienolides are the two classes of CTS. Both are created from scratch in both plants and animals, and a few species sequester CTS from their hosts or prey. It is almost clear that this sequester developed as a defence against 80 predators. Because they bind to the extracellular surface of the transmembrane protein Na+, K+ 81 ATPase (NKA) and prevent Na + and K+ 82 from crossing the membrane, CTS are poisonous. This leads to the disruption of electrochemical gradients, which leads to the dysregulation of 83 physiological systems. Although the NKA is substantially conserved 84 among animals, six 85 taxonomic groups of insects that are specialised on cardenolide-containing plants have independently evolved NKA insensitivity to cardenolides. In many situations, CTS intake causes predators to reject prey and learn to avoid them, which was decoded by Brower et al. (1982) after more than 40 years of research. The chemical and pharmacological underpinnings of the monarch butterfly, Danaus plexippus chemical’s defence were uncovered by Brower et al. (1982), who focused on the caterpillars’ consumption of milkweed plants (Asclepias) and sequestration of cardenolides. When Asclepias-fed monarchs were given to blue jays, Cyanocitta cristatata, the birds all responded by throwing up after consuming them. As a result, the birds stopped assaulting the monarchs in subsequent 94 encounters. Research on resistant predators was also pioneered by Brower et al. (1982), who presented the first proof of avian and rodent species that were resistant to the toxic effects of CTS. They were the first to postulate that physiological resistance arose in the progenitors of avian and rodent predators of monarchs, themes we explore in sections 4.2 and 5.2, 99, respectively, and that resistant predators had probably experienced alterations to their gustatory systems. But 30 years later, despite the identification of 101 expected resistance-granting

12.4

Chemical Defence

389

Table 12.1 The plants’ compounds which are involved in the interaction between prey and predators Plant Plantago lanceolata

Chemical Iridoid glycosides

Prey Junonia coenia

Predator Formica ptanipiles

Tomato

Chlorogenic acid

Manduca sexta

Podisus maculiventris

Crotalaria spp.

Pyrrolizidine alkaloids

Utetheisa ornatrix

Ceraeochrysa cubana

Romalea microptera and Taeniopoda eques

Vertebrate predators

Lubber

Bean, cowpea, tomato, cucumber, cabbage, pear, sweet pepper, grape, and potato Grape

Methyl salicylate

Methyl salicylate + (Z)-3-hexenyl acetate

Prey



Isothiocyanates

– Adonis vernalis

Digitalis purpurea

References Dyer and Deane Bowers (1996) Traugott and Stamp (1996) Eisner et al. (2000) Whitman and Vincent (2008)

Anthocoris nemoralis

Drukker et al. (2000)

James (2003a, b)

Plutella xylostella

Chrysopa nigricornis, Geocoris pallens, Stethorus punctum picipes, and Syrphidae Chrysoperla carnea

Cardenolides

Lygaeus equestris

Chrysoperla carnea

Cardenolides

Horvathiolus superbus

Chrysoperla carnea

(E)-2-hexenal, hexenoic acid, (E)2-decenal, tridecane, tetradecane, and pentadecane Aliphatic glucosinolates

Euschistus heros

Podisus nigrispinus

Pokharel et al. (2020) Pokharel et al. (2020) Lima et al. (2021)

Brevicoryne brassicae

Chrysoperla carnea

Sun et al. (2021)

Sun et al. (2019)

390

12

Offense and Defence Mechanism of Insect Predators

genetic changes, the functional significance of NKA in the CTS resistance of these bird and 100 rodent predators has not been explored (Mohammadi et al. 2022).

12.5

Cryptic Colouration

By avoiding areas where predators are more prevalent, prey can decrease their chances of being observed, being identified due to lack of movement and cryptic appearance, being subdued and devoured due to physical and chemical defences, and being identified due to mimicry or masquerade. Aposematic prey alerts predators who hunt by sight to their protection by utilising obtrusive warning colouring. Interestingly, birds’ decisions to attack chemically protected insect larvae were influenced more by the predators’ level of hunger than by the hue of the aposematic signal (Stevens et al. 2010). The mantis Tenodera aridifolia was shown to exhibit defensive reactions to approaching items. Three distinct behaviours—fixation, evasion, and cryptic reaction—were displayed by the mantis. The cryptic response involved either a quick extension of the forelegs forward or a rapid retraction of the forelegs under the prothorax. The mantis responded less frequently when its vision was blocked, which suggests that the visual cues produced by an approaching object were what caused the cryptic response. For objects that were on a direct collision trajectory as they approached, the response rate of the cryptic replies was highest. Reduced reaction was observed for horizontal deviance from the direct collision course. The cryptic reaction’s response rate grew as the object’s approaching velocity increased, and it reduced when the object ceased arriving at a farther distance from the mantis (Yamawaki 2011).

12.5.1 Aposematism Many species use warning, or aposematic, colour to alert potential predators to their unprofitability. Prey with aposematic colours is quite noticeable to the eye. Conspicuousness is strongly supported by empirical evidence as enhancing the aposematic signal’s potency. Aposematism is the blending of a visible colouring with a repulsive physical or chemical defence, such as a poison. Aposematism has been extensively studied in insects, including the seven-spot ladybird (Coccinella septempunctata), which produces the toxin coccinelline and uses black spots on a red background to advertise, and the common wasp (Vespula vulgaris), whose yellow-and-black stripes serve as a warning of its venomous sting. Aposematic prey rarely simply uses visual cues to indicate their unpalatability; they also use additional nonvisual cues such sounds, scents, and the release of bitter substances. These are believed to serve as “go-slow” cues that encourage predators to lower their assault rates on prey that is more likely to be guarded. According to this theory, it has been repeatedly discovered that naive predators’ bias against novel foods or foods with visual characteristics typically associated with aposematism, such as conspicuousness or

12.5

Cryptic Colouration

391

a red or yellow colour, such as red or yellow, increases when a novel sound, odour, or bitter-tasting compound is presented (Rowland et al. 2013). Prudic et al. 2007 used a mantid and its victim to explain it. Regardless of prey contrast treatment, adult Chinese praying mantids, Tenodera aridifolia sinensis (Mantidae), attacked all milkweed bugs, Oncopeltus fasciatus (Lygaeidae). Mantids oriented more quickly towards palatable bugs with high contrast (0.57 luminance contrast index) than palatable bugs with low contrast (0.19 luminance contrast index) during their initial encounters with milkweed bugs (Prudic et al. (2007). Shield bugs are known for releasing protective fluids when disturbed, and many species use warning colours to highlight this protection (Staddon et al. 1987; Evans et al. 1990; Wink et al. 2000). In order to determine whether there is evidence of selection for colour uniformity across life stages, our second goal is to quantify differences in colour techniques used throughout a person’s lifetime. In order to capitalise on predator learning processes, Medina et al. (2020) predicted that animals with warning colours will have similar colorations throughout their life cycles. In 134 species of shield bugs, Medina et al. (2020) measure colour differences between nymphs and adults and examine if colour similarity between nymphs and adults of the same species is more than would be predicted by chance. Additionally, we study how colour variations between life stages affect predator attack rates and learning procedures using naive predator tests with live shield bug prey (the cotton harlequin beetle) in both the aviary and the wild (Medina et al. 2020).

12.5.2 Green-Brown Polymorphism Orthopterans frequently exhibit green-brown polymorphism, while the variables that control body colours vary from species to species. There has been a wealth of knowledge gained regarding the regulation of colour polymorphism in acridoid grasshoppers: Some grasshoppers adapt their body colour to environmental parameters such temperature, humidity, and background colour when there is low density. Conocephalus maculatus, a katydid, exhibits green-brown colour polymorphism in its adult stage. Conocephalus maculatus adult green-brown polymorphism is partially influenced by genetic variables, and environmental factors like temperature, photoperiod, background colour, and humidity have no impact on the rate at which each morph develops (Oda and Ishii 1998). Conocephalus maculatus recently has shown colour polymorphism in the nymphal stage as well as in the adult stage, according to Oda and Ishii (2001). Although all hatchlings were green regardless of populations or conditions, the rate of green morph (G rate) dropped to 40–80% in the adult stage. Additionally, during one of the successive moults, 10% of animals that had previously gone brown returned to the green morph. Throughout the nymphal stage, the G rate of offspring produced by a green female and a green male tended to be higher than that produced by a brown female and a brown male, although environmental factors including temperature, humidity, and background colour had little to no impact on nymphal colour morph. The current findings demonstrate that Conocephalus maculatus nymphs and adults’ colour-morph determination is

392

12

Offense and Defence Mechanism of Insect Predators

mostly influenced by genetic variables and barely influenced by environmental ones (Oda and Ishii 2001).

12.6

Masquerade

Making sure predators mistake them for non-prey is an alternate form of camouflage. This is accomplished through morphological modifications that make them resemble inedible things found in the environment. The purpose of prey species that gain from disguising themselves as inedible objects is not to avoid detection; rather, it is to impede recognition systems that make the predator hold off on attacking the disguiser after being discovered. Predators that masquerade as inedible objects have been postulated as an adaptation strategy in a wide variety of species. The adult female Extatosoma tiaratum (about 12 cm long) hangs inverted among the vegetation, her procryptic abdomen folded over her back, and she is relatively thick, covered with cuticular spines (Bian et al. 2016). Because predatory insects mimic inedible substances, their prey may mistake them for such things. This would be regarded as an aggressive masquerade example and might boost hunting success. Again, this needs to be validated, but it raises intriguing issues of how the selection pressures put on aggressive masqueraders by their predators and prey interact to shape the evolution of their outward appearance (Skelhorn 2015).

12.7

Parental Care

One of the best examples of an altruistic attribute that evolved to improve the fitness of the care recipients (children) at the expense of the care donor is parental care (parents). In a variety of insect taxa, parental care has repeatedly evolved. It can occur during various stages of an offspring’s development, last anywhere from a few minutes to several years, and involve the mother, the father, or both parents. The remarkable variety of parental care shown by insects is described in depth in this chapter. We first discuss parental care practices that take place prior to oviposition, then we go on to forms that happen between oviposition and egg hatching, and we finish with post-hatching parental care practises that take place prior to and/or following nutritional independence. This section review highlights why insects are excellent biological models for enhancing our general comprehension of its evolution, diversification, and underlying physiological and genetic mechanisms by showing how they represent a perfect example of the variety of parental care that can be found in animals (Meunier et al. 2022).

12.7.1 Types There are two types of parental care: pre- and post-ovipositional. Pre-oviposition care comprises choosing an oviposition site, creating a nest, supplying a lot of food,

12.7

Parental Care

393

and altering the environment. It also includes providing the eggs with nutrition and chemical or structural defences. Post-ovipositional care comprises viviparity, progressive provisioning, egg and offspring attendance or transport, and care after nutritional independence. A form of post-ovipositional parental care has evolved in more than 160 families and 19 orders of insects. This type of parental care then allowed the emergence of sub sociality (simple parent-offspring associations), and finally very complex eusocial societies with overlapping generations, reproductive division of labour, and cooper brood care. In insects, parental care is quite uncommon and is thought to have developed in response to harsh circumstances when the benefits to offspring fitness are great. In predatory insect species, intra-specific predation can also play a significant role in selection. Parental care primarily has three costs: 1. Parents may have energetic costs due to fewer feeding chances while providing care or due to direct calorific expenditure while providing care, which may lower future fertility. 2. Care may also reduce parent survival (or “survival costs”), maybe by making caring individuals more visible to predators. 3. Third, giving care may limit opportunities for mating.

12.7.2 Reduviidae The genus Rhinoceros contains both species with uniparental male care and species with uniparental female care, making it a viable candidate for such a research. Using laboratory and field estimates of the costs and benefits of parental care for two sympatric species of Rhinoceros, the paternal-caring Rhynocoris tristis and the maternal-caring Rhynocoris carmelita, Manica and Johnstone (2004) parameterised a theoretical model and correctly predicted which sex should care. Locally, portions of the leguminous crop Stylosanthes are related with the presence of male caring Rhinoceros, notably Rhynocoris tristis, where the bugs are supposedly present in great densities (Thomas 1994). In Rhynocoris tristis, males guard eggs while also being polygynous (Manica and Johnstone 2004), mating with various females and taking batches of eggs from them. On the other hand, Rhynocoris carmelita exhibits female care, and males are never connected to the eggs. Insect predators and parasitic wasps are mostly to blame for egg mortality in both species. Guarding males’ filial cannibalism prevents Rhynocoris tristis from hatching more eggs (Gilbert and Manica 2009). Additionally, it was claimed that female Rhynocoris carmelita paid the full energetic cost of care while defending Rhynocoris tristis males engaged in filial cannibalism, eating eggs in order to maintain weight. In comparison to non-guarding male and female conspecifics, male Rhynocoris tristis that was guarded suffered survival costs. Rhynocoris tristis has a very high population density, and females favour males who are already guarding eggs (a preference previously observed in fish), which reduced the cost of promiscuity associated with

394

12

Offense and Defence Mechanism of Insect Predators

Fig. 12.1 Rhynocoris marginatus females also showed egg-guarding (a) and egg-carrying (b) behaviour

paternal care and contributed to the difference in care patterns between the two species (Gilbert et al. 2010). Female Pisilus tipuliformis kept watch over their egg batches until the young had hatched. When given the option to choose between this and other batches, they picked the ones they wanted based on how the eggs were arranged in each batch. Brooding Pisilus tipuliformis females still favoured their own batches after 1 day of connection with substitute batches, but after two or more days of interaction with substitute batches, they showed no discernible preference for either batch (Parker 1965). Atopozelus pallens (Heteroptera: Reduviidae) adults have been observed guarding their eggs and nymphs from predators on Pithecellobium dulce (Fabaceae) plants near Santiago de Cali, Colombia. This is a common habit for species of this genus (Tallamy et al. 2004). Adult predators of Atopozelus opsimus showed parental care by standing over their eggs and nymphs (Dias 2013; de Matos et al. 2019). Harpactorine instars frequently pretend to be dead. The harpactorine and ectrichodiine instars also roll their bodies into balls and lie immobile by withholding their cephalic and thoracic appendages. Reduviids with acutely curved rostrums (Peiratinae) have a better-developed nodding of the head and rubbing of the rostrum against the transversely striated prosternal groove to make a distinctive sound than reduviids with straight rostrums, which produce an obscure sound. Additionally, the rostrum extends, watery saliva is thrown out, and stinging takes place. In the Harpactorinae and Reduviinae, the dorsolateral abdominal scent gland, also known as Brindley’s gland, emits a volatile secretion with a distinctive odour (Ambrose and Kumar 2016). Female Rhynocoris marginatus demonstrated both egg-carrying and egg-guarding behaviour in Figs. 12.1a,b. The female Bactrodinae (Reduviidae) Bactrodes exhibit behaviours that we interpret as maternal egg guarding, according to a new investigation (Weirauch et al. 2021). Furthermore, they postulate that female egg guarding may be a reaction to the scarcity of sticky plants and a defence mechanism against cannibalism, rather than or in addition to reducing egg predation or egg parasitism (Weirauch et al. 2021).

12.7

Parental Care

395

12.7.3 Dermaptera According to Caussanel (1970), female Labidura riparia earwigs clean their eggs with their front tarsi. Other species of earwig females use their mouthparts to manipulate eggs, and this “licking” could be seen as cleaning activity (Lamb 1976). However, it is thought that adult egg-tenders use antibacterial secretions and mechanical spore removal to safeguard their young against microbial infection (Costa 2006). To protect their eggs from predators and parasites, true bugs (Hemiptera, Elasmucha dorsalis), earwigs, burying beetles, and dung beetles stretch their wings to grow larger. They then attack predators vigorously to drive them away from the region (Kudo et al. 1989) and eject them from their nests (Scott 1990). For instance, a recent experiment in the earwig Anisolabis maritima showed that females who attend eggs prevent conspecifics from eating the eggs (Miller et al. 2011). In a laboratory experiment, Butnariu et al. (2013) examined the maternal behaviour of the predator of the maize pest Spodoptera frugiperda, Doru lineare (Dermaptera: Forficulidae) (Lepidoptera: Noctuidae). The 29.7 eggs deposited by females needed 7–10 days to hatch. Mothers looked for a suitable nest to clean and protect the eggs and nymphs in. Females adopted eggs from other females because they did not perceive them as strange. Egg-cleaning, nest-cleaning, and a reduced viability of eggs not cared for by females were additional signs of maternal behaviour (Butnariu et al. 2013). Anisolabis maritima, an earwig, affects the microorganisms on the egg’s surface examining, the microbiomes of mothers and their eggs to see if any specific bacteria are consistently passed from female care to eggs. Using 16S rRNA bacterial DNA sequencing, microbiomes were examined, and the results showed that the richness and diversity of bacterial operational taxonomic units (OTUs) were considerably higher in female attended eggs than in unattended eggs. Adult females’ core microbiomes contained bacteria that may have anti-fungal properties; these bacteria were identified in higher concentrations and relative abundance on eggs in environments where females were permitted to take care of the eggs (Greer et al. 2020).

12.7.4 Hymenoptera Predatory wasp contacts with the thorn bug treehopper Umbonia crassicornis (Hemiptera: Membracidae) ranged from 0 to 94 (mean 27.0) (Cocroft 2002). The author also noted that the findings of his research suggested that not all kids benefit from maternal defence equally and simultaneously. With regard to the percentage of current offspring that profit from a parental investment, parental resources can generally be thought of as falling along a continuum, from giving a single offspring food to sounding an alarm that alerts the entire brood. Female Umbonia crassicornis defence falls somewhere in the middle of the two: while it does lower the overall success rate of wasp predators, at any given time, females are only able to successfully defend the offspring in their local surroundings. Therefore, this species’

396

12

Offense and Defence Mechanism of Insect Predators

maternal defence is a finite resource (Cocroft 2002). When exposed to diseases, workers of the ant Formica selysi enhance their egg grooming behaviour, and female earwigs (Dermaptera) engage in a distinctive egg grooming behaviour when caring for their clutch (Costa 2006).

12.7.5 Coleoptera There are two different types of parental care classified among ground beetles (Carabidae). The first type, also known as brutfürsorge, egg-watching, or egg attendance, denotes a stringent reliance on the parent for the survival of the early stages. However, there is only passive protection and no direct interaction between eggs or larvae and adults. The second sort of parental care, known as “offspring nursing” or “brutpflege”, entails the direct care of eggs and larvae, such as cleaning and turning the eggs or feeding the larvae with secretions. Currently, one of these two modes of parental care distinguishes over 30 species of ground beetles from 10 genera. The majority of species (with the exception of Carterus calydonius) are members of the Pterostichini tribe, in which egg attendance is the only form of parental care. Only in Carterus calydonius is offspring nursing documented. Only the female takes care of the young in every situation (Kolesnikov and Karamyan 2019).

12.8

Hibernation

Natural enemies benefit greatly from the resources provided by non-crop environments. Many natural enemies overwinter in non-crop habitats, where they may spring up and overrun arable areas. In order for natural enemies to keep up with the growth of insect populations, spring colonisation assures annual repopulation of the crop. Therefore, the availability of non-crop habitats may be essential for the effectiveness of biological conservation. The presence of hibernation sites close to crop fields enables effective crop colonisation during the early growing season, which may lead to effective pest control. Although lady beetles are considered to be an efficient group of aphid predators, they require uncultivated environments in the landscape for hibernating, such as hedgerows, woodland borders, and grass tussocks. Adult Coccinella transversoguttata spend their winters either alone in neighbouring mountain valleys or in rubbish and leaf debris (Cooke 1963). Numerous predators, including coccinellids, have been seen to hibernate according to the season. According to research done in Japan, the coccinellid Harmonia axyridis typically produces one generation in the spring and seldom two. Only one generation appeared in the autumn. In late November, adults started flying to their hibernation locations. Diapausing populations congregated under the shadows of white buildings, with numbers sharply rising after December (Sakurai et al. 1993). It might range from weeks to years. For instance, Coccinellidae longevity is influenced by prey

12.8

Hibernation

397

synchrony and dormancy mechanisms since species whose adults spend a lot of time dormant may live for a year or perhaps two. Exochomus quadripustulatus and Coccinella nigritus are both known to exhibit similar behaviour in certain regions of Pakistan and India (Nakumata 1984). The majority of multivoltine animals possess this capacity to adjust to changing environmental conditions (Hodek 1967). Three traits are typically linked to such behaviour: 1. Usually ephemeral rather than sessile, the host species. 2. Such organisms nearly always experience protracted periods of hibernation or diapause. 3. The aggregation site is where most mating takes place. In Nabide, Nabis rugosus, the final species, hibernates as adults (Roth and Reinhardt 2003), whereas Nabicula flavomarginata and Nabis major do so in the egg stage (Pericart 1987). The year substantially impacted Nabicula pseudoferus. The species remained in the alfalfa field until late October in 1993, possibly to hibernate there (Roth 2003). After hibernation, female stink bug egg production attracts noticeably more attention. The fertility and egg production of P. bioculatus females necessarily decline when the host, the L. decemlineata, is replaced with both different insect species and artificial nutritional medium (says Adams 2009). Earwigs have also been observed hibernating. Lower productivity may be explained by their vulnerability to tillage during hibernation and below-ground brood care. Abundances of Forficula auricularia in organically managed orchards in southern Sweden (Moerkens et al. 2012). Adult Coleomegilla maculatas hibernate in vast groups in the grass and at the roots of big, solitary trees covered in dead leaves. This microhabitat protects against temperature variations (Jean et al. 1990). Because of their early activity and wide dietary range, generalised predators, such as Carabidae, Staphylinidae, and Dermaptera, may play a significant role in pest suppression in the spring (Geiger et al. 2009). This behaviour was also displayed in a certain season of the year by the seven-spot lady beetle (Coccinella septempunctata) and the adonis’ lady beetle (Adonia variegate) (Coccinellidae) (Ahmadov and Hasanova 2016), spotted stink bugs in two. The Colorado potato beetle, Leptinotarsa decemlineata, is the most significant and commercially significant invasive species. Perillus bioculatus (Hemiptera: Pentatomidae), which exhibits hibernation behaviour, is a predator of this species (Elisovetcaia et al. 2020). Lipid reserves enable baseline metabolism throughout hibernation, as well as spring dispersal and regeneration of the female reproductive system. They also give the essential energy for autumn migration to hibernation sites (Hodek 1973). Animals accumulate large amounts of lipid reserves prior to hibernation, which vary depending on temperature. Hypothesising that carbohydrates may be crucial energy sources during hibernation when predator lipid levels are depleted and prey are equally lean, researchers studied the ground beetle Anchomenus dorsalis to see how macronutrient balance and lipid regeneration occurred during hibernation (Noreika et al. 2016).

398

12

Offense and Defence Mechanism of Insect Predators

Fedyay and Markina (2020) noted that 114 species—or 69.51% of the heteropterofauna—hibernate at the imago stage, with the majority of these species belonging to the families Lygaeidae (37 species), Miridae (14) and Pentatomidae (22 species), Nabidae, Tingidae, and Rhopalidae (seven species each). Winter occurs in 42 species at the egg stage, mostly in the Miridae family (34 species), but also in fewer Lygaeidae, Nabidae, and Rhopalidae species (two or three species each). Only four species (2.44%) in the Pentatomidae (three species) and Reduviidae (one species) families hibernate during the larval stage. At the imago and larvae, eggs, and larvae stages, two species from each group hibernate. According to Froeschner (2019), hibernation can occur in the egg, juvenile, or adult stages in his book on the Family Reduviidae. For instance, Zelus renardii adults exhibit this behaviour (Lozano and de Dios 2018; Lozano et al. 2018). When fed Galleria mellonella larvae, between 30% and 55% of Perillus bioculatus females did not begin egg-laying after emerging from hibernation. We blended the imago of the Leptinotarsa decemlineata (Coleoptera: Chrysomelidae) and the larvae of the Galleria mellonella for P. bioculatus after hibernations. This is significant because the imago Colorado potato beetle, which is well protected by “armour”, cannot be attacked by all adults following hibernation since the stink bugs are inhibited. As a result of mixed feeding, it was discovered that, firstly, P. bioculatus oviposition started two weeks earlier than it would have if Perillus had just eaten Galleria mellonella larvae. Second, the proportion of oviparous females in P. bioculatus dramatically increased, and as a result, the average number of eggs laid by one female and the population as a whole rose (Elisovetcaia et al. 2020). Authors also introduced several species of the Chrysomelidae family of leaf beetles to Perillus bioculatus laboratory cultures over the course of several years, including Entomoscelis suturalis, Entomoscelis adonidis, Chrysolina herbacea, and Chrysolina coerulans, to increase egg laying after hibernation and in the early spring breeding season.

12.9

Camouflaging

A powerful illustration of Darwin’s theory of evolution through natural selection is camouflage. In response to the selection pressure from predators, many types of camouflage have evolved in a variety of animals. The most well-known instances of visual camouflage involve creatures whose appearance shields them from a predator’s eyes. Such creatures are particularly well-known. This is often accomplished by matching an organism’s body colour and/or pattern to the background it is seen against, which keeps the creature unnoticed. The animal must choose a suitable background in addition to the morphological modifications required to evade detection for background matching to be successful. Additionally, improper body position might lessen the effectiveness of camouflage patterns. Many people believe that camouflage is an essential part of both predator and prey survival strategies. Two response variables—mean predator search time (ST) (63 studies) and predator attack rate (AR) of camouflaged prey—are compared while comparing various

12.9

Camouflaging

399

camouflage tactics and predator and prey kinds (28 studies). Overall, camouflage reduced the prey AR by 27.34% while increasing the predator ST by 62.56%. The concealment technique that boosted predator ST (295.43%) the greatest was masquerade. Disruptive coloration and background matching did not differ from one another. Motion camouflage reduces AR on prey but does not boost ST. No proof that eyespot improves ST and lowers AR by predators was discovered. Caterpillars were the kind of prey that most significantly impacted the extent of camouflage’s effect, despite the fact that the various predator types did not differ from one another. We also discuss potential inconsistencies or redundancies among techniques, predator and prey types, and possible evolutionary factors that may have caused camouflage to become a highly successful anti-predator adaption (says de Alcantara Viana et al. 2022).

12.9.1 Mantids Crypsis is the process in which mantids blend into their surroundings. Some species do this by mimicking dead leaves, grass, tree bark or flowers (Green 2014). Pseudocreobotra wahlbergi (Mantodea: Hymenopodidae), (C): Ghost mantid resembling dead leaves, Phyllocrania paradoxa (Mantodea: Hymenopodidae) and a mantid usually found in grasslands (D) Galepsus sp. (Mantodea: Tarachodidae). The Orchid mantis (Hymenopus coronatus (Mantodea: Hymenopodidae) resembles the Orchid flower (Phalaenopsis amabilis (Asparagales: Orchidaceae). However, in an experiment, the Orchid mantis was compared to a common Asystasia intrusa (Scrophulariales: Acanthaceae) flower and the results indicated that the mantis attracted more hymenopteran pollinators than the flowers themselves (O’Hanlon et al. 2014). According to O’Hanlon et al. (2014), orchid mantids were the first mantid species that used mimicry as a hunting strategy. Mimicry is usually used as a defence mechanism to avoid predation (McMonigle 2013). Not only do mantids resemble their backgrounds, but some species also have elongated bodies, position themselves to imitate the position of flowers, grasses, and twigs (O’Hanlon et al. 2014). The wings of mantid species that make use of camouflage usually aids in mimicry and often have bright patterns (McMonigle 2013). Mantids can change colour to match their surroundings; however, this is not as pronounced as in some reptile species, i.e., chameleons (Green 2014). Mantids can only change to different shades of a colour which enables them to resemble their immediate environment more accurately. However, this does not only indicate that they can change colour but also that mantids are aware of colour (Green 2014). Observations have even been made of mantids repositioning themselves to increase the level of camouflage which increases the level of disguise that they achieve (Green 2014). Raptorial forelegs are prey-capture structures that have evolved repeatedly among a variety of insect lineages. Presently, raptorial forelegs occur in predatory insects such as mantid lacewings (Neuroptera), mantises (Mantodea), water scorpions (Hemiptera) and dance flies (Diptera). Raptorial forelegs were also present during

400

12

Offense and Defence Mechanism of Insect Predators

Fig. 12.2 Acanthaspis pedestris fourth instar nymph with natural camouflaging

the Middle and Late Triassic in the extinct group Titanoptera, a predaceous group of large orthopteroid insects.

12.9.2 Reduviids In reduviids, camouflaging is very common among the nymphal stages of certain genera of the subfamilies Reduviinae, Cetherinae, Salyavatinae, Stenopodainae, Triatominae, and Sphaeridopinae (Fig. 12.2a). However, for the first time, (Rédei 2012) recorded camouflaging in adult P. impenetrabilis aults. Immature stages of Reduvius personatus and some other Reduviidae are known to camouflage themselves with a range of materials found in their environment. Even though this behaviour has been observed in several species, camouflaging structures have never been studied in a comparative way. This study documents for the first time the structure that is involved in the application of camouflaging material, i.e., the hind tarsal fan, and reveals structures that assure the fastening of the camouflaging material, i.e., anchor setae and trichomes, in eight species representing five subfamilies of Reduviidae. Whereas anchor setae assure the attachment of camouflaging material by their mechanical properties, short-projection trichomes, long-projection trichomes, and grouped trichomes are here proposed to secrete a sticky substance for this purpose. Primary homology hypotheses on the three types of trichomes are proposed. At least in some species, short-projection trichomes appear to be responsible for the fastening of the camouflaging layer close to the integument, whereas long-projection trichomes may hold the outer layer of camouflaging material in place (Weirauch 2006). Salyavata variegata nymphs capture termite workers (Nasutitermes corniger) at repair sites after attracting them into vulnerable positions with the empty carcasses of previous prey (McMahan 1982). The nymphs of the West African assassin bugs Paredocla and Acanthaspis spp. disguise themselves with a cover of dust, sand and soil particles (“dust coat”) and additionally pile a “backpack” of larger objects, such as empty prey corpses and plant parts, on their abdomen (Brandt and Mahsberg 2002). Acanthaspis cincticrus is an assassin bug with a specialised camouflaging

12.9

Camouflaging

401

behaviour to ambush ants in the nymphal stages (Kou et al. 2017). Louis (1974) noted that Authenta looked to be entirely arboreal on cocoa farms in Ghana and further speculated that Authenta fulvipennis’ dark brown colouring helped conceal it among the canopy’s leaves and twigs (Swanson 2018). To know the importance of camouflaging, Ambrose and Sahayaraj (1991) conducted an experiment with decamouflaging and crowing using Acanthaspis pedestris. The effect of crowding is observed in 3 categories, such as, (a) one individual, (b) two individuals and (c) four individuals under two different conditions (a) camouflaged and (b) decamouflaged. The findings are highlighted here: Acanthaspis pedestris exhibits both natural and corpse camouflaging. The camouflaging efficiency is greater in the I instar and gradually diminishes in the older instars and it totally disappears in the adults. Crowding causes a decrease in the body weight and the particle-carrying capacity. Both crowding and decamouflaging affects the ethology of camouflaging and the prey capturing time. Increased nymphal mortality and decreased stadial period are observed as functions of crowding and decamouflaging. Decamouflaging increases the adult longevity of Acanthaspis pedestris in one individual male and two individual females. Crowding causes an increase in the adult longevity of male in the 2 individuals’ group and causes a decrease in the 4 individuals’ group of decamouflaged individuals. Sex ratio of camouflaged individuals is biased. But in the decamouflaged one individual group the sex ratio is unbiased. The male biased sex ratio of the normal individuals becomes female biased as a function of crowding. Adaptive features: The anchor setae mechanically hold the material in place while the trichomes, which are made up of glandular units and hair-like projections, fix the particles with the sticky secretion of the glands. These two different types of structures facilitate attachment of the camouflaging particles to the body surface in reduviid larvae. Weirauch (2006) provided a thorough description of these features in numerous species from diverse subfamilies. After thoroughly macerating the larva of Porcelloderes impenetrabilis, the integument of the abdomen was studied under a microscope, and neither surface revealed any glandular units. However, it is highly likely that the soil particles on the body were glued together by a sticky secretion of unknown origin given how tightly the soil particles were adhered to one another and how challenging it was to remove them mechanically, even after the specimen was soaked in water, alcohol, moisturisers, or detergents. Both the dorsal and ventral surfaces of the abdomen are adorned with significantly raised carinae in Porcelloderes impenetrabilis adults. The posteromedian angles of the epipleurites II–VI, the posterior margins of all visible abdominal sternites, the anterior, lateral, and posterior margins of synergised I–III, the fusion lines between mediotergites I–II and II–III, the posterior and lateral margins of mediotergites IV–VI, the lateral margins of mediotergite VII, and the lateral margins of the mediotergite IV On the head, pronotum, and abdomen, areas enclosed by setigerous tubercles are smooth. In these deep, smooth patches surrounded by tubercles, the camouflaging material (soil) is undoubtedly deposited and mechanically fixed by the high tubercles themselves as well as by their serrate apical setae. The mud coating on larvae is thinner and mostly concentrated on the head and pronotum because the abdomen of the larva

402

12

Offense and Defence Mechanism of Insect Predators

is poorly sclerotised and lacks such prominent carinae. Species that engage in camouflaging behaviour in some reduviid genera have a fan of long setae. It is in charge of accumulating and loading camouflaging material onto the body and is located near the tip of the apical tarsal segment of the hind leg. Other genera lack a tarsal fan like this (Weirauch 2006). None of the adults or larvae of Porcelloderes impenetrabilis sp. n. have any tarsal application structures like this. It is yet unknown how this species applies the camouflaging substance on its body (Rédei 2012). Trichomes in the nymphal stages of various reduviid species for which camouflaging behaviour has previously been observed (Ambrose and Sahayaraj 1990; Weirauch 2006). Bugaj-Nawrocka et al. (2022) reported that they haven’t looked into whether these structures can act as camouflage, though, by secreting a sticky substance. Additionally, they did not see Psyttala horrida nymphs utilising the substrate as a cover in our lab settings.

12.9.3 Pentatomidae Supputius cincticeps (Asopinae), a generalist predator, consuming Cladomorphus phyllinus (Phasmatodea), a species that exhibits an unusual phenotypic known as the “walking stick” body form as camouflage. We make a radical suggestion: the camouflage may have developed as a result of the introduction of a plant morph gene via an insect vector to the insect genome (Costa 2019).

12.10 Recommendations for Future Works The compounds have the potential to be an innovative and sustainable source of insecticides for the management of agricultural pests. Exact role of various predators’ hibernating behaviour should be identified. Very few works on male or female or both genders prenatal care need to be done. Various anatomical features should be studied using electron microscopes. Plant defences frequently mediate whether competing chewing and sucking herbivores indirectly benefit or harm one another.

References Adams TS (2009) Ann Entomol Soc Am 93:529–535 Ahmadov BA, Hasanova SS (2016) The role of coccinellids (Coccinellidae) in suppression of the number of Aphis craccivora Koch, 1854, and Aphis gossypii glover, 1877. J Entomol Zool Stud 4:234–237 Ambrose DP, Kumar AG (2016) Reduviid predators. Ecofriendly pest management for food security. Academic, Cambridge, MA, pp 217–257 Ambrose DP, Sahayaraj K (1990) Effect of space on the post embryonic development and predatory behaviour of Ectomocoris tibialis Distant (Heteroptera: Reduviidae). Uttar Pradesh J. Zool. 10(2):162–169

References

403

Ambrose DP, Sahayaraj K (1991) Camouflaging behaviour and the effect of decamouflaging on behaviour and development of Acanthaspis pedestris Stal in solitary and crowded conditions (Insecta, Heteroptera, Reduviidae). Mitteilungen aus dem museum für naturkunde in berlin. Zoologisches museum und institut für apezielle Zoologie, Berlin 67(2):325–337 Bian X, Elgar MA, Peters RA (2016) The swaying behavior of Extatosoma tiaratum: motion camouflage in a stick insect? Behav Ecol 27(1):83–92 Bonacci T, Massolo A, Brandmayr P, Brandmayr TZ (2006) Predatory behaviour on ground beetles (coleoptera: carabidae) by Ocypus olens (Müller) (coleoptera: Staphylinidae) under laboratory conditions. Entomol News 117(5):545–551 Brandt M, Mahsberg D (2002) Bugs with a backpack: the function of nymphal camouflage in the west African assassin bugs Paredocla and Acanthaspis spp. Anim Behav 63(2):277–284 Brower LP, Seiber JN, Nelson CJ, Lynch SP, Tuskes PM (1982) Plant-determined variation in the cardenolide content, thin-layer chromatography profiles, and emetic potency of monarch butterflies, Danaus plexippus reared on the milkweed, Asclepias eriocarpa in California. J Chem Ecol 8(3):579–633 Bugaj-Nawrocka A, Danielczyk A, Sułkowska I, Chłond D (2022) Psyttala horrida (Stal, 1865) (Hemiptera: Reduviidae: Reduviinae)—a morphological study of eggs and nymphs. Insects 13(11):1014 Büsse S, Hörnschemeyer T, Gorb SN (2017) The head morphology of Pyrrhosoma nymphula larvae (Odonata: Zygoptera) focusing on functional aspects of the mouthparts. Front Zool 14 (1):1–13 Butnariu AR, Pasini A, Reis FS, Bessa E (2013) Maternal care by the earwig Doru lineare Eschs. (Dermaptera: Forficulidae). J Insect Behav 26(5):667–678 Caussanel C (1970) Principales exigencies écophysiologiques du forfi cule des sables, Labidura riparia (Dermaptera, Labiduridae). Ann Soc Entomol Fr 6:589–612 Cocroft RB (2002) Antipredator defense as a limited resource: unequal predation risk in broods of an insect with maternal care. Behav Ecol 13(1):125–133 Cooke WE (1963) Ecology of the pea aphid in the blue mountain area of eastern Washington and Oregon. U.S. Department of Agricurtural Technology, Washington, DC, p 1287 Costa JT (2006) The other insect societes. Harvard University Press, Cambridge, MA, p 812 Costa J (2019) First report of predation by a stink bug (Supputius cincticeps Stal) on a walking-stick insect (Cladomorphus phyllinus gray), with reflections on evolutionary mechanisms for camouflage. Acta Bioló Parana 48:1–2 de Alcantara Viana JV, Vieira C, Duarte RC, Romero GQ (2022) Predator responses to prey camouflage strategies: a meta-analysis. Proc R Soc B 289(1982):20220980 de Matos MF, Vinicius de Abreu DA, Lemes PG, Zanuncio AJV, Zanuncio JC (2019) Atopozelus opsimus (Hemiptera: Reduviidae) preying on Mastigimas anjosi (Hemiptera: Calophyidae), a pest of tropical cedar, Cedrela fissilis (Meliaceae). Fla Entomol 102(2):447–450 Dias TKR (2013) Atopozelus opsimus (Hemiptera: Reduviidae): presasalternativas, comportamento parental e predação sobre pragas exóticas. xiv, 101 f. Tese de doutorado: Universidade Estadual Paulista, Júlio de Mesquita Filho, Faculdade de Ciências Agronômicas de Botucatu, Botucatu, São Paulo, Brazil Drukker B, Bruin J, Sabelis MW (2000) Anthocorid predators learn to associate herbivoreinduced volatiles with presence or absence of prey. Physiol Entomol 25:260–265 Dyer LA, Deane Bowers M (1996) The importance of sequestered iridoid glycosides as a defense against an ant predator. J Chem Ecol 22(8):1527–1539 Eigenbrode SD (2004) The effects of plant epicuticular waxy blooms on attachment and effectiveness of predatory insects. Arthropod Struct Dev 33(1):91–102 Eisner T, Eisner M, Rossini C, Iyengar VK, Roach BL, Benedikt E, Meinwald J (2000) Chemical defense against predation in an insect egg. Proc Natl Acad Sci 97(4):1634–1639 Elisovetcaia D, Derjanschi V, Agaseva I, Nefedova M (2020) Some results of breeding a predatory stink bug of Perillus bioculatus F. (Hemiptera, Pentatomidae) in the Republic of Moldova. In: BIO web of conferences, 21:00024. EDP Sciences, Les Ulis

404

12

Offense and Defence Mechanism of Insect Predators

Evans C, Staddon B, Games D (1990) Analysis of gland secretions of Pentatomoidea (Heteroptera) by gas chromatography-mass spectrometry techniques. In: Chromatography and isolation of insect hormones and pheromones. Springer, Berlin, pp 321–328 Fedyay IO, Markina TY (2020) Some features of seasonal cycles of semi-hard-winged (Heteroptera) urbocenoses of the city of Kharkov. Biodiver Ecol Experiment Biol 21:104–109 Froeschner RC (2019) Family Reduviidae Latreille, 1807: the assassin bugs. In catalog of the Heteroptera, or true bugs, of Canada and the continental United States. CRC, Boca Raton, FL, pp 616–651 Geiger F, Wäckers FL, Bianchi FJ (2009) Hibernation of predatory arthropods in semi-natural habitats. BioControl 54(4):529–535 Gilbert J, Manica A (2009) Brood conspicuousness and clutch viability in male-caring assassin bugs (Rhinocoris tristis). Ecol Entomol 34:176–182 Gilbert JD, Thomas LK, Manica A (2010) Quantifying the benefits and costs of parental care in assassin bugs. Ecol Entomol 35(5):639–651 Green T (2014) Praying mantis: ultimate care guide. BLEP Publishing, Greenville, p 132 Greer JA, Swei A, Vredenburg VT, Zink AG (2020) Parental care alters the egg microbiome of maritime earwigs. Microb Ecol 80(4):920–934 Hodek I (1967) Bionomics and ecology of predaceous Coccinellidae. Annu Rev Entomol 12:79– 104 Hodek I (1973) Biology of coccinellidae. Academia, Prague Ito K, Ioku Y (2022) A novel type of counterattack against predatory thrips in Schizotetranychus brevisetosus (Acari: Tetranychidae). Int J Acarol 48:1–7 James DG (2003a) Field evaluation of herbivore-induced plant volatiles as attractants for beneficial insects: methyl salicylate and the green lacewing, Chrysopa nigricornis. J Chem Ecol 29:1601– 1609 James DG (2003b) Synthetic herbivore-induced plant volatiles as field attractants for beneficial insects. Environ Entomol 32:977–982 Jean C, Coderre D, Tourneur JC (1990) Effects of temperature and substrate on survival and lipid consumption of hibernating Coleomegilla maculata lengi (coleoptera: Coccinellidae). Environ Entomol 19(6):1657–1662 Karban R, LoPresti E, Pepi A, Grof-Tisza P (2019) Induction of the sticky plant defense syndrome in wild tobacco. Ecology 100:e02746 Kersch-Becker MF, Kessler A, Thaler JS (2017) Plant defenses limit herbivore population growth by changing predator–prey interactions. Proc R Soc B Biol Sci 284(1862):20171120 Kolesnikov FN, Karamyan AN (2019) Parental care and offspring survival in Pterostichus anthracinus (coleoptera: Carabidae): an experimental study. Eur J Entomol 116:33–41 Kou F, Li H, Li S, Xun H, Zhang Y, Sun Z, Cai W (2017) Temporal transcriptomic profiling of the ant-feeding assassin bug Acanthaspis cincticrus reveals a biased expression of genes associated with predation in nymphs. Sci Rep 7(1):1–11 Kudo S, Sato M, Ohara M (1989) Prolonged maternal care in Elasmucha dorsalis (Heteroptera: Acanthosomatidae). J Ethol 7:75–81 Lamb RJ (1976) Parental behavior in the dermaptera with special reference to Forficula auricularia (Dermaptera: Forficulidae) 1. Can Entomol 108(6):609–619 Lambert AM (2007) Effects of prey availability, facultative plant feeding, and plant defenses on a generalist insect predator. Arthropod Plant Interact 1(3):167–173 Lamon B, Topoff H (1981) Avoiding predation by army ants: defensive behaviours of three ant species of the genus Camponotus. Anim Behav 29(4):1070–1081 Lima BSA, Martínez LC, Plata-Rueda A (2021) Interaction between predatory and phytophagous stink bugs (Heteroptera: Pentatomidae) promoted by secretion of scent glands. Chemoecology 31:209–219. https://doi.org/10.1007/s00049-021-00341-9 Lozano BR, de Dios MAG (2018) The invasive species Zelus renardii (Kolenati, 1857) (Hemiptera, Reduviidae) in Spain and comments about its global expansion. Trans Am Entomol Soc 144 (3):551–558

References

405

Louis D (1974) Biology of Reduviidae of cocoa farms in Ghana. Am Midl Nat 91(1):68–69 Lozano BR, Ruiz MB, de Dios MAG (2018) The invasive species Zelus renardii (Kolenati, 1857) (Hemiptera, reduviidae) in Spain and comments about its global expansion. Trans Am Entomol Soc 144:551–558 Lucas E, Coderre D, Brodeur J (1997) Instar-specific defense of Coleomegilla maculata lengi (Col.: Coccinellidae): influence on attack success of the intraguild predator Chrysoperla rufilabris (Neur.: Chrysopidae). Entomophaga 42(1):3–12 Lucas C, Brossette L, Lefloch L, Dupont S, Christides JP, Bagnères AG (2018) When predator odour makes groups stronger: effects on behavioural and chemical adaptations in two termite species. Ecol Entomol 43(4):513–524 Manica A, Johnstone R (2004) The evolution of male care with overlapping broods. Am Nat 164: 517–530 McMahan EA (1982) Bait-and-capture strategy of a termite-eating assassin bug. Ins Soc 29:346– 351. https://doi.org/10.1007/BF02228761 McMonigle O (2013) Keeping the praying mantis. Coachwhip Publications, Greenville, p 202 Medina I, Vega-Trejo R, Wallenius T, Esquerre D, Leon C, Perez DM, Head ML (2020) No link between nymph and adult coloration in shield bugs: weak selection by predators. Proc R Soc B 287(1929):20201011 Merivee E, Märtmann H, Must A, Milius M, Williams I, Mänd M (2008) Electrophysiological responses from neurons of antennal taste sensilla in the polyphagous predatory ground beetle Pterostichus oblongopunctatus (Fabricius 1787) to plant sugars and amino acids. J Insect Physiol 54(8):1213–1219 Meunier J, Körner M, Kramer J (2022) Parental care. In: Reproductive strategies in insects. Routledge, Abingdon, pp 337–357 Michaud JP, Grant AK (2003) Intraguild predation among ladybeetles and a green lacewing: do the larval spines of Curinus coeruleus (Coleoptera: Coccinellidae) serve a defensive function? Bull Entomol Res 93(6):499–505 Miller JS, Rudolph L, Zink AG (2011) Maternal nest defense reduces egg cannibalism by conspecific females in the maritime earwig Anisolabis maritima. Behav Ecol Sociobiol 65:1873–1879 Moerkens R, Leirs H, Peusens G, Beliën T, Gobin B (2012) Natural and human causes of earwig mortality during winter: temperature, parasitoids, and soil tillage. J Appl Entomol 136:490–500. https://doi.org/10.1111/j.1439-0418.2011.01676.x Mohammadi S, Yang L, Bulbert M, Rowland HM (2022) Defence mitigation by predators of chemically defended prey integrated over the predation cycle and across biological levels Nakumata K (1984) Visual orientation of a ladybeetle, Coccinella septempunctata L. (Coleoptera; Coccinellidae), towards its prey. Appl Entomol Zool 19:82–86 Noreika N, Madsen NE, Jensen K, Toft S (2016) Balancing of lipid, protein, and carbohydrate intake in a predatory beetle following hibernation, and consequences for lipid restoration. J Insect Physiol 88:1–9 O’Hanlon JC, Holwell GI, Herberstein ME (2014) Pollinator deception in the orchid mantis. Am Nat 183:126–132 Oda K, Ishii M (1998) Factors affecting adult color polymorphism in the meadow grasshopper, Conocephalus maculatus (Orthoptera: Tettigoniidae). Appl Entomol Zool 33:455–460 Oda KI, Ishii M (2001) Body color polymorphism in nymphs and adults of a katydid, Conocephalus maculatus (orthoptera: Tettigoniidae). Appl Entomol Zool 36(3):345–348 Parker AH (1965) The maternal behaviour of Pisilus tipuliformis Fabricius (Hemiptera: Reduviidae). Entomol Exp Appl 8(1):13–19 Pericart J (1987) Hemiptres Nabidae d Europe occidentale et du Maghreb. Faune de France 71:1– 185 Petie R, Muller M (2007) Curvature facilitates prey fixation in predatory insect claws. J Theor Biol 244(4):565–575. https://doi.org/10.1016/j.jtbi.2006.09.004

406

12

Offense and Defence Mechanism of Insect Predators

Pokharel P, Sippel M, Vilcinskas A, Petschenka G (2020) Defense of milkweed bugs (Heteroptera: Lygaeinae) against predatory lacewing larvae depends on structural differences of sequestered cardenolides. Insects 11(8):485 Prudic KL, Skemp AK, Papaj DR (2007) Aposematic coloration, luminance contrast, and the benefits of conspicuousness. Behav Ecol 18(1):41–46 Rédei D (2012) Porcelloderes impenetrabilis gen. & sp. n. from Tanzania, an assassin bug camouflaging in the adult stage (Hemiptera: Heteroptera: Reduviidae: Physoderinae). Afr Invertebr 53(2):601–613 Roth S (2003) Spatial and temporal pattern of colonization of Nabidae (Heteroptera) in alfalfa (Medicago sativa). J Appl Entomol 127(4):221–227 Roth S, Reinhardt K (2003) Facultative sperm storage in response to nutritional status in a female insect. Proc R Soc B: Biol Sci 270(suppl_1):S54–S56 Rowland HM, Ruxton GD, Skelhorn J (2013) Bitter taste enhances predatory biases against aggregations of prey with warning coloration. Behav Ecol 24(4):942–948 Sakurai H, Kumada Y, Takeda S (1993) Seasonal prevalence and hibernating diapause behaviour in the lady beetle, Harmonia axyridis. Res Bull Fac Agric Gifu Univ 58:51–55 Scott MP (1990) Brood guarding and the evolution of male parental care in burying beetles. Behav Ecol Sociobiol 26:31–39 Skelhorn J (2015) Masquerade. Curr Biol 25(15):R643–R644 Sporer T, Körnig J, Beran F (2020) Ontogenetic differences in the chemical defence of flea beetles influence their predation risk. Funct Ecol 34(7):1370–1379 Staddon B, Thorne M, Knight D (1987) The scent glands and their chemicals in the aposematic cotton harlequin bug, Tectocoris diophthalmus (Heteroptera, Scutelleridae). Aust J Zool 35:227–234. https://doi.org/10.1071/ZO9870227 Stevens M, Mappes J, Sandre SL (2010) The effect of predator appetite, prey warning coloration and luminance on predator foraging decisions. Behaviour 147(9):1121–1143 Sun R, Jiang X, Reichelt M, Gershenzon J, Pandit SS, Vassão DG (2019) Tritrophic metabolism of plant chemical defenses and its effects on herbivore and predator performance. elife 8:e51029 Sun R, Jiang X, Reichelt M, Gershenzon J, Vassao DG (2021) The selective sequestration of glucosinolates by the cabbage aphid severely impacts a predatory lacewing. Pest Manag Sci 94(4):1147–1160 Swanson DR (2018) A new species of African assassin bug in the genus Authenta (Heteroptera: Reduviidae: Harpactorinae). Proc Entomol Soc Wash 120(4):903–912 Tallamy DW, Walsh E, Peck DC (2004) Revisiting paternal care in the assassin bug, Atopozelus pallens (Heteroptera: Reduviidae). J Insect Behav 17:431–436 Thomas LK (1994) The evolution of parental care in assassin bugs. University of Cambridge, Cambridge Traugott MS, Stamp NE (1996) Effects of chlorogenic acid-and tomatine-fed caterpillars on the behavior of an insect predator. J Insect Behav 9(3):461–476 Van Wyk J, Krimmel BA, Pearse IS (2019) Does entrapped pollen feed predatory arthropods as indirect resistance to herbivores? Ecology 100:e02867 Wang Y, Brożek J, Dai W (2020) Comparative morphology of the mouthparts in three predatory stink bugs (Heteroptera: Asopinae) reveals feeding specialization of stylets and sensilla. Insects 11(11):762 Weirauch C (2006) Anatomy of disguise: camouflaging structures in nymphs of some reduviidae (Heteroptera). Am Mus Novit 3542:1–18 Weirauch C, Zhang G, Forero D, Bérenger JM (2021) Living on a sticky trap: natural history and morphology of Bactrodes assassin bugs (Insecta: Hemiptera: Reduviidae: Bactrodinae). J Nat Hist 55(5–6):341–363 Weseloh RM, Hare JD (2009) Predation/predatory insects. In: Encyclopedia of insects. Academic, Cambridge, MA, pp 837–839 Wheeler AG Jr, Krimmel BA (2015) Mirid (Hemiptera: Heteroptera) specialists of sticky plants: adaptations, interactions, and ecological implications. Annu Rev Entomol 60:393–414

References

407

Whitman DW, Vincent S (2008) Large size as an antipredator defense in an insect. J Orthoptera Res 17:353–371 Wink M, Grimm C, Koschmieder C, Sporer F, Bergeot O (2000) Sequestration of phorbolesters by the aposematically coloured bug Pachycoris klugii (Heteroptera: Scutelleridae) feeding on Jatropha curcas (Euphorbiaceae). Chemoecology 10:179–184. https://doi.org/10.1007/ PL00001820 Yamawaki Y (2011) Defence behaviours of the praying mantis Tenodera aridifolia in response to looming objects. J Insect Physiol 57(11):1510–1517 Yamawaki Y (2017) Decision-making and motor control in predatory insects: a review of the praying mantis. Ecol Entomol 42(s1):39–50

Venomous and Other Body Fluids in Insect Predators

13

Contents 13.1 13.2 13.3

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Salivary Gland, Venomous Saliva Collection, and Quantification . . . . . . . . . . . . . . . . . . . . 13.3.1 Salivary Gland . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3.2 Venomous Saliva Collection Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3.3 Saliva Genders on Venom Quantity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3.4 Impact of Prey Deprivation on Venom Quantity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.3.5 Influence of Prey on VS Yield . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.4 Venom Chemical Composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.4.1 Proteinaceous Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.4.2 Enzymes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.4.3 Genes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.5 Non-Proteinaceous Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6 Salivary Venom of Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.1 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.2 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.3 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.4 Diptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.5 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.6 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.6.7 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.7 Other Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.8 Biological Activities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.8.1 Pesticidal Activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.8.2 Primary Metabolites Modulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.8.3 Immunomodulatory Activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.8.4 Inhibition of Haemocyte Aggregation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.8.5 Spreading Inhibitory Behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.8.6 Anti-Microbial and Cytotoxic Activities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.9 Physiological Role . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.10 Future Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

410 410 411 411 415 416 417 418 419 419 421 424 425 426 426 428 429 431 432 432 433 434 434 436 437 438 438 439 439 440 440 440

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_13

409

410

13.1

13

Venomous and Other Body Fluids in Insect Predators

Introduction

The ability of predatory insects to prey on other insect pests by injecting harmful salivary gland contents has been demonstrated in nymphs, adults, and/or any stage of development. These secretions, which frequently come from strange origins, seem to enhance the effectiveness of the better biological control agent. When administered to invertebrates, heteropteran venoms are known to result in paralysis, tissue liquefaction, and death; when administered to vertebrates, they can result in pain, occasionally neurotoxic effects, and/or death. It is also well known that the venom of assassin bugs has a potent cytolytic effect on both insect and mammalian cells. Many of these substances are novel products with a variety of invertebrate preys that they are harmful to. Various lineages use mouthpart-associated venoms to paralyse and pre-digest food while hunting, including the True Bugs (Heteroptera), robber flies (Asilidae), and larvae of many Neuroptera, Coleoptera, and Diptera.

13.2

Chemistry

Various separate toxins (including primary and secondary components) or salivary secretions with a variety of distinct and frequently powerful physiological activities typically make up insect venoms. Venom is simply one type of chemical protection (or offence), but it is important to think about the other forms it can take because many of them serve the same purpose from a predator’s ecologically relevant point of view. Valid and generally acknowledged definition is: 1. The compounds in the venom, known as “toxins”, interfere with another organism’s physiological or biochemical processes (Fry et al. 2009). 2. The venom is created or stored in a gland; it has a specific delivery system (piercing and sucking mouthparts, or rostrum), which is utilised to convey the venom to another organism via an injury; and it serves as a weapon of predation or defence (Nelsen et al. 2014). 3. Venom is a biological substance produced by an organism that contains molecules (or “toxins”) that interfere with the physiological or biochemical processes of another organism. These molecules, which have evolved in the venomous organism to benefit it once they are introduced to the other organism, interfere with the other organism’s physiological or biochemical processes. A specific delivery system actively transfers the venom from one organism to another through an injury once it is produced, stored, or both in a specialised structure (Arbuckle 2017). There are two basic classifications of symptoms that many predatory varieties of venom often generate in prey:

13.3

Salivary Gland, Venomous Saliva Collection, and Quantification

411

1. changing haemolymph consistency and haemocyte agglutination as well as interfering with nerve function, which results in paralysis (and frequently death later), 2. Kill the prey by rapidly inducing loss of consciousness by injecting toxins in the start of shock. Alkaloids, terpenes, polysaccharides, biogenic amines, organic acids, and amino acids have been listed as chemical components of insect venoms (Blum and Hermann 1978); however, peptides, oligopeptides, and proteins may potentially be key constituents (Schmidt 1982; Calvete et al. 2009).

13.3

Salivary Gland, Venomous Saliva Collection, and Quantification

13.3.1 Salivary Gland Assassin bugs (Hemiptera: Reduviidae) and robber flies (Diptera: Asilidae) are two families with mouthpart-associated venom glands that have recently been described in detail based on 3D reconstructions of gland architecture from magnetic resonance imaging (MRI) spectroscopy and other techniques (Drukewitz et al. 2018; Walker et al. 2018b). Both groups have multi-compartment complicated gland systems and muscle valves that regulate the passage of venom. Heteroptera: The anterior major gland, posterior main gland, and accessory gland make up the heteropteran venom gland’s typical three compartments. The AMG and PMG were discovered to be active secretory tissues in the harpactorine assassin insect species Pristhesancus plagipennis (Walker et al. 2018b), Catamirus brevipennis (Sahayaraj et al. 2010), and Rhynocoris marginatus (Kumar et al. 2012; Sahayaraj and Balasubramanian 2016). It consists of a pair of primary glands (PG) with two lobes and a pair of accessory glands present in the reduviid salivary gland complex. The main glands extend to the anterior crop region and are located on either side of the oesophagus. The anterior lobe of the principal gland (APG), which is a short, triangular structure, and the posterior lobe of the principal gland make up the principle gland, which is simply bilobed (PPG) (Fig. 13.1), whereas the each complex venom produced by the AMG and PMG contains a unique combination of proteins and peptides. Instead of the anterior side of the posterior main gland, the posterior lobe is extremely nodulous on the posterior side. As constrictions, these nodules are distinctive (C). A tracheal supply goes to the major and auxiliary/ accessory glands (AG) as well as their ducts. A well-developed, compartmentalised hilus (HI) with valve apertures for the control of secretions released from various lobes of the main and accessory glands is present at the intersection of the anterior and posterior lobes. The principal duct (PD) and accessory duct (AD) arise from the gland in the hilus area, which has a deeper constriction. The PMG’s venom was very protease-rich and quickly paralysed and killed prey insects, suggesting that it may have two functions: capturing prey and extra-oral

412

13

Venomous and Other Body Fluids in Insect Predators

Fig. 13.1 Anterior major gland (AMG), posterior main gland (PMG), accessory gland (AG), constructions (C), compartmentalised hilus (HI), principal duct (PD), and accessory duct (AD) of Rhynocoris marginatus

digesting. We have argued that the venom produced by the AMG has a defensive purpose since, in the absence of electrostimulation, bugs could be provoked to create it. The purpose of AMG venom hasn’t been identified definitively, though. The hilus is a two-chambered structure that connects the AMG and PMG to one another as well as to venom ducts that flow to the mouthparts and AG. The 166 muscle fibres that make up the sphincter valves that surround the basal lamina of the major glands are known as neuromuscular junctions, and they are located where the AMG and PMG meet the outer chamber of the hilus (Walker et al. 2018b). According to external stimuli, it is believed that these structures enable the regulated release of venom from either the AMG or PMG. The muscle-driven venom pump inside the cranium is likely used to further manage injection. Heteropterans employ a proboscis made of greatly expanded maxillary and mandibular stylets to inject venom. The inner mandibular stylets are asymmetric and interlock to form a double-barrelled needle with separate, dedicated channels for venom injection and food intake, in contrast to the outer maxillary stylets, which have a cutting and potentially anchoring function (Walker et al. 2016).

13.3

Salivary Gland, Venomous Saliva Collection, and Quantification

413

Syrphidae: The venom system of the asilid fly Eutolmus rufibarbis exhibits some functionally convergent similarities with that of assassin bugs despite displaying some quite dissimilar morphological adaptations. The thoracic glands, so named after Kahan (1964) who discovered that homogenates of these glands caused paralysis and death when injected into insects, produce paralysing venom. Within the labium, there is yet another pair of glands known as the “labial glands” (Whitfield 1925). It is unclear how these glands operate or if they play a part in envenomation. The ties between these names, which do not imply similarity to any other insect glands, and such relationships are murky. The thoracic glands, in any case, constrict to create ducts, which fuse just before they enter the head capsule and proceed to a salivary/venom pump (Drukewitz et al. 2018). This salivary pump, which consists of multiple muscle groups and a one-way ring valve, empties into a narrow canal in the hypopharynx. It opens close to the end of the hypopharynx, an elongated, strong, and pointed structure ideal for venom injection (Whitfield 1925). Both assassin bugs and robber flies have mouthparts that have evolved convergently to develop stabbing structures with distinct channels for venom injection and food intake. However, in each group, a different anatomical substrate is used to produce this similar functional arrangement: In Asilidae, food is taken in by a tube created by the hypopharynx and concave labium, and the venom apparatus is located on the tip of an expanded, elongated hypopharynx. The two tubes develop in Reduviidae and other Heteroptera as distinct compartments surrounded by asymmetric, interlocking maxillary stylets (Walker et al. 2018a, c). Hymenoptera: An illustration of a non-oral venom injection structure is the hymenopteran venom sting, also known as the aculeus. The sting, venom reservoir, and tubular venom glands that develop at the base of the ninth gastral segment, as well as the Dufour’s gland, make up the aculeate venom apparatus and related components. Secretory tubular glands that open into the venom reservoir make up the venom gland. Due to their ancestry in sex-accessory glands, which are internalised epidermal structures, hymenopteran venom gland secretory cells are type 3 epidermal glands with an end apparatus as opposed to the venom gland secretory cells of reduviids or asilids, which are made up of columnar cells (Walker et al. 2018a, c). Neuropera: Take a look at the cross-section of a lacewing larva’s basal region of the jaw, which includes the mandible (md) and maxilla (mx), which join to form the food canal (fc), which is used to absorb liquid food from prey into the gut. The venom gland, which is a gland located within the maxilla, has been extensively characterised by numerous morphologists. This gland, which is located in the mediodorsal region of the maxilla’s base, is made up of column-shaped cells with big nuclei that resemble secretory cells (Gaumont 1976). The gland’s lumen becomes thinner as it travels anteriorly along a “venom channel” and emerges at a pore near the maxilla’s tip. The maxilla, which is shown here with venom gland (vg) and venom channel (vc) labels, is one potential site of toxins that paralyse prey (Canard 2001). These researchers came to the conclusion that the gland plays a role in prey capture based on the physical configuration of the gland and its delivery canal as well as the fact that it is present in venomous taxa but not non-venomous ones.

414

13

Venomous and Other Body Fluids in Insect Predators

Table 13.1 Venom glands and their injection tool utilised by various predatory insects Predator True bugs

Taxonomy Hemiptera: Heteroptera

Stages All

Venom glands Salivary glands

Wasps, ants Antlions and allies Larval rove Robber flies

Hymenoptera: Apocrita

Adult female Larvae Larvae

Modified sex accessory Maxillary gland or alimentary canal Alimentary canal?

Injection tool Maxillary stylets Modified oviposition Maxillae/ mandibles Mandibles

Adult

Thoracic glands

Hypopharynx

Neuroptera Coleoptera: Staphylinoidea and Hydrophiloidea Diptera: Asilidae

Additionally, it has been proposed that the larvae’s paralysing venom may originate in the alimentary canal. Feeding is ingested by larval neuropterans through a food canal made of their mandibles and maxillae interlocking. This food channel is a paired construction, unlike robber flies or assassin bugs, and each “pincer” is made up of the maxilla and mandible on opposite sides of the body. The food canal appears near the pincer’s tip (i.e. close to where the maxillary gland empties, in the part that penetrates into prey). A study by Henry provides evidence that paralytic factors may be produced in the stomach (Henry 1977). A “jaw extract” (likely made up of the maxillary glands) had very marginal effects, in contrast to the instant paralysis brought on by the head and prothorax extracts. Each group of predatory insects has a distinct form of gland that is utilised for hunting and feeding (Table 13.1.) Diptera: Many True Flies are predators, especially when they are still larval. There is proof that several of these inject bioactive venom into prey or use it to frighten away predators. The venom produced by robber flies is the dipteran venom that is most well-known (Asilidae). Given that thoracic gland extracts quickly paralyse target insects when injected into them, venom most likely originates from these glands. Before succumb to paralysis, prey impaled by the mandibles make a few furious movements. A tube in the mandible that exits near to the tip is used to administer venom. This canal connects to a gland in the head that is completely independent from the alimentary canal and is located at the anterior margin of the cibarial pump. Coleoptera: The mandibles of the Lampyridae, Dytiscidae, Gyrinidae, and Carabidae include tubes or grooves that are used to inject venom into impaled prey (Balduf 1935). Although Silphidae and Hydrophilidae lack these tubes, reports suggest that they inject comparable venoms into wounds caused by their mandibles (Balduf 1935). Glands are classified by structure as multicellular complex acinose, multicellular simple tubular, and unicellular; by position as general, hypodermal, ocular, antennal, intra-mandibular, mandibular, mandibulo-maxillary, maxillary, intra-cardonal and labial. Some open outside the extra-oral cavity (Pradhan 1939). Since none of these kinds of coleopteran larvae have salivary glands, like other coleopteran larvae, it has been hypothesised that the alimentary canal or a nearby structure produces the venom that causes paralysis and death (Balduf 1935). If so, a

13.3

Salivary Gland, Venomous Saliva Collection, and Quantification

415

single delivery tube is most likely used for both venom injection and food absorption (Table 13.1). For Lampyridae, Dytiscidae, Gyrinidae, Carabidae, and Silphidae, there is greater proof of a true poisonous nature than for Hydrophilidae.

13.3.2 Venomous Saliva Collection Methods Insect predators frequently employ venom for protection and predation. Reduviids’ anterior and posterior lobes each contain varying amounts of saliva (Maran et al. 2011). Rhynocoris kumarii, Rhynocoris fuscipes, and Rhynocoris marginatus have the most saliva, respectively. No matter whatever reduviid species, the anterior lobe has less saliva than the posterior lobe (Fig. 13.2). The VS/true venom was collected using three distinct techniques: (a) Manual milking technique in hymenoptera and reduviids (Muthukumar 2011; Sahayaraj et al. 2006a, b; Sahayaraj and Kanna 2009); Piek and Spanjer 1986; Fox et al. 2018). (b) Electric stimulation techniques in hymenoptera (Feás et al. 2021) and reduviidae (Barbosa et al. 1999; Corzo et al. 2001; Sahayaraj et al. 2006a). (c) Whole gland extraction in ants, wasps, and reduviids (Prem Mathi Maran 1999; Ambrose and Maran 1999; Quistad et al. 1994; Rivers et al. 2006). (d) Fischer et al. (2020) collected venom from the reduviids Platymeris biguttatus and Psytalla horrida using the following techniques. Venom produced by subjecting the insects to various stresses, such as light harassment, cold stress, and more strenuous harassment of caged bugs. The insects were divided into plastic enclosures for light harassment and probed with forceps, but they were not held and were permitted to flee. Individual bugs rarely attacked the forceps; instead, a tiny droplet of saliva would come from the tip of the proboscis, which could then be collected with a pipette tip and placed in a chilled 1.5-mL

250

Anterior lobe

Posterior lobe

Protein (ìg/ìl)

200 150 100 50 0

Rhynocoris fuscipes

Rhynocoris kumarii

Rhynocoris marginatus

Reduviids

Fig. 13.2 Salivary venom from the anterior and posterior glands of three Rhynocoris species from India was tested for its protein content (Source: Maran et al. (2011))

416

13

Venomous and Other Body Fluids in Insect Predators

Eppendorf tube. The bugs were subjected to 20 °C for 3 min to create cold stress, which resulted in salivation. The droplet that appeared at the tip of the proboscis was also gathered and put into a different, already-cooled tube. A coldanesthetised assassin insect with the proboscis inserted into a pipette tip was mounted to a foam cuboid for severe harassing stress. The bug was softly tapped and lightly squeezed with forceps when it was fully awake, which caused salivation. (e) The following techniques were used to induce the fire ant to release venom from the sting while it was being held by the pedicel using forceps (Lai et al. 2008). First approach: Capillary collection is method one. Major workers of the two social forms of S. invicta were used to extract pure venom by capillary action in a 5 mL microcapillary tube. Three workers’ worth of venom from each colony were combined and stored in n-hexane at -208 °C for GC-MS analysis. Twelve colonies each of monogyne and polygyne were sampled. Second approach: full-body soaking. Also used to gather venom was wholebody soaking. One glass vial containing 1 mL of n-hexane and 25 workers from each colony was stored at -208 °C. The solvent was taken out and used for GC-MS analysis after 7 days. Eight colonies each of monogyne and polygyne were examined.

13.3.3 Saliva Genders on Venom Quantity In the paired main gland, a little amount of venomous saliva is produced. Reduviids capture a variety of prey items each day; therefore, it seems to reason that they would rigorously regulate the amount of VS emitted in accordance with the type of prey. Too much VS injected into smaller preys may be physiologically costly and deplete venom stores, leaving the reduviid exposed to predators or unable to handle following prey. The quantity of VS injected into various prey objects by reduviids has not been attempted to be measured. The amount of VS secreted also depends on the gender of the animal; female predators like Rhynocoris marginatus and Catamirus brevipennis secrete more VS (Sahayaraj et al. 2006a). By using the discontinuous milking method, Sahayaraj and Kanna (2009) made a comparable observation in Catamirus brevipennis. The posterior lobe produced more saliva overall than the anterior lobe. Rhynocoris fuscipes and Rhynocoris kumarii, two of the three predators, both exhibited the maximum volume of saliva (4.9 μL) in the posterior lobe, followed by Rhynocoris marginatus (4.6 μL). Rhynocoris kumarii (3.4 μL), Rhynocoris marginatus (3.0 μL), and Rhynocoris fuscipes (2.8 μL) had the highest and lowest volumes of saliva in the anterior lobe, respectively (Maran et al. 2011) (Fig. 13.2).

13.3

Salivary Gland, Venomous Saliva Collection, and Quantification

417

13.3.4 Impact of Prey Deprivation on Venom Quantity The milking method was used to examine how malnutrition affected Catamirus brevipennis’s (Hemiptera: Reduviidae) salivary gland’s ability to produce venom. The findings show that females produced more venom than males. After 2 days of hunger, the venom output increased in both males and females (Sahayaraj and Kanna 2009). Similar to continuous (CMM) and discontinuous (DMM) food deprivation to Rhynocoris fuscipes (1, 3, 5, and 7 days), In both the continuous and discontinuous prey deprivation of both manual milking and electrical stimulation approaches, the male Rhynocoris fuscipes predators significantly secreted more VS than the females. 3-day deprived males released more VS during continuous fasting, with milking producing the highest levels (2.6 mg/100 mg of animal wet weight) and electric stimulation producing the lowest levels (1.3 mg/100 mg of animal wet weight), respectively. Five-day starved males milked more VS (4.1 mg/100 mg of animal weight) using the discontinuous starvation in the milking method than three-day starved males (3.2 mg/100 mg of animal weight) using the electric stimulation method. In the continuous milking procedure, 100% survival (SR) and venom milking rate (VMR) were seen. A 20% SR and 30% VMR were recorded during discontinuous manual milking (DMM), a 30% SR and 20% VMR were recorded during continuous electrical stimulation (CES), and a 20% SR and 30% VMR were recorded during discontinuous electrical stimulation (DES). In contrast, Ambrose and Maran (1999) previously showed that an increase in hunger and VS accumulation in the salivary gland results in a reduction in the size and quantity of the VS and salivary gland. Predators typically do not use their VS when starving, which causes VS to build up in the salivary gland. This suggests that on the third day of fasting, the predator secretes the greatest amount of VS. Reabsorbing water is what allows the predator to survive, and on the sixth day of famine, this is accomplished with only a very small amount of VS (0.02 mg/100 mg of the animal weight). In contrast to prior research by Sahayaraj et al. (2006a), which demonstrated that female predators like Rhynocoris marginatus and Catamirus brevipennis produce more VS than males, the third day of fasting was determined to be the best time for VS collection. Muthukumar (2011) made an effort to quantify VS when Rhynocoris marginatus was subjected to continuous prey deprivation for 1, 2, 3, 4, 5, 6, and 7 days and discontinuous prey deprivation for 1, 3, 5, 7, and 9 days. The time it took for Podisus nigrispinus and Brontocoris tabidus (Heteroptera: Pentatomidae) to attack for the first time and kill a sixth-instar Thyrinteina arnobia (Lepidoptera: Geometridae) larva, as well as the number of attacks within a 4-h period, were assessed. The Thyrinteina arnobia caterpillar’s first attack by the predator Podisus nigrispinus took a different amount of time during the 24-, 36-, and 48-hour periods without sustenance. Podisus nigrispinus developed more quickly at 36 and 48 h than it did at 24 h. When Brontocoris tabidus went 48 h without eating, it took less time for it to make its initial attack than when it went 24 or 36 h without eating. For the two animals, the first attack occurred after 24, 36, and 48 h without meals (Pires et al. 2022).

418

13

Venomous and Other Body Fluids in Insect Predators

13.3.5 Influence of Prey on VS Yield Reduviids are said to move more quickly when given small prey items while they are starving (Prem Mathi Maran 1999). Corcyra cephalonica (1.2 mg) and H. armigera milked considerably less VS than Spodoptera litura fed Rhynocoris fuscipes. Corcyra cephalonica (48.8 mg) had negligibly higher protein content in VS than S. litura and Helicoverpa armigera. According to the study on predatory potential, Rhynocoris fuscipes pinned Helicoverpa armigera eight times and paralysed it in 32.4 min. In comparison to Corcyra cephalonica (3.6 min; 3.1) and S. litura (16.2 min; 5.2), it was much greater. The one-day deprived female VS milked in the DES had a lot of protein. Feeding Spodoptera litura VS was secreted in greater amounts by Rhynocoris fuscipes than by Helicoverpa armigera and Corcyra cephalonica fed groups. To paralyse the Spodoptera litura rather than the Corcyra cephalonica and Helicoverpa armigera, the predator uses the greatest amount of VS. Spodoptera litura-fed continuous manual milking can be used to optimise and utilise VS for the highest VS production (Fig. 13.3). The amount of salivary venom produced by male and female Rhynocoris marginatus was evaluated in connection to prey deprivation durations of 1, 2, 3, 4, 5, 6, and 7 days as well as durations of 1, 3, 5, 7, and 9 days (manual and electric stimulation of milking methods). Periods of prey deprivation have a direct impact on venom yield. Results revealed important details on both genders and different types of prey deprivation. After one to 4 days without prey, Rhynocoris marginatus produced more salivary venom. In comparison to electric stimulation (0.92 mg/ insect), manual milking caused male Rhynocoris marginatus to produce greater venom (1.16 mg/insect). The female Rhynocoris marginatus’ rate of venom milking

Posterior lobe A m o u n t ( µ l

Anterior lobe

9 8 7 6 5 4 3

2

)

1 0

Rhynocoris fuscipes

Rhynocoris kumarii

Rhynocoris marginatus

Reduviids

Fig. 13.3 Amount venom (μL) collected from posterior and anterior lobes of three reduviid predators (After Maran et al. 2011)

13.4

Venom Chemical Composition

419

was significantly lowered by a lack of prey. Males almost always produced and spent more money (Muthu Kumar 2010).

13.4

Venom Chemical Composition

The primary salivary glands’ lobes’ secretions have diverse compositions, resulting in the front lobe’s secretion being liquid and homogeneous and the posterior lobe’s secretion are being gelatinous with granules of various sizes. Predators’ saliva contains some protein, but it also contains non-proteinaceous substances.

13.4.1 Proteinaceous Components Despite making up only 0.1% of all transcripts, the Rhynocoris iracundus venom toxin transcripts accounted for 29% of the overall relative expression (292 venom toxin transcripts). Both male and female salivary venom shows similar kind of proteins ranging from 2.2 to 36.2 kDa (Fig. 13.4). Notably, only two toxin families accounted for a significant fraction of the relative venom toxin expression: haemolysins (4 transcripts representing 7% total expression) and venom family

Fig. 13.4 SDS-PAGE analysis of Rhynocoris marginatus venom saliva showing similar molecular weight proteins both in male (MS) and female (FS)

420

13

Venomous and Other Body Fluids in Insect Predators

17 (3 transcripts representing 11% total expression). The S1 protease (38 total, 10 full-length/28 fragment), trypsin (28, 15/33), orphan venom-family (25, 11/14), secreted hypothetical protein (14, 6/8), venom family 13 (11, 0/11), and chitinases (9, 3/6), comprised the majority (>50%) of transcripts related to venom proteins. Only 15% of the expression of all toxins was reflected by the relative expression of these groups as a whole. Other highly expressed toxin families include the orphan venom family (2% overall expression), redulysin (1%), and venom family 11 (1%), in addition to the haemolysin and venom family 17 previously mentioned (Rügen et al. 2021). The transcriptomes (TPM > 3) of the glandular lobes of AMG, PMG, and AG from Platymeris biguttatus projected 458 possible secreted proteins, 368 potential secreted proteins, and 475 potential secreted proteins, respectively. Venom family 8-like peptide Pr8a (TPM 149437.6) and venom haemolysin-like protein 2 were the two most prevalent proteins in AMG (TPM 143250.5). Venom redulysin 2 (TPM 76012.5) and venom protein family 1 protein 1 (TPM 98782.8) were substantially expressed in PMG. The two most prevalent proteins in AG were transferrin (TPM 344146.7) and venom triabin 1 (TPM 143997.7). Seven proteins from the AMG were classified as venom haemolysin-like (total TPM 334907.9), followed by the venom protein family (TPM 187617.31, which contains 17 proteins), and cystatin domain (TPM 187617.31). (TPM 77736.2, four proteins). Serine proteases (TPM 246613.5, 50 proteins), redulysin, and the venom protein family were all found and classed in the PMG as having rather high concentration (TPM 128013.6, six proteins). The putative reduvenom family (TPM 226581.2, encompassing 27 proteins) and redulysin (TPM 192890.4, 2 proteins) were the next most prevalent protein families in the AG, followed by the transferrin-like domain (TPM 344146.74, 1 protein) (Gao et al. 2022). Disulphide-rich peptides, CUB domain-containing proteins, cystatins, homologues of trialysin, trypsin-like proteases, various catabolic enzymes, serpins, a triabin-like protein, bacterial permeability-increasing-like protein, and novel protein families were among the protein families present in the venom (Walker et al. 2017). Protein kinase, inositol-phosphate phosphatase, M12A-like metalloproteases, cathepsin B, peptidase S10, hexosaminidase, and nuclease were among the putative enzymes found in the venom. There were 17 proteins present, and they were grouped into eight families based on their proposed functions, as follows: Families 1, 2, and 4 were unique, but proteins in families 3 and 5–8 exhibited homology to unidentified and anticipated protein sequences from hemipterans and other insects (Walker et al. 2017). Venom families, Cystatin (protease inhibition/other), Ptu1 family peptides (neurotoxin), Triabin-like, Redulysin (Cytolysis Protease) (Cytolysis Protease) M12A 2 27–30 46 29-60 Proteolysis S1 protease Fibrillin Proteolysis 1 30 Unknown Serpin (inhibition of proteases), Cathepsin B (proteolysis), Inositol-phosphate phosphatase, Lipase (catabolism of lipids), Histidine phosphatase (phosphatase), Nuclease (catabolism of nucleic acids), Protease S1 CUB 19 45–70 Proteolysis 47.4 Unknown Protein Kinase 1 S10 Peptidase 1 50.6 Proteolysis 67.8 Hexosaminidase 1 Transferrin, catabolism of biopolymers (Walker et al. 2017). Ptu1 is a neurotoxic and an N-type calcium channel blocker.

13.4

Venom Chemical Composition

421

Galleria mellonella larvae succumbed to Platymeris biguttatus piercings soon. After 5 min, the area where Platymeris biguttatus bit started to get darker. The corpse mostly dissolved and turned black after 30 min of the attack (characterised by the elongated body). Similar to the real attack treatment, the PMG extracts also quickly killed the majority of the larvae, with an estimated LD50 of 20.4 g/g after 30 min. This value was calculated using a probit model from the mortalities of 0, 19, 1, 23, 8, 42, 8, and 71, 4 at concentration gradients of 0.4, 0.8, 1.2, 1.6, and 2.0 g/L. Ele’s LD50 was calculated as 12.5 g/g using the mortalities of 0%, 23%, 8%, 80%, 9%, 90%, and 95.2% at doses ranging from 0.4 to 2.0 g/L. (Gao et al. 2022). Galleria mellonella larvae exposed to PMG injection experienced an immediate fatality. Similar to the actual attack treatment, PMG injection caused the larvae to quickly cease moving and die. Within 10 min, the carcass turned to liquid. Contrary to the actual attack treatments, the corpse remained white throughout the 30-min observation and did not become dark. In contrast, Galleria mellonella larvae treated with AG and AMG did not experience any evident tissue degradation or death. Moving distance in 10 min for AG-treated larvae was 17.5 cm (11.9 cm) and for the control group was 55.9 cm (21.1 cm), respectively, whereas the mobility of larvae treated with AMG extract was not substantially different from the control group (Gao et al. 2022).

13.4.2 Enzymes One of the most significant digestive enzymes, proteases plays key roles in the transformation of proteins into oligo- and dipeptides. Based on the amino acids in their active site and the location where they interact with protein molecules, these enzymes are categorised. By severing intracellular connections, proteinases (endopeptidases) are responsible for the first digestion of proteins. A variety of proteinases are required to break these bonds because peptide chains vary. According to their active site, these proteinases have been divided into three primary subclasses: serine, cysteine, and aspartic proteinases. The proteinases, phospholipases, trypsin-like enzymes, esterases, and serine proteases are the most prevalent enzymes in salivary secretion (Cohen 1995). Serine proteases, which are present in numerous animals, have sparked a wide range of research interests because to their various physiological roles in processes like digestion, immunological response, cellular differentiation, and prothrombin activator (Cohen 1993; Evangelin et al. 2014). Reduviidae: Zelus renardii contains trypsin-like genes for protein digestion since trypsin-like enzyme with a molecular weight of approximately 27 kDa was found in its salivary glands (Cohen 1993). Salivary gland protease activity known as Zelus renardii has been described as a trypsin-like enzyme involved in pre-oral digestion (Cohen 1993). Pristhesancus plagipennis, an Australian assassin beetle, secretes saliva that contains 127 enzymes, 69 of which are linked to proteolysis. In contrast, just three of these enzymes have putative roles in lipid catabolism, one in nucleic acid catabolism, and 10 proteins are linked to cytolysis (Walker et al. 2017).

422

13

Venomous and Other Body Fluids in Insect Predators

The salivary gland complex (SGC), anterior midgut (AMG), and posterior midgut (PMG) of Zelus renardii were investigated for their proteinases. Additionally, a salivary trypsin-like enzyme with a molecular weight of roughly 27 kDa was discovered (Cohen 1993). The female midgut and female accessory gland of Rhynocoris fuscipes have the highest levels of amylase. Male and female accessory glands both had higher levels of lipase activity (41.68 mequ/min/g and 42.81 mequ/min/g, respectively), and the foregut had the highest levels (49.70 mequ/min/g). Both males and females exhibit invertase activity in their foreguts, but it was shown that invertase activity is stronger in the posterior lobe of the main gland in each gender. High levels of trehalase activity and high levels of trehalose activity are present in the posterior lobes of the main gland. The VS (23.4 mg tyrosine/g protein/min) and the hindgut both have significant protease activity levels. The greatest acid phosphatase activity is found in the AG of both males and females. Both males and females contain high levels of enzyme in their foreguts, measuring 7.6 and 7.8 mmol/g/1 h, respectively. The maximal amount of phospholipase A2 enzyme activity found in female poisonous saliva is 18.3 nm/min/mg proteins. The hyaluronidase enzyme is very active in the Rhynocoris fuscipes main gland’s anterior lobe. The female poisonous saliva had the highest degree of trypsin-like enzyme activity (18.1 units/mg). Important enzymes that contributed to the venomous saliva of Rhynocoris fuscipes’ toxin nature include phospholipase and hyaluronidase (Cohen 2000). It has been demonstrated before that the salivary organs of Rhynocoris marginatus contain a variety of physiologically active chemicals. Currently, amylase, invertase, trehalase, protease, acid phosphatase, alkaline phosphatase, phospholipase, lipase-trypsin, hyaluronidase, and esterase were detected in the venom, venom gland, and alimentary canal of Rhynocoris marginatus using conventional colorimetric methods. The alimentary canal lacked phospholipase and hyaluronidase. Esterase, phospholipase, and lipase were the main hydrolases found. Saliva contains significant levels of phospholipase, lipase, and hyaluronidase activity, which spreads enzymes and aids in pre-oral digestion. High levels of lipase and protease activity are present in the anterior and posterior lobes of the major and auxiliary glands of Rhynocoris marginatus, respectively. The different enzymes found in Rhynocoris marginatus saliva are listed in Table 13.2. The amounts of enzymes varied between male and female individuals. When partially digested food passes through the foregut and midgut, these enzymes are mostly used for the digestion of protein and fats. In the hindgut, alkaline phosphatase levels were elevated (Muthu Kumar 2010). Protease activity was found to be higher than other enzymes in the salivary gland complex of Catamirus brevipennis (Sahayaraj et al. 2007, 2010). The venomous saliva of Rhynocoris fuscipes was discovered to have significant protease, trypsinlike activity, and phospholipase activity. These enzymes aid the animal’s liquefaction of its prey (Cohen 1995). The other EOD enzymes spread very quickly as a result of phospholipase’s spreading behaviour. A significant portion of the molecular components in assassin bug venom have cytolytic and enzymatic properties that contribute to tissue liquefaction (Walker et al. 2017). In addition to causing

13.4

Venom Chemical Composition

423

Table 13.2 Quantitative enzyme analysis of saliva of Rhynocoris marginatus Enzyme Amylase (μG maltose released/min/mg protein) Lipase (meq/min/g) Protease (μg tyrosine/mg protein/min) Acid phosphatase (μmol /mg/h-1) Invertase (μg glucose released/min/mg protein) Alkaline phosphatase (μmol/mg/h-1) Trehalase (μg glucose released/min/mg protein) Esterase (mmoles/min/mg protein) Trypsin (U/mg) Hyaluronidase (units/mg protein) Phospholipase (nM/min/mg protein)

Male 0.37 27.95 10.01 3.82 0.291 14.14 1.06 16.40 10.35 25.39 21.08

Female 0.42 29.43 11.18 4.19 0.39 16.56 1.17 21.79 13.40 27.60 40.49

paralysis, G. mellonella was strongly melanised by Rhynocoris iracundus venom (Rügen et al. 2021). Pentatomidae: Podisus maculiventris’ proteolytic activity significantly decreased throughout the period of 96 h of fasting. After 48 h of fasting, proteolytic activity was 90% of what it was for newly fed insects, similar to that of non-starved insects. However, beyond this point, values decreased to roughly 65% at 72 h and 11% at 96 h following the start of hunger (Bell et al. 2005). In the salivary glands of the Podisus nigrispinus, proteases including lipase and trypsin-like protease were found (Oliveira et al. 2006). In Brontocoris tabidus (Heteroptera: Pentatomidae), both lobes of the main salivary gland contained lipases, -amylase, and trypsin-like enzymes (Azevedo et al. 2007). The primary salivary gland of the predator Brontocoris tabidus (Heteroptera: Pentatomidae) contains lipases, a-amylase, and enzymes that resemble trypsin (Haddad et al. 2010). Using 2% azocasein as the substrate, samples taken from the salivary glands of the animal Andrallus spinidens demonstrated proteolytic activity (Hyodo et al. 2014; Zibaee et al. 2012). Heteroptera: Pentatomidae predators Brontocoris tabidus and Podisus nigrispinus salivary enzymes revealed differences in amylase activity, with the predator exhibiting a higher amount of this enzyme. Lipase and protease activity were equal between these predators (Pires et al. 2022). Miridae: Investigated were the protease activity in the green mirid Creontiades dilutes midgut, salivary glands, and secretory saliva. There were no variations in the levels of protease activity between male and female mirids, adult mirids and third instar nymphs, or between fed and fasted mirids. However, the saliva and salivary glands had more protease activity than the midgut. As evidenced by inhibitor selectivity, basic pH maxima, and the hydrolysis of N-benzoyl-l-tyrosine p-nitroanilide and N-succinyl-ala-ala-pro-leu p-nitroanilide, chymotrypsin-like serine proteases predominated in the salivary glands. Midgut extracts had an acidic pH optimum (pH 4), indicating that acidic proteases predominate. However, both aprotinin and E-64 significantly reduced protease activity, indicating the existence of both serine and cysteine proteases in the midgut of the green mermaid (Colebatch

424

13

Venomous and Other Body Fluids in Insect Predators

et al. 2001). The predatory plant insect Deraeocoris nebulosus (Miridae) salivary glands had trypsin-like enzyme, -glucosidase, and pectinase activity (Boyd et al. 2002). Boyd et al. (2002) also talked about the stylet feature of Deraeocoris nebulosus. It was discovered that the salivary gland of the Miridae species Deraeocoris nigritulus produces digestive enzymes such as pectinase, a chymotrypsin- and trypsin-like enzyme (Boyd 2003). Neuroptera: Most Neuroptera larvae—including lacewings, antlions, owlflies, and allies—are poisonous predators of other insects (Tauber et al. 2009a, b). Given that food may only be consumed in liquid form (as is the case with reduviids, asilids, and other animals), venom is also thought to have a liquefying effect. The gland’s lumen becomes thinner as it travels anteriorly along a “venom channel” and emerges at a pore near the maxilla’s tip. The allied non-venomous orders Megaloptera and Raphidioptera, which are predatory neuropteran families and include the basal Nevrorthidae, do not contain it. These researchers came to the conclusion that the gland plays a role in prey capture based on the physical configuration of the gland and its delivery canal as well as the fact that it is present in venomous taxa but not non-venomous ones.

13.4.3 Genes Important metabolic pathways in controlling information and protein digestion include neuroactive ligand-receptor interaction, wnt signalling pathway, glutathione metabolism, arginine and proline metabolism, and others. A number of genes work together to control these processes. In the nymphs and adults of Acanthaspis cincticrus, predation-related SDEGs were enriched in these pathways (Kou et al. 2017). To map the genetic environment of A. cincticrus during various life phases, we assessed the gene expression levels. There were nine transcriptome comparisons we made, including egg vs. first instar, first instar vs. second instar, second instar vs. third instar, third instar vs. fourth instar, fourth instar vs. fifth instar, first instar vs. fifth instar, fifth instar vs. adult male, fifth instar vs. adult female, and male vs. female adults (Kou et al. 2017). A total of 13,479 SDEGs were obtained as a result, and a heat map was created by hierarchically clustering them. Each stage of life in this heat map had a distinct gene expression profile with a large cluster of highly expressed genes in common. Maternal genes including Kruppel, hunchback, nanos-like protein, serine protease snake-like, protein takeout-like, maelstrom, pumilio homolog 1-like, and Staufen were present in many essential embryogenesis-related genes at the egg stage. Acanthaspis cincticrus had high expression unigenes in both its nymph and adult stages. These included cuticular proteins, histone H2A, cytochrome P450s, ecdysone-induced protein, and juvenile hormone binding proteins in the nymph stages, and vitellogenin and sperm flagellar protein 1-like in the adults (Kou et al. 2017). The following procedures were used by Fischer et al. (2020) to harvest the venom from the reduviids Platymeris biguttatus and Psytalla horrida. By using nextgeneration sequencing (RNA-Seq), the protein composition of PMG and AMG

13.5

Non-Proteinaceous Components

425

Table 13.3 Peptides of predators with activities and references Predators Peirates turpis

Peptides Ptu1

Activity Neurotoxic activity

Rhynocoris marginatus

RmIT-1 (3.79 kDa), RmIT2 (9.7 kDa), and RmIT-3 (10.94 kDa) Ponericins G, L, W

Insecticidal

Insecticidal

Orivel et al. (2001)

Solenopsins

Insecticidal, cytolytic, and antibacterial activities

Blum et al. (1958), MacConnell et al. (1971)

Pachycondyla goeldii Solenopsis sp.

References Bernard et al. (2001) Sahayaraj et al. (2013a, b)

venom was determined, enabling the identification and quantification of venomassociated transcripts. The list of peptides recovered from various predators is presented in Table 13.3.

13.5

Non-Proteinaceous Components

Gas chromatography analysis of the venom alkaloids of Solenopsis invicta revealed five to six piperidine alkaloids, including trans-2-methyl-6-n-undecylpiperidine (trans C11), trans-2-methyl-6-(cis-4-n-tridecenyl) piperidine (trans C13:1), trans-2methyl-6-n-tridecylpiperidine (trans C13), trans-2-methyl-6-(cis-8-n-heptadecenyl) piperidine (trans C17:1)). Lai et al. (2012) found that some predatory ants displayed alkaloids and other secondary metabolites from plants. For instance, the venom of the Solenopsis invicta comprises six main alkaloids, whereas the venom of the Solenopsis geminata contains alkaloids that are largely saturated C11 and appear in both cis and transforms (from trans C11 to C17). Furthermore, compared to the monogynous Solenopsis invicta, the proportions of unsaturated alkaloids in the polygynous S. invicta’s venom were noticeably higher (Lai et al. 2012). In Anticarsia gemmatalis, Martínez et al. (2016) discovered non-proteinaceous substances with insecticidal activity in the saliva of Podisus nigrispinus (Hemiptera: Pentatomidae). As non-proteinaceous extract components of Podisus nigrispinus, N, N-dimethylaniline and 1,2,5-trithiepane fractions were discovered (Martínez et al. 2016). According to studies, N, N-dimethylaniline stimulates phagocytosis in Anthonomus grandis (Coleoptera: Curculionidae) and acts as a strong poison in Melophagus ovinus (Diptera: Hippoboscidae) by speeding up body cuticle penetration (Hedin et al. 1968). EDAX evaluation of Rhynocirs fisces’ dried venom is depicted in the figure. It has few components. Table 13.4 and Fig. 13.5 provide information on the elements’ composition, weight, and atomic percentage.

426

13

Table 13.4 Composition of elements, their weight, and atomic percentage

Venomous and Other Body Fluids in Insect Predators

Elements Si K SK Cl K KK Ca K Cu K Zn K Zr L

Weight% 0.46 5.11 6.20 8.56 0.69 3.07 1.70 8.01

Atomic% 2.20 21.27 23.35 29.23 2.29 6.46 3.47 11.72

Fig. 13.5 EDAX spectrum of salivary venom from Rhynocoris fuscipes

13.6

Salivary Venom of Predators

13.6.1 Hymenoptera In contrast to other insect groups, the venoms of the Hymenoptera have received the most attention. Some aculeate venom also contains non-proteinaceous poisons that serve as essential functioning components. By inhibiting excitatory ligand-gated ion channels at neuromuscular junctions, the philanthotoxins, polyamines discovered in the venom of the beewolf Philanthus triangulum (Crabronidae), paralyse victims (Piek et al. 1982; Eldefrawi et al. 1988). Another illustration comes from the jewel wasp Ampulex compressa, which stings cockroaches and injects venom into their 408 ganglia and brain. The inhibitory transmitter aminobutyric acid (GABA), as well as the GABA receptor 410 agonists taurine and alanine, are present in high concentrations (10–30 mM) in the venom of jewel wasps (Weisel-Eichler et al. 1999; Moore et al. 2006). The ensuing activation of inhibitory GABAergic transmission aids in the envenomation-induced hypokinetic state that permits the wasp

13.6

Salivary Venom of Predators

427

grub to eat the surviving cockroach. Additionally, some ants contain venoms that are mainly composed of non-proteins. Fire ants (Solenopsis sp., Myrmecinae) have venom that is primarily made up of hydrophobic piperidines known as solenopsins that have insecticidal, cytolytic, and antibacterial activities (Blum et al. 1958), whereas ants in the subfamily Formicinae have venom that is primarily made up of formic acid, which they spray through an acidopore to defend themselves. The majority of the species in the subfamily Ponerinae use their venom to catch practically all of the prey they come across. Ponericins G (87%), W (92%), or L (96%) were identified in the venom of the predatory ant Pachycondyla goeldii based on the first most prevalent amino-terminal amino acid (Orivel et al. 2001). Peptides made up the venom of Pseudomyrmex termitarius, Pseudomyrmex penetrator, and Pseudomyrmex gracilis. However, P. penetrator’s venom was not as strong as that of other species. Pseudomyrmex termitarius venom contains 87 linear peptides, while the venoms of Pseudomyrmex gracilis and Pseudomyrmex penetrator each contain 23 and 26 disulphide-bonded peptides (Touchard et al. 2014). Ninety percent to ninety five percent of the water-insoluble alkaloids in fire ant venom are species-specific and have functions such as antibacterial, insecticidal, haemolytic, and histamine release (Lai et al. 2012). In a previous study (Lai et al. 2010), it was determined whether the stings of the fire ants Solenopsis geminata, Solenopsis invicta (monogyne form), and Solenopsis invicta (polygyne form) were poisonous to Spodoptera litura larvae. The fire ant larvae caused severe spasms and paralysis within seconds of being stung. After 24 h, the Solenopsis geminata, Solenopsis invicta, and 86.67% of the Solenopsis invicta stung Spodopter litura larvae that had received venom all died. The results also revealed that Solenopsis geminata, Solenopsis invicta (monogyne form), and Solenopsis invicta (polygyne form) were in order of increasing susceptibility to fire ant venom in Spodopter litura (Lai et al. 2010). Only portion C of the Pachycondyla goeldii venom was found to have insecticidal effects. The domestic cricket, Acheta domesticus, was resistant to the insecticidal effects of all 10 synthesised peptides. With toxicity (LD50 in mg venom/gram injected individual), four were extremely active against crickets (G1—82.7, G3— 122.3, W3-desK—129.6, and W4—88.2) and 130 mg peptide/g animal weight. Only two of these four peptides—W3-desK (484) and W4 (613.7)—had an impact on Pachycondyla goeldii employees. Although the amount of toxicity against ants was very modest, this finding raises the possibility that they have developed immunity to their own venom (Orivel et al. 2001). Yoshida et al. reported a novel insight (2001). According to Yoshida et al. (2001), German cockroaches (Blattella germanica) are quickly rendered paralysed when the endosymbiont Enterobacter aerogenes is grown up in culture and introduced into the saliva of Myrmeleon bore larvae. One of the insecticidal proteins was isolated from culture broth, and we discovered that it moved as a single band at a spot corresponding to a relative molecular mass of around 63 K on a denaturing SDS-polyacrylamide gel. This toxin’s partial amino-acid sequencing revealed that it was a GroEL homologue (Yoshida et al. 2001).

428

13

Venomous and Other Body Fluids in Insect Predators

Plutella xylostella larvae treated with the venom of Solenopsis geminata all perished after 17 min; however, some Plutella xylostella with the venom of Solenopsis invicta perished 130 min (for monogynous Solenopsis invicta) and 100 min (for polygynous S. invicta) later. Plutella xylostella fourth-instar larvae symptoms brought on by fire ant venom. (A) The dorsal thoracic area of a larva was treated with one droplet of fire ant venom. (B) Within 2 h, a larva’s dorsal thorax turned black. The venom of the Ponerinae ant Pachycondyla striata was examined for its protein composition. Pachycondyla striata venom contains a complex variety of proteins, of which 43 have been identified. Classical venom proteins (phospholipase A, hyaluronidase, and aminopeptidase N) are among the proteins found in the samples.

13.6.2 Pentatomidae With an LC50 = 2.04 L and LC90 = 3.27 L, in particular, the ether extract from Podisus nigrispinus saliva caused mortality in Anticarsia gemmatalis larvae (Martínez et al. 2016). N-dimethylaniline (separated components) had an LC50 = 136.1 nL and LC90 = 413.8 nL, suggesting that it could be the cause of toxicity in Podisus nigrispinus saliva. However, Anticarsia gemmatalis larval mortality needs larger quantity of induced N (Martínez et al. 2016). N, N-dimethylaniline (DMA) and 1,2,5-trithiepane are poisons found in the predatory stink insect Podisus nigrispinus that kill prey. These substances were tested on Spodoptera frugiperda caterpillars, although their exact mode of action is yet unknown. All substances are poisonous and harm midgut cells, causing apoptosis and necrosis. Apical protrusions with the nucleus exposed to the lumen, vesicular rough endoplasmic reticulum, autolysosomes, and electron-dense lipid droplets are the key characteristics of these cells. The investigated biogenic substances 1,2,5trithiepane, DMA, and Podisus nigrispinus saliva extract are poisonous and adversely affect Spodoptera frugiperda caterpillar development (Campos et al. 2020). Adult Podisus nigrispinus preys on a variety of insects by inserting mouthparts and regurgitating the prey’s salivary gland contents, which quickly paralyses and kills the prey. However, it is unknown what exactly is in Podisus nigrispinus’ saliva that kills its prey. Martínez et al. (2014) assessed the ultrastructure and cytochemistry of the predator’s salivary glands as a preliminary step in identifying the component of the saliva of Podisus nigrispinus. There are two major salivary glands (PSG) and two tubular auxiliary glands in the salivary system of the Podisus nigrispinus (AG). Saliva from the Podisus nigrispinus demonstrated positive reactions for carbohydrates, proteins, and acid phosphatase in several glandular system locations. According to Martínez et al. (2014), the cytochemical and ultrastructural characteristics point to a role for the major and auxiliary salivary glands in the production of saliva’s proteins. However, in another study, non-proteinaceous substances from the saliva of Podisus nigrispinus in Anticarsia gemmatalis have also shown insecticidal action (Martínez et al. 2016).

13.6

Salivary Venom of Predators

429

13.6.3 Reduviidae Assassin bugs like Rhynocoris carmelita and Platymeris rhadamanthus have been known to paralyse prey hundreds of times their own size, according to Edwards (1961), who made the observation more than 50 years ago (over a time scale of seconds). The venoms of predaceous heteropterans and other venomous species have evolved along strongly convergent lines, as shown by the neurotoxic action of these peptides (Bernard et al. 2001; Corzo et al. 2001). Only a small number of families’ venoms have shown neurotoxic action to yet, but the vast majority of heteropteran venoms have never been examined using methods that can recognise and characterise neurotoxins. Therefore, a large number of neurotoxins may be present in the venom of real predatory bugs. In India, Rhynocoris marginatus is a common and potentially lethal reduviid predator of numerous economically significant pests. To create various concentrations, the venomous saliva (VS) was collected using the milking method and diluted with HPLC grade water (200, 400, 600, 800, and 1000 ppm). The VS from Rhynocoris marginatus was discovered to be hazardous, with LD50 values of 768 and 929 ppm at 48 and 96 h for microinjection and oral toxicity experiments, respectively, in Spodoptera litura third instar. After administering VS to the host for 96 h, the level of hydrolase and detoxifying enzymes drastically lowered in a dosedependent manner (Sahayaraj and Muthukumar 2011). Numerous insects belonging to seven orders of insects tested harmful to the saliva. Application of 0.1 ml of 1% saliva of Platymeris saliva to the heart and gut of a fifth-instar Platymeris nymph following removal of the abdominal terga did not significantly affect their rhythmic contractions over a period of more than 3 h, proving that their own species is immune. By mixing 1% saliva solutions with an equal volume of Platymeris haemolymph beforehand, the toxicity of these solutions towards the Periplaneta heartdorsum preparation is not diminished. Although the haemolymph is not deadly on its own, the crop’s contents do keep the saliva’s toxicity alive for at least 2 days. Arthropods are not poisoned by enteral or external applications of saliva; instead, the saliva must enter the haemocoel. Without suffering any negative effects, Periplaneta drank 1% saliva solution, while Calliphora larvae were not affected by a 6-h soaking in 1% solution. In the author’s experience, an assassin bug bite leaves a long-lasting necrotic pit and produces severe regional agony and oedema. The dry saliva powder causes respiratory problems similar to those brought on by snake venom, oedema, vasodilatation, excessive mucous production, and irritation of the eye and nose membranes (Stanic 1956). The left prothoracic pleuron was used to inject an amount of saliva similar to that supplied during natural predation in order to calculate the LD50 of Platymeris saliva for Periplaneta americana. Groups of ten rats received injections of 10% saliva that were diluted over time. The LD50 was 10–25 mg/kg for 18 h at 280 °C (Edwards 1961). The deadly saliva of predatory reduviids yielded three new peptides. They are composed of 34–36 amino acid residues and were discovered via mass spectrometry and HPLC analysis. They are members of the four-loop Cys scaffold structure class

430

13

Venomous and Other Body Fluids in Insect Predators

and are comparatively comparable to the calcium channel blockers produced by marine cone snails called conotoxins. The shortest peptide, Ptu1, was chemically created (sPtu1) and eluted alongside its natural version. The sPtu1’s circular dichroism spectrum revealed a high concentration of turns similar to the conotoxins GVIA and MVIIA. Electrophysiological tests showed that the N-type calcium channels produced in BHK cells are reversibly blocked by sPtu1 (Corzo et al. 2001). RmIT-1 (3.79 kDa), RmIT-2 (9.7 kDa), and RmIT-3 (10.94 kDa) (Rhynocoris marginatus Insect Toxin) are tiny molecular weight peptides that were isolated from this insect (Sahayaraj et al. 2013a, b). The paralysing capability of Rhynocoris fuscipes salivary gland extract against particular pests was previously examined by Maran (2000). In the experimental category, there was a noticeable rise in protein content (50%) and a drop in carbohydrate (21%) and fat (46%) amounts. In the oral toxicity research, the VS decreased Spodoptera litura relative growth rate, approximate digestibility, and efficiency of conversion of ingested and digested food. Blood cells that have been isolated from Spodoptera litura fifth stadium larvae exhibit spreading activity that is affected by salivary venom, which prevents haemoglobin from aggregating. The findings demonstrated that VS toxins induced death, altered Spodoptera litura levels of macromolecule quantity and digestive enzymes, and altered nutritional indices. We came to the conclusion that Rhynocoris marginatus’s VS is poisonous to the prey species Spodoptera litura (Sahayaraj and Muthukumar 2011). Individuals H. armigera and Spodoptera litura injected with the lowest amounts (200 and 400 ppm) initially showed no reaction, but after 90 min the following sequence was seen: wiggling, restless movement, rapid mandibular mastication, lateral fall, and eventually motionlessness. With higher doses, these symptoms appeared to appear sooner (30–40 min later). Within 96 h, the maximum amount of monitoring, none of the control injections of 1.0 L HPLC-grade water resulted in a fatality or envenomation symptoms (Sahayaraj and Vinothkanna 2011). When third instar Spodoptera litura larvae were given VS injections, only 26.67% of the larvae died within 24 h (F = 9.42; df1,18; p 0.01). However, at 96 h, Helicoverpa armigera larvae had died with an LD50 of 846.35 ppm/larvae, while 64.29% of Spodoptera litura larvae had perished (F = 19.43; df1,18; P). Both pests died as a result of dose-dependent VS of R. fuscipes. In Helicoverpa armigera and Spodoptera litura larvae, VS caused less than 50% mortality at 24 and 48 h of observation. In Spodoptera litura, oral administration of VS caused respective mortalities of 71.43% and 69.23% (F = 19.42; df1,18; P and Helicoverpa armigera) (Sahayaraj and Vinothkanna 2011). Galleria mellonella larvae and pupae were paralysed and melanised in a lab setting by the venom of the Rhynocoris iracundus (Rügen et al. 2021). Three larvae were completely paralysed within 1 min, and after 4 h, even though melanisation had begun, the effect gradually lessened. Furthermore, 1 h after injection, larvae started to get quite squishy. Except for one larva that only received 50% of the typical venom dose, all G. mellonella pupae and larvae studied perished (Rügen et al. 2021).

13.6

Salivary Venom of Predators

431

13.6.4 Diptera Many True Flies are predators, especially when they are young (Wiegmann et al. 2011). There is proof that several of these inject bioactive venom into prey or use it to ward off predators (von Reumont et al. 2014). The venom produced by robber flies is the dipteran venom that is well described. Adult Asilidae pursue flying prey that is first rendered unconscious by venom before being eaten. Since thoracic gland extracts quickly paralyse target insects when injected, they are a plausible source of venom (Musso et al. 1978). Robber flies Eutolmus rufibarbis and Machimus arthriticus thoracic gland protein extracts and transcriptome analyses indicate that their venom contains a complex blend of proteins and peptides (Drukewitz et al. 2018). The non-asilid genomes contain 15 of the 30 predominate venom proteins, while the remaining 15 highly expressed venom proteins seem to be specific to robber flies (Drukewitz et al. 2019). The U-Asiliditoxin Mar-1a peptide from Machimus arthriticus has been shown to have neurotoxic action among these. The main structure of Mar-1a suggests that it adopts the ICK fold, which has been frequently utilised in animal venoms (Undheim et al. 2015). Mar-1a generated by solid-phase synthesis caused disorientation and paralysis when injected into the ocelli of bees, indicating that venom peptides contribute to the asilid venom’s reported neurotoxicity. The majority of the other proteins in venom, many of which have higher molecular masses, still have unclear activities. Comparatively few of the discovered proteins include enzyme annotations, as opposed to other species that use salivary venoms that may be used to liquefy prey (Walker et al. 2017). Many other asiloidean families’ larvae, including those of the Asilidae, are predatory. Possibly with the aid of salivary gland fluids pumped into the prey through grooves on the surface of the mouth hooks, these larvae kill animals swiftly and effectively (Sinclair 1992). Another dipteran group with highly bioactive venoms is the Tabanomorpha, which includes march flies, horse flies, and allies. In Tabanomorpha, as opposed to Asilidae, the adults consume nectar or blood while the larvae are venomous predators. Typically found in water or mud, tabanid larvae are fierce predators on tiny vertebrates like toads and invertebrates with strong chemical defences like bombardier beetles (Nowicki and Eisner 1983). Before becoming paralysed, prey impaled by the mandibles make a few furious movements (Teskey 1969). In addition, when larvae bite humans, it causes discomfort, irritability, and itching (Jackman et al. 1983). A tube in the mandible that exits near to the tip is used to administer venom. This canal connects to a gland in the head that is completely independent from the alimentary canal and is located at the anterior margin of the cibarial pump (Woodley 1989). Although Tabanomorpha and Asiloidea belong to the same basal lineage of the suborder Brachycera (Wiegmann and Yeates 2017), they inject their venom using different anatomical features and are poisonous at different periods of their lives. While larval (but not adult) Tabanomorpha inject venom through mandibular canals, adult Asilidae inject venom produced in thoracic glands through a channel in the hypopharynx. Due to the absence of these mandibular venom tubes in Rhagionidae, Tabanidae, Athericidae, and Pelecorhynchidae

432

13

Venomous and Other Body Fluids in Insect Predators

(Courtney et al. 2000), it is possible that Tabanomorpha only employ venom from these former families, and that their use of venom evolved independently of that of Asilidae. Alternatively, the usage of salivary gland-derived venom may be common among Brachycera predatory larvae, such as diverse Tabanomorpha, Asiloidea, and Empidoidea, but has branched into wildly dissimilar forms in Tabanidae and Asilidae.

13.6.5 Dermaptera Earwigs’ foul quinoid secretions can be released as spray or droplets, and the latter is enhanced by rotating the abdomen (Eisner et al. 2005). While the American species Doru taeniatum contains methyl- and 2,3-dimethyl-1,4-benzoquinones that are partially dissolved in an organic pentadecane-phase and in a co-occurring aqueous phase, the European earwig Forficula auricularia produces methyl- and ethyl-1,4benzoquinones along with their corresponding hydroquinones (Schildknecht and Weis 1960). Always, discharges are quinoid solutions (Eisner and Aneshansley 2000). Protective glands at the base of the larval abdomen are absent; hence, it is unknown if the paired larval glands at the base of the pincers are actual defensive glands (Vosseler 1890).

13.6.6 Neuroptera Numerous neuropteran species (Berothidae, Mantispidae, and Chrysopidae) lay their eggs on top of a thin stalk known as an egg pedicel, which is made of a gelatinous fluid produced by the female accessory glands. In some species, including Ceraeochrysa smithi, females secrete an oily secretion that contains oleic acid, butanal, decanal, pentadecanal, and isopropyl myristate and coats the stalks (Eisner et al. 1996, 2005). The eggs of spoon-winged lacewings (Nemopteridae) lay hard eggs that may be carried into their colonies by foraging, granivorous ants, where the lacewing larvae are apparently predatory. Antlion (Myrmeleontidae) eggs are coated with a sticky secretion, for adhesion to sand and soil. Silky lacewing (Psychopsidae) eggs are covered with a material that is (Tauber et al. 2009a, b). Another common neuropteran defence is the ability to lay eggs on stalks, sometimes in conjunction with stalk fluids. The oily egg stalk liquid is made up of a mixture of C4–28 straight-chain aldehydes, oleic and linoleic acids, and isopropyl myristate. Most Neuroptera larvae—including lacewings, antlions, owlflies, and allies—are poisonous predators of other insects (Tauber et al. 2009a, b). They adhere. Antlions are an example of a sit-and-wait predator whose larvae inject poison into their prey through their mouthparts to immobilise or kill it. They use a variety of hunting techniques, from active seeking to ambush, most notably pitfall traps, which were popularised by several antlions (Myrmeleontidae) and served as the model for the fictitious sarlacci in the Star Wars universe. When impaled by the pincer-like

13.6

Salivary Venom of Predators

433

mouthparts, prey typically stops moving within seconds and remains paralysed even after being promptly severed from the neuropteran larva (Canard 2001). Due to the fact that food can only be consumed in liquid form, venom is also thought to have a liquefying effect. Neuropteran larvae (such those of the Osmylidae, Chrysopidae, Ascalaphidae, Hemerobiidae, and Myrmeleontidae) regurgitate toxins into their food during capture to paralyse and kill it. Toxin generation was noticed in the larvae of ant lions (genus Myrmeleon; Myrmeleontidae), which feed on the liquefied internal organs of their insect prey. These cause the prey’s paralysis and eventual death and are obtained from both Myrmeleon-larvae and their bacterial symbionts. A paralytic 165–167 kDa polypeptide that was even more effective as a paralysing agent than tetrodotoxin was also shown to be a larval toxin (Yoshida et al. 1999). (Matsuda et al. 1995). Additionally, chemicals made by bacteria isolated and cultivated from the M. bore larval fore- and midguts, which are separate from the hindgut, were discovered. Molecular chaperone-producing Bacillus and Enterobacter species from insect larvae may create a paralysing toxin that is a homolog of GroEL, a protective heat shock protein (Yoshida et al. 2001). Current literature cannot clearly identify the origin of the poison that causes neuropteran larvae to paralyse their prey. On the one hand, the venom gland, a gland located within the maxilla, has been extensively characterised by several morphologists (Canard 2001; Randolf et al. 2014). These researchers came to the conclusion that the gland plays a role in prey capture based on the physical configuration of the gland and its delivery canal as well as the fact that it is present in venomous taxa but not non-venomous ones. Additionally, the symbiotic bacteria produce a 34 kDa insecticidal sphingomyelinase C, which most likely functions by degrading phospholipids (Nishiwaki et al. 2004). Last but not least, a Bacillus sphaericus isolate from Myrmeleon bore crops yielded the novel insecticidal 53 kDa toxin known as sphaericolysin (Nishiwaki et al. 2007). Additionally, it has been proposed that the larvae’s paralysing venom may originate in the alimentary canal. Feeding is ingested by larval neuropterans through a food canal made of their mandibles and maxillae interlocking. This food channel is a paired construction, unlike robber flies or assassin bugs, and each “pincer” is made up of the maxilla and mandible on opposite sides of the body. The food canal appears near the pincer’s tip (i.e. close to where the maxillary gland empties, in the part that penetrates prey). Different poisons and compounds have been identified from various predator components (Table 13.5).

13.6.7 Coleoptera Many lineages of coleopteran insects have been shown to have poisonous larvae, including the firefly (Lampyridae), diving beetles (Dytiscidae), whirligig beetles (Gyrinidae), ground beetles (Carabidae), tiger beetles (Cicindelidae), water beetles (Hydrophilidae), and (Silphidae). None of these groups, like other coleopteran larvae, have salivary glands; hence, it has been hypothesised that the alimentary

434

13

Venomous and Other Body Fluids in Insect Predators

Table 13.5 Various biochemicals isolated from different parts of the predators with suitable references Order Neuroptera

Forficula auricularia Podisus nigrispinus Podisus nigrispinus Neuroptera Fire ant

Body part involved Female accessory glands Saliva Salivary gland Saliva Larval body Venom

Chemicals secreted Oleic acid, butanal, decanal, Pentadecanal, isopropyl myristate

References Eisner et al. (1996, 2005)

Methyl-and ethyl-1,4benzoquinones N, N-dimethylaniline 1,2,5trithiepane N, N-dimethylaniline (DMA), 1,2,5-trithiepane ALMB-toxin Water-insoluble alkaloids

Schildknecht and Weis (1960) Martínez et al. (2016) and Campos et al. (2020) Campos et al. (2020) Yoshida et al. (1999) Lai et al. (2012)

canal or a nearby structure produces the venom that causes paralysis and death. If so, a single delivery tube is most likely used for both venom injection and food absorption. For Lampyridae, Dytiscidae, Gyrinidae, Carabidae, and Silphidae, there is greater proof of a true poisonous nature than for Hydrophilidae.

13.7

Other Fluids

Bioadhesives are released from glands on the fore- and mid-legs of harpactorine Reduviidae, which includes the vast and widespread genus Zelus (leafhopper bugs). The subfamily Harpactorinae of resin bugs, including the tribes Apiomerini, Diaspidiini, and Ectinoderini, collect sticky plant resins that are then applied to the forelegs to aid in prey acquisition. It has recently been determined that each group’s behaviour evolved independently. The oily substance produced by the cushion-like secretory tissue and the tiny tenet hairs on the tibial pads of the forelegs of Haematorrhophus nigroviolaceus, Pirates affinis, demonstrate that these traits increase the ability to sustain static tension and improve the gripping efficiency of the legs (Haridass and Ananthakrishnan 1980).

13.8

Biological Activities

As a defence strategy, venom can be transferred actively by stings or injections or passively through bristles, spines, or hairs after contact (Schmidt 1982). Venom is a toxin created by glands and injected into another organism using specific equipment. It can render the prey unconscious or kill them (Blum and Hermann 1978; Schmidt 1982). Lepidoptera, Hymenoptera, and Hemiptera species’ venoms have been isolated, identified, and tested on various insects and vertebrates (Sahayaraj and Muthukumar 2011). Several predatory Hemiptera species, including Belostoma

13.8

Biological Activities

435

lutarium (Belostomatidae), Deraeocoris nebulosus (Miridae), and Podisus maculiventris, have had their eating habits and extra-oral digestion examined (Pentatomidae). Pentatomidae predators pierce their prey’s skin with their mouthparts and inject saliva, which quickly paralyses and kills it (Cohen 1990), after which the body’s contents are consumed. Several insect species, notably Galerita lecontei (Coleoptera: Carabidae) and Polistes sulcifer (Hymenoptera: Vespidae), have been found to exhibit deadly and sub-lethal effects when exposed to venom components with low molecular weight and short carbon chains (Bruschini and Cervo 2011). In contrast to P. nigrispinus, these insects have specialised venom glands that manufacture venom instead of salivary glands. The primary purposes of the venoms produced by predatory cimicomorphans and pentatomomorphans are to immobilise, kill, and liquefy their invertebrate victims. Rapid paralysis is one of the main effects of venoms, and it has been observed in animals envenomated by Reduviidae, asopine Pentatomidae, predaceous Miridae (Cohen 1996), and predaceous Lygaeidae (Slater and Carayon 1963), but it may be far more common. We discovered that the assassin bug Pristhesancus plagipennis’ PMG poison quickly paralyses crickets within a few seconds (Walker et al. 2018b). This discovery along with the assassin bug venom’s great potency (Edwards 1961) and occasionally reversible paralytic effects (Zerachia et al. 1973) point to the presence of particular neurotoxins. However, Ptu1, an inhibitor cystine knot (ICK) peptide from the assassin bug Peirates turpis that blocks CaV2.2 voltage-gated calcium channels, is the sole heteropteran toxin with a recognised neurotoxic effect (Bernard et al. 2001; Corzo et al. 2001). Perhaps, as has been seen in Hymenoptera, membrane-disrupting poisons contribute to prey paralysis (Robinson et al. 2018) During the prey capture of the domestic cricket, Acheta domesticus, Orivel and Dejean (2001) examined the paralytic and fatal effects of the venom of 12 Pachycondyla generalist predators. Given the tight evolutionary relatedness of the species, it is remarkable that the observed values covered such a wide range. Despite being used for various objectives, these venoms had a similar physiological impact. They resulted in a swift, dosedependent, and reversible paralysis that was followed by a second, slower paralysis that, when it was finished, was permanent and caused death in less than 4 days. Given that necrosis was frequently seen in deceased animals, this finding raises the possibility that comparable poisons, as well as neurotoxins and histolytic substances, exist. Comparisons based on the species’ nesting locations revealed substantial differences in fatality and paralysis after 2 h, with the venoms of arboreal species being more effective than those of ground-dwelling species due to their higher potency and relatively quick action. This behaviour could be seen as an adaptation to living in trees, where there are more opportunities for prey to flee than there are on the ground or in the leaf litter. The liquefaction of prey tissues may be aided by other elements found in heteropteran venoms. Proteases from the C1A, A1A, and M12 families, particularly those from the prevalent S1 family, as well as enzymes like hyaluronidase, chitinase, and nuclease are likely to break down biopolymers and aid in the spread of additional

436

13

Venomous and Other Body Fluids in Insect Predators

toxins and liquefaction. Additionally, the discovery of venom proteins in both reduviids and belostomatids annotated as gelsolins and fasciclins/periostins raises the possibility of more specialised prey liquefaction mechanisms: By disassembling cytoskeletons, gelsolin, the most effective actin depolymerisation protein known (Sun et al. 2001), may help; fasciclins/periostins, which feature cell adhesion domains (Clout et al. 2003), may interfere with cell attachment. However, based only on their sequence, many of the most prevalent proteins in heteropteran venoms cannot have any plausible function assigned to them. These include the heteropteran venom proteins grouped into families 1–34 and the CUB domain proteins (Walker et al. 2017, 2018b). The majorities of these are not proteases and are found in the venom of P. plagipennis that is obtained through electrostimulation (Walker et al. 2017).

13.8.1 Pesticidal Activity While the investigated species’ anterior major gland (AMG) venom, which was obtained through harassment, did not paralyse prey, the posterior main gland (PMG) venom, which can be evoked by electrostimulation, did (Walker et al. 2018a, c). However, other assassin bug species also exhibited the paralysing effects of AMG venom (Haridass and Ananthakrishnan 1981). Arthropod venoms have insecticidal properties against numerous economically significant pests (Escoubas et al. 1995; Schwartz et al. 2012; Smith et al. 2013; Daly and Wilson 2018). When administered orally to Spodoptera litura, Rhynocoris fuscipes salivary venom dramatically reduced corrected mortality. In all categories, the mortality of the larvae was noted within 24 h. An extended period of exposure also revealed an increased mortality up to 96 h with a lower LD50 value of 861.60 ppm. Fifty percent of mortality was recorded even at 48 h of approximately 53.33% and 60.00% in the 800 and 1000 ppm (LD50 = 890.13 ppm/larva). However, a dose-dependent response was not seen in the oral toxicity bioassay at the 72-h period, where 50% mortality was recorded, and at the 96-h period, where the mortality was 21%, 43%, 35.71%, 50.00%, 71.43%, and 71.43% for concentrations of 200, 400, 600, 800, and 1000 ppm salivary venom, respectively. In the microinjection bioassay, the Helicovera armigera larva shows a somewhat greater mortality rate than in the oral administration bioassay. Those who received VS injections passed away within 24 h and had an LD50 value of 06.24 ppm. The LD50 value at 96 h was 846.35 ppm/ larva, with a maximum larval morality of 84.62%. At 96 h of observation, the LD50 result for the oral toxicity bioassay method was 899.91 ppm/larva, with 71.43% adjusted mortality (Vinoth Kanna 2011). Utilising microinjection and oral administration bioassays, the toxicological, macromolecular, enzymatic, nutritional, and inhibition of haemocyte aggregation and spreading effects of the salivary venom (SV) of the reduviid predator, Rhynocoris marginatus, were assessed against Spodoptera litura third instar larvae. To acquire various quantities, the salivary venom was collected using the milking method and diluted with HPLC grade water (200, 400, 600, 800 and 1000 ppm). For

13.8

Biological Activities

437

Table 13.6 LD50 and LT50 for various pests due to the venom activity of different predatory insects with references Predator Platymeris biguttatus

Podisus nigrispinus N,Ndimethylaniline Pachycondyla goeldii-G1

Pest Galleria mellonella larvae Anticarsia gemmatalis larvae Anticarsia gemmatalis larval Acheta domesticus

Pachycondyla goeldii-G3

Acheta domesticus

Pachycondyla goeldii-W3

Acheta domesticus

Pachycondyla goeldii-W4

Acheta domesticus

Solenopsis geminata

Spodoptera litura larvae Spodoptera litura larvae Spodoptera litura larvae Plutella xylostella Plutella xylostella Plutella xylostella

Podisus nigrispinus

Solenopsis invicta (M) Solenopsis invicta (P) Solenopsis geminata Solenopsis invicta (M) Solenopsis invicta (P)

LD50/LT50 20.4 μg/g

82.7 mg peptide/g animal weight 122.3 mg peptide/g animal weight 129.6 mg peptide/g animal weight 88.2 mg peptide/g animal weight 6.53 min (LT50)

References Gao et al. (2022) Martínez et al. (2016) Martínez et al. (2016) Orivel et al. (2001) Orivel et al. (2001) Orivel et al. (2001) Orivel et al. (2001) Lai et al. (2010)

14.25 min (LT50)

Lai et al. (2010)

29.52 min (LT50)

Lai et al. (2010)

7.28 min (LT50) 21.08 min (LT50) 32.88 min (LT50)

Lai et al. (2012) Lai et al. (2012) Lai et al. (2012)

2.04 μL 136.1 nL

microinjection and oral toxicity experiments, the LD50 values were 0.78 g/larvae and 0.86 g/larvae at 24 and 96 h, respectively (Muthu Kumar 2010). Table 13.6 presents the LD/LC50 and LT50 for various pests caused by the venom activity of several predatory insects.

13.8.2 Primary Metabolites Modulation Results from Muthu Kumar (2010) showed that the amount of the enzyme decreased in a dose-dependent manner. In the experimental category, protein content rises by 50% while the quantities of carbohydrates (21%) and lipids (46%) drop. In the oral toxicity research, the SV decreased Spodoptera litura relative growth rate, approximate digestibility, and efficiency of conversion of consumed and digested food. The outcome demonstrated that SV toxins induced death, affected Spodoptera litura macromolecular and digestive enzyme levels, and impacted nutritional indices. Salivary venom prevents haemoglobin from aggregating and alters the way haemoglobin spreads. It is possible to isolate responsible peptides and use them in the pest management programme (Muthu Kumar 2010).

438

13

Venomous and Other Body Fluids in Insect Predators

13.8.3 Immunomodulatory Activity The salivary venom-mixed Spodoptera litura haemolymph was less viscous than the control haemolymph, according to total haemoglobin testing. Helicoverpa armigera, however, exhibits a large increase in haemocyte death and the haemolymph’s colloidal character changes to one that is watery. As the incubation duration rose, the intact haemocyte population in the control decreased marginally, reaching 20.30% and 12.41%, respectively, at 60-min incubation in Spodoptera litura and Helicoverpa armigera. However, the haemocyte population gradually grew to 800 ppm (35.20%) of salivary venom. Haemolysis was seen in the salivary venom of Spodoptera litura and Helicoverpa armigera in the VS dose-dependent and timedependent component (Vinoth Kanna 2011). Six polypeptides with molecular weights ranging from 151 to 10 kDa were found after the salivary venom’s impact on the Spodoptera litura’s haemolymph protein profile was determined. Rhynocoris fuscipes was added at concentrations of 200 ppm (139 to 3 kDa), 400 ppm (140 to 3 kDa), 600 ppm (136 to 5 kDa), 800 ppm (140 to 4 kDa), and 1000 ppm (139 to 5 kDa) VS. At 400, 600, 800, and 1000 ppm concentrations of salivary venom-treated haemolymph, the polypeptide with a 60 kDa molecular weight was lost. The haemolymph of Helicoverpa armigera included six polypeptides, ranging in size from 184 to 21 kDa. It changed in the haemolymph after exposure to Rhynocoris fuscipes poisonous saliva at 200 ppm (179 to 10 kDa), 400 ppm (179 to 15 kDa), 600 ppm (171 to 10 kDa), 800 ppm (169 to 3 kDa), and 1000 ppm (176 to 19 kDa). Additionally, the 160 kDa polypeptide vanished in the salivary venom-treated groups, but a new polypeptide (21 kDa) developed at 400 ppm of Rhynocoris fuscipes salivary venom (Vinoth Kanna 2011).

13.8.4 Inhibition of Haemocyte Aggregation In the control group, throughout the 30-min period, about 66.67% of the plasmatocytes (PL) and 71.79% of the granulocytes (GC) of Spodoptera litura aggregated. The PL and GC aggregation was greatly reduced by the addition of VS during the 30 min in the 200 ppm, and it was further reduced in the 1000 ppm to 16.41% and 55.88% for PL and GC, respectively. Haemocyte aggregations in the 1000 ppm category considerably decreased in PL (45.2%) and GC (35.7%) after 240 min of incubation. The aggregation was much higher at 1000 ppm in the PL (33.33%) and GC than it was at 30 min in the PL (80.00%) and GC (83.33%). At 240 min, the aggregation in the 1000 ppm of PL (73.07%) and GC (66.67%) dramatically increased (Vinoth Kanna 2011). Sahayaraj and Muthukumar (2011) also reported that Rhynocoris marginatus venom had a similar effect on Spodoptera litura. Spodoptera litura haemoglobin aggregated in a monolayer after being treated by Rhynocoris marginatus venom. The plasma membrane and the enlarged pseudopods both broke down as a result.

13.8

Biological Activities

439

13.8.5 Spreading Inhibitory Behaviour About 85.7% of the PL (600 ppm) spread and 22.2% of the GC (1000 ppm) spread were inhibited after 30 min of incubation. The spreading inhibition rose in GC (57.79%) and reduced insignificantly in PL (30.73%) as the incubation duration was extended (to 240 min). Helicoverpa armigera PL and GC considerably increased the spreading inhibitory percentage, with 58.3% and 64.7% in 1000 ppm at 30-min inhibition. When the incubation time was extended to 240 min, the spreading inhibition percentage dramatically dropped (for PL and GC, this was 45.3% and 57.2%, respectively) (Vinoth Kanna 2011).

13.8.6 Anti-Microbial and Cytotoxic Activities At a concentration of 30 mg (dry weight of venom)/mL, the crude venom of the predatory ant Pachycondyla goeldii demonstrated potent activity against both Grampositive (S. aureus 209P) and Gram-negative (E. coli RL65) bacterial strains. Nevertheless, it exhibited little efficacy against the fungus Aspergillus niger and Botrytis cinerea (Orivel et al. 2001). By using bacterial growth inhibition experiments, Rhynocoris iracundusvenom demonstrated particular antibacterial efficacy against Escherichia coli, but not against Listeria grayi and Pseudomonas aeruginosa (Rügen et al. 2021). Two prey species, Spodoptera litura and Helicoverpa armigera, were tested for immunosuppressive activity by Rhynocoris fuscipes venomous saliva (VS). By evaluating changes in total haemocyte counts and haemocyte behaviours, immunosuppression was evaluated (aggregation and spreading). For Spodoptera litura and Helicoverpa armigera, newly ecdysed fifth stadium larvae showed total haemocyte counts of 1.285 × 107 cells/mL and 1.397 × 107 cells/mL, respectively. A concentration- and time-dependent decrease in haemocyte counts was seen for both prey species when isolated haemocytes were exposed to saliva from Rhynocoris fuscipes. Cell death thought to involve both cytolytic and apoptotic pathways were blamed for the decline. Both lepidopteron species’ haemocytes’ behaviour was modified by saliva. After being introduced into 96-well plates, plasmatocytes and granular cells treated with saline (PBS) or left untreated quickly produced aggregations that clung to and spread throughout the plastic surfaces. However, both cell types’ ability to aggregate and spread was decreased after being exposed to reduviid saliva, and these events were both concentration- and time-dependent (Ayyachamy et al. 2016). The Alamar blue assay, which is based on resazurin, was used to examine the cytotoxic effects of Rhynocoris iracundus venom on S2 cells. S2 cells’ viability was significantly reduced when co-incubated with 174 g/mL of venom, resulting in 99% cell death over the course of a 4-h incubation period. In as little as 30 s, S2 cells treated to 174 g/mL venom displayed morphological disruption by changing their form, resulting in lysis and cell death (Rügen et al. 2021).

440

13.9

13

Venomous and Other Body Fluids in Insect Predators

Physiological Role

1. Considering the proteinaceous nature of the zoophagous Heteroptera, the salivary glands (also known as venom glands) release proteins and peptides that aid in the capture of prey in addition to enzymes for the breakdown of animal tissue (Cohen 1998; Edwards 1961; Walker et al. 2018a, c). 2. Initially, rather than the action of neurotoxins, the fast paralysis of insects attacked by predatory assassin bugs [Rhinocoris carmelita and Platymeris rhadamanthus] was attributed to the rupture of cell membranes by digestive enzymes (Edwards 1961). 3. The majority of heteropterans feed on insect bodily fluid that has been liquefied by a predator (Undheim et al. 2017). 4. The Cav2.2 voltage-gated calcium channels are reversibly inhibited by the cystine knot (ICK) motif, which is inhibited by the peptide Ptu1 from Peirates turpis, Peirates plagipennis, and Platymeris rhadamanthus (Corzo et al. 2001). 5. Reduviid posterior major gland extracts failed to induce paralysis but resulted in death after a few hours, suggesting that the former venom is employed for prey immobilisation while the PMG secretes digesting enzymes. 6. Haemolysin-like proteins, protease inhibitors, and a number of new and uncharacterised proteins make up the majority of Peirates plagipennis (Walker et al. 2018a, c). 7. The venom of fire ants is haemolytic and releases histamine. 8. Primary metabolites like glucose, protein, and lipoproteins are modulated by venomous saliva.

13.10 Future Recommendations • The age of the individuals, food intake, and any other parameters that need to be examined for all predators can affect the insect saliva enzyme composition. • Venom and other fluids involved in paralysis should have their chemistry examined.

References Ambrose DP, Maran SPM (1999) Quantification, protein content and paralytic potential of saliva of fed and prey deprived reduviid Acanthaspis pedestris Stål (Heteroptera: Reduviidae: Reduviinae). Indian J Environ Sci 3(1):11–16 Arbuckle K (2017) Evolutionary context of venom in animals. In: Evolution of venomous animals and their toxins, vol 24. Springer, Cham, pp 3–31 Ayyachamy VK, Sahayaraj K, Rivers DB (2016) Anti-aggregation and cytolytic behaviour of venomous saliva of Rhynocoris fuscipes (Fab.)(Hemiptera: Reduviidae) in response to its prey hemocytes. J Entomol Res Soc 18(3):1–13

References

441

Azevedo DDO, Zanuncio JC, Zanuncio JS Jr, Martins GF, Marques-Silva S, Sossai MF, Serrão JE (2007) Biochemical and morphological aspects of salivary glands of the predator Brontocoris tabidus (Heteroptera: Pentatomidae). Braz Arch Biol Technol 50(3):469–477 Balduf WV (1935) The bionomics of entomophagous coleoptera. J. Swift, St. Louis Barbosa SE, Diotaiuti L, Soares RPP, Pereira MH (1999) Differences in saliva composition among three Brazilian populations of Panstrongylus megistus (Hemiptera, Reduviidae). Acta Trop 72 (1):91–98 Bell HA, Down RE, Edwards JP, Gatehouse JA, Gatehouse M (2005) Digestive proteolytic activity in the gut and salivary glands of the predatory bug Podisus maculiventris (Heteroptera: Pentatomidae); effect of proteinase inhibitors. Eur J Entomol 102(2):139 Bernard C, Corzo G, Mosbah A, Nakajima T, Darbon H (2001) Solution structure of Ptu1, a toxin from the assassin bug Peirates turpis that blocks the voltage-sensitive calcium channel N-type. Biochemistry 40:12795–12800 Blum S, Hermann HR (1978) Venoms and venom apparatuses of the Formicidae: Dolichoderinae and Aneuretinae. In: Arthropod venoms. Springer, Berlin, Heidelberg, pp 871–894 Blum MS, Walker JR, Callahan PS, Novak AF (1958) Chemical, insecticidal, and antibiotic properties of fire ant venom. Science 128(3319):306–307 Boyd DW Jr (2003) Digestive enzymes and stylet morphology of Deraeocoris nigritulus (Uhler) (Hemiptera: Miridae) reflect adaptations for predatory habits. Ann Entomol Soc Am 96(5):667– 671 Boyd DW Jr, Cohen AC, Alverson DR (2002) Digestive enzymes and stylet morphology of Deraeocoris nebulosus (Hemiptera: Miridae), a predacious plant bug. Ann Entomol Soc Am 95(3):395–401 Bruschini C, Cervo R (2011) Venom volatiles of the paper wasp social parasite Polistes sulcifer elicit intra-colonial aggression on the nest of the host species Polistes dominulus. Insect Soc 58:383–390. https://doi.org/10.1007/s00040-011-0155-0 Calvete JJ, Sanz L, Angulo Y, Lomonte B, Gutiérrez JM (2009) Venoms, venomics, antivenomics. FEBS Lett 583(11):1736–1743 Campos JM, Martinez LC, Plata-Rueda A, Weigand W, Zanuncio JC, Serrão JE (2020) Insecticide potential of two saliva components of the predatory bug Podisus nigrispinus (Heteroptera: Pentatomidae) against Spodoptera frugiperda (lepidoptera: Noctuidae) caterpillars. Toxin Rev 41:1–12 Canard M (2001) Natural food and feeding habits of lacewings. In: Whittington AE, PK ME, New TR (eds) Lacewings in the crop environment. Cambridge University Press, Cambridge, pp 116–129 Clout NJ, Tisi D, Hohenester E (2003) Novel fold revealed by the structure of a FAS1 domain pair from the insect cell adhesion molecule fasciclin I. Structure 11(2):197–203 Cohen AC (1990) Feeding adaptations of some predaceous Hemiptera. Ann Entomol Soc Am 83(6):1215–1223 Cohen AC (1993) Organization of digestion and preliminary characterization of salivary trypsinlike enzymes in a predaceous heteropteran Zelus renardii. J Insect Physiol 39:823–829 Cohen AC (1995) Review: extra-oral digestion in predaceous terrestrial arthropod. Ann Rev Entomol 40:85–103 Cohen AC (1996) Plant feeding by predatory Heteroptera: evolutionary and adaptational aspects of trophic switching. In: Alomar O (ed) Zoophytophagous heteroptera: implication for life history and integrated pest management. Thomas Say Publications in Entomology, Lanham, MD, pp 1–7 Cohen AC (1998) Solid-to-liquid feeding: the inside (s) story of extra-oral digestion in predaceous Arthropoda. Am Entomol 44(2):103–117 Cohen AC (2000) Feeding fitness and quality of domesticated and feral predators, effects of longterm rearing on artificial diet. Biol Control 17(1):50–54

442

13

Venomous and Other Body Fluids in Insect Predators

Colebatch G, East P, Cooper P (2001) Preliminary characterisation of digestive proteases of the green mirid, Creontiades dilutes (Hemiptera: Miridae). Insect Biochem Mol Biol 31 (4–5):415–423 Corzo G, Adachi-Akahane S, Nagao T, Kusui Y, Nakajima T (2001) Novel peptides from assassin bugs (Hemiptera: Reduviidae): isolation, chemical and biological characterization. FEBS Lett 499(3):256–261 Courtney WG, Sinclair BJ, Meier R (2000) Morphology and terminology of Diptera larvae. In: Papp L, Darvas B (eds) Contributions to a manual of Palaearctic diptera. Science Herald, Budapest, pp 85–163 Daly NL, Wilson D (2018) Structural diversity of arthropod venom toxins. Toxicon 152:46–56 Drukewitz S, Fuhrmann N, Undheim E, Blanke A, Giribaldi J, Mary R et al (2018) A dipteran’s novel sucker punch: evolution of arthropod atypical venom with a neurotoxic component in robber flies (Asilidae, Diptera). Toxins 10:29 Drukewitz SH, Bokelmann L, Undheim EA, von Reumont BM (2019) Toxins from scratch? Diverse, multimodal gene origins in the predatory robber fly Dasypogon Diadema indicate a dynamic venom evolution in dipteran insects. GigaScience 8(7):giz081 Edwards JS (1961) The action and composition of the saliva of an assassin bug Platymeris rhadamanthus Gaerst. (Hemiptera, Reduviidae). J Exp Biol 38(1):61–77 Eisner T, Aneshansley DJ (2000) Defense by foot adhesion in a beetle (Hemisphaerota cyanea). Proc Natl Acad Sci 97(12):6568–6573 Eisner T, Attygalle AB, Conner WE, Eisner M, MacLeod E, Meinwald J (1996) Chemical egg defense in a green lacewing (Ceraeochrysa smithi). Proc Natl Acad Sci U S A 93:3280–3283 Eisner T, Eisner M, Siegler M (2005) Secret weapons. Belknap Press, Cambridge Eldefrawi AT, Eldefrawi ME, Konno K, Mansour NA, Nakanishi K, Oltz E, Usherwood PN (1988) Structure and synthesis of a potent glutamate receptor antagonist in wasp venom. Proc Natl Acad Sci 85(13):4910–4913 Escoubas P, Palma MF, Nakajima T (1995) A microinjection technique using Drosophila melanogaster for bioassay-guided isolation of neurotoxins in arthropod venoms. Toxicon 33 (12):1549–1555 Evangelin G, Bertrand H, Muthupandi M, John WS (2014) Venomous saliva of non-haematophagous reduviid bugs (Heteroptera: Reduviidae): a review. Biolife 2:615–626 Feás X, Vidal C, Vázquez-Tato MP, Seijas JA (2021) Asian hornet, Vespa velutina Lepeletier 1836 (Hym.: Vespidae), venom obtention based on an electric stimulation protocol. Molecules 27 (1):138 Fischer ML, Wielsch N, Heckel DG, Vilcinskas A, Vogel H (2020) Context-dependent venom deployment and protein composition in two assassin bugs. Ecol Evol 10(18):9932–9947 Fox EG, Xu M, Wang L, Chen L, Lu YY (2018) Speedy milking of fresh venom from aculeate hymenopterans. Toxicon 146:20–123 Fry BG, Roelants K, Champagne DE, Scheib H, Tyndall JDA, King GF, Nevalainen TJ, Norman JA, Lewis RJ, Norton RS, Renjifo C, de la Vega RCR (2009) The toxicogenomic multiverse: convergent recruitment of proteins into animal venoms. Annu Rev Genomics Hum Genet 10: 483–511 Gao F, Tian L, Li X, Zhang Y, Wang T, Ma L, Li H (2022) Proteotranscriptomic analysis and toxicity assay suggest the functional distinction between venom gland chambers in twin-spotted assassin bug, Platymeris biguttatus. Biology 11(3):464 Gaumont J (1976) L’appareil digestif des larves de Planipennes. Ann Sci Nat Zool Biol Anim 18: 145–250 Haddad V, Schwartz EF, Schwartz CA, Carvalho LN (2010) Bites caused by giant water bugs belonging to Belostomatidae family (Hemiptera, Heteroptera) in humans: a report of seven cases. Wilderness Environ Med 21:130–133 Haridass ET, Ananthakrishnan TN (1980) Functional morphology of the fossula spongiosa in some reduviids (Insecta—Heteroptera—Reduviidae). Proc Indian Acad Sci 89:457–466

References

443

Haridass ET, Ananthakrishnan TN (1981) Functional morphology of pylorus and rectal glands in Reduviidae (Insecta—Heteroptera). Proc Anim Sci 90:483–493 Hedin PA, Miles LR, Thompsonand AC, Minyard JP (1968) Constituents of cotton bud. Formulation of boll weevil feeding stimulant mixtures. J Agric Food Chem 16(3):505–513 Henry CS (1977) The behaviour and life histories of two north American ascalaphids. Ann Entomol Soc Am 70:179–195 Hyodo D, Himuro C, Fujisaki K (2014) Prey size affects the costs and benefits of group predation in nymphs of the predatory stink bug Andrallus spinidens (Heteroptera: Pentatomidae). J Ethol 32:173–178 Jackman R, Nowicki S, Aneshansely DJ, Eisner T (1983) Predatory capture of toads by 876 fly larvae. Science 222:515–516 Kahan D (1964) The toxic effect of the bite and the proteolytic activity of the saliva and stomach contents of the robber flies (Diptera: Asilidae). Isr J Zool 13:47–57 Kou F, Li H, Li S, Xun H, Zhang Y, Sun Z, Cai W (2017) Temporal transcriptomic profiling of the ant-feeding assassin bug Acanthaspis cincticrus reveals a biased expression of genes associated with predation in nymphs. Sci Rep 7(1):1–11 Kumar SM, Sahayaraj K, Cohen A (2012) Gross morphology and histology of head and salivary apparatus of the predatory bug, Rhynocoris marginatus. J Insect Sci 12(1):19 Lai LC, Huang RN, Wu WJ (2008) Venom alkaloids of monogyne and polygyne forms of the red imported fire ant, Solenopsis invicta in Taiwan. Insect Soc 55(4):443–449 Lai LC, Chang YYHKH, Wu WJ, Huang RN (2010) Comparative toxicity of three fire ant (Hymenoptera: Formicidae) venoms to Spodoptera litura larvae. Sociobiology 56(3):653 Lai LC, Kuo TC, Huang RN, Wu WJ (2012) The insecticidal activities of fire ant (Hymenoptera: Formicidae) venoms against Plutella xylostella (lepidoptera: Plutellidae) larvae. J Econ Entomol 105(5):1591–1596 MacConnell JG, Blum MS, Fales HM (1971) The chemistry of fire ant venom. Tetrahedron 27(6): 1129–1139 Maran SP (2000) Chosen reduviid predators prey interaction: nutritional and kairomonal chemical ecology (Insecta: Heteroptera: Reduviidae). Ph.D. thesis. Tirunelveli, India: Manonmaniam Sundaranar University Maran SPM, Selvamuthu K, Rajan K, Kiruba DA, Ambrose DP (2011) In: Ambrose DP (ed) The salivary protein profile and paralytic potential of three species of Rhynocoris (Hemiptera: Reduviidae) to three insect pests. Insect pest management, a current scenario. Entomology Research Unit, Palayamkottai, pp 346–361 Martínez LC, do Carmo Queiroz Fialho M, Zanuncio JC, Serrão JE (2014) Ultrastructure and cytochemistry of salivary glands of the predator Podisus nigrispinus (Hemiptera: Pentatomidae). Protoplasma 251(3):535–543 Martínez LC, Fialho MDCQ, Barbosa LCA, Oliveira LL, Zanuncio JC, Serrão JE (2016) Stink bug predator kills prey with salivary non-proteinaceous compounds. Insect Biochem Mol Biol 68: 71–78 Matsuda K, Suzuki H, Nakanishi F, Shio K, Komai K, Nishimura K (1995) Purifi cation and characterization of a paralytic polypeptide from larvae of Myrmeleon bore. Biochem Biophys Res Commun 215:167–171 Moore EL, Haspel G, Libersat F, Adams ME (2006) Parasitoid wasp sting: a cocktail of GABA, taurine, and β-alanine opens chloride channels for central synaptic block and transient paralysis of a cockroach host. J Neurobiol 66(8):811–820 Musso JJ, Garnier R, Legier F (1978) Comparison of toxicity of venom of some asilids 976 (Diptera: Brachycera) on locusts. Ann Soc Entomol Fr 14:177–184 Muthu Kumar S (2010) Rhynocoris marginatus fab. Salivary venom characterization and its zootoxic effects on Spodoptera litura_Fab. Ph.D. thesis, Manonmaniam Sundaranar University, p 139

444

13

Venomous and Other Body Fluids in Insect Predators

Muthukumar S (2011) Rhynocoris marginatus (Fab.) Salivary Venom Characterization and its Zootoxic effects on Spodoptera litura (Fab.). Ph. D. Thesis, Manonmanium Sundaranar University, Tirunelveli, India, pp 15–42 Nelsen DR, Nisani Z, Cooper AM, Fox GA, Gren ECK, Corbit AG, Hayes WK (2014) Poisons, toxungens and venoms: redefining and classifying toxic biological secretions and the organisms that employ them. Biol Rev 89:450–465 Nishiwaki H, Ito K, Otsuki K, Yamamoto H, Komai K, Matsuda K (2004) Purifi cation and functional characterization of insecticidal sphingomyelinase C produced by Bacillus cereus. Eur J Biochem 271:601–606 Nishiwaki H, Nakashima K, Ishida C, Kawamura T, Matsuda K (2007) Cloning, functional characterization, and mode of action of a novel insecticidal pore-forming toxin, sphaericolysin, produced by Bacillus sphaericus. Appl Environ Microbiol 73:340–3411 Nowicki S, Eisner T (1983) Predatory capture of bombardier beetles by a tabanid fly larva. Psyche 90(1–2):119–122 Oliveira JA, Oliveira MGA, Guedes RNC, Soares MJ (2006) Morphology and preliminary enzyme characterization of the salivary glands from the predatory bug Podisus nigrispinus (Heteroptera: Pentatomidae). Bull Entomol Res 96(3):251–258 Orivel J, Dejean A (2001) Comparative effect of the venoms of ants of the genus Pachycondyla (Hymenoptera: Ponerinae). Toxicon 39(2-3):195–201 Orivel J, Redeker V, Le Caer JP, Krier F, Revol-Junelles AM, Longeon A, Rossier J (2001) Ponericins, new antibacterial and insecticidal peptides from the venom of the ant Pachycondyla goeldii. J Biol Chem 276(21):17823–17829 Piek T, Spanjer WILLEM (1986) Chemistry and pharmacology of solitary wasp venoms. In: Venoms of the Hymenoptera: biochemical, pharmacological and behavioural aspects. Elsevier, Amsterdam, pp 161–307 Piek T, Spanjer W, Veldsema-Currie RD, Van Groen T, De Haan N, Mantel P (1982) Effect of venom of the digger wasp Philanthus triangulum F. on the sixth abdominal ganglion of the cockroach. Comp Bioch Physiol C Comp Pharmacol 71(2):159–164 Pires EM, Nogueira RM, Serrão JE, Zanuncio JC (2022) Attack efficacy by Podisus nigrispinus and Brontocoris tabidus on Thyrinteina arnobia with different pre-release strategies. https://doi.org/ 10.21203/rs.3.rs-1497570/v1 Pradhan S (1939) Glands in the head capsule of coccinellid beetles with a discussion on some aspects of gnathal glands. J Morphol 64(1):47–66 Prem Mathi Maran S (1999) Chosen reduviid predators-prey interaction nutritional and pheromonal chemical ecology (Insecta: Heteroptera: Redividae), Doctoral dissertation, Manonmaniam Sundaranar University, Tamil Nadu Quistad GB, Nguyen Q, Bernasconi P, Leisy DJ (1994) Purification and characterization of insecticidal toxins from venom glands of the parasitic wasp, Bracon hebetor. Insect Biochem Mol Biol 24(10):955–961 Randolf S, Zimmermann D, Aspöck U (2014) Head anatomy of adult Nevrorthus apatelios and basal splitting events in Neuroptera (Neuroptera: Nevrorthidae). Arthropod Syst Phylogeny 72:111–136 Rivers DB, Uckan F, Ergin E (2006) Characterization and biochemical analyses of venom from the ectoparasitic wasp Nasonia vitripennis (Walker)(Hymenoptera: Pteromalidae). Arch Insect Biochem Physiol 61(1):24–41 Robinson SD, Mueller A, Clayton D, Starobova H, Hamilton BR, Payne RJ, Undheim EA (2018) A comprehensive portrait of the venom of the giant red bull ant, Myrmecia gulosa, reveals a hyperdiverse hymenopteran toxin gene family. Sci Adv 4(9):eaau4640 Rügen N, Jenkins TP, Wielsch N, Vogel H, Hempel BF, Süssmuth RD, Tonk M (2021) Hexapod assassins’ potion: venom composition and bioactivity from the Eurasian assassin bug Rhynocoris iracundus. Biomedicine 9(7):819 Sahayaraj K, Balasubramanian R (2016) Gut enzyme profile. In: Artificial rearing of reduviid predators for pest management. Springer, Singapore, pp 103–120

References

445

Sahayaraj K, Kanna AV (2009) Starvation impact on venom quantity of a reduviid predator, Catamirus brevipennis (servile) (Hemiptera: Reduviidae). Entomon 34(2):119–121 Sahayaraj K, Muthukumar S (2011) Zootoxic effects of reduviid Rhynocoris marginatus (fab.) (Hemiptera: Reduviidae) venomous saliva on Spodoptera litura (fab.). Toxicon 58:415–425 Sahayaraj K, Vinothkanna A (2011) Insecticidal activity of venomous saliva from Rhynocoris fuscipes (Reduviidae) against Spodoptera litura and Helicoverpa armigera by microinjection and oral administration. J Venom Anim Toxins incl Trop Dis 17:486–490 Sahayaraj K, Borgio JF, Muthukumar S, Anandh GP (2006a) Antibacterial activity of Rhynocoris marginatus (fab.) and Catamirus brevipennis (servile)(Hemiptera: Reduviidae) venoms against human pathogens. J Venom Anim Toxins incl Trop Dis 12:487–496 Sahayaraj K, Kumar SM, Anandh GP (2006b) Evaluation of milking and electrict shock methods for venom collection from hunter reduviids. Entomon 31(1):65–68 Sahayaraj K, Sankaralinkam K, Balasubramanian S (2007) Prey influence on the salivary gland and gut enzymes qualitative profile of Rhynocoris marginatus (fab.) and Catamirus brevipennis (Serville) (Heteroptera: Reduviidae). J Entomol 4:331–336 Sahayaraj K, Kanna AV, Kumar SM (2010) Gross morphology of feeding canal, salivary apparatus and digestive enzymes of salivary gland of Catamirus brevipennis (servile)(Hemiptera: Reduviidae). J Entomol Res Soc 12(2):37–50 Sahayaraj K, Muthukumar S, Rivers D (2013a) Biochemical and electrophoretic analyses of saliva from the predatory reduviid species Rhynocoris marginatus (fab.). Acta Biochim Pol 60(1):91 Sahayaraj K, Subramanium M, Rivers D (2013b) Biochemical and electrophoretic analyses of saliva from the predatory reduviid species Rhynocoris marginatus (fab.). Acta Biochim Pol 60:91–97 Schildknecht H, Weis KH (1960) VI. Mitteilung über Insektenabwehrstoffe* Zur Kenntnis des Pygidialdrüsen-Sekretes vom gemeinen Ohrwurm, Forficula auricularia. Z Für Naturforsch B 15(11):755–757 Schmidt JO (1982) Biochemistry of insect venoms. Annu Rev Entomol 27(1):339–368 Schwartz EF, Mourão CB, Moreira KG, Camargos TS, Mortari MR (2012) Arthropod venoms: a vast arsenal of insecticidal neuropeptides. Pept Sci 98(4):385–405 Sinclair BJ (1992) A phylogenetic interpretation of the Brachycera (Diptera) based on 1081 the larval mandible and associated mouthpart structures. Syst Entomol 17:233–252 Slater JA, Carayon J (1963) Ethiopian Lygaeidae IV: a new predatory lygaeid from Africa with a discussion of its biology and morphology (Hemiptera: Heteroptera). Proc R Entomol Soc Lond A38:1–11 Smith JJ, Herzig V, King GF, Alewood PF (2013) The insecticidal potential of venom peptides. Cell Mol Life Sci 70(19):3665–3693 Stanic M (1956) Allergenic properties of venom hyper-sensitiveness in man and animals. In: Buckley EE, Proges NW (eds) Venoms. American Association for the Advancement of Science, Washington, DC Sun D, Leung CL, Liem RK (2001) Characterization of the microtubule binding domain of microtubule actin crosslinking factor (MACF): identification of a novel group of microtubule associated proteins. J Cell Sci 114(1):161–172 Tauber CA, Tauber MJ, Albuquerque GS (2009a) Neuroptera: (lacewings, antlions). In: Resh VH, Cardé RT (eds) Encyclopedia of insects, 2nd edn. Academic, San Diego, pp 695–707 Tauber CA, Tauber MJ, Albuquerque GS (2009b) Neuroptera: (lacewings, antlions), 1103. In: Resh VH, Cardé RT (eds) Encyclopedia of insects, 2nd edn. Academic, San Diego, pp 695–707 Teskey HJ (1969) Larvae and pupae of some north American Tabanidae (Diptera). Mem Entomol Soc Can 101:5–147 Touchard A, Labrière N, Roux O, Petitclerc F, Orivel J, Escoubas P, Dejean A (2014) Venom toxicity and composition in three Pseudomyrmex ant species having different nesting modes. Toxicon 88:67–67 Undheim EA, Mobli M, King GF (2015) Toxin structures as evolutionary tools: using conserved 3D folds to study the evolution of rapidly evolving peptides. BioEssays 38(1122):539–548

446

13

Venomous and Other Body Fluids in Insect Predators

Undheim EA, Fry BG, King GF (2017) Melt with this kiss: paralyzing and liquefying venom of the assassin bug Pristhesancus plagipennis (Hemiptera: Reduviidae). Mol Cell Proteomics 16: 552–566 Vinoth Kanna A (2011) Rhinoceros fuscipes fab venomous saliva biological and immunomodulatory activity against insect pest and mice. Ph.D. thesis, Manonmaniam Sundaranar University, p 74 von Reumont BM, Blanke A, Richter S, Alvarez F, Bleidorn C, Jenner RA (2014) The first venomous crustacean revealed by transcriptomics and functional morphology: remipede venom glands express a unique toxin cocktail dominated by enzymes and a neurotoxin. Mol Biol Evol 31(1):48–58 Vosseler J (1890) Die Stinkdrüsen der Forflculiden. Arch Mikrosk Anat 36(1):565–578 Walker AA, Weirauch C, Fry BG, King GF (2016) Venoms of heteropteran insects: a 1139 treasure trove of diverse pharmacological toolkits. Toxins 8:4 Walker AA, Madio B, Jin J, Undheim EA, Fry BG, King GF (2017) Melt with this kiss: paralyzing and liquefying venom of the assassin bug Pristhesancus plagipennis (Hemiptera: Reduviidae). Mol Cell Proteomics 16(4):552–566 Walker AA, Mayhew ML, Jin J, Herzig V, Undheim EAB, Sombke A, Fry BG, Meritt DJ, King GF (2018a) The assassin bug Pristhesancus Plagipennis produces two distinct venoms in separate gland lumens. Nat Commun 9:755 Walker AA, Robinson SD, Yeates DK, Jin J, Baumann K, Dobson J, Fry BG, King GF (2018b) Entomo-venomics: the evolution, biology and biochemistry of insect venoms. Toxicon 154:15– 27. https://doi.org/10.1016/j.toxicon.2018.09.004 Walker AA, Mayhew ML, Jin J, Herzig V, Undheim EAB, Sombke A et al (2018c) 1136 the assassin bug Pristhesancus plagipennis produces two distinct venoms in separate 1137 gland lumens. Nat Commun 9:755 Weisel-Eichler A, Haspel G, Libersat F (1999) Venom of a parasitoid wasp induces 1141 prolonged grooming in the cockroach. J Exp Biol 202:957–964 Whitfield FGS (1925) The relation between the feeding-habits and the structure of the mouth-parts in the Asilidæ (Diptera). Proc Zool Soc Lond 95:599–638 Wiegmann BM, Yeates DK (2017) Phylogeny of Diptera. In: Kirk-Spriggs AH, Sinclair BJ (eds) Manual of Afrotropical diptera, pp 253–256 Wiegmann BM, Trautwein MD, Winkler IS, Barr NB, Kim J-W, Lambkin C et al (2011) Episodic radiations in the fly tree of life. Proc Natl Acad Sci U S A 108:5690 Woodley NE (1989) Phylogeny and classification of the “Orthorrhaphous” Brachycera. In: McNpine JF, Wood DM (eds) Manual of nearetic diptera. Ottawa Agric Can, Ottawa, pp 1371–1395 Yoshida NH, Sugama S, Gotoh K, Matsuda K, Nishimura, Komai K (1999) Detection of ALMBtoxin in the larval body of myrmeleon bore by anti-N-terminus peptide antibodies. Biosci Biotechnol Biochem 63:232–234 Yoshida N, Oeda K, Watanabe E, Mikami T, Fukita Y, Nishimura K, Komai K, Matsuda K (2001) Chaperonin turned insect toxin. Nature 411:44 Zerachia T, Bergmann F, Shulov A (1973) Pharmacological activities of the venom of the predaceous bug Holotrichius innesi H. (Heteroptera, Reduviidae). In: Kaiser E (ed) Animal and plant toxins, vol 143–146. Goldman, Munich Zibaee A, Hoda H, Fazeli-Dinan M (2012) Role of proteases in extra-oral digestion of a predatory bug, Andrallus spinidens. J Insect Sci 12(1):5

Prey Record of Various Predators

14

Contents 14.1 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.2 Orthoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.3 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.4 Hymenoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.5 Diptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.6 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.7 Hetroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.8 Future Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

448 460 462 466 470 473 476 478 501

Many field studies reveal the predator’s prey record, which is available in Chap. 2. However, for the easy understanding of international readers, in this chapter, a list of predator’s names, its family, location, prey record, its stage preference with citations are presented in Tables 14.1, 14.2, 14.3, 14.4, 14.5, 14.6, 14.7, 14.8, 14.9, 14.10, 14.11 and 14.12 below.

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_14

447

Coleoptera

Staphylinidae Staphylinidae

Staphylinidae

Staphylinidae

Staphylinidae

Staphylinidae

Dalotia coriaria Oligota kashmirica benefica

Oligota oviformis

Oligota pygmaea

Paederus fuscipes

Paederus fuscipes

Lab

Indonesia

India

Spodoptera frugiperda

Nilaparvata lugens

Oligonychus coffeae

Stethorus

Frankliniella occidentalis Tetranychus urticae

Bradysia coprophila

Rhagoletis pomonella

Anastrephafraterculus

Tetranychus urticae

Musca domestica, Carpophilus hemipterus

Prey Hylemya brassicae

4th and 5th instar nymphs Eggs, first instar larvae





Adults Larve

Larvae





Prey stage Eggs and larvae Eggs and early instars –

Maruthadurai et al. (2022)

García et al. (2012) Silva et al. (1968) Allen and Hagley (1990) Echegaray et al. (2015) Li et al. (2020) Shimoda et al. (1997) Badgley and Fleschner (1956) Perumalsamy et al. (2009) Karindah (2011)

Miller and Williams (1983)

References Read (1962)

14

USA

Lab Japan

USA

Lab

Staphylinidae

Staphylinidae

Belonochus rufipennis

Lab

Dalotia coriaria

Staphylinidae

Atheta coriaria

USA

Lab

Staphylinidae

Atheta coriaria

Location Lab

Dinothenarus badipes

Family Staphylinidae

Predator Aleochara bilineata

Table 14.1 Prey record and their preferred stages of pests by various Coleopteran predators

14.1

448 Prey Record of Various Predators

Staphylinidae

Staphylinidae

Staphylinidae

Staphylinidae

Staphylinidae

Staphylinidae

Staphylinidae

Staphylinidae

Staphylinidae

Staphylinidae Staphylinidae

Paederus fuscipes

Paederus memnonius

Paederus littorarius

Philonthus flavolimbatus

Philonthus longicornis

Philonthus longicornis

Philonthus sericans

Philonthus hepaticus

Philonthus ventralis

Paederus fuscipes Platystethus cornutus, Atheta (=Xenota) mucronata, Anotilus inustus, Gauropterus fulgidus, Cordalia obscura, Ocypus olens, Lepidophallus hesperius, Tasgius (=Paratasgius) ater, Phloenomus minimus

Malaysia Citrus orchards

USA

USA

USA

Egypt

USA

USA

Lab

Egypt

Manisa

Rice pests Ceratitis capitata

Haematobia irritans

Haematobia irritans

Haematobia irritans

Diptera, Coleoptera and Lepidoptera

Haematobia irritans

Haematobia irritans

Agrotis ipsilon, Pegomyia mixta, Scrobipalpa ocellatella, Cassida vittata, Lixus junci Chrysoteuchia topiaria

Aphis illinoisensis, Aphis gossypii

– Pupae

Eggs

Eggs

Eggs

Young ones

Eggs

Eggs

Eggs





Coleoptera (continued)

Haase-Statz (1997) Hu and Frank (1997) Hu and Frank (1997) Abd-Elgayed and El-Khouly (2019) Hu and Frank (1997) Hu and Frank (1997) Hu and Frank (1997) Manley (1977) Urbaneja et al. (2006)

Anlaş et al. (2021) Hassan (2021)

14.1 449

Staphylinidae

Atheta coriaria

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae Carabidae

Abax ater

Agonum dorsale

Agonum dorsale

Agonum dorsale

Amara aenea

Amara aenea

Anisodactylus sanctaecrucis

Bembidion lampros

Bembidion guttula

Calosoma sycophanta Calosoma calidum

Carabidae

Family Staphylinidae

Predator Tachyporus hypnorum

Table 14.1 (continued)

Lab Lab

Lab

Lab

Lab

Metopolophium dirhodum, Heteromurus nitidus Lymantria dispar Rhagoletis pomonella

Codling moth Apple maggot Cabbage root fly

Codling moth Apple maggot Rhagoletis pomonella

Rhopalosiphum padi

Mamestra brussicae

Soil-dwelling and Litter-dwelling prey Mamestra brussicae

Nilaparvata lugens

Prey Rhopalosiphum padi, Sitobion avenae Eggs of nitidulids

Pupae

Larvae pupae Immature stages –

Larvae pupae

Larvae

Larvae

Fourth and fifth instar nymphs

Prey stage –

Guillemain et al. (1997) Vasconcelos et al. (1996) Vasconcelos et al. (1996) Bilde and Toft (1994) Hagley et al. (1982) Allen and Hagley (1990) Hagley et al. (1982) Coaker and Williams (1963) Mundy et al. (2000) Weseloh (1988) Allen and Hagley (1990)

References Kyneb and Toft (2004) El-Shafie et al. (2017) Karindah (2011)

14

Lab

Lab

lab

Lab

Lab

Lab

Indonesia

Location Denmark

450 Prey Record of Various Predators

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Carabidae

Harpalus rufipes

Harpalus affinis

Loricera pilicornis

Nebria brevicollis

Notomus gravis

Ophionea nigrofasciata

Poecilus cupreus

Prerostichus melanarius

Pterostichus melanarius

Poecilus cupreus

Poecilus cupreus

Pterostichus cupreus

Pterostichus melanarius

Pseudophonus rufipes

Lab

Lab

Lab

Lab

Lab

Lab

Lab

Lab

Lab

Australia

Lab

Belgium

Lab

Lab

Ceratitis capitata

Metopolophium dirhodum, Heteromurus nitidus Deroceras reticulatum

Arion lusitanicus, Deroceras reticulatum Arion lusitanicus, Deroceras reticulatum Spodoptera littoralis

Rhopalosiphum padi, Acheta domestica Mamestra brussicae

Rivula atimeta

Deroceras reticulatum

Helicoverpa armigera

Codling moth Apple maggot Isotomurus palustris

Mamestra brussicae

3rd instar larvae

40 mg



Larvae

Eggs

Eggs

Larvae



Eggs



Larvae

Larvae pupae –

Larvae

Coleoptera (continued)

Vasconcelos et al. (1996) Hagley et al. (1982) Pollet et al. (1987) Burgess et al. (2002) Nash et al. (2008) Van den Berg et al. (1992) Lang and Gsödl (2001) Vasconcelos et al. (1996) Oberholzer and Frank (2003) Oberholzer and Frank (2003) Meissle et al. (2005) Mundy et al. (2000) McKemey et al. (2001) Monzó et al. (2011)

14.1 451

Lab

Turkey

Carabidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Adalia decempunctata

Adalia bipunctata

Adalia fasciatopunctata revelieri

Turkey

Aphis affinis, Aphis nerii, Aphis pomi, Aphis punicae, Hyalopterus amygdali, Hyalopterus pruni, Brachycaudus divaricatae, Dysaphis plantaginea, Myzus cerasi, Ovatus mentharius

Aphis affinis, Aphis nerii, Hyalopterus pruni, Brachycaudus schwartzi, Hyadaphis tataricae, Myzus cerasi, Myzus persicae, Ovatus mentharius, Aphis punicae Aphis fabae, Aphis nerii, Aphis pomi, Aphis punicae Ephestia kuehniella

Rhagoletis pomonella

Codling moth Apple maggot Rhagoletis mendax

Sitodiplosis mosellana Sitodiplosis mosellana Ceratitis capitata

Prey Blueberry maggot

Aslan and Uygun (2005)

Aslan and Uygun (2005) De Clercq et al. (2005) Aslan and Uygun (2005)







Eggs

Larvae pupae

– – Pupae

References Renkema et al. (2012) Kromp (1999) Kromp (1999) Urbaneja et al. (2006) Hagley et al. (1982) Renkema et al. (2013) Allen and Hagley (1990)

Prey stage –

14

Lab

Lab

Carabidae

Lab

Carabidae

Carabus nemoralis, Poecilus lucublandus, Pterostichus mutus Harpalus aeneus, Harpalus pennsylvanicus, Pterostichus melanarius Adalia bipunctata

Citrus orchards

Carabidae Carabidae Carabidae

Pterostichus melanarius Platynus dorsalis Peudophonus rufipes, Harpalus distinguendus Stenolophus comma

Location Canada

Family Carabidae

Predator Pterostichus melanarius

Table 14.1 (continued)

452 Prey Record of Various Predators

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae Coccinellidae

Coccinellidae

Coccinellidae

Coccinella undecimpunctata

Micraspis nr.crocea

Micraspis crocea

Micraspis sp.

Oenopia kirbyi

Menochilus sexmaculatus

Propylea dissecta

Nephus nigricans Hippodamia variegata

Chilocorus bipustulatus

Coccinella septempunctata Turkey

Turkey

Turkey Turkey

Lab

Lab

Lab

Indonesia

Mindanao

Lab

Lab

Aphis acetosae, Aphis affinis, Aphis craccivora, Aphis euphorbiae, Aphis fabae, Aphis fabae, Aphis fabae subsp., cirsiiacanthoidis, Aphis fabae, Aphis frangulae, Aphis nerii, Aphis ruborum, Hyalopterus amygdali, Hyalopterus pruni

Brachycaudus helichrysi Aphis craccivora, Aphis fabae, Aphis fabae, Aphis frangulae, Aphis gossypii, Aphis nerii, Aphis ruborum, Hyalopterus pruni, Rhopalosiphum maidis, Brachycaudus helichrysi, Brachycaudus cardui, Myzus cerasi, Ovatus mentharius, Phorodon humuli Aphis acetosae

Aphid

Macrosiphum rosae

Nilaparvata lugens

Scotinophara coarctata

Rivula atimeta

Tetranychus urticae





– –



Fourth and fifth instar nymphs –

Eggs

Eggs

Eggs

Coleoptera (continued)

Aslan and Uygun (2005) Aslan and Uygun (2005)

Aslan and Uygun (2005)

Gaikwad et al. (2022) Chaudhary et al. (2022) Chaudhary et al. (2022)

Farag et al. (2022) Van den Berg et al. (1992) Batay-an et al. (2007) Karindah (2011)

14.1 453

Family Coccinellidae

Coccinellidae

Coccinellidae Coccinellidae Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Predator Coccinella septempunctata

Coccinella septempunctata

Coccinella septempunctata Coccinella septempunctata Coccinella undecimpunctata

Coccinula quatuordecimpustulata

Coccinella transversalis

Cycloneda sanguinea limbifer

Cheilomenes sexmaculata

Cryptolaemus montrouzieri

Chilocorus nigritus

Chilocorus kuwanae

Delphastus catalinae

Table 14.1 (continued)

North America

Albania Italy

Indian

Lab

Bemisia tabaci, Trialeurodes vaporariorum

Scale insects

Scale insects

Planococcus citri

Peregrinus maidis

Myzus persicae

Aphis fabae, Hyadaphis coriandri, Brevicoryne brassicae, Acyrthosiphon pisum Aphis fabae, Acyrthosiphon pisum Aphis glycines Aphis fabae, Aphis frangulae, Aphis gossypii, Aphis nerii, Hyalopterus pruni Aphis affinis, Aphis euphorbiae, Hyalopterus pruni, Acyrthosiphon cyparissiane, Ovatus mentharius Spodoptera frugiperda

Prey Aphis craccivora

Sharanabasappa et al. (2019) Duarte et al. (2014) Dharavath et al. (2022) Alloui-Griza et al. (2022) Roy and Migeon (2010) Roy and Migeon (2010) Roy and Migeon (2010)









Eggs and nymphs –



Aslan and Uygun (2005)

Xue et al. (2009) Aslan and Uygun (2005)

References Katayama and Suzuki (2003) Arshad et al. (2017)







Prey stage –

14

Sorghum—India

Lab

India, Corn

Turkey

Lab Lab Turkey

Lab

Location Lab

454 Prey Record of Various Predators

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Eriopis connexa

Exochomus nigromaculatus

Exochomus quadripustulatus

Hippodamia variegata

Hyperaspis maindroni

Harmonia octomaculata

Harmonia axyridis Harmonia axyridis Hippodamia convergens

Hippodamia axyridis

Hippodamia variegata

Hippodamia variegate

Hippodamia variegata

Harmonia quadripunctata

Hyperaspis femorata

Hyalopterus amygdali

Hyalopterus amygdali

Turkey Turkey

Aphis fabae

Diuraphis noxia

Aphis gossypii

Ephestia kuehniella

Dysaphis plantaginea Aphis glycines Aphid

Spodoptera frugiperda

Phenacoccus manihoti

Aphis fabae, Acyrthosiphon pisum

Macrosiphum euphorbiae, Tetranychus evansi Aphis affinis, Aphis frangulae, Ovatus mentharius Cinara cedri

Lab

Wheat

France

Apple Lab America

India, Corn

India

Lab

Turkey

Turkey

Lab









Eggs

Eggs















Coleoptera (continued)

Sarmento et al. (2007) Aslan and Uygun (2005) Aslan and Uygun (2005) Kayahan et al. (2021) Sreedevi et al. (2020) Sharanabasappa et al. (2019) Dib et al. (2020) Xue et al. (2009) Roy and Migeon (2010) Roy and Migeon (2010) Hosseini et al. (2019) Behnazar and Madadi (2015) Farhadi et al. (2010) Aslan and Uygun (2005) Aslan and Uygun (2005)

14.1 455

Family Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Predator Hyperaspis quadrimaculata

Hyperaspis repensis

Myrrha octodecimguttata

Nephus reunioni

Nephus includens, Nephus bisignatus

Oenopia conglobata

Oenopia oncina

Platynaspis luteorubra

Propylaea quatuordecimpunctata

Propylea japonica

Propylea japonica

Psyllobora vigintiduopunctata

Table 14.1 (continued)

Turkey

Lab

Lab

Lab

Chromaphis juglandicola

Ephestia kuehniella

Cotton aphid

Aphis acetosae, Aphis affinis, Aphis nerii, Aphis pomi, Hyalopterus pruni, Myzus cerasi, Cinara cedri Mimeur Aphis craccivora, Brachycaudus cardui Aphis fabae, Aphis fabae, Uroleucon (=Uromelan) jaceae Aphis affinis, Rhopalosiphum maidis

Planococcus citri

Myrrha octodecimguttata, Dysaphis plantaginea, Myzus cerasi, Ovatus mentharius Scale insects

Prey Aphis acetosae, Brachycaudus helichrysi Aphis fabae

Roy and Migeon (2010) Kontodimas et al. (2007) Aslan and Uygun (2005) Aslan and Uygun (2005) Aslan and Uygun (2005) Aslan and Uygun (2005) Ouyang et al. (2012) Hamasaki and Matsui (2006) Aslan and Uygun (2005)

– –





Eggs











References Aslan and Uygun (2005) Aslan and Uygun (2005) Aslan and Uygun (2005)

Prey stage –

14

Turkey

Turkey

Turkey

Africa, Portugal, France, Greece, Albania Spain Greece

Turkey

Turkey

Location Turkey

456 Prey Record of Various Predators

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Coccinellidae

Rhyzobius forestieri

Stethorus gilvifrons

Scymnus subvillosus

Scymnus rubromaculatus

Scymnus quadriguttatus

Scymnus pallipediformis

Scymnus levaillanti

Scymnus mimulus

Scymnus interruptus

Scymnus frontalis

Scymnus apetzi

Turkey

Turkey

Turkey

Turkey

Turkey

Turkey

Turkey

Turkey

Turkey

Australia, Italy, France, Greece Albania Turkey

Aphis affinis, Brachycaudus helichrysi, Myzus persicae, Ovatus mentharius, Pterochloroides persicae

Rhopalosiphum maidis, Chaitophorus leucomelas, Pemphigus bursarius Aphis fabae subsp. mordvilkoi, Macrosiphoniella artemisiae, Chaitophorus leucomelas, Pemphigus bursarius Aphis gossypii, Rhopalosiphum padi

Aphis affinis, Aphis fabae, Aphis pomi, Aphis punicae, Aphis ruborum Aphis ruborum, Amphorophora rubi, Chaitophorus leucomelas, Pemphigus bursarius Aphis affinis, Aphis euphorbiae, Aphis sambuci, Brachycaudus helichrysi, Ovatus mentharius, Uroleucon (=Uromelan) jaceae Chaitophorus leucomelas, Pemphigus bursarius Aphis gossypii

Aphis affinis

Scale insects

Aslan and Uygun (2005) Aslan and Uygun (2005) Aslan and Uygun (2005)



Aslan and Uygun (2005) Aslan and Uygun (2005) Aslan and Uygun (2005)



Coleoptera (continued)

Aslan and Uygun (2005) Aslan and Uygun (2005)

– –

Aslan and Uygun (2005)





Aslan and Uygun (2005)







Roy and Migeon (2010)



14.1 457

Lab

Coccinellidae

Coccinellidae

Coccinellidae

Coccinella septempunctata

Hippodamia variegata

Parlatoria date scale

Palmaspis phoenicis

Coccinellidae

Coccinellidae

Red date palm scale

Pineapple mealy bug

El-Shafie et al. (2017) El-Shafie et al. (2017) El-Shafie et al. (2017)

Khattawi et al. (2022) El-Shafie et al. (2017)







Macrosiphum miscanthi, Schizaphis graminum, Aphis maidis, Empoasca kerri Nipaecoccus viridis Aphiscraccivora, Aphis gossypii, Aphis nerii, Lipaphis erysimi, Myzuspersicae, Uroleuconcompositae Macrosiphum rosae

Zarghami et al. (2016) Omkar and Srivastava (2003)



White flies

Coccinellidae

Lab

India

References Roy and Migeon (2010) Roy and Migeon (2010) Inayat (2011)

Prey stage –

Prey Bemisia tabaci

14

Pharoscymnus setulosus, Pharoscymnus ovoideus, Pharoscymnus pharoides, Pharoscymnus horni, Rhyzobius lophanthae

Coccinella undecimpunctata, Coccinella quinquepunctata, Cocinella septempunctata, Chilocorus bipustulatus, Scymnus bipunctata, Scymnus pictus Scymnus punetillum, Pharoscymnus anchorago, Pharoscymnus varius Pharoscymnus numidicus

India

Coccinellidae

Coccinella septempunctata, Cheilomenes sexmaculata, Hippodamia variegata Nephus arcuatus

Asia

Coccinellidae

Serangium parcesetosum

Location Europe

Family Coccinellidae

Predator Serangium parcesetosum

Table 14.1 (continued)

458 Prey Record of Various Predators

Cybocephalidae

Cybocephalus nigriceps, Cybocephalus pullus, Cybocephalus aegyptiacus, Cybocephalus micans Teretrius pulex

Histeridae

Anthicidae Cicindelidae

Anthicus unicolor Megacephala affinis

Turkey

Date palm bostrichid

Diuraphis noxia Onthophagus landolti, Onthophagus marginicollis Parlatoria scale

Essigella californica

Anthicidae

Ischyropalpus nitidulus

Lab

Peregrinus maidis

Chrysomelidae

Sorghum—India

Ommatissus lybicus

Coccinellidae

Exochomus nigripennis, Cheilomenes propinqua Monolepta signata



Eggs and nymphs –

El-Shafie et al. (2017)

Parlatoria scale

El-Shafie et al. (2017) Dharavath et al. (2022) Landwehr (1977) Bacci (2009) Young (1980)

14.1 Coleoptera 459

460

14.2

14

Prey Record of Various Predators

Orthoptera

Table 14.2 Prey record and stages preferred by Orthopterans Predator Gryllus pennsylvanicus Anaxipha longipennis

Family Gryllidae

Location Canada

Prey Phyllotreta cruciferae Trigonidiidae Rice pests Trigonidiidae Tamil Cnaphalocrocis Nadu-India medinalis, Rivula atimeta, Melanitis leda ismene Trigonidiidae Rivula atimeta

Prey stage References Adult Burgess and Hinks (1987) India Chitra et al. (2000) Egg Canapi et al. (1988)

Trigonidiidae Indonesia

Fourth and fifth instar nymphs Eggs

Nilaparvata lugens

Trigonidiidae Philippines Cnaphalocrocis medinalis, Marasmia patnalis Trigonidiidae Philippines Rivula atimeta

Metioche bicolor Metioche vittaticollis

Eggs

Eggs

Trigonidiidae Philippines Cnaphalocrocis medinalis, Marasmia patnalis Trigonidiidae Mindanao Scotinophara coarctata Trigonidiidae Pakistan Chilo suppressalis Trigonidiidae Pakistan Chilo suppressalis Trigonidiidae Tamil Lepidopterous Nadu-India rice pests Trigonidiidae India Rice pests Trigonidiidae India Rice pests Philippines Cnaphalocrocis medinalis Marasmia patnalis Bradley Rivula atimeta

Eggs

Tamil Lepidopterous Nadu-India insect pests of rice Philippines Cnaphalocrocis medinalis, Marasmia patnalis

Eggs

Eggs

Van den Berg et al. (1992) Karindah (2011)

De Kraker et al. (2000)

Van den Berg et al. (1992)

Nymphs

Batay-an et al. (2007) Iqbal (2020)

Nymphs

Iqbal (2020)

Eggs

Chitra et al. (2000)

India India Egg

Chitra et al. (2000) Chitra et al. (2000) De Kraker et al. (2000)

Eggs

Van den Berg et al. (1992) Chitra et al. (2000)

Eggs

(continued)

14.2

Orthoptera

461

Table 14.2 (continued) Predator

Family

Location Mindanao Indonesia

Conocephalus, Conocephalus longipennis

Conocephalus maculatus

Chlorobalius leucoviridis

Prey Scotinophara coarctata Brown Plant Hopper

Tettigoniidae Philippines Nilaparvata lugens Philippines Sogatella furcifera Philippines Nephottetix virescens Philippines Scirpophaga incertulas Philippines Chilo suppressalis Philippines Marasmia patnalis Philippines Hydrellia philippina Philippines Leptocorisa oratorius Philippines Yellow stemborer Anaxipha sp. and Metioche sp. Tettigoniidae Philippines Nilaparvata lugens, Sogatella furcifera, Nephotettix virescens, Scirpophaga incertulas, Chilo suppressalis, Marasmia patnalis, Hydrellia philippina, Leptocorisa oratorius Tettigoniidae New Leptocorysa Guinea oratorius Philippines Scirpophaga innotata Tettigoniidae Austra Cicada prey

Prey stage References Eggs Batay-an et al. (2007) Karindah (2011) Adults and nymphs Nymphs Rubia et al. (1990) and adults Nymphs Rubia et al. (1990) and adults Nymphs Rubia et al. (1990) and adults Adults Rubia et al. (1990) Adults

Rubia et al. (1990)

Adults

Rubia et al. (1990)

Adults

Rubia et al. (1990)

Adults

Rubia et al. (1990)

Egg

Pena et al. (1987)

Rubia et al. (1990) Nymphs and adults

Eggs and nymphs Eggs

Rothschild (1970) and Pitkin (1980) Litsinger et al. (2006) Marshall and Hill (2009)

Neuroptera

Chrysopidae

Chrysopidae

Chrysopidae

Chrysopa nigricornis

Chrysopa oculata

Chrysopidae

Chrysopa formosa

Lab

Chrysopidae

Chrysopidae

Lab

Chrysopidae

Chrysopa cubana

Lab

Lab

Lab

Lab

Lab

Mesolecanium nigrofasciatum, Planococcus citri, Pseudococcus comstocki Phenacoccus gossypii, Pseudococcus maritimus

Aulacaspis citri

Chrysomphalus aonidum, Planococcus citri

Planococcus citri

Lepidosaphes beckii

Chrysomphalus aonidum

Saccharicoccus sacchari

Fiorinia theae

Fiorinia theae

Maconellicoccus hirsutus

Drosicha stebbingi

Prey Rastrococcus invadens

























Prey stage –

References Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004)

14

Chrysopa sanchezi

Lab

Chrysopidae

Chrysopa claveri

Lab

Chrysopidae

Ceraeochrysa claveri

Lab

Chrysopidae

Lab

Chrysopidae

Chrysoperla bicarnea

Lab

Chrysopidae

Chrysopa scelestes

Location Lab

Family Chrysopidae

Predator Borniochrysa squamosus

Table 14.3 Prey record and their preferred stages of pests by various neuropterans

14.3

462 Prey Record of Various Predators

Chrysopidae

Chrysopidae

Chrysopidae

Chrysoperla nipponensis

Chrysoperla orestes

Lab

Lab

Lab

Brazil

Chrysopidae

Chrysoperla harrisii

Brazil

Chrysopidae

India

Chrysopidae Lab

Lab Lab

Chrysopidae Chrysopidae

Chrysopidae

Lab

Chrysopidae

Chrysoperla externa

Lab

Chrysopidae

Chrysoperla carnea

Lab

Chrysopidae

Pseudococcus maritimus

Ferrisia virgate, Rastrococcus iceryoides

Aulacaspis citri

Fiorinia theae

Phyllocnistis citrella

Anticarsia gemmatalis, Diatraea saccharalis

Parlatoria cinerea, Selenaspidus articulatus

Ferrisia virgata

Acyrthosiphon pisum Aphis fabae, Ephestia kuehniella

Spodoptera littoralis, Rhopalosiphum padi

Ceroplastes sinensis, Parthenolecanium persicae, Protopulvinaria pyriformis, Pulvinaria horii, Pulvinaria tenuivalvata, Pulvinaria vitis, Pulvinaria floccifera, Saissetia oleae, Aonidiella orientalis, Aspidiotus nerii, Chrysomphalus aonidum, Diaspidiotus perniciosus, Lepidosaphes tapleyi, Parlatoria blanchardi, Unaspis yanonensis Helicoverpa armigera







Eggs

Eggs



– Nymps Eggs –

Neuroptera (continued)

Hassanpour et al. (2011) Meier and Hilbeck (2001) Zhu et al. (2005) Alghamdi and Sayed (2017) Krishnamoorthy and Mani (1989) Miller et al. (2004) Degasperi et al. (2010) Ribeiro et al. (2007) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004)

– –

Miller et al. (2004)



14.3 463

Lab Lab

Lab

Soybean Field cage Lab Lab

Chrysopidae

Chrysopidae

Chrysopidae Chrysopidae

Chrysopidae Chrysopidae

Chrysopidae

Chrysopide

Chrysopidae

Chrysopidae

Chrysopodes collaris

Chrysopa vulgaris aeqyptiaca Chrysopa oculata Chrysoperla nipponensis

Chrysopa pallens Chrysopa pallens

Chrysoperla rufilabris

Korea-Field Lab

Lab

Lab

Lipaphis erysimi, Myzus persicae, Brevicoryne brassicae, Aphis craccivora, Aphis spiraecola Aphis fabae, Aphis nerii, Aphis pomi, Hyalopterous pruni, Macrosiphum rosae, Myzus persicae

Pseudoplusia includen

Bemisia tabaci, Aphis gossypii, Sitotroga cerealella, Helicoverpa zea

Chrysomphalus aonidum, Icerya aegyptiaca, Maconellicoccus hirsutus Acyrthosiphon pisum Aphis craccivora, Rhopalosiphum maidis, Corcyra cephalonica Aphis gossypii Acyrthosiphon pisum, Frankliniella occidentalis

Coccus hesperidum, Chrysomphalus aonidum, Fiorinia theae, Pseudococcus comstocki Icerya purchasi

Prey Parasaissetia nigra, Parthenolecanium pruinosum, Pulvinaria vitis, Aonidiella aurantii, Chrysomphalus aonidum, Hemiberlesia lataniae, Lepidosaphes beckii Icerya pattersoni





Young once Eggs Larvae

– –

– –









Prey stage –

Richman et al. (1980) Halder and Rai (2016) Pappas et al. (2007)

Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Miller et al. (2004) Zhu et al. (2005) Shafique et al. (2019) Lee et al. (2000) Sarkar et al. (2019) Legaspi et al. (1994)

References Miller et al. (2004)

14

Chrysoperla zastrowi sillemi Dichochrysa prasina

Lab

Chrysopidae

Chrysoperla rufilabris

Lab

Chrysopidae

Chrysoperla pudica

Location Lab

Family Chrysopidae

Predator Chrysoperla plorabunda

Table 14.3 (continued)

464 Prey Record of Various Predators

Sorghum—India

Neuropteran

Peregrinus maidis

Reticulitermes speratus Robber flies

Lab Lab

Berothidae Nemopteridae

Isoscelipteron okamotonis Palmipenna aeoleoprera, Palmipenna cf. pilicornis Micromus timidus

Nasonovia ribisnigri, Frankliniella occidentalis

Phenacoccus manihoti, Phenacoccus madeirensis, Pseudococcus jackbeardsleyi Heliothis zea, Heliothis virescens

Planococcus citri

Planococcus citri

Aphis craccivora, Rhopalosiphum maidis, Corcyra cephalonica Planococcus citri, Maconellicoccus hirsutus

Thaumastocoris peregrinus

Lab

Cotton

India

Lab

Ephestia kuehniella, Aphis fabae, Aphis punicae, Macrosiphum rosae Helicoverpa armigera, Aphis gossypii

Lab

Hemerobiidae

Hemerobius bolivari

Chrysopidae

Chrysopidae

Chrysopidae

Chrysoperla carnea

Lab

Chrysopidae

Chrysopa lacciperda [=Plesiochrysa lacciperda] Anisochrysa basalis [=Mallada basalis] Plesiochrysa ramburi

Chrysopidae

India

Chrysopidae

Mallada boninensis

Chrysopa Carnea, Chrysopa rufilabris

India

Chrysopidae

Plesiochrysa ramburi

Field

Chrysopidae

Mallada desjardinsi

Lab

Chrysopidae

Dichochrysa tacta

Eggs and nymphs

– –

Nymphs

Eggs Earley learvae















Shrestha and Enkegaard (2013) Garcia et al. (2013) Komatsu (2014) Picker et al. (1992) Dharavath et al. (2022)

Sayed and Alghamdi (2017) Kabissa et al. (1996) Shafique et al. (2019) Krishnamoorthy and Mani (1989) Krishnamoorthy and Mani (1989) Krishnamoorthy and Mani (1989) Sattayawong et al. (2016) Lingren et al. (1968)

14.3 Neuroptera 465

Hymenoptera

Vespidae Vespidae Vespidae

Vespidae

Vespidae Vespidae Vespidae Vespidae

Vespa velutina Ischocyttarus flavitarsis

Polybia fastidiosuscula

Vespula gennanica

Polybia (=Trichothorax) sericea

Vespula vulgaris, Vespula rufa consobrina

Ropalidia brevita

Polistes cf. olivaceus

Solenopsis geminata

Formicidae

Vespidae

Vespa velutina

India Corn

India Corn

Field

Lab

Lab

Brazil

Prey Leucoptera coffeella, Cassida rubiginosa, Honeybees, flies and social wasps Honeybees Lepidoptera Lepidoptera Diptera Orthoptera Ceratitis capitata, Alabama argillacea, Spodoptera fugiperda, Elasmopalpos lignosellus, Pseudoplusia includens, Chlosynelacinia saundersii, Neodiprion swainei, Spodoptera frugiperda, Spodoptera frugiperda, Forficula auricularia, Ceratitis capitata, Ceratitis capitata, Aedia leucomelas, Spodoptera litura, Acheta domesticus, Diatraea saccharalis, Aphis craccivora, Rice pests, Diatraea saccharalis, Belenois solilucis, Diatraea saccharalis, Oryctes rhinoceros, Setora nitens, Pteroma pendula, Anastrepha sp., Rhagoletis pomonella, Anastrepha fraterculus, Anastrepha suspensa, Anastrepha sp., Z. curcubitae, Spodoptera frugiperda, Elaeidobius kamerunicus, Anastrepha sp. Anastrepha fraterculus, Anastrepha suspensa, Anastrepha ludens, Anastrepha suspensa, Anastrepha sp., Halyomorpha halys, Diatraea saccharalis, Heteracris, Larvae

Larvae



Caterpillars All Nymphs

– Caterpillar

Hendrichs and Hendrichs (1998) Filho et al. (2009) Smirnoff (1961) Firake and Behere (2020) Firake and Behere (2020) Kurczewski (1968)

Schenk and Bacher (2002) Rome et al. (2021) Ueno (2014) McPheron and Mills (2007) Brügger et al. (2019)

– –

References Fernandes (2012)

Prey stage –

14

Nearctic vespine, Vespula maculifrons

Creeping thistle

Vespidae

Field

Location Coffee crops

Family Vespidae

Predator Brachygastra lecheguana, Polistes versicolor, Polybia occidentalis, Polybia paulista, Polybia scutellaris, Protonectarina sylverae, Synoeca cyanea Polistes dominulus

Table 14.4 Prey record and their preferred stages of pests by various Hymenopterans

14.4

466 Prey Record of Various Predators

Sugarcane

Lab Rice

Formicidae Formicidae

Formicidae Formicidae

Pachycondyla

Brachymyrmex incisus, Camponotus crassus, Camponotus rufipes, Camponotus novogranadensis, Camponotus melanoticus, Nylanderia fulva Tetramorium caespitum, Lasius niger

Aphaenogaster obsidiana, Cardiocondyla stambouloffi, Crematogaster antaris, Crematogaster subdentata, Crematogaster warburgi, Messor alexandri, Camponotus oasium, Camponotus xerxes, Cataglyphis auratus, Cataglyphis albicans var. auratus, Cataglyphis lividus, Cataglyphis nodus var. drusa, Cataglyphis semitonsus, Formica glauca, Lasius alienus, Lasius neglectus, Lepisiota frauenfeldi, subsp. Karavievi,

Sweet potato Lab

Formicidae

Formicidae

Messor structor, Pheidole pallidula, Formica fusca, Messor bouvieri, Lassius niger, Linepithema humile, Plagiolepis pygmaea, Lasius fuliginosus, Hypoponera eduardi, Camponotus pilicornis, Messor barbarus, Cardiocondyla elegans, Lasius grandis Tetramorium bicarinatum

Coffee and orange orchards Citrus orchards Physokermes hellenicus, Callosobruchus chinensis*



Hymenoptera (continued)

Katayama and Suzuki (2003) Ghahari et al. (2009)



Egg



Suenaga (2017) Orivel and Dejean (2001) De Oliveira et al. (2012)

Urbaneja et al. (2006)

Eggs

Pupae

Eskafi and Kolbe (1990)

14.4 467

Sugarcane

Formicidae Formicidae

Formicidae Formicidae Formicidae Formicidae Formicidae

Myopopone castanea

Oecophylla smaragdina

Oecophylla smaragdina

Ectatomma brunneum

Formica fusca

Pachycondyla striata

Lab

Lab

Lab

Lab

Prey





Larvae and pupae –

Larvae

Larvae

Eggs

Prey stage

De Oliveira et al. (2012) Eisawi et al. (2022) De Oliveira et al. (2012) Tobing and Kuswardani (2018) Falahudin (2013) Pierre and Idris (2013) Fernandes et al. (2012) Monteith (1972) Abeijon et al. (2019)

References

14

Lab

Oil palm

Sudan

Sugarcane

Location

Formicidae

Formicidae

Lepisiota karavievi, Plagiolepis maura, Polyrhachis lacteipennis, Linepithema humile, Tapinoma festae, Tapinoma karavievi, Tapinoma simrothi subsp. karavievi, Messor darianus, Messor denticulatus, Messor medioruber, Messor sultanus, Monomorium areniphilum, Monomorium pharaonis, Monomorium venustum, Pheidole megacephala, Solenopsis wolfi, Tetramorium caespitum, Tetramorium taurocaucasicum Solenopsis saevissima

Axinidris acholli, Tapinoma carininotum, Technomyrmex moerens Linepithema humile, Dorymyrmex brunneus

Family

Predator

Table 14.4 (continued)

468 Prey Record of Various Predators

Lab

Formicidae Formicidae

Formicidae Formicidae Formicidae Formicidae Formicidae Formicidae Formicidae Ponerinae Crabronidae Encyrtidae Ants

Pheidole megacephala

Pheidole megacephala

Pheidole oxyops

Pogonomyrmex naegelli

Odontomachus brunneus

Solenopsis geminata Solenopsis invicta

Solenopsis saevissima

Crematogaster scutellaris

Ectatomma brunneum

Stizus continuus

Microterys lunatus, Pseudorhopus testaceus

Camponotus rufipes, Monomorium minimum, Dorylus labiatus

Sugarcane

Lab

Lab

Lab Lab

Lab

Lab

Lab

Ghana Corn Oil palm

Lab











– –















– Hennessey (1998) Fernandes et al. (2012) Clausen et al. (1965) Koffi et al. (2020) Muhammad Luqman et al. (2018) Fernandes et al. (2012) Abeijon et al. (2019) Hennessey (1998) Aluja (1994) Hennessey (1998) Fernandes et al. (2012) Castracani et al. (2017) De Oliveira et al. (2012) Polidori et al. (2009) Papanastasiou et al. (2018) Aslam et al. (2006)

Hymenoptera

* Storage pest

Conifers

Formicidae

Pheidole magacephala

Lab

Formicidae

Pheidole gertrude

Lab

Formicidae

Paratrechina parvula

14.4 469

Diptera

Aphis grossulariae Cavariella Macrosiphum rosae

Soybean

Soybean Willow Hogweed Rose

Asilidae Asilidae Asilidae Syrphidae

Syrphidae Syrphidae

Cardiocladius australiensis

Cardiocladius oliffi

Chironomus thummi

Allograpta obliqua, Paragus haemorrhous, Toxomerus marginatus, Syrphus rectus, Eupeodes americanus, Sphaerophoria contigua, Eupeodes volucris Eupeodes volucris, Paragus hemorrhous, Toxomerus marginatus Episyrphus balteatus

Palaearctic

Afrotropical

Australasian

Grassland Afrotropical

Asilidae Asilidae

Aphis glycines

Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009)







Kaiser et al. (2007)

Joern and Rudd (1982) Werner and Pont (2003) Werner and Pont (2003) Werner and Pont (2003) Werner and Pont (2003) Kaiser et al. (2007)

Kaiser et al. (2007) Pollock (2021)

References Kaiser et al. (2007) Faria et al. (2007)





Pupae

Pupae

Larvae

Larvae

Larvae

Prey stage

14

Aphis glycines

Simulium pontina

Austrosimulium immatures Simulium damnosum

Ageneotettix deorum Blackfly

soybean Mexico

Cecidomyiidae Asilidae

Prey Aphis glycines Cochliomyia macellaria Aphis glycines Pogonomyrmex

Aphidoletes aphidimyza Cerotainiops abdominalis, Saropogon combustus, Saropogon pritchardi Proctacanthus milbertii Cardiocladius africanus

Location soybean Lab

Family Chamaemyiidae Calliphoridae

Predator Leucopis glyphinivora Chrysomya albiceps

Table 14.5 Prey record and their preferred stages of pests by various Dipterans

14.5

470 Prey Record of Various Predators

Syrphidae

Syrphidae Syrphidae Syrphidae Syrphidae

Syrphus ribesii

Allograpta exotica Metasyrphus corollae Metasyrphus confrater

Episyrphus balteatus+

Myzus

Wild cherry

Aphis gossypii



– – –

























(continued)

Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009) Almohamad et al. (2009) Arcaya et al. (2017) Adams et al. (1987) Verma and Sharma (2006) Hong and Hung (2010)

Diptera

Lab

Lab Lab

Aphis craccivora Sitobion avenae Macrosiphum rosae

Aphis sambuci

Elder

Rose

Microlophium carnosum Macrosiphum rosae

Rhopalosiphum

Bird-cherry

Nettle

Hyalopterus

Reed

Cavariella

Aphis fabae

Spindle

Hogweed

Aphis fabae

Thistle

Drepanosiphum

Brevicoryne brassicae

Cabbage

Sycamore

Schizoneura

Elm

14.5 471

Cabbage Lab Virginia apple orchards Onion

Syrphidae Syrphidae Syrphidae

Syrphidae

Syrphidae

Syrphidae Syrphidae Syripidae

Heringia calcarata

Sphaerophoria macrogaster

Melanostoma scalare

Eupeodes corollae Eupeodes corollae Copestylum caudatum

Lab Lab Mexico

Onion

Location Broad bean

Family Syrphidae

Predator

Table 14.5 (continued)

Thrips tabaci, Frankliniella occidentalis Thrips tabaci, Frankliniella occidentalis Myzus persicae Aphis craccivora Monoctenus sanchezi

Brevicoryne brassicae Episyrphus balteatus Eriosoma lanigerum

Prey Aphis fabae

Sekine et al. (2022)

Sekine et al. (2022)

Lillo et al. (2021) Zheng et al. (2019) Ordaz-Silva et al. (2021)

– – –

References Amiri-Jami and Sadeghi-Namaghi (2014) van Rijn et al. (2006) Noël et al. (2022) Short and Bergh (2004)



– – –

Prey stage –

472 14 Prey Record of Various Predators

Dermaptera

Anastrepha suspensa Cosmopolites sordidus Cotton boll weevil

Guava

Brazil

Anisolabididae

Anisolabididae

Forficulidae

Forficula auricularia, Euborellia moesta, Euborellia annulipes Euborellia annulipes

Euborellia annulata

Forficula auricularia

Halotydeus destructor Aphis pomi

Lab

Ostrinia furnacalis

Lab

Philippines





Diatraea saccharalis Diamondback

Eggs larvae

Plutella xylostella

Lab

Lab



Plutella xylostella

Pupae

Pupae

Prey stage –

Lab

Anthonomus grandis

Ceratitis capitata

Brontispa longissima Ostrinia furnacalis

Coconut Rice Philippines citrus orchards

Chelisochidae Chelisochidae

Prey Euwallacea interjectus

Chelisoches morio Proreus simulans

Location Orchard

Family

Predator Anisolabella marginalis

Table 14.6 Prey record and their preferred stages of pests by various Dermapterans

14.6

Dermaptera (continued)

Hennessey (1998) Klostermeyer (1942) Lemos et al. (1998, 1999) Lemos et al. (2003a, b) Nunes et al. (2019a, b) da Silva et al. (2020) de Souza et al. (2022) Nunes et al. (2019a, b) Situmorang and Gabriel (1988) Weiss and McDonald (1998) Asgari (1966)

References Jiang and Kajimura (2020) Ramos et al. (2019) Barrion and Litsinger (1985) Urbaneja et al. (2006)

14.6 473

Family

Forficulidae

Forficulidae

Forficulidae Forficulidae

Forficulidae

Forficulidae

Predator

Forficula pubescens

Doru luteipes

Doru luteipes Doru lineare Doru lineare

Doru taeniatum

Doru luteipes

Table 14.6 (continued)

– – Eggs Spodoptera frugiperda Spodoptera frugiperda, Dalbulus maidis Spodoptera frugiperda

Brazil Lab



Diatraea saccharalis, Spodoptera frugiperda

Brazilian maize



Spodoptera frugiperda, Helicoverpa zea

Lab



Spodoptera frugiperda

Naranjo-Guevara et al. (2017a, b) Silva et al. (2022) Sueldo et al. (2010) Mariani et al. (1996) Jones et al. (1988) and Lastres (1990) Lanza Reis et al. (1988) and Cruz (1995) Naranjo Guevara et al. (2017a, b) Sueldo et al. (2010) Badji et al. (2004) Moreira et al. (2022)

References Manyuli et al. (2008) Jana et al. (2021) Dib et al. (2020) Carroll et al. (1985) Madge and Buxton (1976) Dib et al. (2020) Happe et al. (2018)

14

Lab

Lab Lab

Egg Caterpillar – – Egg



Dysaphis plantaginea Eriosoma lanigerum

Apple Organic orchards in Germany Maize Spodoptera frugiperda, Diatraea saccharalis, Helicoverpa zea, Ascia monuste orseis Caliothrips phaseoli Spodoptera frugiperda Spodoptera frugiperda

– – – –

Aphids and psyllids Harmonia axyridis, Dysaphis plantaginea Woolly apple aphids Phorodon humuli

Orchards Lab Orchard

Prey stage –

Prey Aphis craccivora

Location Eastern Africa

474 Prey Record of Various Predators

Forficulidae

Labiduridae

Labiduridae

Labiduridae

Forficulidae

Chelisochidae

Doru lineare

Nala lividipes

Labidura riparia

Labidura truncata

Chelisoches morio

Chelisoches morio

Guava

Dacus curcurbitae

Trichoplusia ni Phthorimaea operculella Helicoverpa punctigera, Pterohelaeus nr. darlingensis, Teleogryllus commodus Rhynchophorus ferrugineus

Cabbage Lab

Coconut

Melanotus communis

Lab

Ostrinia furnacalis

Spodoptera frugiperda

Diatraea saccharalis

Lab Argentina— maize Philippines

Spodoptera frugiperda

Lab

Pupae

Eggs

Eggs

Abraham and Kurian (1973) Nishida (1955)

da Silva et al. (2022) Fenoglio and Trumper (2014) Romero Sueldo, and Cuezzo (2001) Situmorang and Gabriel (1988) Shepard et al. (1973) Strandberg (1981) Horne and Edward (1995)

14.6 Dermaptera 475

476

14.7

14

Prey Record of Various Predators

Hetroptera

Table 14.7 Prey record and their preferred stages of pests by various predators of Pentatimideae of Heteroptera Predator Amyotea malabaricus Antilochus coquebertii

Andrallus spinidens

Location India

Prey Melanitis leda ismene

Prey stage

References Bal (2013)

Lab

Dysdercus koenigii



Lab Lab

Dysdercus cingulatus Dysdercus spp.

Nymphs Nymphs

India

Dysdercus koenigii

Nymphs

India Corn Lab

Spodoptera frugiperda



Galleria mellonella Ephestia kuehniella Pamara mathias Sesamia inferens Spodoptera litura Tenebrio molitor



Chauthani and Misra (1966), Muhammad et al. (2020), Saeed et al. (2020) and Sahayaraj and Fernandez (2017, 2021) Kohno et al. (2004) Kohno (2003) and Kohno et al. (2002) Sahayaraj and Fernandez (2017) Firake and Behere (2020) Mohaghegh and AmirMaafi (2007) Bal (2013)

Spodoptera frugiperda



Amsacta albistriga, Spodoptera exigua, Spodoptera litura, Athalia lugens proxima, Thosea cervina, Utethesia pulchella, Hyblaea puera, Semiothesia pervolgata, Eurema hecabe, Catopsilia pyranthe, Helicoverpa armigera, Thiacidas postica Spodoptera frugiperda





Lygocoris communis, Lygocoris communis Anticarsia gemmatalis

Firake and Behere (2020) Arnoldi et al. (1991)

Eggs

Godfrey et al. (1989)

India

Brontocoris tabidus Eocanthecona furcellata

Podisus maculiventris

India Corn India

India Corn Apple orchard Lab

Zanuncio et al. (2000) Firake and Behere (2020) Bal (2013)

(continued)

14.7

Hetroptera

477

Table 14.7 (continued) Predator

Podisus sagitta Podisus nigrispinus Podisus nigrispinus, Supputius cincticeps Podius maculiventris, Podius nigrispinus

Location Lab

Prey Pyrrhalta viburni

Prey stage –

Lab Lab

Ephestia kuehniella Preys

Eggs –

Lab Lab Soybeans field cages Mexico

Spodoptera frugiperda Epilachna varivestis Epilachna varivestis

– Caterpillar Third instar

Spodoptera frugiperda

Eggs and larvae

Lab

Spodoptera exigua

De Clercq et al. (2000)

Lab soybean

Spodoptera frugiperda Anticarsia gemmatalis

Zanuncio et al. (2008) De Castro et al. (2013)

Lab

Spodoptera exigua



Lab

Tenebrio molitor, Alabama argillacea Chrysodeixis chalcites Spodoptera exigua Peridroma saucia

Mohaghegh et al. (2001) Lemos et al. (2003a, b)





De Clercq et al. (1998) De Clercq (2001) Ables and McCommas (1982) Costa et al. (2019)



Bal (2013)

Tomato Podisus maculiventris Supputius cincticeps Zicrona caerulea

USA cotton Lab India

Cladomorphus phyllinus Sdoptera litura, Semiothesa pervolgata, Catopsilia pyranthe, Eurema hecabae, Anticarsia irrorata, Helicoverpa armigera, Pelopidas mathias

Larvae

References Desurmont and Weston (2008) Oliveira et al. (2004) Wiedenmann et al. (1994) Avery et al. (2022) O'Neil (1989) Wiedenmann and O'Neil (1992) Hoballah et al. (2004)

Prey record of Longlegged fly, Mantidfly, Robber fly, Variable lady beetle, Rambur's forktail damselfly, Eastern pondhawk dragonfly (Fig. 14.1a–h), Ambush bugs and Black bee assassin (Fig. 14.2), Pink (twelve-) spotted lady beetle, Multicolored Asian lady beetle and Western red-bellied tiger beetle (Fig. 14.3), Florida predatory stink bug (a), Predatory stink bug (b and c) and Antilochus coqueberti (d) (Fig. 14.4), reduviids—Rhynocoris kumarii (Fig. 14.5), Rhynocoris fuscipes (Fig. 14.6), and Rhynocoris longifrons (Fig. 14.7) is also provided here.

478

14.8

14

Prey Record of Various Predators

Future Recommendations

It is imperative to record all available records on prey and/or host record of all predatory insects. It is very useful to carry out biological control works of students, research scholars, faculties, extension workers, policy makers and whoever have been working on plant protection and production. Table 14.8 Prey record and their preferred stages of pests by various predators of Nabidae of Heteroptera Predator Himacerus sp.

Location Lab

Prey Halyomorpha halys

Himacerusapteru

Metatetranvchusulmi

Hyphantria apterus

England orchards Ukraine

Prey stage Egg and 1st nympal instar –

Lymantria dispar



Hyphantria apterus

Poland

Lymantria dispar



Nabis alternatus

Lab

Nabis alternatus

Lab

Lygus sp.



Nabis americoferus

Lab

Spodoptera exigua, Lygus hesperus Myzus persicae, Ephestia kuehniella, Trichoplusia ni Collops vittatus



Dei and Nikitenko (1980) Dei and Nikitenko (1980) Atim and Graham (1984) Perkins and Watson (1972) and Tamaki et al. (1978) Propp (1982)



LaFlair (2022)



Acyrthosiphon kondoii, Acyrthosiphon pisum, Sidnia kinbergi, Calocoris norvegicus Epilachna varivestis, Anthonomus grandis



Nielson and Henderson (1959) Siddique and Chapman (1987)

Lygus lineolaris, Lygocoris communis



Nabis ferus

Lab

Nabis kinbergii

Lab

Nabis roseipennis

Lab

Nabicula subcoleoptrata

Apple orchard





References Bulgarini et al. (2019)

Collyer (1953)

Wiedenmann and O'Neil (1990) Arnoldi et al. (1991) (continued)

14.8

Future Recommendations

479

Table 14.8 (continued) Predator Nabis sp.

Location Lab

Prostemma sp Alloeorhynchus sp Pagasa sp Pagasa species Geocoris bullatus, Nabis alternatus Reduviolus americoferus

Lab

Prey Trifolium pratense, Ceratagallia agricola Blissinae sp. Rhyparochromina sp.

Prey stage – –

Lab Lab

Chinch bugs Lygus sp.

– –

Lab

Potato leafhopper nymphs and pea aphids Beet leafhopper Potato leafhopper



Lab



– Eggs and larvae Eggs and larvae

Lab Lab

Colorado potato beetle Heliothis zea

Tropiconabis capsiformis

Lab

Lepidoptera

Nabis americoferus, Nabis roseipennis Reduviolus roseipennis, Tropiconabis capsiformis, Hoplistoscelis deceptivus Tropieonams eapsiformis adult Tropieonabis eapsiformis nymph Nams roseipennis adult Nabis roseipennis nymph Hoplistoseeis deeeptivus adult

Lab

Green coverworm



Soybean Field cage

Pseudoplusia includen

Larvae

Lab

Anticarsia gemmatalis

Eggs

Reduviolus roseipennis

References Stasek et al. (2018) Harris (1928) and Pericart (1987) Reinert (112) Tamaki et al. (1978) Flinn et al. (1985) Knowlton (1935) and Martine and Pienkowski (1982) Parshley (1923) Donahoe and Pitre (1977) Carty et al. (1980) and Samson and Blood (1980) Sloderbeck and Yeargan (1983) Richman et al. (1980)

Godfrey et al. (1989)

480

14

Prey Record of Various Predators

Table 14.9 Prey record and their preferred stages of pests by various predators of Anthocoridae of Heteroptera Predator

Location

Prey

Prey stage

Anthocoris nemoralis Bilia castanea

Italy, Pear

Cacopsylla (Psylla) pyri



Brassica nigra India Brassica nigra India Okra, chili, rose, brinjal India Coconut India Lab

Lipaphis erysimi



Lipaphis erysimi



Tetranychus urticae



Opisina arenoosella



Frankliniella occidentalis



Syria

Gynaikothrips uzeli



Lab

Frankliniella occidentalis



Cucumber France Ficus retusa

F. occidentalis



Hawaii

Gynaikothrips uzeli

Lab

Megalurothrips sjostedti



Lab

Frankliniella occidentalis



Lab

Spodoptera exigua

Lab

Whitefly



Lab

Spodoptera exigua



Strawberry France Lab

F. occidentalis



Frankliniella occidentalis, Anagasta kuehniella

Greenhouses

Thrips

Nymphs, adults eggs –

Field

Frankliniella occidentalis



Maize USA

Ostrinia nubilalis



Potato Iran

Tetranychus urticae

Orius albidipennis Blaptostehus pallescens Cardiastethus exiguus Dicyphus tamaninii Montandoniola indica Macrolophus pygmaeus Orius majusculus Montandoniola moraguesi Montandoniola confusa Orius albidipennis

Orius laevigatus

Orius laevigatus Orius insidiosus

Orius insidiosus Orius minutus

Gynaikothrips ficorum

References Ballal et al. (2016) Ballal et al. (2016) Ballal et al. (2016) Ballal et al. (2016) Ballal et al. (2016) Blaeser et al. (2004) Ali and Streito (2019) Blaeser et al. (2004) Ballal et al. (2016) Ballal et al. (2016) Ballal et al. (2016) Gitonga et al. (2002) Blaeser et al. (2004) AragónSánchez et al. (2018) Fernández et al. (2012) AragónSánchez et al. (2018) Ballal et al. (2016) Calixto et al. (2013) Ballal and Yamada (2016) Funderburk et al. (2000) Ballal et al. (2016) Ballal et al. (2016)

(continued)

14.8

Future Recommendations

481

Table 14.9 (continued) Prey stage

Predator

Location

Prey

Orius majusculus

Lab

Dysaphis plantaginea

Lab

Frankliniella occidentalis



Tomato

Macrosiphum euphorbiae



Lab

Echinothrips americanus, Thrips setosus



Orius sauteri

Lab

Frankliniella occidentalis



Orius strigicollis Orius tantillus

Egg plant Taiwan greenhouses

Thrips



Thrips



F. occidentalis

Orius tristicoler

Onion Australia Lab

Collops vittatus



Apple Lab

Aphid Amphibolus venator* Peregrinator biannulipes* Joppeicus paradoxus* Nezara viridula Dictyophara europea Acheta domesticus Ephestia cautella

– –

Tribolium confusum*



Orius laevigatus, Orius majusculus

Orius minutus Xylocoris flavipes

Lab

GroundnutUSA Japan Bangladesh

Lab

Medicine storage place Warehouse

Lab Yam Benin

* Storage pest

Tribolium confusum* Tribolium castaneum* Cryptolestes pusillus* Tribolium castaneum* Tribolium confusum* Crytolestes pusillus* Rhizopertha dominica* Trogoderma granarium* Stegobium paniceum* Lasioderma serricorne* Tribolium castaneum*, Oryzaephilus surinamensis*, Ahasverus advena*, Typhaea stericorea*, Carpophilus dimidiatus* Acanthoscelides obtectus, Callosobruchus maculatus Dinoderus porcellus*

References Helgadóttir et al. (2017) Blaeser et al. (2004) Alvarado et al. (1997) Mouratidis et al. (2022) Blaeser et al. (2004) Ballal et al. (2016) Ballal and Yamada (2016) Ballal et al. (2016) Nielson and Henderson (1959) Qin (1985) Imamura et al. (2008)



Davranoglou (2011)



Ballal et al. (2016) Ballal et al. (2016) Rahman et al. (2009) Ahmed et al. (1991)

Awadallah et al. (1986) Brower and Press (1992)

Berger et al. (2017) Loko et al. (2013)

482

14

Prey Record of Various Predators

Table 14.10 Prey record and their preferred stages of pests by various predators of Miridae of Heteroptera Predator Cyrtopeltis (=Nesidiocoris tenuis Deraeocoris lutescens Dicyphus maroccanus Dicyphus tamaninii

Location and crop

Lab Tomato—Spain

Rhopalosiphum padi Ephestia kuehniella Tuta absoluta Ephestia kuehniella Vegetable crop pests

Cage study

Whitefly

Dortus primarius

Tomato

Dicyphus errans

Lab

Europe Lab

Germany, Spain

Distantiella theobroma

Prey stage

Nymphs Eggs Eggs

Frankliniella schultzei

Dicyphus maroccanus Dicyphus hesperus

Dicyphus tamaninii

Prey Ephestia kuehniella

Dicyphus errans



Ephestia kuehniella

Eggs

Echinothrips americanus, Trialeurodes vaporariorum, Tuta absoluta Tuta absoluta



Aphis gossypii, Tetranychus urticae, Frankliniella occidentalis, Trialeurodes vaporariorum Aphis gossypii



Germany, Spain

Frankliniella occidentalis

Lab

Pheidole megacephala, Camponotus species, Oecophylla longinoda





References Sánchez (2009) Lamine et al. (2005) Abbas et al. (2014) Albajes et al. (1996) Gabarra et al. (1995) Varshney and Budhlakoti (2022) Ingegno et al. (2013) Alma et al. (2010) Ingegno et al. (2017)

Ingegno et al. (2013) Ghabeish et al. (2010)

Saleh and Sengonca (2000, 2001a, b) Castane et al. (1997) and Zegula and Sengonca (2000) Williams (1954) and Marchart (1968) (continued)

14.8

Future Recommendations

483

Table 14.10 (continued) Predator

Engytatus nicotianae Macrolophus caliginosus Macrolophus pygmaeus Macrolophus caligin nosus Nesidiocoris tenuis

Location and crop Lab

Prey Pheidole megacephala, Camponotus species, Oecophylla longinoda

Green house

Bactericera cockerelli

Lab

Myzus persicae, Tetranychus urticae Acyrthosiphon pisum

Lab Lab Lab

Trialeurodes vaporariorum Spodoptera litura

Lab

Ephestia kuehniella

Lab

Whiteflies, thrips

Tomto

Helicoverpa armigera

Lab

Trialeurodes vaporariorum Bemisia tabaci

Lab

Tomato Tomato

Tuta absoluta Ephestia kuehniella Bemisia tabaci

Lab

Bemisia tabaci

Lab

Tuta absoluta

Lab

Myzus persicae

Lab

Whitefly

Tomatoes, India

Lagenaria siceraria

Lab

Whitefly

Prey stage

Aphids

Eggs

Eggs

References Williams (1954) and Marchart (1968) Veronesi et al. (2022) Foglar et al. (1990) Duran Prieto et al. (2016) Hamdan (2006) Wei et al. (1998) Urbaneja et al. (2005) Calvo et al. (2009) Devi et al. (2002) Arzone et al. (1990) CarneroHernández et al. (2000) Mollá et al. (2014) Arnó et al. (2010) Ziaei Madbouni et al. (2017) Sánchez et al. (2014) Duarte et al. (2014) Fernández et al. (2012) Singh and Rizvi (1993) and Hameed et al. (1975) Fernández et al. (2012) (continued)

484

14

Prey Record of Various Predators

Table 14.10 (continued) Predator Nesidiocoris volucer

Location and crop Tomato

Prey Trialeurodes vaporariorum, Ephestia kuehniella, Bemisia tabaci, Thrips parvispinus Lygus lineolaris, Lygocoris communis Anticarsia gemmatalis

Prey stage



References Marquereau et al. (2022)

Arnoldi et al. (1991) Godfrey et al. (1989) Barbosa et al. (2019) Callan (1975) Van Doesburg (1964) Myers (1935)

Phymata pennsylvanica Spanogonicus albofasciatus Thaumastocoris peregrinus Termatophylidea sp. Termatophylidea opaca

Apple orchard

Termatophylidea pilosa

Trinidad

Tupiocoris cucurbitaceus Tupiocoris cucurbitaceus Tupiocoris cucurbitaceus

Argentina— Tomato

Tupiocoris cucurbitaceus Tytthus parviceps

Argentinian tomato Sorghum—India

Tuta absoluta

Eggs

Peregrinus maidis

Eggs and nymphs

Dicyphus tamaninii, Macrolophus caliginosus Deraeocoris sp. Campylomma nicolasi Macrolophus pygmaeus, Nesidiocoris tenuis Macrolophus basicornis, Engytatus varians Nesidiocoris tenuis, Macrolophus pygmaeus

Tomato

Macrosiphum euphorbiae

Alvarado et al. (1997)

India-Lab

Bemisia tabaci

Kapadia and Puri (1991)

Italy

Tuta absoluta

Nannini et al. (2012)

Lab

Tuta absoluta, Neoleucinodes elegantalis Tuta absoluta, Ephestia kuehniella

Martínez et al. (2022)

Lab

Eucalyptus benthamii

Lab

Selenothrips rubrocinctus Selenothrips ruhrocinctu

Coco plant

Tomato, Pepper Tobacco

Lab

Eggs

Selenothrips (=Heliothrips) rubricinctus Trialeurodes vaporariorum Whiteflies Sitotroga cerealella Bemisia tabaci

Eggs

Del Pino et al. (2009) Lois et al. (2022) Orozco Muñoz et al. (2012) Cáceres et al. (2022) Dharavath et al. (2022)

Mollá et al. (2014) (continued)

14.8

Future Recommendations

485

Table 14.10 (continued) Predator Campyloneuropsis infumatus, Engytatus varians Macrolophus pygmaeus, Nesidiocoris tenuis Macrolophus basicornis, Engytatus varians, Campyloneuropsis infumatus Nesidiocoris tenuis, Macrolophus pygmaeus, Dicyphus maroccanus Nesidiocoris tenuis, Macrolophus praeclarus Cyrtopeltis = Engyatus varians Campyloneuropsis infumatus, Engytatus varians, Macrolophus basicornis Macrolophus basicornis, Engytatus varians, Campyloneuropsis infumatus Campyloneuropsis infumatus, Engytatus varians, Macrolophus basicornis

Location and crop Tomtto in Europe

Prey Tuta absoluta

Prey stage –

Tomtto

Tuta absoluta



Urbaneja et al. (2009)

Tomtto

Tuta absoluta

Eggs and larvae

Silva et al. (2016)

Lab

Tuta absoluta, Bemisia tabaci



Pérez-Hedo et al. (2015)

Tomato Tobacco

Tuta absoluta

Eggs

Martínez et al. (2014)

Brazil—Tomato

Ephestia kuehniella

Eggs

Bueno et al. (2018)

Tomato

Tuta absoluta

Eggs

van Lenteren et al. (2016)

Brazil—Tomato

Ephestia kuehniella

Eggs

Bueno et al. (2018)

References van Lenteren et al. (2018)

486

14

Prey Record of Various Predators

Table 14.11 Prey record and their preferred stages of pests by various predators of Phymatidae, Thaumastocoridae, Lygaeidae, or Geocoridae of Heteroptera Predator Phymata pennsylvanica (Phymatidae) Thaumastocoris peregrinus (Thaumastocoridae) Geocoris amabilis

Prey stage –

Location Apple orchard

Prey Lygus lineolaris, Lygocoris communis

Lab

Eucalyptus benthamii

Barbosa et al. (2019)

Syria

Gynaikothrips uzeli

Geocoris erythrocephalus Geocoris punctipes

Cabbage-India

Brevicoryne brassicae

Lab

Anticarsia gemmatalis

Eggs

Geocoris punctipes

Cotton

Heliothis zea, Heliothis virescens

Eggs Earley learvae Eggs

Ali and Streito (2019) Rajan et al. (2018) Godfrey et al. (1989) Lingren et al. (1968)

Lab

Helicoverpa zea, Acyrthosiphum pisum Helicoverpa zea, Aphis gossypii Helicoverpa zea

Geocoris punctipes

Lab

Anasa tristis

Geocoris punctipes sonoraensis

Lab

Collops vittatus

Geocoris floridanus

Lab

Helicoverpa zea, Spodoptera exigua

Geocoris ochropterus, Geocoris ochropterus

India

Thrips

India

Lab

Mylloceous viridanus, Calocoris anguslatus Ragmus importunitus, Coptosoma crubrarria, Aphid gossypii, Aphis nerii, Pseudococcus sp. Dachylopius nipae Helicoverpa armigera

Lab

Sitotroga cerealella

Lab

Sitotroga cerealella

Caged cotton

Eggs Adults Eggs

Eggs Young larvae

References Arnoldi et al. (1991)

Eubanks and Denno (2000) Tillman and Mullinix (2003) Ruberson et al. (2001) Fair and Braman (2019) Nielson and Henderson (1959) Torres et al. (2004) Kumar and Ananthakrishnan (1985) Bal (2013)

Eggs

Varshney et al. (2018) Varshney and Ballal (2017a, b) Ballal (2017) (continued)

14.8

Future Recommendations

487

Table 14.11 (continued) Predator

India-Lab

Prey Oecophylla smaragdina, Bombyx mori, Aphis gossypii Bemisia tabaci

Geocoris superbus

Lab

Mealybug

Orius tantills

India

Geocoris varius

Lab

Thrips, aphids and mites Ephestia kuehniella

Geocoris varius

Watermelon

Geocoris varius, Geocoris proteus Geocoris varius, Geocoris proteus

Lab

Geocoris punctipes, Geocoris uliginosus

Location Lab

Lab

Lab Soybean Field cage

Thrips, spider mites and aphids Ephestia kuehniella Tetranychus urticae, Frankliniella occidentalis, Aphis gossypii, Helicoverpa armigera Spodoptera frugiperda Pseudoplusia includen

Prey stage

References Ngoc Bao Chau et al. (2021)



Kapadia and Puri (1991) Varshney and Ballal (2017a, b) Bal (2013)

Eggs

Yumika et al. (2012) Fukao et al. (2012) Oida and Kadono (2012) Oida and Kadono (2011)

Thrips Eggs Adult 2nd instar Adults Eggs

Joseph (2006) Larvae

Bal (2013). Handbook on Major Hemipteran Predators of India: 1–44

Richman et al. (1980)

488

14

Prey Record of Various Predators

Table 14.12 Prey record and their preferred stages of pests by various predators of Reduviidae of Heteroptera Predator Acanthspis pedestris

Location India

Prey Helicoverpa armigera, Spodoptera litura Corcyra cephalonica

Prey stage Larvae

Helicoverpa armigera, Spodoptera litura, Pectinophora gossypiella, Earias insulena Tylotropidus variecornis, Orthacris maindroni Termites

Larvae

India

Coptotermes heimi



India

Mylabris purtulata



India

Helicoverpa armigera, Spodoptera litura, Pectinophora gossypiella, Corcyra cephalonica Corcyra cephalonica



Laphygma exiguae Odontotermes wallonensis Dysdercus koenigii, Dysdercus laetus Oecophylla longinoda Corcyra cephalonicaa, Tribolium confusuma, Anagasta kuehniellaa Dinoderus porcellusa

– –

Lab India

India

Acanthaspis petax Acanthaspis quinquespinosa

India India India India Acanthaspis vitticollis Allaeocrcmum biannulipes

Alloeocranum biannulipes Amphibolus venator

Lab Lab

Yam Benin

Dinoderus porcellusa

Lab

Tribolium confusuma

Larvae

References George and Ravichandran (2021) Sahayaraj et al. (2008) Sahayaraj and Ambrose (1994)



Balakrishnan et al. (2011)



Odhiambo (1958) Ambrose et al. (2008) Ambrose (1988) Sahayaraj (1991)



– – –

Grubs and aduts

Sahayaraj (2007) Butani (1958) Rajagopal (1984) Lakkundi (1989) Louis (1974) Awadallah et al. (1984) Loko et al. (2022) Loko et al. (2013) Youssef and Abd-Elgayed (2015) (continued)

14.8

Future Recommendations

489

Table 14.12 (continued) Predator

Location Lab

Lab Japan

Apiomerus duckei, Apiomerus pilipes, Apiomerus luctuosus Apiomerus crassipes

Brazil

Arilus gallus

Coffee Colombia

Arilus cristatus

Soybean Field cage

Lab

Atopozelus opsimus Catamirus brevipennis

Coranus africana

Coranus siva

Lab India

Prey Plodia interpunctellaa, Tribolium confusuma, Trogoderma granariuma, Lasioderma serricornia, Rhizopertha dominicaa Tribolium confusum, Trogoderma granarium, Corcyra cephalonica, Ephestia cautella, Latheticus oryzae, Alphitobius diaperinus, Tribolium castaneum Orchid bee

Apis mellifera, Chauliognathus marginatus, Podabrus rugulosus, Photinus pyralis Galleria mellonella

Pseudoplusia includen

Glycaspis brimblecombei Spodoptera litura

Prey stage

Nishi et al. (2004)



Marsaro Júnior et al. (2022)



Swadener and Yonke (1973)



GiraldoJaramillo et al. (2021) Richman et al. (1980)

Medium sized larvae – –

Dysdercus cingulatus, Spodoptera litura



Tomto

Bemisia tabaci



Lab

Anagasta kuehniella, Corcyra cephalonica



Lab

Spodoptera littoralis, Agrotis ypsilon Hypothenemus hampei



Coffee India

References Abd-Elgayed and Youssef (2015)



Dias et al. (2012) Sahayaraj et al. (2006) Sahayaraj and Vembudurai (2016) El-Sebaey and Abd El-Wahab (2011) El- Sebaey and El-Bishry (2001) El-Sebaey (2001) Kiran and Venkatesha (2022) (continued)

490

14

Prey Record of Various Predators

Table 14.12 (continued) Prey Corcyra cephalonica, Earias vitella, Leptocoris vericornis, Riptortus clavatus, Dysdercus cingulatus Anagasta kuehniella

Prey stage –

References Claver and Reegan (2010)

Larvae

Spodoptera frugiperda Spartocera dentiventris



Helopeltis antonii



Endochus inornatus

Helopeltis antonii



Endochus africanus

D. theobroma



El-Shazly and Iman (1997) Firake and Behere (2020) da Rocha and Redaelli (2004) Srikumar et al. (2014) Naik and Sundararaju (1982) Collingwood (1971) Sahayaraj and Ambrose (1996) Louis (1974) Bal (2013) Mouly et al. (2018)

Predator Coranus spiniscutis

Location India

Coranus aegpus

Lab

Cosmolestes sp.

India Corn

Cosmoclopius nigroannulatus

Lab

Cydnocoris gilvus

Lab India

Nymphs

Ectomocoris tibialis

Lab

Dysdercus cingulatus



Eugubinus annulatus Irantha armipes Isyndus heros

Cocoa Rice India Mango India

Drosophila spp. Leptocorisa acuta Idioscopus nitidulus, Idioscopus niveosparsus, Idioscopus clypealis, Amritodus atkinsoni, Orthaga euadrusalis Cladomorphus phyllinus Hylesia paulex



Harpactor angulosus Harpactor angulosus

Brazil

Harpactor costalis

Tobacco India Ghana Corn Brazil

Haematochares obscuripennis Harpactor angulosus Margasus afzelii Odontogonus dimensis Pisilus tipuliformis Pselliopus latispina

Lab



– –

Spodoptera litura

Larvae

Spodoptera frugiperda Cladomorphus phyllinus Oecophylla longinoda Bathycoelia thalassina Physopelta analis Tetranychus urticae

– – – Eggs – –

Costa et al. (2022) Pereira et al. (2009) Sitaramaiah et al. (1975) Koffi et al. (2020) Costa et al. (2022) Louis (1974) Louis (1974) Louis (1974) Ordaz-Silva et al. (2014) (continued)

14.8

Future Recommendations

491

Table 14.12 (continued) Predator Panthous bimaculatus Peprius nodulipes

Phonoctonus spp

Lab

Prey Corcyra cephalonica, Spodoptera litura Spodoptera frugiperda Helicoverpa armigera, Nezara viridula Helicoverpa spp. Creontiades spp. Pseudoplusia includens Dysdercus spp.

Phonoctonus nigrofasciatus Phonoctonus spp. Phonoctonus lutescens Phonoctonus fasciatus Phonoctonus lasciatus, Phonoctonus subimpictus, Phonoctonus lutescens Pristhesancus plagipennis

Lab

Dysdercus

Nymphs

Martin and Brown (1984) Schaefer and Ahmad (1987) Evans (1962)

Field

Dysdercus spp. O. sexpunctatus

– –

Fadare (1978) Stride (1956a)

Dysdercus superstitiosus Dysdercus superstitiosus



Stride (1956a)



Stride (1956b)

Drosophila sp., Tribolium castaneum, Tenebrio molitor, Biprorulus bibax, Nezara viridula Carpophilus mutilatus, Ahasverus advena, Oryzaephilus mercator, Tribolium castaneum, Araecerus fasciculatus, Necrobia rufipes Oryctes rhinoceros



James (1994)



Zipagan and Pacumbaba (1994)

Grubs

Antony et al. (1980) Vanderplank (1958)

Pristhesancus plagipennis

Location Lab India Ghana Corn Lab

Cotton Australia

Field Lab

Lab

Perigrinator biannulipes

Copra

Platymeris laevicollis

Coconut India

Platymeris rhadamanthus

Oryctes boas, Oryctes monoceros, Solenostethium liligerum, Caura rufiventris, Lerida pugnax, Nezara viridula, Pseudotheraptus wayi

Prey stage – – –

References Muthupandi et al. (2014) Koffi et al. (2020) Grundy and Maelzer (2000)



Grundy (2007)

– –



(continued)

492

14

Prey Record of Various Predators

Table 14.12 (continued) Location cotton Centrafrica Lab

Prey Dysdercus

Prey stage –

References Pierrard (1972)

Tribolium castaneum



Rhinocoris albopunctatus

Lab

Larvae

Rhinocoris bicolor Rhinocoris bicolor, Rhinocoris tropicus Rhynocoris iracundus, Nagusta goedelii

Lab

Heliothis armigera, Earias biplaga, E. insulana, Cryptophlebia leucotreta, Pectinophora gossypiella S. singularis Pisilus tipuliformis

Kwadjo et al. (2008) Niziira (1970)

– –

Louis (1974) Parker (1969)

Lab

Halyomorpha halys

Bulgarini et al. (2019)

Rhynocoris marginatus

India

Helicoverpa armigera, Spodoptera litura, Mythimna separata, Anomis flava Spodoptera litura, Mylabris pustulata, Dysdercus cingulatus Spodoptera litura, Pieris brassicae Helicoverpa armigera, Spodoptera litura Mythimna separata

Egg and 1st nympal instar –

Predator Rhynocoris albopilosus

India

Lab Lab

Lab Groundnut India

India Lab

Aphis craccivora, Mylabris indica, Mylabris pustulata, Spodoptera litura, Helicoverpa armigera, Atractomorpha crenulata, Chrotogonous trachypterus Spodoptera litura, Sylepta derogata, Pericallia ricini, Mylabris indica, Mylabris pustulata, Dysdercus cingulatus

Pradeep et al. (2022)



Ambrose and Claver (1999)



Arshad et al. (2021) George and Srenivasagan (1998) Pravalika et al. (2016) Sahayaraj and Martin (2003)

– – –



Sahayaraj et al. (2016)

(continued)

14.8

Future Recommendations

493

Table 14.12 (continued) Predator

Location Lab

Prey Spodoptera litura

Prey stage –

India

Helicoverpa armigera, Spodoptera litura, Mythimna separata, Anomis flava Diabolocatantops pinguis, Oxya nitidula, Atractomorpha crenulata, Orthacris maindroni, Trilophidia annulata Spodoptera litura



Lab India

Lab India

Lab India

Lab India Lab India

Rhynocoris fuscipes

India

Pots India

Cotton Field India

India Lab

References Ullah et al. (2019) Pradeep et al. (2022)



Sahayaraj et al. (2021)



Sahayaraj (1995) and Sahayaraj and Paulraj (2001) Ambrose and Claver (1996)

Spodoptera litura, Euproctis mollifera, Mylabris pustulata Eutectona machaeralis Amsacta albistriga, Aproaerema modicella, Helicoverpa armigera, Spodoptera litura Phenacoccus solenopsis, Corcyra cephalonica, Dysdercus koenigii Phenacoccus solenopsis, Dysdercus cingulatus Aphis gossypii, Phenacoccus solenopsis, Dysdercus cingulatus, Helicoverpa armigera Spodoptera litura



Mylabris pustulata, Dysdercus cingulatus



– –



Ambrose et al. (2013) Sahayaraj (1999)

Tomson (2021)

– –



Ambrose and Claver (1997) Ambrose and Claver (1999) (continued)

494

14

Prey Record of Various Predators

Table 14.12 (continued) Predator

Location Lab

Rhynocoris fuscipes

India

Rhynocoris kumarii

Lab India

Prey stage –

References Ambrose (1996)



Singh and Gangrade (1976) Sahayaraj et al. (2015) Sahayaraj and Sivakumar (1995) and Claver and Ambrose (2001a) Sahayaraj and Asha (2010) Ambrose (2000)



Lab India

Phenacoccus solenopsis Spodoptera litura

Lab India

Aphis craccivora



Lab India

Helicoverpa armigera, Anomis flava Dysdercus cingulatus



Spodoptera litura, Earias vittella, Corcyra cephalonica Aphis craccivora, Spodoptera litura, Helicoverpa armigera, Chrotogonus trachyterus, Atractomorpha crenulata Aphis gossypii, Dysdercus cingulatus, Phenacoccus solenopsis, Helicoverpa armigera Aphis gossypii, Phenacoccus solenopsis, Dysdercus cingulatus



Lab India

Lab India

Groudnut India

Rhynocoris longifrons

Prey Spodoptera litura, Helicoverpa armigera, Mylabris pustulata, Dysdercus cingulatus, Achaea janata Diacrisia obliqua

Screen house India

Lab India





Claver and Ambrose (2001b) George (2000)



Sahayaraj and Ravi (2007)



Sahayaraj et al. (2020)



Sahayaraj et al. (2012)

(continued)

14.8

Future Recommendations

495

Table 14.12 (continued) Predator

Location Lab India

Prey stage –

References Kumar and Ambrose (2014)

Lab

Prey Spodoptera litura, Helicoverpa armigera, Achaea janata, Mylabris indica, Dysdercus cingulatus Tribolium castaneuma

Rhynocoris squamulosus Rhynocoris iracundus



Europe

Halyomorpha halys



Rihirbus trochantericus Ricolla pallidinervis, Rocconota tuberculigera Sinea diadema

Cashew India Panama

Helopeltis spp.



Oebalus insularis

Adults

Seydou et al. (2021) Bulgarini et al. (2021) Bhat et al. (2013) Ramírez Silva et al. (2022)



Knowlton (1944)

Sinea sanguisuga

Lab

Syrphid fly Pea aphid, Alfalfa weevil larva and caterpillar Laphygma sp.

Larva

Sinea spinipes

Lab

Chlorochroa sayi

Nymph

Sinea confuse

Lab

Collops vittatus



Sycanus annulicornis

Lab



Lab

Crocidolomia pavonana, Tenebrio molitor, Setothosea asigna Helopeltis

Luginbill (1928) Caffrey and Barber (1919) Nielson and Henderson (1959) Sahid and Natawigena (2018)

Sycanus annulicornis

Lab

Setothosea asigna



Sycanus affinis

Lab India

Corcyra cephalonica



Sycanus collaris

Lab

Tenebrio molitor



Lab

Corcyra cephalonica, Spodoptera litura Corcyra cephalonica

Larvae

Lab

Lab India

Caterpilars

Larvae

Kalshoven (1981) Sahid et al. (2018) Satpathy et al. (1975) Maneerat and Sakarind (2018) and Poopat and Maneerat (2021) Rajan et al. (2017) Saurabh et al. (2019) (continued)

496

14

Prey Record of Various Predators

Table 14.12 (continued) Predator Sycanus aurantiacus

Location Lab

Sycanus indagator

Lab

Sycanus galbanus

Lab

Sycanus dichotomus

Lab Lab Lab

Prey Plutella xylostella, Crocidolomia pavonana, Tenebrio molitor Galleria mellonella, Spodoptera frugiperda Galleria mellonella, Helopeltis antonii Metisa plana

Prey stage –

References Yuliadhi et al. (2021)

Larvae

Bass and Shepard (1974)



Plutella xylostella, Corcyra cephalonica Tenebrio molitor



Nitin et al. (2017) Syari et al. (2010, 2011) Zulkefli et al. (2004) Syari et al. (2011) and Yusof and Fairuz (2011) Ahmad and Kamarudin (2016)

Lab

Sphedanolestes himalayensis Sphedanolestes variabilis Sphedanolestes minusculus Tegea atropicta Rhynocoris fuscipes, Endocus inornatus Endochus albomaculatus, Epidaus bicolor, Euagoras plagiatus, Irantha armipes, Panthous bimaculatus, Sphedanolestes signatus Sycanus collaris, Sphedanolestes signatus, E. inornatus, Irantha armipes, Occamus typicus





Lab

Sycanus reclinatus





Lab India

Tenebrio molitor, Corcyra cephalonica Corcyra cephalonica

Larvae

India

Corcyra cephalonica

Larvae

Lab India

Corcyra cephalonica

Larvae

Lab India

Nasutitermes exitiosus Pterophorus lienigianus

– –

Cashew India

Helopeltis antonii



India

Helopeltis antonii



Das et al. (2008) Ambrose et al. (2009) Ambrose et al. (2006) Casimir (1960) Anand and Chandral (2011)

Sundararaju (1984) and Vennison and Ambrose (1990) (continued)

14.8

Future Recommendations

497

Table 14.12 (continued) Predator Vachiria natolica

Location Lab

Prey Anagasta kuehniella

Prey stage Larvae

Vesbius sanguinosus

Lab

Corcyra cephalonica



Vestula lineaticeps

Lab



Zelus tetracanthus, Pselliopus zebra Zelus socius

Mexico

Crematogaster clariventris Monoctenus sanchezi Lygus lineolaris, Lygocoris communis Spodoptera frugiperda Monalonion bondari

– Eggs and larvae –

Dalbulus maidis



Small Coleoptera, Diptera, Hemiptera and other small arthropods Easchistus Chlorochroa ligata Collops vittatus



Egg Nymphs –

Circalifer tenellus



Choristoneura houstonana

Larvae and adult

Tenebrio, Zonocerus uariegatus

Larvae

Zelus longipes

Apple orchard Mexico

Zelus pedestris

cacao

Zelus obscuridorsis Zelus renardii

Corn Argentina Lab

Zelus renardii Zelus renardii Zelus renardii

Lab Lab Lab

Zelus socias

Hediocoris tibialis, Pisilus tipuliformis, Rhinocoris bicolor, Rhinocoris carmelita, Rhinocoris loratus, Sphedanolestes leucocephalus, Vestula lineatice a

Storage pest

Lab



References El-Shazly and Iman (1997) Das and Ambrose (2008) Louis (1974) Ordaz-Silva et al. (2021) Arnoldi et al. (1991) Hoballah et al. (2004) Santos et al. (2022) Virla et al. (2015) Ables (1978)

Clancy (1946) Morrill (1910) Nielson and Henderson (1959) Knowlton (1932) Heinrichs and Thompson (1968) Louis (1974)

498

14

Prey Record of Various Predators

Fig. 14.1 Longlegged fly (Dolichopodidae: condylostylus caudatus) and chironomid midge prey (a), Polistes wasp mimics Mantidfly (Mantispidae: Climaciella brunnea) mimics and tipulid fly prey (b), Robber fly (Asilidae: Promachus sp.) and crane fly (Tipulidae) prey (c1 and c2), Robber fly (Asilidae: Laphria canis or L. sicula) and flatid planthopper prey (d), Robber fly (Asilidae: Atomosia sp.) and alate male ant (Tapinoma sessile) prey (e), Variable lady beetle (Coccinellidae: Coelophora inaequalis) preying on aphids (f), Rambur's forktail damselfly (Coenagrionidae: Ischnura ramburii) and moth prey (g), and Eastern pondhawk dragonfly (Libellulidae: Erythemis simplicicollis) female and conspecific female dragonfly prey (h). (Photos Courtesy by Gerald S. Wegner, USA)

Fig. 14.2 Ambush bugs (Reduviidae: Phymata sp.) with blow fly (a, b) and moth (c) prey, Spined assassin bug, Sinea sp., nymph and speckled sharpshooter (Paraulacizes irrorata) (d) and Black bee assassin bug (Reduviidae: Apiomerus longispinis) with flower scarab beetle prey (e). (Photos Courtesy by Gerald S. Wegner, USA)

14.8

Future Recommendations

499

Fig. 14.3 Pink (twelve-) spotted lady beetle (Coccinellidae: Coleomegilla maculata) larva and hover fly (Syrphidae: Uroleucon sp.) larva preying on aphids (a), multicolored Asian lady beetle (Coccinellidae: Harmonia axyridis) larva preparing to feed on moth eggs (b), and Western red-bellied tiger beetle (Carabidae: Cincindelidia sedecimpunctata) and moth prey (c). (Photos Courtesy by Gerald S. Wegner, USA)

Fig. 14.4 Florida predatory stink bug (Pentatomidae: Euthyrhynchus floridanus) with gulf fritillary butterfly larva prey (a), Predatory stink bug (Pentatomidae: Conquistator mucronatus) and muscoid fly prey (b and c) and Antilochus coqueberti adult feeding on Dysdercus koenigii adults (d). (Photos Courtesy by Gerald S. Wegner, USA)

Fig. 14.5 Rhynocoris kumarii nymphs feeding on hairy caterpillar (a), adults feeding on hairy caterpillar (b), and mealybugs (c and d)

500

14

Prey Record of Various Predators

Fig. 14.6 Rhynocoris fuscipes nymphs feeding on melybugs (a), adults feeding on melybugs (b), red cotton bug (c), Spodoptera litura (d), and Helicoverpa armigera (e)

Fig. 14.7 Rhynocoris longifrons female feeding on Corcyra cephalonica larvae (a), Dysdercus koenigii nymph (b), and nymph feeding on Spodoptera litura lavae (c, d), mealybug (e), and aphid (f)

References

501

References Abbas S, Pérez-Hedo M, Colazza S et al (2014) The predatory mirid Dicyphus maroccanus as a new potential biological control agent in tomato crops. BioControl 59(5):565–574 Abd-Elgayed AA, El-Khouly NMA (2019) Biological parameters of Philonthus longicornis Steph. (Coleoptera: Staphylinidae) preying on snail, Monacha obstructa pfeiffer (Stylommatophora: Hygromiidae). Egypt J Basic Appl Sci 34(11):30–40 Abd-Elgayed AA, Youssef NA (2015) Effect of some stored insect pest species on biological aspects of the predator, Amphibolus venator Klug (Hemiptera: Reduviidae). Ann Agric Sci 60(1):47–51 Abeijon LM, Kruger AP, Lutinski J et al (2019) Can ants contribute to the conservative biological control of the South American fruit fly? Biosci J 35:941–948 Ables JR (1978) Feeding behavior of an assassin bug, Zelus renardii. Ann Entomol Soc Am 71(4): 476–478 Ables JR, McCommas DW Jr (1982) Efficacy of Podisus maculiventris as a predator of variegated cutworm on greenhouse cotton. J Ga Entomol Soc 17(2):204–206 Abraham VA, Kurian C (1973) Chelisoches moris F.(Forficulidae: Dermaptera), a predator on eggs and early instar grubs of the red palm weevil Rhynchophorus ferrugineus F.(Curculionidae: Coleoptera). J Plant Crops 1:147–152 Adams THL, Chambers RJ, Dixon AFG (1987) Quantification of the impact of the hoverfly, Metasyrphus corollae on the cereal aphid, Sitobion avenae, in winter wheat: laboratory rates of kill. Entomol Exp Appl 43:153–157 Ahmad SH, Kamarudin N (2016) Growth and longevity of the insect predator, Sycanus dichotomus Stål (Hemiptera: Reduviidae) fed on live insect larvae. J Oil Palm Res 28:471–478 Ahmed KN, Khatun M, Rahman MM (1991) Biological notes on Xylocoris flavipes (Reuter) (Hemiptera: Anthocoridae). J Asiat Soc Bangladesh Sci 17:65–67 Albajes R, Alomr O, Riudavests J et al (1996) The mirid bug Dicyphus tamaninii: an effective predator for vegetable crops. Bull OILB SROP (France) 19(1):1–4 Alghamdi A, Sayed S (2017) Biological characteristics of indigenous Chrysoperla carnea (Neuroptera: Chrysopidae) fed on a natural and an alternative prey. Asian J Biol 2(2):1–6 Ali AY, Streito JC (2019) First report in Syria of two predatory true bugs: Montandoniola indica (Hemiptera: Anthocoridae) and Geocoris amabilis (Hemiptera: Geocoridae). Mun Entomol Zool 14(2):427–431 Allen WR, Hagley EAC (1990) Epigeal arthropods as predators of mature larvae and pupae of the apple maggot (Diptera: Tephritidae). Environ Entomol 19:309–312 Alloui-Griza R, Cherif A, Attia S et al (2022) Lethal Toxicity of Thymus capitatus Essential Oil Against Planococcus citri (Hemiptera: Pseudococcidae) and its Coccinellid Predator Cryptolaemus montrouzieri (Coleoptera: Coccinellidae). J Entomol Sci 57(3):425–435 Alma CR, Gillespie DR, Roitberg BD et al (2010) Threat of infection and threat-avoidance behavior in the predator Dicyphus hesperus feeding on whitefly nymphs infected with an entomopathogen. J Insect Behav 23(2):90–99 Almohamad R, Verheggen F, Haubruge É (2009) Searching and oviposition behavior of aphidophagous hoverflies (Diptera: Syrphidae): a review. Biotechnol Agron Soc Environ 13(3):467–481 Aluja M (1994) Bionomics and management of Anastrepha. Annu Rev Entomol 39:155–173 Alvarado P, Balta O, Alomar O (1997) Efficiency of four heteroptera as predators ofAphis gossypii andMacrosiphum euphorbiae (Hom.: Aphididae). Entomophaga 42(1):215–226 Ambrose DP (1988) Biological control of insect pests by augmenting assassin bugs (Insecta: Heteroptera: Reduviidae). Bicovas 2:25–40 Ambrose DP (1996) Assassin bugs (Insecta: Heteroptera: Reduviidae) in biocontrol: success and strategies, a review. In: Ambrose DP (ed) Biological and cultural control of Insect pests, an Indian Scenario. Adeline publishers, Tirunelveli, pp 262–284

502

14

Prey Record of Various Predators

Ambrose DP (2000) Suppression of Helicoverpa armigera (Hubner) and Anomis flava (Fabricius) infesting okra by the reduviid predator Rhynocoris kumarii Ambrose and Livingstone in field cages. Pest Manag Hortic Ecosyst 6:32–35 Ambrose DP, Claver MA (1996) Size preference and functional response of the reduviid predator Rhynocoris marginatus Fabricius (Heteroptera: Reduviidae) to its prey Spodoptera litura (Fabricius) (Lepidoptera: Noctuidae). J Biol Contr 10(1/2):29–37 Ambrose DP, Claver MA (1997) Functional and numerical responses of the reduviid predator, Rhynocoris fuscipes F.(Het., Reduviidae) to cotton leafworm Spodoptera litura F.(Lep., Noctuidae). J Appl Entomol 121(1,5):331–336 Ambrose DP, Claver MA (1999) Suppression of cotton leaf worm Spodoptera litura, flower beetle Mylabris pustulata and red cotton bug Dysdercus cingulatus by Rhynocoris marginatus (Fabr.) (Heteroptera: Reduviidae) in cotton field cages. J Appl Entomol 123:225–229 Ambrose DP, Kumar SP, Nagarajan K et al (2006) Redescription, biology, life table, behaviour and ecotypism of Sphedanolestes minusculus Bergroth (Hemiptera: Reduviidae). Entomol Croat 10(1-2):47–66 Ambrose DP, Raja JM, Rajan SJ (2008) Functional response of Acanthaspis quinquespinosa (Fabricius) (Hemiptera: Reduviidae) on Coptotermes heimi (Wasmann). J Biol Contr 22(1): 163–165 Ambrose DP, Rajan XJS, Nagarajan K et al (2009) Biology, behaviour and functional response of Sphedanolestes variabilis Distant (Insecta: Hemiptera: Reduviidae: Harpactorinae), a potential predator of lepidopteran pests. Entomol Croat 13(2):33–44 Ambrose DP, Nagarajan K, Kumar AG (2013) Interaction of reduviid predator, Rhynocoris marginatus (Fabricius)(Hemiptera: Reduviidae) with its prey teak skeletonizer, Eutectona machaeralis Walker (Lepidoptera: Pyralidae) as revealed through functional response. J Entomol Res 37(1):55–59 Amiri-Jami AR, Sadeghi-Namaghi H (2014) Responses of Episyrphus balteatus DeGeer (Diptera: Syrphidae) in relation to prey density and predator size. J Asia Pac Entomol 17(3):207–211 Anand BG, Chandral S (2011) Eco-friendly technology for the management of brinjal pest using reduviids. Int J Appl Bioeng 10(09):21–24 Anlaş S, Yener H, Yağmur EA (2021) Notes on the seasonal dynamics of some Paederinae (Coleoptera: Staphylinidae) species in the vineyards of Manisa, Western Anatolia. J Entomol Res Soc 23(2):121–132 Antony J, Daniel M, Kurian C et al (1980) Attempts on introduction and colonization of the exotic reduviid predator, Platymeris laevicollis Distant for the biological suppression of the coconut rhinoceros beetle Oryctes rhinoceros L. PLACROSYM II: entomology, microbiology, nematology, plant pathology and rodentology of plantation crops/editor-in-chief, CS Venkata Ram Aragón-Sánchez M, Román-Fernández LR, Martínez-García H et al (2018) Rate of consumption, biological parameters, and population growth capacity of Orius laevigatus fed on Spodoptera exigua. BioControl 63:785–794 Arcaya E, Pérez-Bañón C, Mengual X et al (2017) Life table and predation rates of the syrphid fly Allograpta exotica, a control agent of the cowpea aphid Aphis craccivora. Biol Control 115:74– 84 Arnó J, Castañé C, Riudavets J et al (2010) Risk of damage to tomato crops by the generalist zoophytophagous predator Nesidiocoris tenuis (Reuter) (Hemiptera: Miridae). Bull Entomol Res 100(1):105–115 Arnoldi D, Stewart RK, Boivin G (1991) Field survey and laboratory evaluation of the predator complex of Lygus lineolaris and Lygocoris communis (Hemiptera: Miridae) in apple orchards. J Econ Entomol 84(3):830–836 Arshad M, Khan HAA, Hafeez F et al (2017) Predatory potential of Coccinella septempunctata L. against four aphid species. Pak J Zool 49(2):623–627 Arshad M, Ullah MI, Khan RR et al (2021) Demographic parameters of the reduviid predator, Rhynocoris marginatus (Reduviidae: Hemiptera) fed on two lepidopterous insect pests. BioControl 66:227–235

References

503

Arzone A, Alma A, Tavella L (1990) Roulo dei Miridi (Rhynchota: Heteroptera) in the control of Trialeurodes vaporariorum Wetswood (Rhynchota; Aleyrodidae). Bull Zool Agraria Bachicoltura 22:43–51 Asgari A (1966) Untersuchungen liber die im Raum Stuttgart-Hohenheim als wichstigste Priidatoren der grlinen Apfelblattlaus (Aphidula pomi DeG.) auftretenden Arthropoden. Z Angew Zool 53:35–93 Aslam M, Shaheen FA, Ayyaz A (2006) Management of Callosobruchus chinensis Linnaeus in stored chickpea through interspecific and intraspecific predation by ants. World J Agric Res 2(1):85–89 Aslan MM, Uygun N (2005) The Aphidophagus Coccinellid (Coleoptera: Coccinellidae) species in Kahramanmaraş, Turkey. Turk J Zool 29(1):1–8 Atim AB, Graham HM (1984) Predation of Geocoris punctipes by Nabis alternatus. Southwest Entomol 9:227–231 Avery PB, George J, Markle L et al (2022) Choice behavior of the generalist pentatomid predator Podisus maculiventris when offered lepidopteran larvae infected with an entomopathogenic fungus. BioControl 67(2):201–211 Awadallah KT, Tawfik MFS, Abdellah MMH (1984) Suppression effect of the reduviid predator, Allaeocranum biannulipes (Montr, et Sign.) on populations of some storedproduct insect pests. Zeitschrift für Angewandte Entomologie 97(1,5):249–253 Awadallah KT, Tawfik MFS, El-Husseini MM (1986) Bio-cycle of the anthocorid predator Xylocoris flavipes (Reuter) in association with rearing on major pests of stored drug materials. Bulletin de la Société Entomologique d'Égypte 66:27–33 Bacci L (2009) Picanço Marcelo Coutinho, Rosado Jander Fagundes, Silva Gerson Adriano, Crespo André Luiz Barreto, Pereira Eliseu José Guedes, Martins Júlio Cláudio. Appl Entomol Zool 44(1):103–113 Badgley ME, Fleschner CA (1956) Biology of Oligota oviformis Casey (Coleoptera: Staphylinidae). Ann Entomol Soc Am 49(5):501–502 Badji CA, Guedes RNC, Silva AA et al (2004) Impact of deltamethrin on arthropods in maize under conventional and no-tillage cultivation. Crop Prot 23:1031–1039 Balakrishnan P, Nagarajan K, Kumar AG et al (2011) Host preference, stage preference and functional response of Acanthaspis pedestris Stål (Hemiptera: Reduviidae) to its most preferred prey the acridid grasshopper, Orthacris maindroni Boliver. Insect Pest Management, a Current Scenario. Entomology Research Unit, St. Xavier’s College, Palayamkottai, pp 210–217 Bal A (2013) Handbook on major Hemipteran predators of India. Zoological Survey of India Ballal CR (2017) Biological, morphological and life table parameters of the predator, Geocoris ochropterus Fieber. (Hemiptera: Geocoridae), Fed on Sitotroga cerealella (Olivier) eggs. Egypt J Biol Pest Contr 27(2) Ballal and Yamada (2016) Anthocorid predators. In: OMKAR (ed) Ecofriendly pest management for food security. Elsevier, London, pp 329, 727 pp–366 Ballal CR, Yamada K (2016) “Anthocorid predators.” Ecofriendly pest management for food security. Academic Press, pp 183–216 Barbosa L, Santos F, Soliman E et al (2019) Biological parameters, life table and thermal requirements of demographic parameters of the reduviid predator, Thaumastocoris peregrinus (Heteroptera: Thaumastocoridae) at different temperatures. Sci Rep 9:10174 Barrion AT, Litsinger JA (1985) Proreus simulans (Dermaptera: Chelisochidae), a predator of rice leaffolder (LF) and skipper larvae. Int Rice Res Newslett 10(1):25 Bass JA, Shepard M (1974) Predation by Sycanus Indagator on larvae of Galleria mellonella and Spodoptera frugiperda 1. Entomol Exp Appl 17(2):143–148 Batay-an EH, Torrena PS, Estoy AB (2007) Bioecological studies on rice black bug Scotinophara coarctata (Fabricius) in Cotabato, Mindanao, Philippines. In: Rice Black Bugs, p 339 Behnazar T, Madadi H (2015) Functional response of different stages of Hippodamia variegate (Col.: Coccinellidae) to Diuraphis noxia (Hemiptera: Aphididae) on two wheat cultivars. Biocontr Sci Technol 2:1180–1191

504

14

Prey Record of Various Predators

Berger A, Degenkolb T, Vilcinskas A et al (2017) Evaluating the combination of a parasitoid and a predator for biological control of seed beetles (Chrysomelidae: Bruchinae) in stored beans. J Stored Prod Res 74:22–26 Bhat PS, Srikumar KK, Raviprasad TN et al (2013) Biology, behavior, functional response and molecular characterization of Rihirbus trochantericus Stal var. luteous (Hemiptera: Reduviidae: Harpactorinae) a potential predator of Helopeltis spp. (Hemiptera: Miridae). Entomol News 123(4):264–277 Bilde T, Toft S (1994) Prey preference and egg production of the carabid beetle Agonum dorsale. Entomol Exp Appl 73(2):151–156 Blaeser P, Sengonca C, Zegula T (2004) The potential use of different predatory bug species in the biological control of Frankliniella occidentalis (Pergande) (Thysanoptera: Thripidae). J Pestic Sci 77(4):211–219 Brower JH, Press JW (1992) Suppression of residual populations of stored-product pests in empty corn bins by releasing the predator Xylocoris flavipes (Reuter). Biol Control 2(1):66–72 Brügger BP, Alcántara-de la Cruz R, de Carvalho AG et al (2019) Polybia fasticiosuscula (Hymenoptera: Vespidae) foraging activity patterns. Florida Entomol 102(1):264–265 Bueno VHP, Calixto AM, Montes FC et al (2018) Population growth parameters of three Neotropical mirid predators (Hemiptera: Miridae) at five temperatures on tobacco with Ephestia kuehniella Zeller (Lepidoptera: Pyralidae) eggs as food. Isr J Entomol 48:1–22 Bulgarini G, Badra Z, Maistrello L (2019) Predatory ability of wild generalist predators against eggs and first instar nymphs of Halyomorpha halys. IOBC/WPRS Bull 146:106–107 Bulgarini G, Badra Z, Leonardi S et al (2021) Predatory ability of generalist predators on eggs, young nymphs and adults of the invasive Halyomorpha halys in southern Europe. BioControl 66(3):355–366 Burgess L, Hinks CF (1987) Predation on adults of the crucifer flea beetle, phyllotreta cruciferae (goeze), by the northern fall field cricket, gryllus pennsylvanicus burmeister (orthoptera: Gryllidae). Can Entomol 119(5):495–496 Burgess EPJ, Lövei GL, Malone LA et al (2002) Prey-mediated effects of the protease inhibitor aprotinin on the predatory carabid beetle Nebria brevicollis. J Insect Physiol 48(12):1093–1101 Butani DK (1958) Parasites and predators recorded on sugarcane pests in India. Indian J Ent 20: 270–282 Cáceres RIM, Rocca M, Luna MG (2022) Searching behavior and prey preference by the predatory mirid bug Tupiocoris cucurbitaceus, a potential biocontrol agent of tomato crops in Argentina Caffrey DJ, Barber GW (1919) The Grain Bug. Bulletin No. 779. United States Department of Agriculture Calixto AM, Bueno VHP, Montes FC et al (2013) Effect of different diets on reproduction, longevity and predation capacity of Orius insidiosus (Say) (Hemiptera: Anthocoridae). Biocontr Sci Technol 23(11):1245–1255 Callan EM (1975) Miridae of the genusTermatophylidea [Hemiptera] as predators of cacao thrips. Entomophaga 20:389–391 Calvo J, Bolckmans K, Stansly PA et al (2009) Predation by Nesidiocoris tenuis on Bemisia tabaci and injury to tomato. BioControl 54(2):237–246 Canapi BL, Rubia EG, Litsinger JA et al (1988) Predation by sword-tailed cricket Anaxipha longipennis (Serville) (Gryllidae) on eggs of three lepidopterous pests of rice. Int Rice Res Newslett 13(4):40–41 Carnero-Hernández A, Díaz-Hernández S, Amador-Martín S et al (2000) Impact of Nesidiocoris tenuis Reuter (Hemiptera: Miridae) on whitefly populations in protected tomato crops. In: IOBC WPRS Bulletin Carroll DP, Walker JTS, Hoyt SC (1985) European earwigs (Dermaptera: Forficulidae) fail to control apple aphids on bearing apple trees and woolly apple aphids (Homoptera: Aphididae) in apple rootstock stool beds. J Econ Entomol 78:972–974 Carty MT, Shepard M, Turnipseed SG (1980) Identification of predaceous arthropods in soybeans by using auto radiography. Environ Entomol 9:199–203

References

505

Casimir M (1960) Tegea atropicta Stål (Hemiptera, Reduviidae), an unusual predator of termites. In: Proceedings of the Linnean Society of New South Wales, Sydney, vol 85 Castane C, Alomar O, Riudavets J (1997) Biological control of greenhouse cucumber pests with the mirid bug Dicyphus tamaninii. IOBC/WPRS Bull 20:237–240 Castracani C, Bulgarini G, Giannetti D et al (2017) Predatory ability of the ant Crematogaster scutellaris on the brown marmorated stink bug Halyomorpha halys. J Pestic Sci 90:1181–1190 Chaudhary DD, Kumar B, Mishra G (2022) Functional response in coccinellid beetles (Coleoptera: Coccinellidae) is modified by prey-density experience. Can Entomol 154:e11 Chauthani AR, Misra US (1966) Observations on the life history of Antilochus coqueberti, a predator of Dysdercus koenigi. J Econ Entomol 59:767–768 Chitra N, Soundararajan RP, Gunathilagaraj K (2000) Orthoptera in rice fields of Coimbatore. Zoos’ Print J 15(8):309–311 Clancy DW (1946) Natural enemies of some Arizona cotton insects. J Econ Entomol 39:326–328 Clausen CP, Clancy DW, Chock QC (1965) Biological control of the Oriental Fruit Fly (Dacus Dorsalis Hendel) and other Fruit Flies in Hawaii. Agricultural Research Service; US Department of Agriculture, Washington, DC, p 108 Claver MA, Ambrose DP (2001a) Impact of antennectomy, eye blinding and tibial comb coating on the predatory behaviour of Rhynocoris kumarii Ambrose and Livingstone (Het., Reduviidae) on Spodoptera litura Fabr. (Lep., Noctuidae). J Appl Entomol 125(9,10):519–525 Claver MA, Ambrose DP (2001b) Impact of augmentative release of Rhynocoris kumarii Ambrose and Livingstone (Heteroptera: Reduviidae) on Dysdercus cingulatus (Fabricius) (Hemiptera: Pyrrhocoridae) population and damage on cotton. J Biol Contr 15:119–125 Claver MA, Reegan AD (2010) Biology and mating behaviour of Coranus spiniscutis Reuter (Hemiptera: Reduviidae), a key predator of rice gandhi bug Leptocorisa varicornis Fabricius. J Biopest 3(2):437 Coaker TH, Williams DA (1963) The importance of some Carabidae and Staphylinidae as predators of the cabbage root fly, Erioischia brassicae (Bouché). Entomol Exp Appl 6(2):156–164 Collingwood C (1971) China Past and Present (Book Review). Journal of Contemporary Asia 1 (4):66 Collyer E (1953) Biology of some predatory insects and mites associated with the fruit tree red spider mite (Metatetranvchusulmi (Koch)) in south-eastern England. III. Further predators of the mite. J Hortic Sci 28(2):98–113 Costa J, Torres L, Provance DW et al (2019) First report of predation by a stink bug (Supputius cincticeps Stål) on a walking-stick insect (Cladomorphus phyllinus Gray), with reflections on evolutionary mechanisms for camouflage. Acta Biol 48:5–15 Costa J, Chiquetto-Machado PI, Gil-Santana HR (2022) Predation strategies of Harpactor angulosus (Lepeletier & Serville, 1825) (Hemiptera: Reduviidae) on Cladomorphus phyllinus Gray, 1835 (Phasmatodea: Phasmatidae) in captivity. Rev Chil Entomol 48(3) Cruz I (1995) A lagarta do cartucho na cultura do milho. EMBRAPA-CNPMS Circular técnica 21: 1–45 da Rocha L, Redaelli LR (2004) Functional response of Cosmoclopius nigroannulatus (Hem.: Reduviidae) to different densities of Spartocera dentiventris (Hem.: Coreidae) nymphae. Braz. J Biol 64(2):309–316 da Silva NG, Truzi CC, Cardoso CP et al (2020) Temperature-dependent functional response of Euborellia annulipes (Dermaptera: Anisolabididae) preying on Plutella xylostella (Lepidoptera: Plutellidae) larvae. J Therm Biol 93:102686 da Silva HEG, de Oliveira R, De Brito CH (2022) Biological aspects and predatory capacity of Doru luteipes when fed with Spodoptera frugiperda. Rev Caatinga 35:490–497 Das SSM, Ambrose DP (2008) Redescription, biology and behaviour of a harpactorine assassin bug Vesbius sanguinosus Stål (Insecta, Hemiptera, Reduviidae). Pol J Entomol 77:11 Das S, Krishnan SS, Jebasingh V et al (2008) Redescription, postembryonic development and behaviour of a harpactorine assassin bug Sphedanolestes himalayensis Distant (Hemiptera: Reduviidae). Entomol Croat 12(1):37–54

506

14

Prey Record of Various Predators

Davranoglou LR (2011) Zelus renardii (Kolenati, 1856), a new world reduviid discovered in Europe (Hemiptera: Reduviidae: Harpactorinae). Entomol Monthly Mag 147(1766):157 De Castro AA, Corrêa AS, Legaspi JC et al (2013) Survival and behavior of the insecticide-exposed predators Podisus nigrispinus and Supputius cincticeps (Heteroptera: Pentatomidae). Chemosphere 93(6):1043–1050 De Clercq P (2001) Functional response of the predators Podisus maculiventris (Say) and Podisus nigrispinus (Dallas) (Het., Pentatomidae) to the beet armyworm, Spodoptera exigua (Hübner) (Lep., Noctuidae): effect of temperature. J Appl Entomol 125(3):131–134 De Clercq P, Merlevede F, Mestdagh I et al (1998) Predation on the tomato looper Chrysodeixis chalcites (Esper) (Lep., Noctuidae) by Podisus maculiventris (Say) and Podisus nigrispinus (Dallas) (Het., Pentatomidae). J Appl Entomol 122:93–98 De Clercq P, Mohaghegh J, Tirry L (2000) Effect of host plant on the functional response of the predator Podisus nigrispinus (Heteroptera: Pentatomidae). Biol Control 18(1):65–70 De Clercq P, Bonte M, Van Speybroeck K et al (2005) Development and reproduction of Adalia bipunctata (Coleoptera: Coccinellidae) on eggs of Ephestia kuehniella (Lepidoptera: Phycitidae) and pollen. Pest Manag Sci 61(11):1129–1132 De Kraker J, Van Huis A, Van Lenteren JC et al (2000) Identity and relative importance of egg predators of rice leaffolders (Lepidoptera: Pyralidae). Biol Control 19(3):215–222 De Oliveira RDF, De Almeida LC, De Souza DR et al (2012) Ant diversity (Hymenoptera: Formicidae) and predation by ants on the different stages of the sugarcane borer life cycle Diatraea saccharalis (Lepidoptera: Crambidae). Eur J Entomol 109(3):381 de Souza JM, Seleghim AR, da Silva NG et al (2022) Predation of Diatraea saccharalis eggs and neonates by the earwig Euborellia annulipes. Biol Control 172:104953 Degasperi TC, Milano P, Berti Filho E et al (2010) Dispersion and preying capacity of Chrysoperla externa Hagen (Neuroptera: Chrysopidae) on soybean and sugarcane plants infested with Anticarsia gemmatalis and Diatraea saccharalis (Lepidoptera: Noctuidae, Crambidae) eggs in laboratory and in greenhouse. Rev Agric (Piracicaba) 85(1):68–79 Dei EA, Nikitenko GN (1980) The predators and symbionts of the gypsy moth Porthetria dispar L. (Lepidoptera) in the lower cis Dneprov’ya. Vestn Zoo 3:91–92 Del Pino M, Polack A, Gamboa S et al (2009) Tupiocoris cucurbitaceus (Hemiptera: Miridae), population aspects in relation to the control of the whitefly of the greenhouses Trialeurodes vaporariorum (Hemiptera: Aleyrodidae) and the tomato crop under cover. In: Presented in XXXII Congreso Argentino de Horticultura.ASAHO. Salta, Argentina Desurmont G, Weston PA (2008) Predation by Podisus maculiventris (Say) (Hemiptera: Pentatomidae) on viburnum leaf beetle, Pyrrhalta viburni (Paykull) (Coleoptera: Chrysomelidae), under laboratory and field conditions. Environ Entomol 37:1241–1251 Devi PK, Yadav DN, Anand J (2002) Role of Nesidiocoris tenuis Reuter (Hemiptera: Miridae) in natural suppression of tomato fruit borer, Helicoverpa armigera (Hübner) (Lepidoptera: Noctuidae). Pest Manag Hort Ecosys 8:109–113 Dharavath S, Karabhantanal SS, Jagginavar SB (2022) Natural enemies associated with sorghum shoot bug Peregrinus maidis (Ashmead). Indian J Entomol:257–259 Dias TKR, Wilcken CF, Soliman EP et al (2012) Occurrence of Atopozelus opsimus preying on nymphsand adults of Glycaspis brimblecombei. Phytoparasitica 40(2):137–141 Dib H, Siegwart M, Delattre T et al (2020) Does combining Forficula auricularia L.(Dermaptera: Forficulidae) with Harmonia axyridis Pallas (Coleoptera: Coccinellidae) enhance predation of rosy apple aphid, Dysaphis plantaginea Passerini (Hemiptera: Aphididae). Biol Control 151: 104394 Donahoe MC, Pitre HN (1977) Reduviolus roseipennis behavior and effectiveness in reducing numbers of Heliothis zea on cotton. Environ Entomol 6:872–876 Duarte L, Pacheco R, Quiñones M et al (2014) Nesidiocoris tenuis Reuter (Hemiptera: Miridae) and Cycloneda sanguinea limbifer (Casey) (Coleoptera: Coccinelidae): Behaviour and predatory activity on Myzus persicae Sulzer (Hemiptera: Aphididae). Revista De Protección Vegetal 29: 99–105

References

507

Duran Prieto J, Trotta V, Fanti P et al (2016) Predation by Macrolophus pygmaeus (Hemiptera: Miridae) on Acyrthosiphon pisum (Hemiptera: Aphididae): Influence of prey age/size and predator’s intraspecifi c interactions. Eur J Entomol 113:37–43 Echegaray EA, Cloyd RA, Nechols JR (2015) Rove beetle (Coleoptera: Staphylinidae) predation on Bradysia sp. nr. coprophila (Diptera: Sciaridae). J Entomol Sci 50(3):225–237 Eisawi K, Subedi IP, Yode CD et al (2022) Ants (Hymenoptera: Formicidae) increase predation of Belenois solilucis (Lepidoptera: Pieridae) eggs in organic agriculture production systems: a multiple-site field study at Rashad, Sudan. Sociobiology 69(2):e7746 El- Sebaey IIA, El-Bishry MH (2001) Biological aspects and predation of Coranus africana El-Sebaey on the Laboratory preys Anagasta kuehniella and Corcyra cephalonica. Egypt J Bio Pest Cont 11(2):57–62 El-Sebaey IA (2001) Biology and predation rates of the, assassin bug, Coranus africana El-Sebaey (Heteroptera: Reduviidae) on the cotton pests, Spodoptera littoralis (Bosid.) and Agrotis ypsilon Rott. Bull Fac Agric Cairo Univ 52:655–668 El-Sebaey I, Abd El-Wahab HA (2011) Effect of different stages of reduviid predator Corans africana El-Sebaey (Hemiptera: Heteroptera) on the population of bemisia tabaci genn in tomato field at qalubia and bani-swif governorates. Egypt J Agric Res 89(2):413–429 El-Shafie HAF, Abdel-Banat BMA, Al-Hajho MR (2017) Arthropod pests of date palm and their management. CABI Rev:1–18 El-Shazly MM, Iman IA (1997) Development and predation rates of Coranus aegpus (F) AND Vachiria natolica STAL (Hemiptera: Reduviidae: Harpactorinae). Egypt J Biol Pest Contr 7(1): 47–51 Eskafi FM, Kolbe MM (1990) Predation on larval and pupal Ceratitis capitata (Diptera: Tephritidae) by the ant Solenopsis geminata (Hymenoptera: Formicidae) and other predators in Guatemala. Environ Entomol 19(1):148–153 Eubanks MD, Denno FD (2000) Health food versus fast food: The effects of prey quality and mobility on prey selection by a generalist predator and indirect interactions among prey species. Ecol Entomol 25:140–146 Evans DE (1962) The food requirements of phonoctonus nigrofasciatus stål.(hemiptera, reduviidae). Entomol Exp Appl 5(1):33–39 Fadare TA (1978) Efficiency of Phonoctonus spp. [Hemiptera: Reduviidae] as regulators of populations of Dysdercus spp. [Hemiptera: Pyrrhocoridae]. Niger J Entomol 1:45–48 Fair CG, Braman SK (2019) Predation of Anasa tristis (Hemiptera: Coreidae) by Geocoris punctipes (Hemiptera: Geocoridae). J Entomol Sci 54(1):87–92 Falahudin I (2013) Peranan semut rangrang (Oecophylla smaradigna) dalam pengendalian biologis pada perkebunan kelapa sawit. In: Prosiding Konferensi AICIS, vol 12, pp 2604–2618 Farag AA, Ha AER, Refaei AE et al (2022) Effect of cannibalism phenomenon on some biological aspects and predation efficiency on Tetranychus urticae Koch of the ladybird beetle, Coccinella undecimpunctata Linnaeus. J Entomol Zool Stud 10(3):05–01 Farhadi R, Allahyari H, Juliano SA (2010) Functional response of larval and adult stages of Hippodamia variegata (Coleoptera: Coccinellidae) to different densities of Aphis fabae (Hemiptera: Aphididae). Environ Entomol 39:1586–1592 Faria LD, Reigada C, Trinca LA et al (2007) Foraging behaviour by an intraguild predator blowfly, Chrysomya albiceps (Diptera: Calliphoridae). J Ethol 25(3):287–294 Fenoglio MS, Trumper EV (2014) Influence of weather conditions and density of Doru luteipes (Dermaptera: Forficulidae) on Diatraea saccharalis (Lepidoptera: Crambidae) egg mortality. Environ Entomol 36(5):1159–1165 Fernandes FL (2012) Ecology of Vespidae (Hymenoptera) predators in Coffea arabica plantations. Sociobiology 59(4):1269–1280 Fernandes WD, Sant’Ana MV, Raizer J et al (2012) Depredación de larvas de mosca de la fruta Anastrepha (Diptera: Tephritidae) por hormigas en el bosque. Psique 10:83–89 Fernández M, Amor F, Bengochea P et al (2012) Effects of the insecticides methoxyfenozide and abamectin to adults of the whitefly natural enemies Eretmocerus mundus (Mercet)

508

14

Prey Record of Various Predators

(Hymenoptera: Aphelinidae), Orius laevigatus (Fieber) (Hemiptera: Anthocoridae) and Nesidiocoris tenuis Reuter (Hemiptera: Miridae) under laboratory conditions. In: Proceedings of the IOBC/WPRS Working Group Pesticides and Beneficial Organisms, Marbella, Spain, pp 8–11 Filho CCB, Santos GMDM, Resende JJ et al (2009) Foraging behavior of the swarm-founding wasp, Polybia (Trichothorax) sericea (Hymenoptera, Vespidae): prey capture and load capacity. Sociobiology 53(1) Firake DM, Behere GT (2020) Natural mortality of invasive fall armyworm, Spodoptera frugiperda (J. E. Smith) (Lepidoptera: Noctuidae) in maize agroecosystems of northeast India. Biol Control 148:104303 Flinn PW, Hower AA, Taylor RAJ (1985) Preference of Reduviolus americoferus (Hemiptera: Nabidae) for potato leafhopper nymphs and pea aphids. Can Entomol 117:1503–1508 Foglar H, Malausa JC, Wajnberg E (1990) The functional response and preference ofmacrolophus caliginosus [heteroptera: Miridae] for two of its prey: Myzus persicae andtetranychus urticae. Entomophaga 35(3):465–474 Fukao S, Oida H, Tsukamoto T (2012) Multiple control effects by the polyphagous predator, Geocoris varius (Heteroptera: Geocoridae) in presence of multiple pest species. Ann Rep Kanto Pl Prot Soc 59:103–109 Funderburk J, Stavisky J, Olson S (2000) Predation of Frankliniella occidentalis (Thysanoptera: Thripidae) in field peppers by Orius insidiosus (Hemiptera: Anthocoridae). Environ Entomol 29:376–382 Gabarra R, Castane C, Albajes R (1995) The mirid bug Dicyphus tamaninii as a greenhouse whitefly and western flower thrips predator on cucumber. Biocontr Sci Technol 5(4):475–488 Gaikwad MB, Verma SC, Sharma PL et al (2022) Functional response and predatory potential of coccinellid predator, Oenopia kirbyi Mulsant (Coccinellidae: Coleoptera) on rose aphid, Macrosiphum rosae L. (Aphididae: Hemiptera). Int J Trop Insect Sci 42:2233 García M, Ortego F, Castañera P et al (2012) Assessment of prey-mediated effects of the coleopteran-specific toxin Cry3Bb1 on the generalist predator Atheta coriaria (Coleoptera: Staphylinidae). Bull Entomol Res 102(3):293–302 Garcia A, Figueiredo E, Valente C et al (2013) First record of Thaumastocoris peregrinus in Portugal and of the neotropical predator Hemerobius bolivari in Europe. Bull Insectol 66(2): 251–256 George PE (2000) Reproductive performance of a harpactorine reduviid Rhynocoris fuscipes (Fabricius)(Heteroptera: Reduviidae) on three lepidopteran insect pests. Int J Trop Insect Sci 20(4):269–273 George PJ, Ravichandran B (2021) Prey and stage preference of Acanthaspis pedestris Stål (Hemiptera: Reduviidae) on pests of cotton (Gossypium hirsutum). Crop Res 56(5):217–221 George PJE, Srenivasagan R (1998) Predatory efficiency of Rhynocoris marginatus (Fabricius) (Heteroptera: Reduviidae) on Helicoverpa armigera (Hubner) and Spodoptera litura (Fabricius). J Biol Contr 12:25–29 Ghabeish I, Saleh A, Dababat A (2010) Prey preference, interaction with selected natural enemies, and alternative nutritional sources of the mirid bug Dicyphus tamaninii Wagner. Turk J Agric For 34:415–420 Ghahari H, Collingwood CA, Tabari M et al (2009) Faunistic notes on Formicidae (Insecta: Hymenoptera) of rice fields and surrounding grasslands in Northern Iran. Mun Ent Zool 4(1): 184–189 Giraldo-Jaramillo M, Forero D, Benavides P (2021) The Wheel Bug Arilus gallus (Hemiptera: Reduviidae): life history and description of immature stages. Proc Entomol Soc Wash 123(3): 551–563 Gitonga LM, Overhalt WA, Lohr B et al (2002) Functional response of Orius albidipennis (Hemiptera: Anthocoridae) to Megalurothrips sjostedti (Thysanoptera: Thripidae). Biol Control 24:1–6

References

509

Godfrey KE, Whitcomb WH, Stimac JL (1989) Arthropod predators of velvetbean caterpillar, Anticarsia gemmatalis Hübner (Lepidoptera: Noctuidae), eggs and larvae. Environ Entomol 18(1):118–123 Grundy PR (2007) Utilizing the assassin bug, Pristhesancus plagipennis (Hemiptera: Reduviidae), as a biological control agent within an integrated pest management programme for Helicoverpa spp. (Lepidoptera: Noctuidae) and Creontiades spp. (Hemiptera: Miridae) in cotton. Bull Entomol Res 97(3):281–290 Grundy P, Maelzer D (2000) Predation by the assassin bug Pristhesancus plagipennis (Walker) (Hemiptera: Reduviidae) of Helicoverpa armigera (Hübner) (Lepidoptera: Noctuidae) and Nezara viridula (L.) (Hemiptera: Pentatomidae) in the laboratory. Aust J Entomol 39(4): 280–282 Guillemain M, Loreau M, Daufresne T (1997) Relationships beetween the regional distribution of carabid beetles (Coleoptera, Carabidae) and the abundance of their potential prey. Acta Oecologica 18(4):465–483 Haase-Statz S (1997) Evaluation of Paederus Littorarius (Coleoptera: Staphylinidae) as an egg predator of Chrysoteuchia Topiaria (Lepidoptera): Pyralidae in Wisconsin Cranberry Bogs. Gt Lakes Entomol 30(2) https://scholar.valpo.edu/tgle/vol30/iss2/5 Hagley EAC, Holliday NJ, Barber DR (1982) Laboratory studies of the food preferences of some orchard carabids (Coleoptera: Carabidae). Can Entomol 114(5):431–437 Halder J, Rai AB (2016) Suitability of different prey aphids on the growth, development and reproduction of Chrysoperla zastrowi sillemi (Esben-Petersen) (Chrysopidae: Neuroptera). Proc Zool Soc 69(1):89–95 Hamasaki K, Matsui M (2006) Development and reproduction of an aphidophagous coccinellid, Propylea japonica (Thunberg) (Coleoptera: Coccinellidae), reared on an alternative diet, Ephestia kuehniella Zeller (Lepidoptera: Pyralidae) eggs. Appl Entomol Zool 41:233–237 Hamdan AJS (2006) Effect of host-plant species on survival, adult longevity and fertility of predatory bug, Macrolophus caliginosus Wagner [Hemiptera: Miridae]. HURJ 2(2):1–15 Hameed SF, Kashyap NP, Vaidya DN (1975) Host records of insect pests of crops. FAO Plant Prot Bull 23(6):191–192 Happe AK, Roquer-Beni L, Bosch J et al (2018) Earwigs and woolly apple aphids in integrated and organic apple orchards: responses of a generalist predator and a pest prey to local and landscape factors. Agric Ecosyst Environ 268:44–51 Harris HM (1928) A monographic study of the hemipterous family Nabidae as it occurs in North America. Entomol Am 9:1–98 Hassan HM (2021) Role of the rove beetle, Paederus memnonius (Erichson)(Coleoptera: Staphylinidae) in controlling certain sugar beet pests for the first time in Egypt. J Plant Prot Pathol 12(2):117–120 Hassanpour M, Jafar M, Shahzad I et al (2011) Functional response of Chrysoperla carnea (Neuroptera: Chrysopidae) to Helicoverpa armigera (Lepidoptera: Noctuidae): Effect of prey and predator stages. Insect Sci 18:217–224 Heinrichs EA, Thompson HE (1968) Parasites, predators, and other arthropods associated with Choristotseura houstonana (lepidoptera: tortricidae) on juniperus species in kansas1. Can Entomol 100(1):33–39 Helgadóttir F, Toft S, Sigsgaard L (2017) Negative effects of low developmental temperatures on aphid predation by Orius majusculus (Heteroptera: Anthocoridae). Biol Control 114:59–64 Hendrichs MA, Hendrichs J (1998) Perfumed to be killed: interception of Mediterranean fruit fly (Diptera: Tephritidae) sexual signaling by predatory foraging wasps (Hymenoptera: Vespidae). Ann Entomol Soc Am 91(2):228–234 Hennessey MK (1998) Predation on Caribbean Fruit Fly Prepupae and Pupae in Guava and Carambola Grove. USDA, ARS, Tektran, Miami, FL Hoballah ME, Degen T, Bergvinson D et al (2004) Occurrence and direct control potential of parasitoids and predators of the fall armyworm (Lepidoptera: Noctuidae) on maize in the subtropical lowlands of Mexico. Agric For Entomol 6(1):83–88

510

14

Prey Record of Various Predators

Hong BM, Hung HQ (2010) Effect of temperature and diet on the life cycle and predatory capacity of Episyrphus balteatus (De Geer) (Syrphidae: Diptera) cultured on Aphis gossypii (Glover). I. Int Soc Southeast Asian Agric Sci 16:98–103 Horne PA, Edward CL (1995) Phenology and food preferences of Labidura truncata Kirby (Dermaptera: Labiduridae) in western Victoria. Aust J Entomol 34(2):101–104 Hosseini A, Hosseini M, Michaud JP et al (2019) Life history responses of Hippodamia variegata (Coleoptera: Coccinellidae) to changes in the nutritional content of its prey, Aphis gossypii (Hemiptera: Aphididae), mediated by nitrogen fertilization. Biol Control 130:27–33 Hu GY, Frank JH (1997) Predation on the horn fly (Diptera: Muscidae) by five species of Philonthus (Coleoptera: Staphylinidae). Environ Entomol 26(6):1240–1246 Imamura T, Murata M, Miyanoshita A (2008) Biological aspects and predatory abilities of hemipterans attacking stored-product insects. Jpn Agric Res Q 42(1):1–6 Inayat TP (2011) Predation rate in selected coccinellid (Coleoptera) predators on some major aphidid and cicadellid (hemipteran) pests. Int J Agric Biol 13(3):427–430 Ingegno BL, Ferracini C, Gallinotti D et al (2013) Evaluation of the effectiveness of Dicyphus errans (Wolff) as predator of Tuta absoluta (Meyrick). Biol Control 67:246–252 Ingegno BL, Bodino N, Leman A et al (2017) Predatory efficacy of Dicyphus errans on different prey. Acta Hort 1164:425–430 Iqbal S (2020) Insect, pest and disease management in rice. Austin Publication, Irving, TX, p 85 James DG (1994) Prey consumption by ‘Pristhesancus plagipennis’ Walker (Hemiptera: Reduviidae) during development. Aust Entomol 21(2):43–47 Jana N, Michal S, Jana O et al (2021) Forficula auricularia (Dermaptera) in orchards: Monitoring seasonal activity, the effect of pesticides, and the perception of European fruit growers on its role as a predator or pest. Pest Manag Sci 77(4):1694–1704 Jiang ZR, Kajimura H (2020) Earwig preying on ambrosia beetle: evaluating predatory process and prey preference. J Appl Entomol 144(8):743–750 Joern A, Rudd NT (1982) Impact of predation by the robber fly Proctacanthus milbertii (Diptera: Asilidae) on grasshopper (Orthoptera: Acrididae) populations. Oecologia 55(1):42–46 Jones R, Gislstrap F, Andrews K (1988) Biology and life table for the predaceous earwig, Doru taeniatum (Derm.: Forficulidae). Entomophaga 33(1):43–54 Joseph SV (2006) The potential role of heteropteran predators-Geocoris punctipes (Say), G. uliginosus (say) (geocoridae) and Orius insidiosus (Say) (anthocoridae) in warm-season turfgrass. Doctoral dissertation, University of Georgia Kabissa JC, Kayumbo HY, Yarro JG (1996) Seasonal abundance of chrysopids (Neuroptera: Chrysopidae) preying on Helicoverpa armigera (Hubner) (Lepidoptera: Noctuidae) and Aphis gossypii (Glover) (Homoptera: Aphididae) on cotton in eastern Tanzania. J Crop Prot 15(1):5–8 Kaiser ME, Noma T, Brewer MJ et al (2007) Hymenopteran parasitoids and dipteran predators found using soybean aphid after its midwestern United States invasion. Ann Entomol Soc Am 100(2):196–205 Kalshoven LGE (1981) The pest of crops in Indonesia; revised and translated by P.A. Van der Laan. PT. Ichtiar Baru van Hoeve. Jakarta. 701 p Kapadia MN, Puri SN (1991) Biology and comparative predation efficacy of three heteropteran species recorded as predators of Bemisia tabaci in Maharashtra. Entomophaga 36(4):555–559 Karindah S (2011) Predation of five generalist predators on brown planthopper (Nilaparvata lugens Stål). J Entomol Indon 8(2):55–62 Katayama N, Suzuki N (2003) Bodyguard effects for aphids of Aphis craccivora Koch (Homoptera: Aphididae) as related to the activity of two ant species, Tetramorium caespitum Linnaeus (Hymenoptera: Formicidae) and Lasius niger L. (Hymenoptera: Formicidae). Appl Entomol Zool 38(3):427–433 Kayahan B, Kayahan A, Karaca İ (2021) Determination of the preferred two different coccinellid species on different aphid species feeding on broad bean plants. Int J Agric Environ Food Sci 5(2):166–172

References

511

Khattawi S, Kayahan A, Karaca I (2022) Functional and numerical response of Hippodamia variegata (Goeze) (Coleoptera: Coccinellidae) on Macrosiphum rosae (L.) (Hemiptera: Aphididae). Int J Agric Environ Food Sci 6(2):311–318 Kiran R, Venkatesha MG (2022) Record of Coranus siva Kirkaldy (Hemiptera: Reduviidae) on coffee berry borer, Hypothenemus hampei Ferrari (Coleoptera: Curculionidae) in India. Entomon 47(2):179–182 Klostermeyer EC (1942) The life history and habits of the ring-legged earwig, Euborellia annulipes Lucas. J Kans Entomol Soc 15:13–18 Knowlton GF (1932) The beet leafhopper in Northern Utah. Utah Agr Expt Sta Bul 234 Knowlton GF (1935) Beet leafhopper insect predator studies. Proc Utah Acad Sci Arts Lett 12:255– 260 Knowlton GF (1944) Sinea diadema feeding notes. Brooklyn Ent Soc Bul 39:71 Koffi D, Kyerematen R, Eziah VY et al (2020) Natural enemies of the Fall Armyworm, Spodoptera frugiperda (J.E. Smith) (Lepidoptera: Noctuidae) in Ghana. Fla Entomol 103:85–90 Kohno K (2003) Development and reproduction of Antilochus coqueberti (Heteroptera, Pyrrhocoridae), the specific predator of Dysdercus spp. (Heteroptera, Pyrrhocoridae). Appl Entomol Zool 38:103–108 Kohno K, Takahashi K, Sakakibara M (2002) New prey–predator association in aposematic pyrrhocorid bugs, Antilochus coqueberti as a specialist predator on Dysdercus species. Entomol Sci 5:391–397 Kohno K, Bui Thi N, Fujiwara M (2004) Predation of Dysdercus cingulatus (Heteroptera, Pyrrhocoridae) by the specialist predator Antilochus coqueberti (Heteroptera, Pyrrhocoridae). Appl Entomol Zool 39:661–667 Komatsu T (2014) Larvae of the Japanese termitophilous predator Isoscelipteron okamotonis (Neuroptera, Berothidae) use their mandibles and silk web to prey on termites. Insectes Soc 61(2):203–205 Kontodimas DC, Milonas PG, Stathas GJ et al (2007) Life table parameters of the pseudococcid predators Nephus includens and Nephus bisignatus (Coleoptera: Coccinelidae). Eur J Entomol 104(3):407 Krishnamoorthy A, Mani M (1989) Records of green lacewings preying on mealybugs in India. Curr Sci 58(3):155–156 Kromp B (1999) Carabid beetles in sustainable agriculture: a review on pest control efficacy, cultivation impacts and enhancement. Elsevier Science BV, Amsterdam Kumar AG, Ambrose DP (2014) Olfactory response of an assassin bug, Rhynocoris longifrons (Insecta: Hemiptera: Reduviidae) to the hexane extracts of different agricultural insect pests. IRJIS 1(1):12–19 Kumar NS, Ananthakrishnan TN (1985) Geocoris ochropterus Fabr.as a predator of some thrips. Proc Natl Acad Sci India Sect B Biol Sci 51:185–193 Kurczewski FE (1968) Vespula maculifrons (Hymenoptera: Vespidae) preying on the European earwig Forficula auricularia. J N Y Entomol Soc:84–86 Kwadjo KE, Doumbia M, Ishikawa T et al (2008) Morphometrical changes and description of eggs of Rhynocoris albopilosus Signoret (Heteroptera: Reduviidae) during their development. Faunistic Entomol 61:151–155 Kyneb A, Toft S (2004) Quality of two aphid species (Rhopalosiphum padi and Sitobion avenae) as food for the generalist predator Tachyporus hypnorum (Col., Staphylinidae). J Appl Entomol 128(9,10):658–663 LaFlair A (2022) Survey of Native Predatory Hemipterans and Assessment of Nabis americoferus Carayon (Hemiptera: Nabidae) as a Biological Control Agent of Lepidopteran Tomato Pests. Electronic Theses and Dissertations. p 8787 Lakkundi NH (1989) Assessment of reduviids for their predation and possibilities of their utilization in biological control. PhD Thesis, Indian Agricultural Research Institute, New Delhi Lamine K, Lambin M, Alauzet C (2005) Effect of starvation on the searching path of the predatory bug Deraeocoris lutescens. BioControl 50(5):717–727

512

14

Prey Record of Various Predators

Landwehr VR (1977) Ischyropalpus nitidulus (Coleoptera: Anthicidae), a predator of mites associated with Monterey pine. Ann Entomol Soc Am 70(1):81–83 Lang A, Gsödl S (2001) Prey vulnerability and active predator choice as determinants of prey selection: a carabid beetle and its aphid prey. J Appl Entomol 125(1,2):53–61 Lanza Reis L, Jacob Oliveira L, Cruz I (1988) Biologia e potencial de Doru luteipes no controle de Spodoptera frugiperda. Pesqui Agropecu Bras 23(4):333–342 Lastres M (1990) The role of two predators, Doru taeniatum Dorhn and Solenopsis geminata Fabricius, as control agents of Spodoptera frugiperda J.E. Smith, in Honduras. MS Thesis, Texas A and M University, College Station, TX Lee G, Lee S, Choi M et al (2000) Prey consumption and suppression of vegetable aphids by Chrysopa pallens (Neuroptera: Chrysopidae) as a predator. Korean J Appl Entomol 39(4): 251–258 Legaspi JC, Carruthers RI, Nordlund DA (1994) Life-history of Chrysoperla rufilabris (Neuroptera: Chrysopidae) provided sweetpotato whitefly Bemisia tabaci (Homoptera: Aleyrodidae) and other food. Biol Control 4(2):178–184 Lemos WP, Medeiros RS, Ramalho FS (1998) Influência da temperatura no desenvolvimento de Euborellia annulipes (Lucas) (Dermaptera:Anisolabididae), predador do bicudo-do-algodoeiro. An Soc Entomol Brasil 27:67–76 Lemos WP, Medeiros RS, Ramalho FS (1999) Exigências térmicas de Euborellia annulipes (Lucas) (Dermaptera, Anisolabididae) e sua relação com a presa: bicudo-do-algodoeiro. Revta Bras Entomol 43:61–68 Lemos WP, Ramalho FDS, Zanuncio JC (2003a) Age-dependent fecundity and life-fertility tables for Euborellia annulipes (Lucas) (Dermaptera: Anisolabididae) a cotton boll weevil predator in laboratory studies with an artificial diet. Environ Entomol 32(3):592–601 Lemos WP, Ramalho FS, Serrão JE et al (2003b) Effects of diet on development of Podisus nigrispinus (Dallas) (Het., Pentatomidae), a predator of the cotton leafworm. J Appl Entomol 127(7):389–395 Li Y, Cloyd RA, Bello NM (2020) Predation efficacy of rove beetle (Coleoptera: Staphylinidae) adults in response to western flower thrips (Thysanoptera: Thripidae) pupal stage, predator–prey ratio, and searchable area. Entomol Sci 55(3):350–365 Lillo I, Perez Bañón C, Rojo S (2021) Life cycle, population parameters, and predation rate of the hover fly Eupeodes corollae fed on the aphid Myzus persicae. Entomol Exp Appl 169(11): 1027–1038 Lingren PD, Ridgway RL, Jones SL (1968) Consumption by several common arthropod predators of eggs and larvae of two Heliothis species that attack cotton. Ann Entomol Soc Am 61(3): 613–618 Litsinger JA, Alviola AL, Dela Cruz CG et al (2006) Rice white stemborer Scirpophaga innotata (Walker) in southern Mindanao, Philippines. II. Synchrony of planting and natural enemies. Int J Pest Manag 52(1):23–37 Lois M, Carrizo PI, López SN et al (2022) Evaluación de Artemia sp. (Branchiopoda, Artemiidae) como presa alternativa para la cría en laboratorio de Tupiocoris cucurbitaceus (Hemiptera: Miridae), predador de plagas hortícolas. Agriscientia 39:83–92 Loko YL, Dansi A, Tamo M et al (2013) Storage insects on yam chips and their traditional management in Northern Benin. Sci World J 2013:484536 Loko YLE, Toffa J, Gavoedo DM et al (2022) Effect of population density on oviposition, development, and survival of Alloeocranum biannulipes (Hemiptera: Reduviidae) preying on Dinoderus porcellus (Coleoptera: Bostrichidae). J Basic Appl Zool 83(1):1–8 Louis D (1974) Biology of Reduviidae of cocoa farms in Ghana. Am Midl Nat:68–89 Luginbill P (1928) The Fall Armyworm. USDA Technical Bulletin 34 Madge DS, Buxton JH (1976) The evaluation of the European earwig (Forficula auricularia) as a predator of the damson-hop aphid (Phorodon humuli). II. Choice of prey. Entomol Exp Appl 19: 221–226

References

513

Maneerat T, Sakarind P (2018) Potential mass production of assassin bug, Sycanus collaris F. (Hemiptera: Reduviidae) using various stages of Tenebrio molitor L. (Coleoptera: Tenebrionidae) as food. Khon Kaen Agr J 46:195–200 Manley GV (1977) Paederus fuscipes [Col.: Staphylinidae]: a predator of rice fields in west Malaysia. Entomophaga 22:47–59. https://doi.org/10.1007/BF02372990 Manyuli T, Kyamanywa S, Luther GC (2008) Capability of Forficula auricularia linnaeus (Dermaptera: Forficulidae) to prey on Aphis craccivora koch (Homoptera: Aphididae) in eastern and Central Africa. Agron Afr 20(1):49–58 Marchart H (1968) Radiotracer study of the predators on Distantiella theobroma (Distant) (Hemiptera: Miridae). In: Isotopes and radiation in entomology, pp 3–15 Mariani R, Vera L, Virla E (1996) Aportes al conocimiento de Doru lineare (Eschs, 1822) (Dermaptera, Forficulidae), un insecto de importancia agronómica en el Noroeste Argentino. CIRPON Revista de investigación 10(1-4):13–18 Marquereau L, Cottineau JS, Fontaine O et al (2022) Life history parameters and predation capacities of Nesidiocoris volucer: a new biological control agent for tomato crop. Bull Entomol Res 112:584–592 Marsaro Júnior AL, Storck-Tonon D, Halinski R et al (2022) New records of predation of Harpactorinae (Hemiptera: Reduviidae) over Euglossini and Xylocopini bees (Hymenoptera: Apidae) in Brazil. Biol Behav 48(1) Marshall DC, Hill KBR (2009) Versatile aggressive mimicry of cicadas by an Australian Predatory Katydid. PLoS One 4(1):e4185 Martin WR Jr, Brown TM (1984) The action of acephate in Pseudoplusia includens (Lepidoptera: Noctuidae) and Pristhesancus papuensis (Hemiptera: Reduviidae). Entomol Exp Appl 35(1):3–9 Martine DG, Pienkowski RL (1982) Laboratory studies on insect predators of potato leafhopper eggs, nymphs, and adults. Environ Entomol 11:361–362 Martínez MA, Duarte L, Baños HL et al (2014) Predatory mirids (Hemiptera: Heteroptera: Miridae) in tomato and tobacco in Cuba. Rev Protección Veg 29(3):204–207 Martínez LD, de los Angeles Martínez Rivero M, Bueno VHP et al (2022) Predation behaviour and prey preference of two neotropical mirids against two key lepidopteran pests in tomato. Int J Trop Insect Sci 42:815–825 Maruthadurai R, Ramesh R, Veershetty C (2022) Prevalence and predation potential of rove beetle Paederus fuscipes Curtis (Coleoptera: Staphylinidae) on Invasive Fall Armyworm Spodoptera frugiperda in fodder maize. Natl Acad Sci Lett 45(2):119–121 McKemey AR, Symondson WOC, Glen DM et al (2001) Effects of slug size on predation by Pterostichus melanarius (Coleoptera: Carabidae). Biocontr Sci Technol 11(1):81–91 McPheron LJ, Mills NJ (2007) Influence of visual and olfactory cues on the foraging behavior of the paper wasp Mischocyttarus flavitarsis (Hymenoptera: Vespidae). Entomol Gen 30:105–118 Meier MS, Hilbeck A (2001) Influence of transgenic Bacillus thuringiensis corn-fed prey on prey preference of immature Chrysoperla carnea (Neuroptera: Chrysopidae). Basic Appl Ecol 2(1): 35–44 Meissle M, Vojtech E, Poppy GM (2005) Effects of Bt maize-fed prey on the generalist predator Poecilus cupreus L. (Coleoptera: Carabidae). Transgenic Res 14(2):123–132 Miller KV, Williams RN (1983) Biology and host preference of Atheta coriaria (Coleoptera: Staphylinidae), an egg predator of Nitidulidae and Muscidae. Ann Entomol Soc Am 76(2): 158–161 Miller GL, Oswald JD, Miller DR (2004) Lacewings and scale insects: a review of predator/prey associations between the Neuropterida and Coccoidea (Insecta: Neuroptera, Raphidioptera, Hemiptera). Ann Entomol Soc Am 97(6):1103–1125 Mohaghegh J, Amir-Maafi M (2007) Reproduction of the predatory stinkbug Andrallus spinidens (F.) (Heteroptera: Pentatomoidae) on live and frozen prey. Appl Entomol Zool 42:15–20 Mohaghegh J, De Clercq P, Tirry L (2001) Functional response of the predators Podius maculiventris (Say) and Podius nigrispinus (Dallas) (Heteroptera: Pentatomidae) to the beet

514

14

Prey Record of Various Predators

armyworm, Spodoptera exigua (Hubner) (Lepidoptera: Noctuidae): effect of temperature. J Appl Entomol 125:131–134 Mollá O, Biondi A, Alonso-Valiente M et al (2014) A comparative life history study of two mirid bugs preying on Tuta absoluta and Ephestia kuehniella eggs on tomato crops: implications for biological control. Biocontrol 59:175–183 Monteith LG (1972) Status of predators of the adult of apple maggot, Rhagolethis pomonella (Diptera: Tephritidae) in Ontario. Can Ent 104:257–262 Monzó C, Sabater-Muñoz B, Urbaneja A et al (2011) The ground beetle Pseudophonus rufipes revealed as predator of Ceratitis capitata in citrus orchards. Biol Control 56(1):17–21 Moreira LB, Lima LLR, de Sá FE et al (2022) Response of Doru luteipes (Dermaptera: Forficulidae) to insecticides used in maize crop as a function of its life stage and exposure route. Environ Sci Pollut Res Int 30:15010 Morrill AW (1910) Plant-bugs injurious to cotton bolls. USDA Bur Entomol Bull 86:95–100 Mouly R, Shivananda TN, Verghese A (2018) Effect of temperature on minor invertebrate predator reduviid Isyndus heros (Fab.) (Hemiptera: Reduviidae). Curr Sci 115(5):983–986 Mouratidis A, de Lima AP, Dicke M et al (2022) Predator-prey interactions and life history of Orius laevigatus and O. majusculus feeding on flower and leaf-inhabiting thrips. Biol Control 172: 104954 Muhammad Luqman HA, Dzulhelmi MN, Idris AB (2018) The potential natural predators of Elaeidobius kamerunicus Faust, 1878 (Coleoptera: Curculionidae) in Malaysia. Serangga 22:240–252 Muhammad S, Muhammad S, Syed MZ (2020) Survival and fecundity of Antilochus coquebertii (Hemiptera: Pyrrhocoridae) on different stages of Dysdercus koenigii (Hemiptera: Pyrrhocoridae). J Kansas Entomol Soc 92(3):526–535 Mundy CA, Allen Williams LJ, Underwood N et al (2000) Prey selection and foraging behaviour by Pterostichus cupreus L. (Col., Carabidae) under laboratory conditions. J Appl Entomol 124 (9,10):349–358 Muthupandi M, Evangelin G, Horne B et al (2014) Biology of the harpactorine assassin bug, Panthous bimaculatus Distant (Hemiptera: Reduviidae) on three different diets. Halteres 5:11– 16 Myers JG (1935) Notes on cocoa-beetle and cocoa-thrips. Trop Agric 12(1):71 Naik BG, Sundararaju D (1982) In CPCRI Annual Report. Central Plantation Crops Research Institute, Kasaragod, India Nannini M, Atzori F, Murgia G et al (2012) Use of predatory mirids for control of the tomato borer Tuta absoluta (Meyrick) in Sardinian greenhouse tomatoes. EPPO Bull 42(2):255–259 Naranjo-Guevara N, Peñaflor MFG, Cabezas-Guerrero MF et al (2017a) Nocturnal herbivoreinduced plant volatiles attract the generalist predatory earwig Doru luteipes Scudder. Sci Nat 104:77 Naranjo-Guevara N, Peñaflor MFG, Cabezas-Guerrero MF et al (2017b) Nocturnal herbivoreinduced plant volatiles attract the generalist predatory earwig Doru luteipes Scudder. Sci Nat 104(9):1–11 Nash MA, Thomson LJ, Horne PA et al (2008) Notonomus gravis (Chaudoir) (Coleoptera: Carabidae) predation of Deroceras reticulatum Müller (Gastropoda: Agriolimacidae), an example of fortuitous biological control. Biol Control 47:328–334 Ngoc Bao Chau N, Quynh Phuong Anh N, Le Nhu TT et al (2021) Ant and silkworm pupae as convenient diets for the development and reproduction of big-eyed bug Geocoris ochropterus (Hemiptera: Geocoridae). J Asia Pac Entomol 24(1) Nielson MW, Henderson JA (1959) Biology of Collops vittatus (Say) in Arizona, and feeding habits of seven predators of the spotted alfalfa aphid. J Econ Entomol 52(1):159–162 Nishi A, Imamura T, Miyanoshita A, Morimoto S, Takahashi K, Visarathanonth P, Sato K (2004) Predatory abilities of Amphibolus venator (Klug) (Hemiptera: Reduviidae), a predator of storedproduct insect pests. Appl Entomol Zool 39(2):321–326

References

515

Nishida T (1955) Natural enemies of the melon fly, Dacus curcurbitae Coq. in Hawaii. Ann Entomol Soc Am 48:171–178 Nitin KS, Bhat PS, Raviprasad TN et al (2017) Biology, behaviour and predatory efficiency of Sycanus galbanus Distant. Hemiptera: Reduviidae: Harpactorinae recorded in cashew plantations. J Entomol Zool Stud 524:524–530 Niziira ZM (1970) The biology and behavior of Rhinocoris albopunctatus [Hemiptera: Reduviidae]. Ann Entomol Soc Am 63:1224–1227 Noël G, Caetano J, Blanchard S et al (2022) High temperatures adversely affect the hoverfly Episyrphus balteatus (Diptera: Syrphidae) fitness and aphid prey consumption. Turk J Zool 46(2):186–193 Nunes GDS, Dantas TAV, de Souza MDS et al (2019a) Life stage and population density of Plutella xylostella affect the predation behavior of Euborellia annulipes. Entomol Exp Appl 167(6): 544–552 Nunes GDS, Truzi CC, do Nascimento J et al (2019b) Beauveria bassiana (Ascomycota: Hypocreales)–treated Diamondback moth (Lepidoptera: Plutellidae) larvae mediate the preference and functional response of Euborellia annulipes (Dermaptera: Anisolabididae) nymphs. J Econ Entomol 112(6):2614–2619 Oberholzer F, Frank T (2003) Predation by the carabid beetles Pterostichus melanarius and Poecilus cupreus on slugs and slug eggs. Biocontrol Sci Tech 13(1):99–110 Odhiambo TR (1958) Some observations on the natural history of Acanthaspis petax Stll (Hemiptera:Reduviidae) living in termite mounds in Uganda. Proc R Entomol Soc Lond Ser A Gen Entomol 33:167–175 Oida H, Kadono F (2011) Prey consumption by Geocoris varius and G. proteus (Heteroptera: Geocoridae) provided with horticultural major pests in greenhouses. Jpn J Appl Entomol Zool 55(4):217–225 Oida H, Kadono F (2012) Development of Geocoris varius and G. proteus (Hemiptera: Geocoridae) provided with Ephestia kuehniella (Lepidoptera: Pyralidae) eggs. App Entomol Zoo 47(4): 365–372 Oliveira HN, De Clercq P, Zanuncio JC et al (2004) Nymphal development and feeding preference of Podisus maculiventris (Heteroptera: Pentatomidae) on eggs of Ephestia kuehniella (Lepidoptera: Pyralidae) parasitised or not by Trichogramma brassicae (Hymenoptera: Trichogrammatidae). Braz J Biol 64:459–463 Omkar A, Srivastava S (2003) Influence of six aphid prey species on development and reproduction of a ladybird beetle, Coccinella septempunctata. BioControl 48:379–393 O'Neil RJ (1989) Comparison of laboratory and field measurements of the functional response of Podisus maculiventris (Heteroptera: Pentatomidae). J Kans Entomol Soc 62:148–155 Ordaz-Silva S, Chacón-Hernández JC, Hernández-Juárez A et al (2014) Predation of Pselliopus latispina (Hemiptera: Reduviidae) on Tetranychus urticae (Acari: Tetranychidae). Acta Zool Mex 30(3):500–507 Ordaz-Silva S, López-Sánchez IV, Ríos-Velasco C et al (2021) First Records of Zelus tetracanthus and Pselliopus zebra1 and Copestylum caudatum2 as Predators of Monoctenus sanchezi3. Southwest Entomol 46(4):1047–1050 Orivel J, Dejean A (2001) Comparative effect of the venoms of ants of the genus Pachycondyla (Hymenoptera: Ponerinae). Toxicon 39(2-3):195–201 Orozco Muñoz A, Villalba Velásquez V, López SN (2012) Desarrollo de Tupiocoris cucurbitaceus (Hemiptera: Miridae) sobre Bemisia tabaci (Hemiptera: Aleyrodidae) en diversas hortalizas. Fitosanidad 16(3):147–153 Ouyang F, Men X, Yang B et al (2012) Maize benefits the predatory beetle, Propylea japonica (Thunberg), to provide potential to enhance biological control for aphids in cotton. PLoS One 7(9):e44379 Papanastasiou I, Kavallieratos NG, Saitanis CJ et al (2018) Parasitoids and predators of physokermes hellenicus (Hemiptera: Coccomorpha: Coccidae) in Greece. J Econ Entomol 111(3):1121–1130

516

14

Prey Record of Various Predators

Pappas ML, Broufas GD, Koveos DS (2007) Effects of various prey species on development, survival and reproduction of the predatory lacewing Dichochrysa prasina (Neuroptera: Chrysopidae). Biol Control 43(2):163–170 Parker AH (1969) The predatory and reproductive behaviour of Rhinocoris bicolor and R. tropicus (Hemiptera: Reduviidae). Entomol Exp Appl 12(1):107–117 Parshley HM (1923) Records of Nova Scotia Hemiptera-Heteroptera. Proc Acad Entomol Soc 8: 102–108 Pena NB, Manalili K, Shepard BM (1987) Predation efficiency of the short-meadow grasshopper, Conocephalus longipennis de Haan [Philippines]. In: 18 Anniversary and Annual Convention of the Pest Control Council of the Philippines, Davao City (Philippines), 5-8 May 1987 Pereira AI, Zanuncio JC, Gil-Santana HR et al (2009) Harpactor angulosus (Reduviidae: Harpactorinae), A Predator of Neotropical Saturniids, Hylesia spp. in Brazil. Entomol News 120(2):206–212 Pérez-Hedo M, Bouagga S, Jaques JA et al (2015) Tomato plant responses to feeding behaviour of three zoophytophagous predators (Hemiptera: Miridae). Biol Control 86:46–51 Pericart J (1987) Hetnipteres Nabidae d'Europe Occidentale et du Maghreb. Faune Fr 71:185 Perkins PV, Watson TF (1972) Biology of Nabis alternatus (Hemiptera: Nabidae). Ann Entomol Soc Am 65:54–57 Perumalsamy K, Selvasundaram R, Roobakkumar A et al (2009) Life table and predation of Oligota pygmaea (Coleoptera: Staphylinidae) a major predator of the red spider mite, Oligonychus coffeae (Acarina: Tetranychidae) infesting tea. Biol Control 51(1):96–101 Picker MD, Leon B, Londt JGH (1992) Influence of hindwing size in nemopterids (Insecta: Neuroptera: Nemopteridae) on predation by robber flies (Diptera: Asilidae). In: Current Research in Neuropterology: Proc. 4th Int. Symp. Neuropterol. Bagneres-de-Luchon (France), pp 313–318 Pierrard G (1972) Le contrôle de Dysdercus völkeri Schmidt défini par l'acquisition de connaissances de la biologie de l'insecte et de ses dégâts. Thèse de doctorat. Faculté des Sciences Agronomiques de l'Etat de Gembloux, Zoologie et Entomologie Tropicales, Gembloux. p 135 Pierre EM, Idris AH (2013) Studies on the predatory activities of Oecophylla smaragdina (Hymenoptera: Formicidae) on Pteroma pendula (Lepidoptera: Psychidae) in oil palm plantations in Teluk Intan, Perak (Malaysia). Asian Myrmecol 5(1):163–176 Pitkin LM (1980) A revision of Pacific species of Conocephalus Thunberg (Orthoptera: Tettigoniidae). Bull Br Mus Nat His (Ent) 41(5):315–355 Polidori C, Mendiola P, Asís JD et al (2009) Predatory habits of the grasshopper-hunting wasp Stizus continuus (Hymenoptera: Crabronidae): diet preference, predator–prey size relationships and foraging capacity. J Nat Hist 43(47-48):2985–3000 Pollet M, Desender K, Kerckvoorde MV (1987) Prey selection in Loricera pilicornis (Col., Carabidae). Acta Phytopathol Entomol Hung 22(1-4):425–431 Pollock DA (2021) Robber flies (diptera: asilidae) as predators of harvester ant workers (hymenoptera: formicidae: pogonomyrmex) in eastern new mexico and west texas. Southwest Nat 65(1):19–27 Poopat P, Maneerat T (2021) Prey preference and predation efficacy of Sycanus collaris (F.) (Hemiptera: Reduviidae) on Tenebrio molitor (L.) (Coleoptera: Tenebrionidae). Khon Kaen Univ J 49:1047–1058 Pradeep P, Deshmukh SS, Kalleshwaraswamy CM et al (2022) Biology and predation potential of the hemipteran predator, Rhynocoris marginatus (Fab., 1794) on the fall armyworm, Spodoptera frugiperda (JE Smith) (Lepidoptera: Noctuidae). Egypt J Biol Pest Contr 32(1):1–6 Pravalika K, Umamaheswari T, Shanker C (2016) Biology and Stage Preference of Assassin Bug, Rhynocoris marginatus (Fabricius) (Heteroptera: Reduviidae) on cutworm, Mythimna separata (Walker) a pest of cereal crops. Indian J Dryland Agric Res Dev 31(2):36–38 Propp GD (1982) Functional response of Nabis americoferus to two of its prey. Spodoptera exigua and Lygus hesperus. Environ Entomol 11:670–674

References

517

Qin SH (1985) Mass rearing of predatory insects in the field for control of injurious mites in apple orchards. Nat Enemies Insects Kunchong Tiandi 7:117–124 Rahman MM, Islam W, Ahmed KN (2009) Functional response of the predator Xylocoris flavipes to three stored product insect pests. Int J Agric Biol 11(3):316–320 Rajagopal D (1984) Observation on the natural enemies of Odontotermes wallonensis (Wassmann) (Isoptera: Termitidae) in South India. J Soil Biol Ecol 4(2):102–107 Rajan SJ, Suneetha N, Sathish R (2017) Biology and predatory behaviour of an assassin bug, Sycanus collaris (Fabricius) on rice meal moth, Corcyra cephalonica (Stainton) and leaf armyworm, Spodoptera litura (Fabricius). Agric Update 12(5):1181–1186 Rajan SJ, Latha ES, Sathish R (2018) Biology of Big-eyed bug, Geocoris erythrocephalus (Lepeletier & Serville) on cabbage aphid, Brevicoryne brassicae (L.). Int J Curr Microbiol App Sci 7(7):3301–3305 Ramírez Silva JA, Forero D, Santos Murgas A et al (2022) First report of Ricollapallidinends and Rocconota tuberculigera (Hemiptera: Reduviidae: Harpactorinae) predatory Oebalus insularis (Hemiptera: Pentatomidae) in natural reserves of Darien, Panama. Idesia 40(1):123–128 Ramos FLD, De Oño JL, Cosico SMQ (2019) Field evaluation and behavioral response of Black Earwig, Chelisoches morio (Fabricius) (Dermaptera: Chelisochidae) to the Coconut Leaf Beetle, Brontispa longissima (Gestro)(Coleoptera: Chrysomelidae). Ascendens Asia J Multidiscip Res Abs 3(2) Read DC (1962) Notes on the life history of Aleochara bilineata (Gyll.) (Coleoptera: Staphylinidae), and on its potential value as a control agent for the cabbage maggot, Hylemya brassicae (Bouché) (Diptera: Anthomyiidae). Can Entomol 94(4):417–424 Renkema JM, Lynch DH, Cutler GC et al (2012) Predation by Pterostichus melanarius (Illiger) (Coleoptera: Carabidae) on immature Rhagoletis mendax Curran (Diptera: Tephritidae) in semifield and field conditions. Biol Contr 60(1):46–53 Renkema JM, Manning P, Cutler GC (2013) Predation of lowbush blueberry insect pests by ground beetles (Coleoptera: Carabidae) in the laboratory. J Pestic Sci 86:525–532 Ribeiro LJ, Berti Filho E, Macedo LPM et al (2007) Predation on the citrus leafminer Phyllocnistis citrella Stainton (Lepidoptera: Gracillariidae) by Chrysoperla externa (Hagen) (Neuroptera: Chrysopidae). Caatinga 20(2):100–105 Richman DB, Hemenway RC Jr, Whitcomb WH (1980) Field cage evaluation of predators of the soybean looper, Pseudoplusia includens (Lepidoptera: Noctuidae). Environ Entomol 9(3): 315–317 Rome Q, Perrard A, Muller F et al (2021) Not just honeybees: predatory habits of Vespa velutina (Hymenoptera: Vespidae) in France. Annales de la Société entomologique de France (NS) 57 (1):1–11 Romero Sueldo M, Cuezzo F (2001) Presence of Paratrechina silvestrii (Hymenoptera: Formicidae), possible natural enemy of Doru lineare (Dermaptera: Forficulidae) in maize crops. Neotrópica 47:60 Rothschild GHL (1970) Observations on the ecology of the rice earhead bug Leptocorysa oratenius (F) (Hemiptera: Alydidae) in Sarawak (Malaysia Borneo). J Appl Ecol 7:147–167 Roy H, Migeon A (2010) Ladybeetles (Coccinellidae). Chapter 8.4. BioRisk 4:293–313 Ruberson JR, Yeargan KV, Newton BL (2001) Variation in diapause responses between geographic populations of the predator Geocoris punctipes (Heteroptera: Geocoridae). Ann Entomol Soc Am 94(1):116–122 Rubia EG, Ferrer ER, Shepard BM (1990) Biology and predatory behavior of Conocephalus longipennis (de Haan) (Orthoptera): Tettigoniidae. J Plant Prot Res 7(1):47–54 Saeed R, Naveed M, Razaq M et al (2020) Biology of newly recorded predator Antilochus coquebertii of red cotton bug, Dysdercus koenigii Fabricius (Hemiptera: Pyrrhocoridae) from Pakistan. Pak J Zool 52(5):2011–2013 Sahayaraj K (1991) Bioecology, ecophysiology and ethology of chosen predatory Hemipterans and their potential in biological control (Insecta: Heteroptera: Reduviidae). PhD Thesis, Madurai Kamaraj University, Madurai

518

14

Prey Record of Various Predators

Sahayaraj K (1995) Bioefficacy and development of reduviid predator Rhynocoris marginatus (fab) on Spodoptera litura (fab.) infesting groundnut. Int Arachnids Newslett 20:72–74 Sahayaraj K (1999) Effect of prey and their ages on the feeding preference of Rhynocoris marginatus (Fab.). Int Arachis Newslett 19:39–41 Sahayaraj K (2007) Ecotypic variation in the biology of Acanthaspis quinquespinosa Fabricius 1781 (Hemiptera: Reduviidae: Reduviinae) from peninsular India. Egypt J Biol 9:53–59 Sahayaraj K, Ambrose DP (1994) Stage and host preference and functional response of a reduviid predator Acanthaspis pedestris Stal to Four. J Biol Control 8(1):23–26 Sahayaraj K, Ambrose DP (1996) Short term functional response and stage preference of the reduviid predator Ectomocoris tibialis Distant on the cotton stainer Dysdercus cingulatus Fabr. Z Angew Zool 81(2):219–225 Sahayaraj K, Asha A (2010) Biological control potential evaluation of Rhynocoris kumarii Ambrose and Livingstone on Aphis craccivora (Koch). Indian J Agric Sci 44:281–287 Sahayaraj K, Fernandez SM (2017) Life traits and predatory potential of Antilochus coqueberti (Fab.)(Heteroptera: Pyrrhocoridae) against Dysdercus koenigii Fab. J Asia Pac Entomol 20 (4):1314–1320 Sahayaraj K, Fernandez SM (2021) The predation behavior and the prey size preferences of Antilochus coquebertii (Pyrrhocoridae) against Dysdercus koenigii (Pyrrhocoridae). Int J Trop Insect Sci 41:1763–1769 Sahayaraj K, Martin P (2003) Assessment of Rhynocoris marginatus (Fab.) (Hemiptera: Reduviidae) as augmented control in groundnut pests. J Cent Eur Agric 4:103–110 Sahayaraj K, Paulraj MG (2001) Rearing and life table of reduviid predator Rhynocoris marginatus Fab. (Het., Reduviidae) on Spodoptera litura Fab. (Lep., Noctuidae) larvae. J Appl Entomol 125:321–325 Sahayaraj K, Ravi C (2007) Evaluation of reduviid predators and plant products against chosen groundnut pests. Arch Phytopathol Plant Prot 40(4):281–290 Sahayaraj K, Sivakumar K (1995) Groundnut pest and pest stage preference of a reduviid predator Rhynocoris kumarii Ambrose and Livingstone (Heteroptera: Reduviidae). Fresenius Environ Bull 4:263–269 Sahayaraj K, Vembudurai R (2016) Bioefficacy of Catamiarus brevipennis (Servile) against two cotton pests in relation to starvation level. J Asia Pac Entomol 19(2):327–331 Sahayaraj K, Borgio JF, Muthukumar S et al (2006) Antibacterial activity of Rhynocoris marginatus (Fab.) and Catamirus brevipennis (Servile)(Hemiptera: Reduviidae) venoms against human pathogens. J Venom Anim Toxins Incl Trop Dis 12:487–496 Sahayaraj K, Lalitha C, Balasubramaniam R (2008) Biosafety of Metarhizium anisopliae (Metschnikoff) Sorokin on a reduviid predator Acanthaspis pedestris (Hemiptera: Reduviidae). Hexapoda 15(1):46–48 Sahayaraj K, Kalidas S, Tomson M (2012) Stage preference and functional response of Rhynocoris longifrons (Stål)(Hemiptera: Reduviidae) on three hemipteran cotton pests. Braz Arch Biol Technol 55:733–740 Sahayaraj K, Kumar V, Avery P (2015) Functional response of Rhynocoris kumarii (Heteroptera, Reduviidae) on Phenacoccus solenopsis (Hemiptera, Pseudococcidae) in the laboratory. Eur J Entomol 112(1):69–74 Sahayaraj K, Kumar SM, Enkegaard A (2016) Response of the reduviid bug, Rhynocoris marginatus (Heteroptera: Reduviidae) to six different species of cotton pests. Eur J Entomol 113:29–36 Sahayaraj K, Kalidas S, Estelle LYL (2020) Bioefficacy of Rhynocoris longifrons (Stål) (Heteroptera: Reduviidae) against multiple cotton pests under screen house and field conditions. Sci Rep 10(1):1–9 Sahayaraj K, Lakshmanan PD, Loko YLE (2021) Nutritional fitness of a reduviid predator Rhynocoris marginatus (fab.) (Hemiptera: Reduviidae) using biological traits and macromolecules of pests. Int J Trop Insect Sci 41:611–616

References

519

Sahid A, Natawigena WD (2018) Laboratory rearing of Sycanus annulicornis (Hemiptera: Reduviidae) on two species of prey: Differences in its biology and efficiency as a predator of the nettle caterpillar pest Setothosea asigna (Lepidoptera: Limacodidae). Eur J Entomol 115:208–216 Sahid A, Natawigena WD, Hersanti et al (2018) Laboratory rearing of Sycanus annulicornis (Hemiptera: Reduviidae) on two species of prey: differences in its biology and efficiency as a predator of the nettle caterpillar pest Setothos eaasigna (Lepidoptera: Limacodidae). Eur J Entomol 115:208–216 Saleh A, Sengonca C (2000) Untersuchungen über die Raubwanze Dicyphus tamaninii Wagner (Heteroptera: Miridae) als natürlicher Feind von Aphis gossypii Glover (Homoptera: Aphididae). Mitt Biol Bundesanst Land- Forstwirtsch 376:577–578 Saleh A, Sengonca C (2001a) Life table of the predatory bug Dicyphus tamaninii Wagner (Het., Miridae) by feeding on Aphis gossypii Glover (Hom., Aphididae) as prey. J Plant Dis Prot 108:608–615 Saleh A, Sengonca C (2001b) Development and longevity of the polyphagous mirid bug Dicyphus tamaninii Wagner (Het., Miridae) with different ages of Aphis gossypii Glover as a prey. Mitt Dtsch Ges Alg Angew Ent (Giessen) 13:193–196 Samson PR, Blood PRB (1980) Voracity and searching ability of Chrysopa signata (Neuroptera: Chrysopidae), Micromus tasmaniae (Neuroptera: Hemerobiidae), and Tropiconabis capsiformis (Hemiptera: Heteroptera: Nabidae). Aust J Zoo 28:575–580 Sánchez JA (2009) Density thresholds for Nesidiocoris tenuis (Heteroptera: Miridae) in tomato crops. Biol Control 51(3):493–498 Sánchez JA, La-Spina M, Lacasa A (2014) Numerical response of Nesidiocoris tenuis (Hemiptera: Miridae) preying on Tuta absoluta (Lepidoptera: Gelechiidae) in tomato crops. Eur J Entomol 111:387–395 Santos EB, Soares AML, Bahia BL et al (2022) First record of Zelus pedestris (Hemiptera: Reduviidae) preying on the cacao pest Monalonion bondari (Hemiptera: Miridae). Biocont Sci Tech 32(4):511–514 Sarkar SC, Wang E, Wu S et al (2019) Comparison of Acyrthosiphon pisum (Homoptera: Aphididae) and Frankliniella occidentalis (Thysanoptera: Thripidae) as prey for the development and survival of Chrysopa pallens (Neuroptera: Chrysopidae). J Asia Pac Entomol 22(3): 690–692 Sarmento RA, Pallini A, Venzon M et al (2007) Functional response of the predator Eriopis connexa (Coleoptera: Coccinellidae) to different prey types. Braz Arch Biol Technol 50:121– 126 Satpathy JM, Patnaik NC, Samalo AP (1975) Observations on the biology and habits of Sycanus affinis Reut. (Hemiptera: Reduviidae) and its status as a predator. J Bombay Nat Hist Soc 72(2): 589–595 Sattayawong C, Uraichuen S, Suasa-ard W (2016) Larval preference and performance of the green lacewing, Plesiochrysa ramburi (Schneider) (Neuroptera: Chrysopidae) on three species of cassava mealybugs (Hemiptera: Pseudococcidae). Agric Nat Resour 50(6):460–464 Saurabh P, Ganguli JL, Sonalika K et al (2019) Tri-trophic interaction of different diet fed Corcyra cephalonica (Stainton) on feeding efficiency of reduviid predator, Sycanus collaris (Fab.) under laboratory condition. J Biol Contr 33(3):271–273 Sayed S, Alghamdi A (2017) Suitability of four different prey species for Dichochrysa tacta (Neuroptera: Chrysopidae). Biocontr Sci Technol 27(2):200–209 Schaefer CW, Ahmad I (1987) Parasites and predators of pyrrhocoroidea [Hemiptera], and possible control of cotton stainers by Phonoctonus spp. [Themiptera: Reduviidae]. Entomophaga 32(3): 269–275 Schenk D, Bacher S (2002) Functional response of a generalist insect predator to one of its prey species in the field. J Anim Ecol 71(3):524–531 Sekine T, Osaka M, Itabashi T et al (2022) Predation of syrphid larvae (Diptera: Syrphidae) on thrips in onion fields intercropped with barley. App Entomol Zoo:1–7

520

14

Prey Record of Various Predators

Seydou SD, Mamadou DOUMBIA, Eric KK et al (2021) Biology of Rhynocoris squamulosus (Heteroptera: Reduviidae) fed on developmental stages of T. castaneum (Coleoptera: Tenebrionidae). J Anim Plant Sci 49(3):8933–8947 Shafique AM, Arif A, Narejo MUN et al (2019) Pre-imaginal development of different prey species on the life history parameters of different species of green lacewings (Neuroptera: Chrysopidae). Pak J Agric Res 32(2):317–324 Sharanabasappa S, Kalleshwaraswamy CM, Poorani J, Maruthi MS, Pavithra HB, Diraviam J (2019) Natural enemies of Spodoptera frugiperda (JE Smith) (Lepidoptera: Noctuidae), a recent invasive pest on maize in South India. Florida Entomol 102(3):619–623 Shepard M, Waddill VAN, Kloft W (1973) Biology of the predaceous earwig Labidura riparia (Dermaptera: Labiduridae). Ann Entomol Soc Am 66(4):837–841 Shimoda T, Shinkaji N, Amano H (1997) Prey stage preference and feeding behaviour ofOligota kashmirica benefica (Coleoptera: Staphylinidae), an insect predator of the spider miteTetranychus urticae (Acari: Tetranychidae). Exp Appl Acarol 21(10):665–675 Short BD, Bergh JC (2004) Feeding and egg distribution studies of Heringia calcarata (Diptera: Syrphidae), a specialized predator of woolly apple aphid (Homoptera: Eriosomatidae) in Virginia apple orchards. J Econ Entomol 97(3):813–819 Shrestha G, Enkegaard A (2013) The green lacewing, Chrysoperla carnea: Preference between lettuce aphids, Nasonovia ribisnigri, and western flower thrips, Frankliniella occidentalis. J Insect Sci 13:94 Siddique AB, Chapman RB (1987) Effect of prey type and quantity on the reproduction, development, and survival of Pacific damsel bug, Nabis kinbergii Reuter (Hemiptera: Nabidae). N Z J Zool 14(3):343–349 Silva A, Gonçalves CR, Galvão DM et al (1968) Catálogo dos Insetos que Vivem nas Plantas do Brasil: Seus Parasitos e Predadores. Ministério da Agricultura, Rio de Janeiro, Brazil, p 622 Silva DB, Bueno VHP, Montes FC et al (2016) Population growth of three mirid predatory bugs feeding on eggs and larvae of Tuta absoluta on tomato. Biocontrol 61:545–553 Silva LP, Souza IL, Marucci RC et al (2022) Doru luteipes (Dermaptera: Forficulidae) and Orius insidiosus (Hemiptera: Anthocoridae) as nocturnal and diurnal predators of Thrips. Neotrop Entomol:1–10 Singh OP, Gangrade GA (1976) Biology of a reduviid predator, Rhinocoris fuscipes Fabricius on the larvae of Diacrisia obliqua Walker (Lepidoptera: Arctiidae). JNKVV Res J 10:158–160 Singh HM, Rizvi SMA (1993) Evaluation of insecticides for their relative efficacy against two bugs, Metacanthus pulchellus Dallas and Cyrtopeltes tenuis (Reuter) infesting bottlegourd. Int J Plant Prot 21(1):108–109 Sitaramaiah S, Joshi BG, Prasad GR et al (1975) Harpactor costalis Stal. (Reduviidae; Heteroptera)—a new predator of the tobacco caterpillar (Spodoptera litura F.). Sci Cult 41(11):545–546 Situmorang J, Gabriel BP (1988) Biology of Htwo species of predatory earwigs Nala lividipes (Dufour) (Dermaptera: Labiduridae) and Euborellia (Euborellia) annulata (Fabricius) (Dermaptera: Carcinophoridae). Philip Entomol 7:215–238 Sloderbeck PE, Yeargan KV (1983) Comparison of Nabis americoferus and Nabis roseipennis (Hemiptera: Nabidae) as predators of the green coverworm (Lepidoptera: Noctuidae). Environ Entomol 12:161–165 Smirnoff WA (1961) Predators of Larvae of Neodiprion swainei Midd. (Hymenoptera: Tenthredinidae). Can Entomol 93(4):272–275 Sreedevi KOLLA, Veena NV, Joshi SUNIL et al (2020) Record of coccinellid predator, Hyperaspis maindroni Sicard (Coleoptera: Coccinellidae) on invasive cassava mealybug, Phenacoccus manihoti Matile-Ferrero (Hemiptera: Pseudococcidae). J Biol Contr 34(4):303–307 Srikumar KK, Bhat PS, Raviprasad TN et al (2014) Biology, behaviour and functional response of Cydnocorisgilvus Brum. (Heteroptera: Reduviidae: Harpactorinae) a predator of Tea Mosquito Bug (Helopeltisantonii Sign.) on cashew in India. J Threat Taxa 6(6):5864–5870

References

521

Stasek DJ, Rad JN, Crist TO (2018) The functional response and prey preference of generalist Nabis (Hemiptera: Nabidae) predators to leafhopper prey (Hemiptera: Cicadellidae). Can Entomol 150:190–200 Strandberg JO (1981) Predation of cabbage looper, Trichoplusiani, pupae by the striped earwig, Lahidurariparia, and two bird species. Environ Entomol 10:712–715 Stride GO (1956a) On the mimetic association between certain species of Phonoctonus (Hemiptera, Reduviidae) and the Pyrrhocoridae. J Entomol Soc South Afr 19(1):12–28 Stride GO (1956b) On the biology of certain West African species of Phonoctonus (Hemiptera, Reduviidae), mimetic predators of the Pyrrhocoridae. J Entomol Soc South Afr 19(1):52–69 Sueldo MR, Bruzzone OA, Virla EG (2010) Characterization of the earwig, Dorulineare, as a predator of larvae of the fall armyworm, Spodopterafrugiperda: a functional response study. J Insect Sci 10:38 Suenaga H (2017) Ants (Hymenoptera: Formicidae) as potential egg predators of lepidopteran pests of sweet potato, Ipomoea batatas (Solanales: Convolvulaceae), in Japan. Appl Entomol Zool 52(2):197–207 Sundararaju D (1984) Cashew pests and their natural enemies in Goa. J Plant Crop 12:38–46 Swadener SO, Yonke TR (1973) Immature stages and biology of Apiomerus crassipes (Hemiptera: Reduviidae). Ann Entomol Soc Am 66(1):188–196 Syari J, Muhamad R, Norman K et al (2010) Feeding behavior and predatory efficiency of assassin bug, Sycanus dichotomus Stål on oil palm bagworm, Metis aplana Walk. Malays Appl Biol l39:51–55 Syari J, Muhamad R, Norman K et al (2011) Laboratory rearing of Sycanus dichotomus Stål (Hemiptera: Reduviidae), insect predator of oil palm bagworm, Metis aplana Walker (Lepidoptera: Psychidae). Sains Malays 40:1129–1137 Tamaki G, Olsen DP, Gupta RK (1978) Laboratory evaluation of Geocoris bullatus and Nabis alternatus as predators of Lygus. J Entomol Soc BC 75:35–37 Tillman PG, Mullinix BG Jr (2003) Effect of prey species on plant feeding behavior by the big-eyed bug, Geocoris punctipes (Say) (Heteroptera: Geocoridae), on cotton. Environ Entomol 32 (6):1399–1403 Tobing MC, Kuswardani RA (2018) The potential of Myopopo necastanea (Hymenoptera: Formicidae) as a predator for Oryctes rhinoceros Linn. larvae (Coleoptera: Scarabaeidae). In: Journal of Physics: Conference Series, vol 1116. IOP Publishing, Bristol, p 052074 Tomson M (2021) Functional response, host stage preference and development of Rhynocoris fuscipes (Fab.) (Heteroptera: Reduviidae) for two cotton pests. JBiopest 14(1):12–21 Torres JB, Silva-Torres CS, Ruberson JR (2004) Effect of two prey types on life-history characteristics and predation rate of Geocoris floridanus (Heteroptera: Geocoridae). Environ Entomol 33(4):964–974 Ueno T (2014) Establishment of the invasive hornet Vespa velutina (Hymenoptera: Vespidae) in Japan. Int J Chem Environ Biol Sci 2(4):220–222 Ullah MI, Altaf N, Afzal M et al (2019) Effects of entomopathogenic fungi on the biology of Spodoptera litura (Lepidoptera: Noctuidae) and its reduviid predator, Rhynocoris marginatus (Heteroptera: Reduviidae). Int J Insect Sci 11:1–7 Urbaneja A, Tapia G, Stansly P (2005) Influence of host plant and prey availability on developmental time and surviorship of Nesidiocoris tenius (Het.: Miridae). Biocontrol Sci 15(5):513– 518 Urbaneja A, Marí FG, Tortosa D et al (2006) Influence of ground predators on the survival of the Mediterranean fruit fly pupae, Ceratitis capitata, in Spanish citrus orchards. Biocontrol 51 (5):611–626 Urbaneja A, Montón H, Mollá O (2009) Suitability of the tomato borer Tuta absoluta as prey for Macrolophus pygmaeus and Nesidiocoris tenuis. J Entomol 133:292–296 Van den Berg H, Litsinger JA, Shepard BM (1992) Acceptance of eggs of Rivula atimeta, Naranga aenescens [Lep.: Noctuidae] and Hydrellia philippina [Dipt.: Ephydridae] by insect predators on rice. Entomophaga 37(1):21–28

522

14

Prey Record of Various Predators

Van Doesburg PH (1964) Termatophyli deaopaca Carvalho, a predator of thrips (Hem.-Het.). Entomol Ber (Amsterdam) 24:248–253 van Lenteren JC, Hemerik L, Lins JR (2016) Functional responses of three neotropicalmirid predators to eggs of Tuta absoluta on tomato. Insects 7:34 van Lenteren JC, Bueno VHP, Calvo FJ et al (2018) Comparative effectiveness and injury to tomato plants of three neotropicalmirid predators of Tuta absoluta (Lepidoptera: Gelechiidae). J Econ Entomol 111:1080–1086 van Rijn PC, Kooijman J, Wackers FL (2006) The impact of floral resources on syrphid performance and cabbage aphid biological control. IOBC WPRS Bull 29(6):149 Vanderplank FL (1958) The assassin bug, Platymerus rhadamanthus Gerst (Hemiptera: Reduviidae), a useful predator of the rhinoceros beetles Oryctes boas (F.) and Oryctes monoceros (Oliv.) (Coleoptera: Scarabaeidae). J Entomol Soc South Afr 21(2):309–314 Varshney R, Ballal CR (2017a) Biology and feeding potential of Geocoris superbus Montandon (Heteroptera: Geocoridae), a predator of mealybug. J Entomol Zool Stud 5:520–524 Varshney R, Ballal CR (2017b) Biological, morphological and life table parameters of the predator, Geocoris ochropterus Fieber. (Hemiptera: Geocoridae) fed on Sitotroga cerealella (Olivier) eggs. Egypt J Biol Pest Control 27:189–194 Varshney R, Budhlakoti N (2022) Biology and functional response of the predator, Dortus primarius (Distant) (Hemiptera: Miridae) preying on Frankliniella schultzei (Trybom) (Thysanoptera: Thripidae). Egypt J Biol Pest Contr 32(1):1–7 Varshney R, Budhlakoti N, Ballal CR (2018) Functional response of Geocoris ochropterus Fieber (Hemiptera: Geocoridae) to different egg densities of Helicoverpa armigera (Hübner) (Lepidoptera: Noctuidae). Phytoparasitica 46:451–458 Vasconcelos SD, Williams T, Hails RS et al (1996) Prey selection and baculovirus dissemination by carabid predators of Lepidoptera. Ecol Entomol 21(1):98–104 Vennison SJ, Ambrose DP (1990) Biology and behaviour of Sphedanolestes signatus Distant (Insecta: Heteroptera: Reduviidae) a potential predator of Helopeltis antonii Signoret. UP J Zool 10:30–43 Verma JS, Sharma KC (2006) Biology and predatory potential of Metasyrphus confrater (Wiedemann) on aphid, Macrosiphum rosae L. infesting Rosa spp. J Entomol Res 30:31–32 Veronesi ER, Saville DJ, van Koten C et al (2022) Potential of the mirid bug, Engytatus nicotianae, for the biological control of the tomato-potato psyllid in greenhouses. J Crop Prot 156:105941 Virla EG, Melo CM, Speranza S (2015) Preliminary observations on Zelus obscuridorsis (Stål) (Hemiptera: Reduviidae) as predator of the corn leafhopper (Hemiptera: Cicadellidae) in Argentina. Insects 6(2):508–513 Wei D, Xian X, Zhou Z et al (1998) Preliminary study on the functional responses of Cyrtopeltis tenuis to Spodoptera litura. Acta Agri Univ Henanensis 32:55–59 Weiss MJ, McDonald G (1998) European earwig, Forficula auricularia L. (Dermaptera: Forficulidae), as a predator of the redlegged earth mite, Halotydeus destructor (Tucker) (Acarina: Penthaleidae). Aust J Entomol 37(2):183–185 Werner D, Pont AC (2003) Dipteran predators of Simuliid blackflies: a worldwide review. Med Vet Entomol 17(2):115–132 Weseloh RM (1988) Prey preferences of Calosoma sycophanta L. (Coleoptera: Carabidae) larvae and relationship of prey consumption to predator size. Can Entomol 120(10):873–880 Wiedenmann RN, O'Neil RJ (1990) Response of Nabis roseipennis [Heteroptera: Nabidae] to Larvae of Mexican bean beetle, Epilachna varivestis [Coleoptera: Coccinellidae]. Entomophaga 35(3):449–458 Wiedenmann RN, O’Neil RJ (1992) Searching strategy of the predator Podisus maculiventris (Say) (Heteroptera: Pentatomidae). Environ Entomol 21(1):1–9 Wiedenmann RN, Legaspi JC, O’Neil RJ (1994) Impact of prey density and facultative plant feeding on the life history of the predator Podisus maculiventris (Heteroptera): Pentatomidae. In: Alomar O, Wiedenmann RN (eds) Zoophytophagous Heteroptera: implications for life history and integrated pest management. Entomology Society of America, Lanham, pp 95–118

References

523

Williams G (1954) Bull Entomol Res 45(4):723 Xue Y, Bahlai CA, Frewin A et al (2009) Predation by Coccinella septempunctata and Harmonia axyridis (coleoptera: Coccinellidae) on Aphis glycines (Homoptera: Aphididae). Environ Entomol 38(3):708–714 Young OP (1980) Predation by tiger beetles (Coleoptera: Cicindelidae) on dung beetles (Coleoptera: Scarabaeidae) in Panama. Coleopt Bull 34:63–65 Youssef NA, Abd-Elgayed AA (2015) Biological parameters of the predator, Amphibolus venator Klug (Hemiptera: Reduviidae) preying on larvae of Tribolium confusum Duv. (Coleoptera: Tenebrionidae). Ann Agric Sci 60:41–46 Yuliadhi KA, Supartha IW, Wijaya IN et al (2021) The preference and functional response of Sycanus aurantiacus (Hemiptera: Heteroptera: Reduviidae) on three prey types in laboratory conditions. Biodivers J Biol Divers 22(12) Yumika S, Youichi K, Hiroshi OIDA et al (2012) Effect of Agrichemicals on the Polyphagous Predatory Bug, Geocoris varius (Uhler) (Heteroptera: Geocoridae). Jpn J Appl Entomol Zool 56:43–48 Yusof I, Fairuz MO (2011) Demographic parameters reproductive performance of the assassin bug Sycanus dichotomus Stal.fed on mealworm Tenebrio molitor L. J Oil Palm Res 26:974–978 Zanuncio JC, Zanuncio TV, Guedes RNC et al (2000) Effect of feeding on three eucalyptus species on the development of Brontocoris tabidus (Heteroptera: Pentatomidae) fed with Tenebrio molitor (Coleoptera: Tenebrionidae). Biocontr Sci Technol 10:443–450 Zanuncio JC, Silva CADD, Lima ERD, Pereira FF, Ramalho FDS, Serrão JE (2008) Predation rate of Spodoptera frugiperda (Lepidoptera: Noctuidae) larvae with and without defense by Podisus nigrispinus (Heteroptera: Pentatomidae). Braz Arch Biol Technol 51:121–125 Zarghami S, Mossadegh MS, Kocheili F et al (2016) Functional responses of Nephus arcuatus Kapur (Coleoptera: Coccinellidae), the most important predator of spherical Mealybug Nipaecoccus viridis (Newstead). Psyche 2016:9417496 Zegula T, Sengonca C (2000) Entwicklung biologischer Bekämpfungsmethoden gegen den Schadthrips Frankliniella occidentalis (Pergande) durch Verwendung natürlicher Feinde im Unterglasanbau. Mitt Biol Bundesanst Land-Forstwirtsch 376:581–582 Zheng Z, Liu H, Wang X et al (2019) Development and reproduction of the hoverfly Eupeodes corollae (Diptera: Syrphidae). JESES 4(4) Zhu J, Obrycki JJ, Ochieng SA et al (2005) Attraction of two lacewing species to volatiles produced by host plants and aphid prey. Naturwissenschaften 92(6):277–281 Ziaei Madbouni MA, Samih MA, Namvar P et al (2017) Temperature-dependent functional response of Nesidiocoris tenuis (Hemiptera: Miridae) to different densities of pupae of cotton whitefly, Bemisia tabaci (Hemiptera: Aleyrodidae). Eur J Entomol 114:325–331 Zipagan MB, Pacumbaba EP (1994) Perigrinator biannulipes Montrouzier and Sigmoret, a rudiviid predator of insect pests of copra. In: Pest Management Council of the Philippines, Inc. Anniversary and Annual Scientific Meeting. Cagayan de Oro City (Philippines), pp 3–6 Zulkefli M, Norman K, Basri MW (2004) Life cycle of Sycanus dichotomus (Hemiptera: Pentatomidae)—a common predator of bagworm in oil palm. J Oil Palm Res 16(2):50–56

Mass Production of Insect Predators

15

Contents 15.1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.1.1 Levels of Mass Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.1.2 Concepts of Artificial Diet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2 Hemipteran Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.1 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.2 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.3 Lygeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.4 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.5 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.2.6 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.3 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.4 Lacewings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.5.1 Mass of Adults and Eggs and Ovariole Number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.6 Quality Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.7 Future Recommendation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

15.1

525 526 527 530 530 534 534 536 552 552 554 556 559 564 567 568 572

Introduction

In order to better utilise biological control agents (BCAs), advanced raising techniques, affordable replacement meals, and dependable field application techniques are now required due to the commercial use of biological control’s rapid growth. The bulk breeding of BCAs on fictitious hosts has drawn more attention as a result of the widespread use of BCAs. The factitious hosts of BCAs include hemipteran herbivores, dipteran larvae, lepidopteran eggs or larvae, and artificial diets. For the bulk production of parasitoids and predators, Corcyra cephalonica eggs and larvae, a typical factitious host for BCAs, have been particularly effectively developed. Studies have demonstrated the adequacy of food for the

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_15

525

526

15

Mass Production of Insect Predators

bulk mass rising of BCAs across multiple generations (e.g., parasitoid wasps and predatory mites). However, there is not much information on how it was modified for large production in a banking plant system. This approach has been used in numerous researches to assess how insects temperature, the host plant or prey, insecticides, etc., are all influences. It is a promising study method that can analyse the impacts of various food sources and environmental circumstances on BCAs in pest management while properly estimating the field population size. Moreover, a successful mass-rearing programme and the field use of BCAs both benefit from the linkage between the life table and the predation rate. This chapter discusses a number of important aspects of predatory insect-rearing techniques. Review of the importance of natural, fabricated, and artificial feeds for the main species employed in biological control. There are numerous fictitious prey/ foods and artificial diets for generalist animals that are either insect-free or insectbased. The development of artificial diet-based production systems for polyphagous as opposed to oligophagous animals may be more efficient in terms of time and money. To lessen the detrimental impacts of crowding and cannibalism in colonies, more research is required to ascertain how to adjust rearing (population) densities— relative to food quality/quantity, cage size, oviposition, and mating. To begin achieving these goals, effective, standardised mass-rearing techniques are needed, including the following: 1. 2. 3. 4.

The use of low-cost, nutrient-dense diets; Mechanised and space-efficient production systems; Trustworthy storage techniques; and Periodically assessing the quality of natural enemies.

15.1.1 Levels of Mass Production According to Morales-Ramos et al. (2013), predators can be manufactured utilising artificial diets at levels 1–6 from a small number of individuals to millions of individuals. The amount of production and problems grew with each level. 1. Only a small number of people are created in artificial media at level 1, which Arthropods are raised using holidic artificial meals in tiny containers or petri plates. 2. A sustainable population at level 2 reflects a colony size of approximately 100 people at each life stage with artificial diets; however, for the sake of simplification, some chemically determined components are frequently replaced by natural components. 3. Methods to collect eggs in the absence of hosts or prey are frequently developed at this stage, driven by the necessity to generate arthropods in vitro in quantities of several hundred individuals (considerable numbers). The majority of the

15.1

Introduction

527

ingredients in diets are meridic; however, some still might be chemically characterised. 4. Production of thousands of individuals at level 4 drives the advancement of complex meridional diet-based rearing techniques. However, storage and microbiological contamination are frequent. 5. The creation of tens of thousands to hundreds of thousands of level 5 professionals. At this level, the content of the diet is typically oligidic, with few chemically defined components, primarily vitamins and few amino acids, 6. The production capabilities at level 6 should be able to serve millions of people. To achieve this level of production, the majority of raising techniques must be mechanised or drastically simplified. It is necessary to have dependable packaging, shipping, and release procedures.

15.1.2 Concepts of Artificial Diet According to the degree of chemical definition, which ranges from fully chemically defined to partially chemically defined to completely undefined, artificial diets are categorised as holidic, meridic, and oligidic (Cohen and Crittenden 2004). By employing data from chemical analyses of the host or prey, which were utilised to approximate the nutritional requirements of its predator, artificial diets have been created. 1. Meridic artificial diets are those that exclude all insect ingredients. 2. Holidic artificial diets are usually defined chemically The chemical composition of artificial food is unique to each predator. The main components of an artificial diet, however, are carbohydrates (monosaccharides such as arabinose, ribose, xylose, and galactose, and disaccharides such as sucrose and maltose), lipids such as linoleic and linolenic acids and their derivatives (eicosanoids), sterols such as phytosterols from plants and ergosterol from fungi, phospho Proteins (globulins, nucleoproteins, lipoproteins, and insoluble protein), 10 essential amino acids (leucine, isoleucine, valine, threonine, lysine, methionine, histidine, phenylalanine, and tryptophan), vitamins (vitamins C and B), retinol, carotenoids (A), tocoferols (E), calciferol (D). It is included in Table 15.1 and is accessible to different predators in different forms. When necessary, food should be refined based on how it performs. Five steps are proposed by Morales-Ramos et al. (2013) as an approach for diet improvement: 1. The relative amounts of the three main nutritional types—lipid, protein, and carbohydrate—should be determined 2. The proteins’ amino acid composition (chemically defined additions may still be required in the case of amino acids); 3. The proportions of the three different kinds of lipids—saturated, monounsaturated, and polyunsaturated—in the body

528

15

Mass Production of Insect Predators

Table 15.1 Package materials utilised to provide different forms of artificial diet for various predatory insects are listed here with references Predators Perillus bioculatus Podisus maculiventris Lygus hesperus Lygus lineolaris Geocoris punctipes Podisus sagitta Podisus maculiventris Chrysoperla rufilabris Rhynocoris margintus Rhynocoris kumarii Rhynocoris fuscipes

Consistency Liquid or semiliquid Liquid or semiliquid Semiliquid Semiliquid Solid diet Solid diet

Package material Parafilm®

References Rojas et al. (2000)

Mylar®– Parafilm® – – Parafilm® Parafilm®

Wittmeyer et al. (2001)

Solid diet

Parafilm®

Solid diet

Parafilm®

Patana (1982) Cohen (1998) Cohen (1985) De Clercq and Degheele (1992) De Clercq and Degheele (1992) Cohen and Smith (1998a, b)

Liquid

Cotton/parafilm

Sahayaraj et al. (2003)

Liquid Liquid

Cotton/parafilm Cotton/parafilm

Sahayaraj et al. (2003) Sahayaraj et al. (2003)

4. Sugars, digestible polysaccharides, and indigestible polysaccharides (fibre), in relation to one another and 5. The sum of the three primary nutrient categories—lipid, protein, and carbohydrate. According to Eric and Riddick (2009), factors that could guarantee progressive rearing success might include 1. taking advantage of the highly polyphagous nature of some species, 2. subjecting predators to artificial diets for several generations to promote adaptation, 3. creating artificial diets that mimic preferred natural prey in terms of texture and chemical composition, and 4. encasing the diet in Parafilm to resemble natural prey in terms of shape. Encasement may also lessen desiccation and delay diet deterioration. Other fictitious prey can be found, and artificial diets can be developed to ensure the creation of high-quality predators. Juan et al. (2023) say according to the degree of chemical definition—fully chemically defined, partially chemically defined, or completely undefined—artificial diets are categorised as holidic, meridic, and oligidic, respectively. Meridic artificial diets are also described as being free of any insect ingredients. Artificial diets must be developed with the successful mass production of entomophagous arthropods at a cost appropriate for commercial application and a process that can be automated as their ultimate goals if artificial

15.1

Introduction

529

diets are to play a significant role in the mass production of natural enemies. In order to conceptualise the process of developing an artificial diet, we divided it into the six levels necessary to accomplish these objectives. Level 1: In the artificial diet formulation, the arthropod completes development and reproduces. At this stage of life, arthropods are often raised in tiny containers or Petri plates and fed artificial diets that are typically chemically specified (i.e., holidic). At this stage of development, however, there are some cases of meridic and oligidic diets. At this level, artificial media only produces a small number of people. Level 2: In an artificial diet, the arthropod completes its development and reproduces for at least five generations. It is necessary to maintain a viable population in vitro in order to sustain a colony across several successive generations. A viable population is about equal to the colony size of 100 people at each stage of life. For this level of output, several rearing techniques must be streamlined. For instance, artificial diets may still be holidic in composition, but for the sake of simplification, many times some chemically defined components are swapped out for natural ones. Level 3: Enough reproducible arthropods are produced on artificial diets to allow for a statistically meaningful comparison of biological parameters with arthropods raised on natural hosts. The ability to collect eggs in the absence of hosts or prey is frequently developed at this stage due to the requirement to create arthropods in vitro in significant numbers (several hundred individuals). The majority of the ingredients in diets are meridic; however, some still might be chemically characterised. Level 4: Enough arthropods are created by artificial means to allow for experimental field releases and scientifically reliable field assessments. The need for thousands of individuals to be produced for field releases drives the evolution of more complex rearing techniques. It is necessary to establish strategies to improve food mixing and dispensing, oviposition, egg handling, immature feeding, adult emergence, and packing for release. Meridic diets with accommodations for microbial contamination and storage are typical at this level. Level 5: Enough arthrophages are grown through artificial nutrition to be released in a commercial field and control the target pest at least somewhat. Tens of thousands to hundreds of thousands of high-quality individuals must be produced in order to release arthropods in a commercial field for pest management. To keep up that level of production, some level of mechanisation must take place. Diet and rearing methods must be completely dependable, else the business will collapse. At this level, the content of the diet tends to be oligidic, with few chemically defined components, primarily vitamins and few amino acids. The quality of the food at this level must be such that it produces natural adversaries that are at least as effective as those raised on their native hosts or prey. The diet must be prepared in large quantities; hence the formulation needs to be stable enough for short-term storage. Level 6: Arthropod production levels in artificial diets are only constrained by available space, and costs are reasonable for commercialisation. This amount of production should be able to serve millions of people. To achieve this level of production, the majority of raising techniques must be mechanised or drastically

530

15

Mass Production of Insect Predators

simplified. Cheaper substitutes are used in place of pricey components and diet items. At this level, diets are typically steady and oligidic in composition. Mechanised processes and complex sterilisation techniques, such as extrusion and/or flash sterilisation, are used in the mass manufacturing of nutrition. It is necessary to create dependable field release, shipping, and packaging procedures.

15.2

Hemipteran Predators

In diverse agroecosystems, a number of species of predatory Heteroptera from the families Anthocoridae, Miridae, Lygaeidae, Nabidae, Reduviidae, and Pentatomidae (subfamily Asopinae) have been or are being used in programmes to supplement biological management of significant agricultural pests. In small-scale lab cultures, some predatory bugs are raised on their natural prey. However, these systems of animal rearing frequently entail significant labour expenses as well as expensive plant production facilities (such as greenhouses) and may experience discontinuity issues. A variety of factitious hosts, such as the eggs of some lepidopterans like Ephestia kuehniella (Lepidoptera: Pyralidae), Corcyra cephalonica (Lepidoptera: Pyralidae), and Sitotroga cerealella (Lepidoptera: Gelechiidae), and cysts of brine shrimps, Artemia spp., have proven advantageous for continuous mass rearing (Anostraca: Artemiidae). For predatory bugs, other artificial diets have also been suggested, although their use in large-scale production systems is still quite limited. The majorities of these diets are made up of meat and eggs and are packaged in parafilm sachets. It is raised using larvae of Tenebrio molitor (Coleoptera, Tenebrionidae), Helicoverpa armigera, Spodoptera litura, Earias insulana, Earias vitella (Lepidoptera, Noctuidae), and Plutella xylostella (Lepidoptera: Plutellidae).

15.2.1 Anthocoridae The utilisation of eggs from Ephestia kuehniella (Lepidoptera: Pyralidae) as a rearing diet is essential for the commercial production of Orius spp. (Anthocoridae), especially Orius majusculus. Ephestia kuehniella eggs, however, have become a pricey feed, driving up the cost of these important biological control agents. Artificial diet utilisation may result in lower production costs. Establishing a connection between nutritional composition and fitness in this area could aid in the creation of an ideal artificial substitute diet for these predatory insects. The current study’s objective was to assess the effects of Ephestia kuehniella eggs and six artificial diets with varying macronutrient compositions on the growth and reproductive fitness of Orius majusculus. In general, nutrition had no impact on nymphal survival, although unnatural diets caused development to take a little longer. However, compared to the Ephestia kuehniella eggs diet, female body mass and fecundity were significantly lower on all artificial diets, indicating that they were of lower quality. Females raised on artificial meals that were fed the viable food with the highest concentration of fat produced more eggs than those who were fed the diet

15.2

Hemipteran Predators

531

with the highest protein content. The Orius majusculus fed on different diets showed some variation in carcase composition, but these variations did not correspond to the differences seen in the fitness indices assessed (Montoro et al. 2020). Orius spp. have recently made strides in cultivation thanks to artificial meals based on liver rather than insect prey (Arijs and De Clercq 2004). This opens the door for a much larger application of these predators in daily life. Numerous research has examined how artificial diets enriched with embryonic cell lines affected Orius insidiosus (Ferkovich and Shapiro 2004), Orius laevigatus (Bonte and De Clercq 2008), and Orius sauteri (Tan et al. 2013) population performance. Numerous commercially significant insect pests are preyed upon by Orius sauteri, a significant predator. Orius sauteri is challenging to mass rear, which restricts its use in pest management. Here, we evaluated the nutritional value of eggs from Spodoptera litura, Agrotis ypsilon, and Sitotroga cerealella as well as their potential for two generations of laboratory-raised Orius sauteri. Sitotroga cerealella eggs produced the maximum survival and reproduction of Orius sauteri among the examined species when compared to the other two lepidopteran species. The nymphal stage thrived on eggs from Agrotis ypsilon, where it developed more quickly than it did on eggs from Sitotroga cerealella. On the other hand, Spodoptera litura eggs were not a good diet for Orius sauteri and interfered with its growth and reproduction. Sitotroga cerealella eggs displayed benefits across the board when its nutritional content was examined. Protein levels in Orius sauteri fed Spodoptera litura eggs were substantially lower than those in Orius sauteri fed the other examined eggs. The CAT activity in Orius sauteri was markedly increased by Spodoptera litura eggs, which may indicate that some elements from S. litura eggs were detrimental to the growth and reproduction of Orius sauteri. These findings lead us to recommend a mixed diet for the mass rearing of the pirate bug, giving the nymphs and adults eggs from Sitotroga cerealella and Agrotis ypsilon, respectively. This research advances the understanding of artificial diets for the future mass rearing of Orius sauteri and other Orius species (Zhang et al. 2022). Based on the age stage, two-sex life table theory thoroughly assessed the suitability of artificial foods for Orius strigicollis and also determined the harvest rate with cost-effectiveness consideration. The following techniques were used to create three meridic artificial diets and three oligidic diets (Hung et al. 2021): 1. Yeast extract powder was mixed to reverse osmosis water and autoclaved for 25 min at 121 °C to create three meridic artificial diets without insect component (designated as M1, M2, and M3). The weighed chicken egg yolk, honey, sucrose, beef extract, milk powder, and lactalbumin were all successively incorporated into the autoclaved solution once it was cooled (i.e., to below 40 °C). Following a thorough mixing process using magnetic stirrers for 30 min, the diet was aliquoted into 2 ml Eppendorf tubes (Axygen, China) and kept at 20 °C until it was needed. 2. Galleria mellonella (Lepidoptera: Pyralidae), Philosamia cynthia ricini (Lepidoptera: Saturniidae), and pre-pupa of Hermetia illucens (Diptera: Stratiomyidae) pupae and pre-pupa, respectively, were used to construct three oligidic diets with

532

15

Mass Production of Insect Predators

insect components (designated O1, Before usage, all of the pupae and pre-pupae were kept in storage at 20 °C. The air-dried insect carcases were pulverised in a mortar with liquid nitrogen and added to the diet after the pupae and pre-pupae were sterilised with 75% ethanol, rinsed with reverse osmosis water, and ground. It has been determined that Blaptostethus pallescens (Hemiptera: Anthocoridae) is a potential predator of pests that feed via sucking, including thrips, mites, aphids, cotton, and papaya mealybugs. Using UV-irradiated eggs of Corcyra cephalonica (Lepidoptera: Pyralidae) and Sitotroga cerealella (Lepidoptera: Gelechiidae) as prey, laboratory investigations on the biology of Blaptostethus pallescens were done. When raised on Corcyra cephalonica eggs as opposed to Sitotroga cerealella eggs, it was shown that longevity and fertility were much higher. When given Sitotroga cerealella or Corcyra cephalonica eggs as prey, Blaptostethus pallescens nymphs and adults’ daily feeding rates did not significantly change. The adult males and females consumed substantially more Corcyra cephalonica eggs than Sitotroga cerealella eggs when total feeding was taken into account. The current study suggests that Corcyra cephalonica eggs are suitable as host material for the mass rearing of Blaptostethus pallescens based on biological criteria. By using UV-irradiated Corcyra cephalonica eggs as prey, a straightforward mass production technique for Blaptostethus pallescens was standardised. 450,000 nymphs could be produced monthly by keeping 150 ovipositional containers with 50 Blaptostethus pallescens adults each, supplying Corcyra cephalonica eggs as food, and using bits of bean pods as the ovipositional substrate. It was calculated that it would cost Rs. 150 to produce 1000 seven-day-old Blaptostethus pallescens nymphs by factoring in both fixed and variable costs (approximately 2.3 US dollars). The procedure created can be used to produce this predator commercially. The price of producing Blaptostethus pallescens might be further decreased if a production facility for Corcyra cephalonica is also present in the same production unit (Gupta et al. 2013). The type of food is one of the most important aspects in determining how well anthocorids are raised. Artmia franciscana (Anostraca: Artemiidae), a brine shrimp, and a whole-pupa homogenate of the Antheraea paphia (Lepidoptera: Saturniidae) have all been used in artificial feeding methods (Tan et al. 2013). The diet employed had a big impact on how quickly Orius strigicollis nymphs developed and how many survived. With the exception of Diet M1, all artificial diets that were examined had nymphal survival rates of more than 80%. Diet M2 had the shortest nymphal lifespan (15.0 days) and the highest nymphal survival rate (96.3%) of the three meridic diets. The survival rates of Orius strigicollis fed on the three oligidic diets were comparable (83.3–87.7%); however, the paired bootstrap test revealed that Orius strigicollis fed on Diet O3 had a considerably shorter nymphal period (14.5 days) than those in other oligidic-fed groups. Of all the cohorts studied, Orius strigicollis nymphs fed on Diet M2 had the highest survival rate and one of the quickest development times. However, once the females reached adulthood, they did not necessarily have a higher reproduction rate than other adults (Hung et al. 2021). The combined Diet 1 did improve the fertility of adult females, showing that changing the diet affects how Orius strigicollis reproduces. On combined Diet 1, the most eggs were deposited,

15.2

Hemipteran Predators

533

with an average of 54.7 eggs per female. Furthermore, compared to female adults given only Diet M2, eggs deposited by female adults fed the combined Diet 1 demonstrated a significantly greater hatch rate (77.5%) throughout the whole lifespan. However, Orius strigicollis fed on the combined Diet 2 showed no improvement in fertility or hatch rate (Hung et al. 2021). Biological management of aphids, leaf mites, diamondback moths (Plutella xylostella Lepidoptera: Plutellidae), and two-spotted spider mites (Tetranychus urticae) is anticipated to be effective with the native Asian species Orius minutus (Acari: Tetranychidae). The provision of artificial substrates for oviposition Orius minutus lays 13.1 eggs per female on cork, followed by plants (12.3 eggs per female) and rubber bands (4.1 eggs per female). The predatory bug, Orius laevigatus, was raised on three different diets: the CY diet, which consisted of 95% cornmeal and 5% yeast, the CBGY diet, which consisted of 53.3% cornmeal and 26.7% wheat bran and 15% glycerine and 5% yeast, and the WBGY diet, which consisted of 53.3% wheat flour and 26.7% wheat bran and 15% glycerine and 5% yeast (WBGY diet). When the eggs laid by moths raised on the CBGY diet were fed, the development period of the second nymphal stage was shortened. However, the eggs of Ephestia kuehniella raised on the various larval diets did not significantly differ in how long the remaining nymphal stages and overall development period of Orius laevigatus took to mature. The biological characteristics of Orius laevigatus adults were noted. The times before oviposition, during oviposition, and after oviposition were not significantly impacted by the various artificial diets. Additionally, Orius laevigatus adults’ lifespans and fecundity did not statistically change when raised and fed eggs made by moths raised on various artificial diets (Pehlivan 2021). The anthocorid genus Xylocoris has been found in Brazilian maize. It has roughly 60 species, the most of which are found in the northern hemisphere. Although they can also be found in temperate places, these predators are native to hot climates. They have been employed successfully to slow the spread of various Lepidoptera and Coleoptera grain pests by feeding on their eggs, larvae, and pupae. Most species in the genus Xylocoris are predators, particularly of pests that attack stored-grain crops. Plutella xylostella, the diamondback moth, was under the authority of Xylocoris (Lepidoptera: Plutellidae). The two treatments’ net reproduction rates (R0), which were 14.7 for Corcyra cephalonica and 1.5 for Plutella xylostella, were different. The average generation time (T) showed the reverse pattern, with lower values in the Plutella xylostella treatment (26.6 days) and greater values in the Corcyra cephalonica treatment (44.0 days). The time it took for the population to double in size was substantially shorter when the predator was fed eggs from this prey, but the intrinsic rate of population growth (rm) and the finite population growth rate were significantly higher in the Corcyra cephalonica treatment (Vieira et al. 2018). The developmental period and survival rate were noted while feeding potential meals for Orius minutus in order to choose one. We verified that, in comparison to other treatment groups, the group treated with brine shrimp eggs had a longer developmental time, ranging from 5.9 to 7.8 days. The survival rate of the treatment group given brine shrimp eggs with iron coating was 67.5%, which was 5.9 times

534

15

Mass Production of Insect Predators

greater than that of brine shrimp eggs in granular form. This supported the statistical significance of the findings. The treatment group fed granular brine shrimp eggs did not generate any females to lay eggs in the fecundity experiment; hence the fecundity could not be measured. Ephestia cautella and iron-coated brine shrimp eggs were fed to the groups, and fertility was high. In both the E (Ephestia cautella eggs) and I (Iron-coated brine shrimp eggs) treatment groups, the average number of eggs deposited each day was comparable. Our suggested system’s costs were examined and contrasted with those of the traditional mass-rearing system. We used the “industrial standards of insect breeding and specification II” to determine the egg-harvesting cost of the traditional mass-rearing technique. The cost of egg collection was cut by 98.2% compared to the traditional mass-rearing approach, and the cost of diet was down by 48.3%, bringing the whole production cost down by 70.5% (Jun et al. 2022).

15.2.2 Nabidae Nabis pseudoferus produced 165.3 and 14.7 eggs per female per day while it fed on Aphis gossypii, respectively, for oviposition rates (Sadat Mahdavi and Madadi 2016). However, when given the opportunity to feed on egg Ephestia Kuehniella as fake prey, the intrinsic rate of increase (r), the net reproductive rate (R0), the gross reproductive rate (GRR), the finite rate of population increase, and the mean generation time (T ) were estimated to be 0.0470.004 day 1, 12.68 eggs/individual, 35.17 eggs/individual, 1.04 day 1, and 58.14 days, respectively (Madadi et al. 2016). A significant biocontrol agent for sucking pests and caterpillars is Nabis pseudoferus. The life history characteristics, rate of consumption, and cold storage of Nabis pseudoferus were calculated for three diets: Aphis gossypii, Ephestia kuehniella eggs, and Aphis gossypii + Ephestia kuehniella eggs at 25.1 °C, 60.10% R.H., and a photoperiod of 16:8 h. (L:D). The findings showed that, depending on the feeding regimens, the intrinsic rates of increase (r) were 0.033, 0.043, and 0.062 day 1. According to data analysis, Nabis pseudoferus had a net predation rate (C0) of 1347.95 for cotton aphid meals and 2259.62 for diets that included both cotton aphid and Ephstia kuehniella eggs. Furthermore, after being kept in a cold environment at 5 °C for 7, 14, and 21 days, the hatching rates of Nabis pseudoferus eggs were 72.22, 69.23, and 69.23%, respectively. These findings may help to improve the way Nabis pseudoferus is raised (Ahmadi and Madadi 2021).

15.2.3 Lygeidae On ant pupae, Oecophylla smaragdina, a potential food source that is both naturally occurring and commercially available, some of the life parameters of Geocoris ochropterus have been studied, including nymphal development periods, growth (by weight), food consumption, egg-laying period, rate, and percent mortality. To understand the nutritional value of the meal for mass rearing, the moisture content

15.2

Hemipteran Predators

535

and essential nutrients, such as protein, carbohydrate, and lipid, of the coldpreserved ant pupae have been determined. The results of an experiment with variable male company revealed that the total fecundity outperformed the fortnightly mating programme in constant male company by a small margin. Virgins lay fewer eggs when they do not mate (Mukhopadhyay and Sannigrahi 1993).

15.2.3.1 Artificial Diet for Lygeidae A known key predator in numerous agro-ecosystern species is Geocoris punctipes (Crocker and Whitcomb 1980). Numerous studies have focussed on the dietetics, nutrition, and rearing techniques for this predator. However, the described artificial diets for Geocoris punctipes do not support the growth of continuous generations. In order to establish continuous generations of artificial meals that more nearly approximated the chemical makeup of insect eggs, I focussed my efforts after several years of trying to augment basic casein and yeast hydrolysate diets. To raise Geocoris punctipes, an artificial diet made entirely of beef was created. There have been five full continuous generations thus far. Sixth-generation eggs have been laid. Biomass increased starting with the fifth generation. There are more Geocoris punctipes now. For raising Geocoris punctipes, an artificial diet made entirely of beef was created by Cohen (1985). This diet has produced five full generations in a continuous line thus far, at a cost of about $0.63 per 1000 persons. Sixth-generation eggs have been laid. The culture’s biomass increased from about 1.0 mg for the initial instars to almost 1200 mg by the fifth generation. The fact that the population of Geocoris punctipes has grown with each new generation suggests that the current diet and farming practises would be suitable for mass-producing this key predator. Igarashi and Nomura (2013) created two types of artificial meals to raise the polyphagous predator big-eyed insect Geocoris varius (Hemiptera: Geocoridae). However, because Ephestia kuehniella (Lepidoptera; Pyralidae) eggs are employed as the main food supply for mass rearing, Geocoris varius production costs are relatively high. Therefore, the growth rate and reproductive capacity of Geocoris varius fed two different artificial diets based on ground pig and liver were studied. The food was either presented lyophilised or wrapped in Parafilm® to administer the diets. Results from Geocoris varius raised on Ephestia kuehniella eggs were compared to those from the experimental method. Although Geocoris varius-fed artificial meals developed more slowly than Geocoris varius-fed Ephestia kuehniella eggs, these delays are minor and unimportant for mass-rearing programmes. The results suggested that mass rearing of Geocoris varius (Igarashi and Nomura 2013) may be done at a lower cost by using both artificial diets. Additionally, dry power was used to raise Geocoris sp. Igarashi et al. (2013) investigated whether applying a powdered artificial feed to target plants will encourage Geocoris varius colonisation (Geocoridae). According to laboratory tests, 93% of the Geocoris varius individuals were still present on the treated strawberry plants after 72 h, compared to only 30% of the Geocoris varius on the untreated strawberry plants. Additionally, using the powdered food helped Geocoris varius develop from the third to the fifth nymphal stages and increased its average lifespan by 12 days. The findings suggested that even when pest densities are low, the application of

536

15

Mass Production of Insect Predators

powdered artificial feed encourages the development of Geocoris varius released in a greenhouse, enhancing the effectiveness of biological management. Geocoris pallidipennis is mass-raised on artificial feeds. By using a microencapsulated artificial feed, Khanzada et al. (2019) attempted to optimise the mass raising of Geocoris pallidipennis. Insect body, Rhopalosiphum maidis homogenate; micronutrient (fish egg) homogenates; nectar source (pure honey and corn pollen); and nutritional supplements, soluble L-Tyrosine and yeast extract, were the four types of elements we used for the diet compositions. To determine the substances’ most potent components, we tested 25 combinations of the compounds using an orthogonal test. We then screened the top five combinations. The findings demonstrated that numerous biological features of Geocoris pallidipennis were considerably altered by the various substances and their combinations. Only five component combinations were necessary for development to sexual maturity (Khanzada et al. 2019).

15.2.4 Reduviids Ambrose looked at how substrate effect affected the biology of Rhynocoris kumarii in 2000. The findings demonstrate that the stadial periods of Rhynocoris kumarii nymphal instars reared on dry plant litter and green plant shoot were, respectively, 58.5 and 51.46 days and significantly shorter when compared to those reared on tissue and glutting papers (61.5 days), plastic substrate (66.8 days), and sand with stone substrates (65.4 days). In comparison to other substrata like plastic (34.8%), sand with stone (46.6%), tissue and glutting papers (53.7%), and dry plant litter (57.6%), predator survival was higher on green plant shoots (61.4%). On the dry plant litter and green plant shoot substrata, the preoviposition duration was decreased to 20.8 days and 21.4 days, respectively. When Rhynocoris kumarii was raised on four different substrata, including sand and stone, dry plant litter, tissue and glutting paper, and green plant shoots, fresh adult body weights were slightly higher (128.1, 133.2, 138.7, and 142.5 mg, respectively) than when they were raised on untreated plastic substrata (120.4 mg). Fecundity was much greater in the substrate category for green plant shoots (158.9). During the mass upbringing of reduviids, George and Ambrose (2000) suggested taking nymphal cannibalism into account. George (2000) also claims that cannibalism results from a lack of prey in addition to nymphal cannibalism. In mass production, Sahayaraj et al. (2003) recommended taking the gender ratio into account. Grundy et al. (2000) present a containerised mass-rearing technique for Pristhesancus plagipennis in Australia utilising larvae of Tenebrio molitor and Helicoverpa armigera that have been killed in hot water. Additionally, it was looked into how Pristhesancus plagipennis used density throughout the rearing of nymphs and adult oviposition. The raising approach is labour and space efficient, decreases Pristhesancus plagipennisusing cannibalism, and does not require live insect prey. The most advantageous prey for reducing nymphal development time and mortality while yielding insects with the largest body weight were the larvae of the yellow

15.2

Hemipteran Predators

537

mealworm, Tenebrio molitor. The ideal rearing density for Tenebrio molitor-fed larvae was 20–27 nymphs per 5-liter container. With this rearing density, nymphal mortality was reduced to 16–22%. The best balance between egg production and space usage was achieved with 16 adults in a 5-liter container, which was the optimal density for oviposition (Grundy et al. 2000). In pass production, rearing has been done on a large and small scale on Corcyra cephalonica larvae, Rhynocoris marginatus was raised at four different densities (25, 50, 75, and 100). For small-scale laboratory rearing of this predator, 50 predators per rearing container were the ideal density since it resulted in shortened nymphal developmental, preoviposition, and incubation durations, female-biased sex ratios, minimum food requirements, and maximum fecundity and hatchability. The predator density was taken into account when creating a life table. The net reproduction rate was higher in the group of 50 predators, and it fell as predator density rose. The population doubling period of Rhynocoris marginatus (14.9 days) was discovered to be longer in the 100 predator category and to rise as predator density fell (Sahayaraj 2002). A methodology for adult group raising of the predator Rhynocoris fuscipes inside microenvironmental cages (MECs) was devised by Tomson et al. (2017). Results show that Rhynocoris fuscipes observed shorter nymphal developmental period, higher fertility, and longer adult longevity in the MECs than under normal growth circumstances. In the MECs, the reduviid predator liked to conceal itself under stones and fallen leaves (Tomson et al. 2017). The following are some of the disadvantages of using traditional techniques and diets for insect rearing: 1. Reduviids cannot be mass-multiplicated on a commercial scale using the larval card approach. 2. The correct card preparation is the only factor that determines whether the larval card approach succeeds. If the card is improperly prepared, the Cocryra larvae may develop into pupa, which reduviid predators cannot consume, or they may escape, which would make it difficult for them to feed and result in their demise. 3. In order to feed the reduviid predators, these traditional approaches call for the collection of prey insects from various agro-ecosystems. 4. These traditional techniques cannot be used to feed on target insects that have any offensive or defensive mechanisms. 5. To provide a continuous supply of food for the gathered pests, these traditional methods may require that they be raised on their native host plants. Given the foregoing, it is necessary to create an artificial diet for reduviid rearing that is affordable, nourishing, and efficient in producing reduviids in large quantities with predatory efficacy.

15.2.4.1 Artificial Diet Taylor and Schmidt (1996) took the initial step by feeding first-instar spined assassin bugs Sinea diadema water or glucose solutions, which considerably postponed the commencement of conspecific predation. The holidic basis of the meridic diet

538

15

Mass Production of Insect Predators

contains one or more unprocessed or chemically unidentified ingredients (e.g., yeast, liver extract etc.). The results of earlier research by Sahayaraj and Balasubramanian (2016) and Sahayaraj and Sujatha (2011) were mostly consistent with the influence of variance in individual dietary parameters on growth of Rhynocoris marginatus nymphs. Meat pig liver can be used as the major food source for the meridic diet (MD), and many chemical ingredients can be added to the rearing medium to reduce costs. In this study, we developed an oligidic diet that is primarily composed of meat. The liver, blood, and/or serum of the meat were employed as the ingredients’ sources. The meat-based diet examined in this study had a good nymphal survival rate and was able to support Rhynocoris marginatus for numerous generations without providing any insect food. Sahayaraj et al. (2006) created an oligidic diet for growing this bug for the first time based on a comparison of its real and made-up preys.

15.2.4.2 Consistency of Artificial Diet In general, artificial meals were produced and used for hemipteran predator rearing in liquid, semi-liquid, gel, and solid form. Rhynocoris marginatus is categorised as a liquid feeder by the Hemiptera. Over the course of five decades, reduviid biologists have raised reduviid predators on a range of wild and laboratory hosts. Additionally, reduviid predators have rostrums that are piercing and sucking in nature. Long tubes with a little hole at the front make up the rostrum. As a result, the liquid fluid enters the tube easily, travels down the oesophagus, and eventually arrives at the crop. These insects, along with numerous other predatory hemipterans, attack solid meals, which they liquefy using digestive enzymes and specific mechanical action (Sahayaraj and Muthukumar 2011; Kumar et al. 2012). However, a solid meal is also inappropriate for raising predators with sucking mouthparts since it is quickly polluted and dried up (Tan et al. 2013). As a result, we were forced to create the liquid diet that is described in the pre-eminent section. To try to ascertain the various requirements and specify the meal composition for entomophagy’s, various analytical methodologies were used. Although respiration is not the main factor, diet presentation is a crucial factor for predators. Earlier research from our lab established the ideal ratios of the primary components for a meridic diet based on its impact on Rhynocris kumarii growth and reproductive ability (Sahayaraj et al. 2006). Rhynocris kumarii could easily consume a promoted liquid diet, but due to the liquid feed’s nutritional inadequacy, there was a significant rate of predator death. Additionally, the liquid versions of the previously tried artificial diet did not satisfy the requirements for reduviid development and reproduction. We created a meridic diet utilising cotton balls for the mass rearing of Rhynocoris marginatus based on these findings. Previous research demonstrated that the growth of Reduviids in the artificial diet was significantly influenced by meat-based ingredients like beef and hog liver (Sahayaraj et al. 2006). Similar to this, it was known that Reduviids metabolised Protein X, Pork Blood, and Meridian Egg Yolk from artificial meals utilised in their mass rearing for nutrition and energy (Sahayaraj et al. 2006). Chemical compounds like sodium chloride and potassium chloride were also added to the diet as essential nutrients. The current

15.2

Hemipteran Predators

539

Table 15.2 Diets 1–4 in various combinations, either with the insects Spodoptera litura or Bombyx mori Ingredients Insect whole body Beef liver (g) Pork liver Milk powder (g) Honey (ml) Acetic acid 10% (ml) Ascorbic acid (mg) Egg yolk (g) Sucrose (mg) Vitamin complex (mg) Wesson’s salt (mg) Streptomycin (mg) NaCl (mg) KCl (mg) Protein X (mg) Water (ml) Table 15.3 Diets 5 and 6 with a variety of foods and varying effects

Diet 1 Spodoptera litura 3.0 – 2.0 15 2.7 – 2 200 40 – 25 5 5 250 100

Diet 2 Bombyx mori 3.0 – 2.0 15 3.7 – 2 200 40 – 25 5 5 250 100

Ingredients Beef liver (g) Honey (ml) Acetic acid -10% (ml) Egg yolk (g) Sucrose (mg) Vitamin complex (mg) Streptomycin (mg) NaCl (mg) KCl (mg) Water (ml)

Diet 3 Bombyx mori 10.0 – 2.0

Diet 4 Bombyx mori – 10.0 2.0

10 20 200

10 20 200

20

20

100

100

Diet 5 30.0 2.5 1.85 2.0 100 5.0 3.75 2.15 2.15 100

Diet 6 60.0 5.0 3.7 4.0 200 10.0 7.5 5.0 5.0 100

chapter aims to establish the ideal meridian diet for Rhynocris marginatus according to the predator’s developmental and reproductive condition. We also looked into the impact of diet on natural prey as compared to when Rhynocris marginatus was fed on natural prey. In 2002, we created two meridic diets—diets 1 and 2—based on the whole-body diets of Bombyx mori and Spodoptera litura, respectively, and observed how Rhynocoris marginatus fed on them. Below is a list of the ingredients for diets 1–4 (Table 15.2). We substituted Wesson’s salt mixture and ascorbic acid for the insect components, milk powder, Protein X, and individual salts in the diet. The following components are included in the diets 3 and 4 in varying amounts and produced Diet 5 and Diet 6 (Table 15.3):

540

15

Mass Production of Insect Predators

Diets 3 and 4’s complete omission of Milk Powder and Protein X also applied to diets 5 and 6. Due to the addition of the beef liver as the only component, the quantities of the other components were also decreased. Rhynocoris marginatus can grow into nymphal instars on these diets, but reproduction is not supported. Diet 5 contains 0.46 mg/ml of total carbohydrates, 0.34 mg/ml of total protein, and 0.23 mg/ml of total lipid, according to biochemical analysis of the diet. Diet 6 has higher total fat (0.29 mg/ml), total protein (0.46 mg/ml), and total carbohydrate (0.6 mg/ml) concentrations (Table 15.3). When formaldehyde was absent from the diet, diet contaminations were frequent, according to Sahayaraj and Balasubramanian (2016) research. After the formaldehyde was added, over 70% of diet contaminations were avoided. Additionally, it was indirectly protecting Rhynocoris marginatus nymphs and adults from Aspergillus flavus infection. Additionally, eliminating yeast from diets 5 and 6 may help stop contamination and infection. Our findings also demonstrated that yeast extract supplies a significant amount of these components, as their addition to the base diet fosters the growth and reproduction of predators (diet 5). Rhynocoris marginatus emergences as adults were significantly increased by the yeast extract compared to the liver diet (Xie et al. 1997). Insects grown on a meat diet were smaller, lighter in weight, and had longer embryonic and nymphal development times than those grown on a control diet. Grenier and Clercq (2003) and Castañé and Zapata (2005) both report and explore this prevalent trait (2003). In keeping with our experience from 2003, we created a set of diets [diet 7–diet 12] in 2006 that included cow liver in addition to pork blood and pork blood serum (Table 15.4). The meatbased meridic diet utilised in this study was made using a modified version of the technique employed by Sahayaraj et al. (2006). Table 15.4 lists the origin and additional elements of the oligidic diets. A nitrogenous source (protein/free amino acids), fats, carbohydrates, vitamins (Vitamins B and C, Vitamin A), minerals, preservatives, and antimicrobial protection agents are all components of artificial diets. In the cells and bodily fluids (hemolymph, salivary secretion) of insects, Na and K play a role in pH regulation. We added minerals to the food since insects cannot produce them. All living things need chloride, but insects especially need it for enzyme activities and the potential of their membranes. Additionally, minerals stimulate appetite just like other nutrients like sugar, certain amino acids, lipids, and ascorbic acids. The nutritional value is also improved by the use of mineral salts. As indicated by Hou and Hsiao (1979), we employed streptomycin as an antibacterial agent in the food because it has been used as an antibiotic in people. It holds out a lot of hope for antimicrobial therapy for insects as well.

15.2.4.3 Preparation Pig blood and liver were used as source ingredients, and they were dried in a hot air oven for 25–30 min at 60 °C. They were well pounded using a mortar and pestle before being placed in the refrigerator for usage within a month. 100 °C was used to boil 100 ml of pure water for 20 min. Then, 10 ml of hot water was added to the milk powder (Lactogen, Nestle, Mumbai, India), and the mixture was allowed to cool.

15.2

Hemipteran Predators

541

Table 15.4 Composition of numerous oligidic diet substances, including source ingredients prepared for 100 ml and varied potents of various chemicals Components (in mg/ml) Source ingredient (g) Pork blood (g) Blood serum (ml) Sucrose (mg) Yeast extract (water soluble) Yeast (g) Milk powder (g) Egg yolk (g) Honey (ml) Vitamin (multivit mg) Vitamin E Vitamin C Cacein (mg) Cholesterol (mg) Acetic acid (10%) (ml) Nacl (mg) Streptromycin (mg) Formaldehyde 40% (ml)

Diets Diet 7 60 – – 200 – – – 5.0 10 – – – – 3.7 5.0 7.5 –

Diet 8 5 – – 200 – – – 4.0 5 200 – – – – – 5.0 100 –

Diet 9 5 – – – – 2 5 4.0 5 200 – – 5 200 – 5.0 100 –

Diet 10 5 5 – 500 – 2 5 4.0 5.0 200 200 2.5 4 200 2.5 – 100 –

Diet 11 5 5 20 500 5 – 5 5 5.0 200 200 2.5 4 200 2.5 – – 1

Diet 12 5 – 20 500 5 – 5 5 5.0 300 200 2.5 4 200 2.5 – – 1

The remaining 90 ml of water was mixed with the appropriate amount of acetic acid (Glaxo, Gujarat, India), liquid honey (Dabur Narendrapur, West Bengal, India), dried egg yolk (Fine chemicals Ltd, Mumbai), and water-soluble yeast extract (Fine chemicals Ltd, Mumbai). After 10 min, the temperature was lowered to 40 ° C, and the source components—multivitamin, vitamins C and E, and streptomycin—were added and thoroughly mixed. These ingredients were provided by Sarabairaman Ltd. in Vadodara, India. The milk powder mixture was then added and thoroughly mixed. The prepared diet was well blended, then allowed to cool at room temperature before being filtered through Whatman No. 1 filter paper to create a liquid diet. Filtered liquid diet was kept in the fridge in 125 ml reagent bottles for use over the course of more than 2–3 weeks. We kept the diet in the refrigerator for the sake of convenience, figuring that storage would not appreciably deteriorate the diet’s components. Additionally, it was discovered that during the course of our investigation, ingredients and their diets remained unchanged after storage. However, the surplus yeast that was added to the diet after it had been stored for a few weeks at a time gave it a distinctive fragrance. We have created more than 17 diets so far through trial and error. However, we identified the components of meridic diets that are advantageous for Rhynocoris marginatus development and fecundity. Pork liver is a source ingredient in each of the aforementioned diets. There were other components, but they were mixed in at various ratios and concentrations. Sahayaraj and Balasubramanian (2016)

542

15

Mass Production of Insect Predators

highlighted the benefits and restrictions of reduviid predator artificial rearing in their book, “Artificial raising of reduviid predators for pest management.”

15.2.4.4 Object Preference This research is very helpful in choosing the right thing to feed the predator’s artificial diet by identifying which object the predator is able to locate. Four items, including cotton, a paraffin capsule, commercial foam, and a cavity micro slide, were initially tried. Preference was based on the percentage of predators supplying the food with a certain object, approaching time, and consumption time. To begin with, we experimented with four items (cotton, capsule, cavity slide, and foam) for giving Rhynocoris marginatus life stages OD. Adult predators (50.5%), the first stadium (48.5%), the fifth stadium (33.3%), and chose to feed the artificial food supplied in cotton. Additionally, it took adults 3.3 and 7.1 min to approach and eat the food found in cotton, compared to 5.1 and 7.2 min for the first and fifth nymphal stages and 4.9 and 3.13 min for adults. Individually or in groups, all life stages can easily consume the artificial diet included in the cotton (Fig. 15.1). Since they liked cotton, the reduviid implanted its rostrum with ease and sucked the artificial diet without any restrictions. In comparison to the other three tested things, it was the cheapest further coastward. Another item that reduviids favour is the parafilm capsule (33.3%, 23.3%, and 25.5% for the first stadium, fifth stadium, and adults, respectively) (Fig. 15.2). The feeding schedule for Rhynocoris marginatus varies with its life stages (2.9, 1.7, and 2.4 min for first stadium, fifth stadium and adults, respectively). Reduviid feeds alone or in groups as it follows when it encounters an insect prey, as evidenced by the observation that the para film crumbled after some time of feeding. Furthermore, the colour of the parafilm has no impact on how time is captured. 15.2.4.5 Feeding Behaviour on Meridic Diet While the reduviids were receiving a meridic diet, the following feeding patterns were observed: Reduviids can be observed in five different ways: (1) moving forward across the feeding area, (2) standing still, (3) rapidly moving their legs across their body

Fig. 15.1 Reduviid predator feeding behaviours, adults of Rhynocoris marginatus individually feeding artificial diet (beginning (a) and final feedings (b)), as well as group feeding in red-coloured parafilm (c) and red and black-coloured parafilm (d)

15.2

Hemipteran Predators

543

Fig. 15.2 Rhynocoris marginatus nymph (a) and adults ((b) and (c) in group), Rhynocoris fuscipes ((d) approaching and (e) feeding), Rhynocoris longiferons (f), Panthous bimaculatus adult (g) feeding on artificial diet given in cotton

surface and antennae, (4) inserting their mouthparts into the diet packet but not feeding for very long, and (5) inserting their mouthparts into the diet packet and feeding for a very long time. The reduviid predators frequently engage in the behaviours indicated above. The reduviids were drawn to eat on the meridic food by the signalling compounds as well as other factors like texture and consistency. Reduviids were the subject of preliminary behavioural experiments to better understand how they approached the meridic diets. Rhynocoris marginatus adults, nymphs, and laboratoryemerged adults were all used in the study. In general, the predator’s approaching time got shorter as it got older. For instance, feeding Rhynocoris marginatus artificial food based on Bombyx mori required only 1.4 min in the first stadium and 5.4 min in the fifth stadium. Rhynocoris marginatus tries to bury its rostrum further into the cotton during feeding (the feeding arena). It was not accurate for all of the diets examined here, though. Another finding we made was that adults consumed the artificial diet in less time than nymphs, which indicates that adults have more metabolic activity than nymphs. Protein-X was added to the pork liver-based diet (diet 3) to raise the fake diet’s protein content. Fortification was carried out at four different concentrations, namely 5, 10, 15, and 20%. We found that adding 5% protein-X (w/v) to the

544

15

Mass Production of Insect Predators

diet increased Rhynocoris marginatus nymphs’ feeding time (16.59 and 22.24 min for the first and second stadiums, respectively) compared to adults (16.24 min). We evaluated water against the fake diet since water is a crucial part of the meridian diet. According to the findings, Rhynocoris marginatus drank more artificial food than water. It backs with Edney (2012) assertion that many insects ingest water to survive. Water intake and artificial diets 1, 2, and 3 were correlated, with the correlation coefficients for these diets being extremely positive.

15.2.4.6 Feeding Behavioral on Meridic Diet in Relation to Starvation When given an artificial diet, starvation affects the approaching time, sucking time, and weight increase of Rhynocoris marginatus male and female. As the predator grew older, animal weight gradually decreased (Table 15.5). Males started the diet quicker than females did. Females consumed more food at 24 and 72 h of famine because they took longer to eat. But after 48 h of starvation, this predator proved that this was untrue (Table 15.5). Instead of 10 g of beef liver, in diet 5 and 6, 30 and 60 g of beef liver was incorporated respectively and recorded the feeding activity of Rhynocoris marginatus nymphs and adults (Sahayaraj and Sujatha 2011) in relation to 24, 48, 72, and 96 h starvations. In second, third, fourth and fifth stadium, the consumption time gradually increased while the food deprivation period prolonged from 24 h to 96 h both at 30 and 60 g beef liver added diets (Table 15.6). Calculated Access Proportion Index (API) showed that it was higher (API = 0.8–1.0) both in first and second stadium than in third, fourth, and fifth stadium (API = 0.6–1.0). According to Hill (1989), sugar is a crucial component that encourages the development of eggs. Similar recommendations for yeast and sugar for optimum egg production were made by McEwen and Kidd (1995). Additionally, honey is a crucial element in fertility. In order to produce viable eggs in adults, McEwen and Kidd (1995) and Kubota and Shiga (1995) conducted analyses on a honey and yeast autolysate mixture. Rhynocoris marginatus was able to breed from diets 4–6, therefore the lack of reproduction in other diets can’t be attributed to the honey. The artificial diet of milk, eggs, sweets, and yeast used by Sahayaraj et al. (2006) to grow Rhynocoris marginatus adults was found to be favourable for fecundity. According to Norioka et al. (1984) increased fecundity was seen in diets with higher egg yolk content (15.5% amino acids). Additionally, Rhynocoris marginatus Table 15.5 Feeding behaviour of field-collected Rhynocoris marginatus adults on artificial diets Sex Female Male Female Male Female Male

Starvation (h) 24 48 78

Approaching time (min) 1.2 5.6 12.7 8.5 5.5 2.7

Sucking time (min) 59.6 6.5 11.8 12.3 26.1 20.8

Weight gain (mg) 30.6 8.1 24.2 6.3 21.0 4.7

15.2

Hemipteran Predators

545

Table 15.6 Feeding time (min) of Rhynocoris margiantus life stages starved for 24, 48, 72, and 96 h then provided with artificial diet prepared using 30 and 60 g beef liver as source ingredients Life stages First stadium Second stadium Third stadium Fourth stadium Fifth stadium Adult

Beef-liver quantity (g) 30 60 30 60 30 60 30 60 30 60 30 60

Starvation time (h) 24 48 17.7 2.8 6.1 6.2 2.6 4.4 2.6 4.0 1.5 3.7 1.1 4.7 2.3 3.3 0.6 2.6 2.3 6.3 2.8 4.8 – – – –

72 7.6 4.4 8.8 6.3 7.0 5.0 5.3 3.6 7.7 5.0 6.1 5.2

96 12.1 6.3 10.5 14.1 10.1 15.6 14.5 6.8 7.9 5.8 – –

– indicates no observation was recorded

multiplied in the presence of blood serum, casein, cholesterol, acetic acid, vitamin E, and vitamin C. We therefore postulated that each of these components is crucial for this reduviid’s development and reproduction. Choice and Non-choice Test Against Artificial Diet Rhynocoris marginatus feeding behaviour was assessed using both the choice test and the non-choice test in order to determine the acceptability of the oligidic diets. In the olfactometer, water and one to three concentrations of the simulated diets (1–3) were used for the laboratory tests. Adult Rhynocoris marginatus species were then added to the glass olfactometer after being deprived for 24 h. For each experiment, fifteen replications were created. In each trial, the approaching time (AT), consumption time (CT), and reduviid feeding preference were constantly monitored for an hour using a visual method. All life stages consistently preferred the meal based on beef liver (diet 3), then the diet based on Spodoptera litura (diet 1). This was calculated based on the percentage of time the predator spent feeding its prey. Although diets 1 and 2 comprise insect diet, the predator favoured the meal that had cow liver as a source item. It is encouraging to see that using vertebrate body parts as a source material makes it possible to produce reduviid predators in large quantities. On the basis of these discoveries, we continued our research. Best Strain Selection Rhynocoris marginatus nymphs and adults that were taken from the stock culture in the lab were fed an artificial diet and given laboratory hosts. Artificial diets were favoured by more than 72% of adults, followed by the fifth stadium, the fourth stadium, the third stadium, the second stadium, and the first stadium. In a different study, diets containing 30 g of beef liver and 60 g of beef liver, respectively, were

546

15

Mass Production of Insect Predators

Fig. 15.3 Female Rhynocoris marginatus adult feeding adult Dysdercus cigulatus

given to Rhynocoris marginatus nymphs and adults as opposed to Corcyra cephalonica. All other life stages except the second stadium highly selected diet 5 (Sahayaraj and Sujatha 2011). Stage Preference of Reduviid Fed with Artificial Diet By using a choice experiment as suggested by Holling (1966), a stage preference research of Rhynocoris marginatus with various life stages of Dysdercus cingulatus, Spodoptera litura, and Corcyra cephalonica independently was conducted. First, second, third, fourth, and fifth instar stages of Dysdercus cingulatus were placed into a Petri dish along with Rhynocoris marginatus, and the predatory behaviour was visually recorded continuously for 6 h. Stages of successfully obtained, dispatched, and devoured prey were noted as the reduviid’s favourite stages. Predators for the third, fourth, and fifth nymphal stages as well as adults of Dysdercus cingulatus were available. For both Corcyra cephalonica and Spodoptera litura, all five of the prey’s nymphal instars were given to Rhynocoris marginatus at different life stages. For the predator’s various life stages, ten replications were kept. The studies evaluating the biological control potential employed the pests’ preferred life stages. Within a patch, Reduviid encountered hosts at various stages of development. The profitability of those possible attack-prone stages may vary. The stage preference research for Rhynocoris marginatus in relation to Spodoptera litura, Corcyra cephalonica, and Dysdercus cingulatus life phases. Rhynocoris marginatus life stages favoured various stages of the pests investigated, according to studies on prey stage choice. According to the results, adult and fifth instar predators had a better chance of catching the large-sized Spodoptera litura (iv and v instars) and Corcyra cephalonica (v instars) larvae. Rhynocoris marginatus liked diverse nymphal instars of lepidoptera larvae, while second and third instar reduviid preferred second instar Dysdercus cingulatus nymphs, and the remaining reduviid life stages chose only adult Dysdercus cingulatus (Fig. 15.3). Food Preference Index (FPI) Additional combinations of various items were added to the artificial diet, including oligidic diet (T3), oligidic diet + Corcyra cephalonica (T4), and Corcyra cephalonica + water (T2). One millilitre of OD was given every 2 days. To maintain

15.2

Hemipteran Predators

547

hygienic conditions, the cotton ball was removed and thrown away after 2 days. Additionally, the 100- and 500-ml size containers were cleaned with 0.2% sodium hypochlorite before the insect was placed inside. From each treatment, 30 male and female Rhynocoris marginatus were randomly chosen, weighed, and placed in an olfactometer covered with muslin material. Among the insects, females took longer to sucke (59.6 min) and gained greater weight (30.6 mg) in all of the testes. Feeding Behavior with Live Preys The potential for biocontrol was assessed using the Rhynocoris marginatus life stages that arose in the lab, with the exception of the first instar. Dysdercus cingulatus preferred stages (5/container) were added to the container containing cotton twig and let 10 min to acclimate. The feeding events, such as approaching time, handling time, and the locations selected by the reduviid for feeding, were then continually monitored for 6 h while a Rhynocoris marginatus was kept in the same container. A predator’s weight gain and the number of prey items ingested were noted after 24 h. For the predator’s various life stages, ten replications were kept. Approaching Time (AT) To research the feeding habits on oligidic diets, fieldcollected reduviids were used. Field-collected Rhynocoris marginatus adults gravitated for the oligidic diets after 24 h of fasting. Adults who had been recruited from the field 48 h earlier showed 30% reactions to milk-based oligidic diets, which escalated to 60% within an hour. 70% of the predators approached the diet when their hunger levels reached 72 h. Reduviids failed to approach the diet after 2 h when 5% insect sources such Mylabris indica, M. pustulata, and Dysdercus cingulatus were added to the milk-based diet. Antenna stretching, cat pacing in the direction of the food, restless activity leading to flight antenna brushing, leg brushing, and repeated movement in the direction of the food source were all seen during feeding. However, the predators were unable to use an insect source to approach the cotton that was milk-based in their diet. Table 15.5 further showed that as the fasting period was extended from 24 to 72 h, Rhynocoris marginatus food consumption gradually reduced. Similar reductions were made in approaching and sucking times. In comparison to men, women eat more milk-based oligidic food. Significantly more time was spent sucking the diet by women than by men. According to Sahayaraj and Balasubramanian (2016), the following sequence of predatory behaviours occurs: (a) Arousal: the predator becomes restless and straightens its legs while intently observing the prey. Approach: the predator advances towards the prey while pointing its rostrum and antennae forward. (b) Capture: the predator closes up on the prey, captures it, and holds it firmly by its forelegs, adjusting its pace of movement to that of the prey. (c) Rostral probing: the predator uses its rostrum to investigate various regions of the prey’s body,

548

15

Mass Production of Insect Predators

Table 15.7 Site preference of Rhynocoris marginatus during feeding on three pests Predator life stages II III IV V Male Female

Dysdercus cingulatus Thoracic pleural membrane Eye, thoracic pleural membrane Thoracic and abdomens pleural membrane Thoracic and abdomens pleural membrane Head neck muscle, tergum Neck membrane, tergum

Spodoptera litura Thoracic pleural membrane, tergum Thorax pleural membrane, tergum Thoracic and abdomen pleural membrane Thoracic pleural membrane, tergum Thoracic and abdomens pleural membrane Thoracic pleural membrane, tergum

Corcyra cephalonica Sternum, neck membrane Neck membrane, sternum Neck and abdomens pleural membrane Neck and abdomens pleural membrane Neck and abdomens pleural membrane Neck and abdomens pleural membrane

(d) When a predator paralyses its prey, the victim loses agility and becomes motionless, suggesting that the paralysis has taken place. (e) Feeding: the predator carries the prey to a remote location and extracts the body’s contents by sticking its rostrum into one or more locations. and (f) Post-predatory behaviour: The predator cleans its rostrum, antennae, and forelegs before dropping the carcase. The time spent by the predators on each of the first three events (1–3) was totalled and used as a general indicator of their success in finding and seizing the prey (termed approaching time, AT). Similar to this, the length of time the predators spent handling and consuming the prey for each of the following three episodes (4–6) was added up (termed handling time, HT). Our research shows that Rhynocoris marginatus faced its antenna towards the prey during feeding time. After establishing this ideal orientation position, the reduviid aroused and displayed the approach rostral probing, injection of toxic saliva for paralysing, sucking the prey content, and post-predatory behaviour that was observed in this study. Additionally, Rhynocoris kumarii, P. bimaculatus, and S. collaris life stages fed with fake diets based on meat displayed the approaching behaviour. Results indicate that reduviids approach the diet with their antennae forward (Sahayaraj 2012), suggesting that they recognise it visually and chemically, as was the case for Rhynocoris marginatus (Sahayaraj et al. 2006) when it was being fed. Catching the Prey When the victim was within striking distance, the predator extended its rostrum and caught the animal, ideally in the abdomen. When the prey is little and less active, the predator will firmly keep its forelegs over the prey to trap it. The predator would elevate its antennae, stretch its rostrum, and pin the prey at a chosen location if the prey was agile (Table 15.7). The predator was observed using its antennae to probe the immobile prey, possibly testing the prey with its inserted rostrum. However, the ensuing predatory behaviours, including

15.2

Hemipteran Predators

549

pinning and paralysis, were unaffected by the acts. To determine the best location for sucking, the extended rostrum was put into the trapped prey. When compared to predators who were starved of prey, nourished predators took longer to catch and pin victims. Paralysing After successfully capturing the victim, the predator used the prey’s toxic salivary fluid to immobilise it. When compared to predators that were starved of food, the fed predators took longer to paralyse the prey. Rostral Probing and Sucking After paralysing the prey, the predator drained the predigested body fluids from the prey by inserting the rostrum at various locations on the prey’s body. During eating, Rhynocoris marginatus regularly switched where it sucked. One of the antennae was maintained facing the prey in an upright position while the other was kept facing the prey in a drooping yet extended position, presumably to test different prey ratios. Rostral probing helped to further the goal. Predators frequently choose the abdomen region as the site for sucking rather than the cephalic and thoracic regions (Table 15.7). The predator’s forelegs were discovered on the prey as it was sucking. Predator took shorter time sucking after eating. The consumption of available artificial diet may be the cause of the lower prey ingestion. Post-Feeding Behaviour Satisfied, the predator began cleaning the prey’s rostrum and antennae, drawing and grooming the area in between the forelegs, and then washing the antenna, fore tibial pad, and hind tibia. Feeding Habits in Opposition Against Three Pests Table 15.7 displays the first Rhynocoris marginatus attack site on three pests. Table 15.8 made it abundantly evident that different life stages of reduviids favoured a specific location for paralysing and sucking the prey’s contents. Reduviid considerably favoured the thoracic pleural membrane of the pests, followed by the abdominal pleural membrane, for both paralysing and eating the victim. From the second instar to the fifth nymphal stage, the approaching time generally decreased (7.0, 2.2, and 2.4 min for Corcyra cephalonica, Spodoptera litura, and Dysdercus cingulatus, respectively). Rhynocoris marginatus, a predator raised on a meridional diet, took longer to approach its prey Corcyra cephalonica and Dysdercus cingulatus than the predator raised on a natural host, Spodopter litura. Predator approaching times between those raised on an artificial diet and those raised on prey were not statistically different at the 5% level. Reduviid raised on an artificial diet came closer to Spodoptera litura than other prey did. Regardless of its prey, Rhynocoris marginatus raised on an oligidic diet managed its maximum time (270.5 min) substantially longer than a predator raised on insect hosts. The findings show a favourable relationship between handling time (250.1 min) and weight gain (48.1 mg). Maximum care was taken in the rearing of Rhynocoris marginatus second, third, fourth, and fifth instar nymphs and adults on insect hosts.

550

15

Mass Production of Insect Predators

Table 15.8 Feeding behaviour (min) and weight gain (mg.) of Corcyra cephalonica (CC) and artificial diet reared Rhynocoris marginatus on Dysdercus cingulatus life stages Instar Approching time Dysdercus cingulatus II 7.0 III 4.7 IV 4.4 V 1.7 Male 4.1 Female 5.0 Artificial diet II 7.3 III 7.3 IV 3.5 V 4.3 Male 4.3 Female 5.4

Handling time

Weight gain

No. of site selected

131.3 167.2 199.5 141.6 128.5 138.7

08.1 18.1 31.0 41.5 25.6 35.2

2.0 2.3 2.4 2.6 1.8 1.6

88.0 259.5 282.8 214.5 144.2 125.0

1.9 11.1 15.8 31.2 35.2 25.6

11.6 3.2 2.2 2.3 3.2 2.6

Table 15.9 Approaching time, handling time (min) and weight gain (mg) of Rhynocoris marginatus life stages on Corcyra cophalonica and oligidic diet Predator Approaching time stage (min) C. cephalonica II 2.2 III 2.0 IV 1.5 V 1.4 Male 1.3 Female 1.6 Artificial diet II 4.6 III 4.7 IV 4.4 V 3.0 Male 4.6 Female 4.0

Handling time (min)

Weight gain (mg)

No. of site selected

130.6 120.0 134.8 157.2 165.0 163.3

7.3 11.7 11.4 22.9 34.3 33.2

1.3 1.1 2.1 2.2 1.8 1.6

227.5 215.0 250.1 270.5 159.0 198.0

8.0 20.5 48.1 43.9 34.6 53.0

1.5 1.8 1.6 1.2 2.2 2.4

Spodoptera litura and Dysdercus cingulatus In third instar Rhynocoris marginatus feeding on Corcyra cephalonica, the shortest consumption time was noted (Table 15.9). Female Rhynocoris marginatus handled Corcyra cephalonica and Spodoptera litura more slowly than male. But when Rhynocoris marginatus was given Dysdercus cingulatus, male took less time (Table 15.8). Regardless of the preys, Rhynocoris marginatus gradually put on weight as it grew older (except adult) (Table 15.9). Table 15.8 show that Spodoptera litura (76.1 mg) was given to

15.2

Hemipteran Predators

551

Table 15.10 Approaching time, handling time (min) and weight gain (mg) of Rhynocoris marginatus life stages on Spodoptera litura and artificial diet Life Approaching time stage (min) Spodoptera litura II 24.3 III 19.5 IV 13.7 V 10.1 Male 11.2 Female 9.6 Artificial diet II 7.4 III 4.5 IV 3.3 V 3.8 Male 5.5 Female 4.5

Handling time (min)

Weight gain (mg)

No. of site selected

126.5 148.0 129.0 159.0 188.7 175.0

8.6 19.0 21.5 54.0 34.5 29.0

1.6 1.8 2.0 2.1 1.5 1.5

230.5 207.0 241.5 241.5 225.0 227.5

6.7 13.3 60.7 63.7 63.7 76.1

2.1 1.8 1.5 1.5 2.0 1.8

Rhynocoris marginatus raised on an artificial diet, which resulted in the adult female gaining the most weight. But when Dysdercus cingulatus was given to this reduviid predator, for instance, the reverse pattern was seen with other pests. Generally speaking, Rhynocoris marginatus raised on an artificial diet ingested more prey than those raised on Corcyra cephalonica and Spodoptera litura. Similar to how Spodoptera litura adult females were the target of the highest predatory rate in Rhynocoris marginatus raised on an artificial diet (2.40). It was extremely low in the third instar of Dysdercus cingulatus (1.36 prey/ predator/day). In this experiment, Rhynocoris marginatus nymphs and adults can eat more Spodoptera litura larvae (76.1 mg/predator) when the predator was raised on an oligidic diet (except from second instar larvae and adult females) (Table 15.10). Similar insect hosts included second and fifth nymphal stages of Rhynocoris marginatus, adults of Spodoptera litura in greater numbers, and second instars of Dysdercus cingulatus. However, Rhynocoris marginatus was raised on artificial feeds or insect hosts, and both life stages absorbed a minimal amount of Corcyra cephalonica. Since Spodoptera litura and Dysdercus cingulatus are found in crops, this reduviid can be utilised as a biological control agent. Additionally, weight gain was observed in predators raised on artificial food, Corcyra cephaonica larvae, and water. According to the findings, this reduviid needs artificial food or water to continue growth and reproduction. Thus, as the predator aged, weight gain rapidly increased. Reduviid was raised with an artificial diet in addition to receiving the lab host Corcyra cephalinica larvae once per week for three successive generations, and weight gain was monitored. Predator put on weight, showing that artificial nutrition supports development. However, in the

552

15

Mass Production of Insect Predators

second generation and subsequently in the third generation, the rate of weight gain somewhat decreased. Three reduviid predators, Sycanus collaris, Panthose bmaculatus, and Panthose bmaculatus, have been raised in lab settings for the first time utilising artificial diets based on meat. Rhynocoris kumarii (10 days) incubation period is less than that of Panthose bmaculatus (15 days) and Sycanus collaris (15 days) (21 days). Compared to other reduviids, Panthose bmaculatus substantially required longer time (101.1 days) and had a higher survival rate (62.9%) to complete the nymphal maturation phase (Sahayaraj 2012).

15.2.5 Miridae The generalist predator Dicyphus tamaninii has been raised on a diet that is primarily composed of meat (Miridae). The diet revealed overall, free amino acid, and lipid nutritional deficits. The predator was tested once more on a modified meal that contained fresh sources of fats and proteins. The reformed diet has enhanced some biological traits of bugs that were inferior in the initial meat diet as compared to those of the traditionally grown insects, such as nymphal development time and fresh weight (Zapata et al. 2005). On a bigger scale, the predatory mirids Macrolophus pygmaeus and Nesidiocoris tenuis (Hemiptera: Miridae) are employed (Calvo et al. 2012; Biondi et al. 2018). Egg development times for the Neotropical mirid species Campyloneuropsis infumatus, Engytatus varians, and Macrolophus basicornis were 9.9, 10.0, and 10.7 days, respectively, with survival rates of 95.7%, 92.3%, and 94.3%. (Van Lenteren et al. 2019). Effects of prey deprivation on the developmental characteristics of the omnivorous predator Dicyphus errans (Lepidoptera: Gelechiidae) (Hemiptera: Miridae). Five treatments of continuous inadequate prey provision during nymphal development and eight treatments of monthly prey provision in various nymphal instars were evaluated. While prey availability in late instars was connected to substantial adult weight gain, Dicyphus errans nymphs fed only in early instars were able to complete their development. The highest levels of prey availability resulted in female nymphs having very fast developmental times and very intense food consumption, which led to the greatest weight gain across all evaluated treatments. It’s interesting to note that Dicyphus errans nymphs fed at half their saturation level grew just like those who had access to a full saturation level. The findings are explored in light of the predatory bug’s coping mechanisms, with a focus on the implications for mass rearing and biological control. The findings could be used to identify and raise the best new prospects to challenge Tuta absoluta (Arvaniti et al. 2021).

15.2.6 Pentatomidae For Podisus maculiventris, the effects of an artificial food free of insects were assessed on the rate of development, life table parameters, and fertility table parameters. The findings showed that feeding an insect-free artificial diet during

15.2

Hemipteran Predators

553

both the nymphal and adult stages resulted in longer developmental times, longer preoviposition times, and significantly lower reproductive rates (R0) and intrinsic rates of increase (r) than feeding larval insect prey during both the nymphal and adult stages. The proportion of fertile females, the quantity of eggs laid by mated females, the reproductive rate, and the intrinsic rate of increase were all significantly increased when feeding larval prey to adults raised as nymphs on an artificial diet, but the mean generation time was not significantly altered. The reproductive rate and intrinsic rate of rise were also markedly reduced when adults raised on larval prey were fed artificial diets. Using the cost of raw materials, fertility table parameters, and doubling time values, the “realised” cost of rearing Podisus maculiventris on the artificial diet was estimated (as the cost to double the population size). The cost of raw materials for raising a colony of Podisus maculiventris on Trichoplusia ni (Lepidoptera: Noctuidae) was only 1.4 times higher than the cost of raw materials for artificial diets needed to raise a colony of the same size. However, owing to the delayed developmental period and decreased reproductive output when raised on an artificial diet, the actual cost of rearing was 3.5 times higher. As well as discussing the benefits and drawbacks of enhancing adult diets with natural prey during the reproductive stage, this article also discusses the economic effectiveness of rearing beneficial insects on an artificial diet that reduces the intrinsic rate of increase of a colony (Wittmeyer and Coudron 2001). Nymphal and adult Podisus maculiventris were fed natural larval prey (Trichoplusia ni) and an insect-free artificial diet (liver and egg-based). The findings demonstrated that females’ ovarian growth and reproductive production were significantly influenced by both nymphal and adult dietary sources. The quantity of vitellogenic and chorionated follicles present within the ovarioles indicates that the adult dietary source had a considerable impact on the process of vitellogenesis. Only in the adult stage, feeding on larval prey increased the pace of ovarian maturation and the quantity of vitellogenic follicles. However, nymphal diet continued to play a considerable influence in the reproductive success of married females. When adults raised as nymphs on larval prey were fed the artificial diet, ovarian maturation rate and fecundity were considerably lowered compared to females fed larval prey during both the nymphal and adult stages. Prey-fed females’ reproductive ability was unaffected by mating with diet-fed males, and diet-fed females’ reproductive capacity was unaffected by mating with prey-fed males (Wittmeyer et al. 2001). Developmental time was prolonged, preoviposition period was prolonged, reproductive rate (R0) and intrinsic rate of increase (r) were significantly lower when fed larval insect prey at both nymphal and adult stages than when fed an insect-free artificial diet during both the nymphal and adult stages. The proportion of fertile females, the quantity of eggs laid by mated females, the reproductive rate, and the intrinsic rate of increase were all significantly increased when feeding larval prey to adults raised as nymphs on an artificial diet, but the mean generation time was not significantly altered. The reproduction rate and intrinsic rate of rise were also significantly decreased when adults raised on larval prey were fed an artificial diet. The price of raw materials, fertility table parameters, and doubling time values was used to calculate the “realised” cost of raising Podisus maculiventris on the artificial diet.

554

15

Mass Production of Insect Predators

The cost of raw materials for raising a colony of Podisus maculiventris on Trichoplusia ni (Lepidoptera: Noctuidae) was only 1.4 times higher than the cost of raw materials for artificial diets needed to raise a colony of the same size. However, owing to the delayed developmental period and decreased reproductive output when raised on an artificial diet, the actual cost of rearing was 3.5 times higher. As well as discussing the benefits and drawbacks of enhancing adult diets with natural prey during the reproductive stage, this article also discusses the economic effectiveness of rearing beneficial insects on an artificial diet that reduces the intrinsic rate of increase of a colony (Wittmeyer and Coudron 2001). For the predator Perillus bioculatus, two artificial meals free of insect components were created. Diet 2 was used to rear Perillus bioculatus for 11 successive generations without having any cumulatively negative consequences on its biology. In Perillus bioculatus, animals raised on artificial diets, development time and preoviposition duration were much longer, and egg viability, survival from egg to adult, and fecundity were significantly lower. After 11 generations of in vitro rearing Perillus bioculatus on diet 2, we saw a significant reduction in developmental time, an increase in survival from egg to adult, and a reduction in the length of the preoviposition phase (Rojas et al. 2000).

15.3

Syrphids

Of the 800 species of palaearctic syrphids, 40% are aphidophagous. Under straightforward laboratory circumstances, it is possible to raise more than 1000 flies per month, with the availability of aphids serving as the syrphid larvae’s only nutritional restriction (Drescher and Tornier 1990). Having trouble in raising some syrphid species: Species with significantly skewed male gender ratios include Episyrphus balteatus. 100% male offspring were produced by larvae obtained from cowpea, cauliflower, and cabbage that were raised in a lab at PDBC between 1997 and 1999. Additionally, certain species, including B. linga. B. fletcheri, E. confrater, and D. aegrota, were unable to reproduce in a lab setting and only produced sterile eggs. Episyrphus balteatus and Eupeodes bucculatus, two species of syrphids, have been raised on an artificial diet consisting of drone honeybee powder (DP) (Iwai et al. 2007). Later, it was discovered that adding linoleic and oleic acids could increase the rate at which adults emerge and their body sizes (Iwai et al. 2009). According to Paragus serratus, artificial nutrition for adults significantly increases oviposition. In this diet, 50% honey is combined with multivitamins, clomiphene citrate, and tocopheryl acetate (Baskaran et al. 2009a). The same team of researchers has also investigated how pollen grains affect Paragus serratus fertility (Baskaran et al. 2009b). The behaviour of its oviposition is induced by the syrphid oviposition was experimentally produced on an inert surface using semiochemicals [E-(β)farnesene, R-(+)-limonene, and (Z)-3-hexenol], honeydews, and “fake honeydews” (10% or 30% aqueous solutions of sucrose, fructose, and glucose). The most effective ovipositional stimulants were found to be E-(β)-farnesene and concentrated mono-sugars (30%). A plastic lamella covered in syrphid eggs was suspended on

15.3

Syrphids

555

aphid-infested plants to gauge the effectiveness of the biological control device. Laboratory and field trials were conducted to test and validate the device. The elimination of populations of 500 aphids in 10 days after the introduction of 15 syrphid eggs showed that the results were encouraging (Leroy et al. 2010). According to Leroy et al. (2010), using such a biological control tool might unquestionably help with biological control to lessen aphid infestations. Artificial Diet In 2007, Iwai and colleagues created an artificial diet for raising Episyrphus balteatus and Eupeodes bucculatus, but their development was subpar compared to those relying on natural prey (aphids). In both species, drone honeybee brood powder (DP) provided the bare minimum nutrients required for larval development. Eupeodes bucculatus did not reach adulthood while Eupeodes bucculatus did on the original DP diet. Eupeodes bucculatus larvae had a higher survival rate on an agargel diet that contained DP, and half of them matured into adults. Physical factors such as diet softness were crucial while raising syrphid larvae. The amount of food ingested by Eupeodes bucculatus third instar stages was counted. The findings showed that the larvae were consuming enough DP. The diet was enhanced by the inclusion of autolysed yeast and sugar. Fatty acids were additionally added to gel diets containing DP in order to enhance larval development and fecundity in Episyrphus balteatus and Eupeodes bucculatus. The emergence rate in Eupeodes bucculatus was greater when each fatty acid was introduced singly to the basic meal than on diets without supplements. In particular, the emergence rate and adult body size were enhanced by the addition of oleic and linoleic acids. The development of Eupeodes bucculatus was most successfully facilitated by the mixture of the two acids. For young and old larvae, respectively, two gel diets with particular amounts of DP, oleic acid, and linoleic acid were created. The concentration of the diet agar was also determined. These emergence rates exceeded 90% and adult body sizes were comparable to aphid-fed controls when these diets were given to larvae of both species. Propionic acid, sodium benzoate, and methyl paraben were the three preservatives that were evaluated to stabilise diets. Propionic acid (0.5%) had no impact on Episyrphus balteatus, but even at a concentration of just 0.1%, Eupeodes bucculatus emerged at a rate that was significantly lower than the control diet (Iwai et al. 2009). Early treatments (storage at 5 °C) or late treatments (storage at 5 °C) were performed on pupae of Eristalinus aeneus and Eristalis tenax (Diptera: Syrphidae, Eristalini) at various stages of development (5, 10, 15, 20, or 30 days). Both treatments’ development halted at 5 °C, although in general, pupae housed at the start of the pupal stage produced better results in terms of survival and adult quality. The pupal growth times of Eristalis tenax and Eristalinus aeneus were effectively prolonged to 18 and 23 days, respectively, without impacting survival and morphology, despite Eristalis tenax weaker cold tolerance (Campoy et al. 2022).

556

15.4

15

Mass Production of Insect Predators

Lacewings

For many years, Chrysoperla spp. have been regarded as significant naturally occurring predators in a variety of cropping systems, including ornamentals, greenhouse crops, vegetables, fruits, nuts, fibre and feed crops, and woodlands. They are also among the most widely used and readily available natural adversaries on the planet. In North America and Europe, two species of Chrysoperla, Chrysoperla carnea and Chrysoperla rufilabris, have been mass-produced and sold for commercial purposes for many years (Wang and Nordlund 1994; Daane et al. 1998). Latin America and Asia employ the species Chrysoperla externa and Chrysoperla nipponensis [=Chrysoperla sinica], respectively (Wang and Nordlund 1994). Members of the Association of Applied Insect Ecologists rated Chrysoperla spp. as unmatched among regularly used, commercially accessible predators in answer to a questionnaire in 1992. In the eastern and midwestern United States, two separate and reproductively isolated species of Chrysoperla carnea known as Chrysoperla plorabunda and Chrysoperla downesi have been employed. Over the past 15 years, Micromus tasmaniae (Neuroptera: Hemerobiidae) has drawn a lot of interest, because of its fecundity (474 no./female) and fecundity rate (11.5 no./day). However, both the fecundity (619 no./female) and fecundity rate (18 no./day) were maximum in Micromus vinaceus. The development of a method for raising Micromus tasmaniae, the Tasmanian brown lacewing, took different densities into account. The larvae were fed the Rhopalosiphum padi oat aphid. Chrysopidae species have been raised in laboratories. Chlysopa carnea, Cluysopa septempunctata, Chlysopa perla, and Chlysopa formosa are examples of common green lacewings. Only C. carnea has been generated in significant numbers (millions) of any of these (Tulisalo 1984). For Chrysoperla carnea, Finney (1948) was the first to design a mass rearing method. The potato tuber moth Phthorimaea operculella’s prepared eggs and larvae were distributed on alternate layers of waxed paper in the process used to produce adults. These layers were kept in cardboard cylinders that were coated in paper. Each cylinder initially contained 50 males and 50 females. The cylinder was replaced three times a week, and the adults were fed. Fibreglass cylinders are used in more recent methods. The cardboard cylinders were replaced with ones that were 10 cm tall and 35 cm in diameter. The cylinders’ tops were sealed with black paper, and the bottoms were fitted with netting. Each cylinder initially included 400 persons, 65–70% of whom were female. The egg sheets were changed daily, and the adults were fed (Morrison and King 1977; Tulisalo 1984). For the purpose of Integrated Pest Management Programs targeting insect pests, Sattar and Abro (Sattar and Abro 2011) were mass-reared Chrysoperla carnea adults. The methods for rearing Chrysoperla carnea have been improved, allowing for mass rearing in the lab and field releases. To avoid the need for anaesthetic or a vacuum sucker, as was the case with previous methods, the cannibalistic larvae were maintained separately in firm gelatine capsules (500 mg). The created pupae were retrieved after the larvae had developed for 10 days, and they were then put in cages

15.4

Lacewings

557

for adult emergence and egg deposition. Eggs collected were applied to the landscape (El-Arnaouty et al. 1998; Ashfaq et al. 2002, 2004). Food and Artificial Diet Artificial feeds for adult green lacewings raised in laboratories have improved the mass-rearing process. Five parts of Food Wheat® (a yeast called Saccharomyces fragilis that has been cultivated on a whey substrate) are mixed with six parts of sugar and ten parts of water to create a semi-artificial nutrition combination that has been created for raising and egg production. The best results in adult rearing have been achieved using this artificial medium, which should be provided continually in the rearing cylinders (Tulisalo 1984). Adult C. carnea lay eggs on oviposition papers during mass raising. The eggs from this substrate have been removed using a variety of techniques. The egg stalks were disintegrated by sodium hypochlorite, according to Finney (1948). To cut the egg stalks, Ridgway et al. (1970) utilised nylon netting, and Morrison and King (1977) employed electrically heated wire (Tulisalo 1984). If there is not enough food available, green lacewing larvae will become cannibalistic, much like any predator. As larvae get older, this cannibalistic inclination gets stronger. Lack of an appropriate and affordable artificial medium is another issue with mass raising. Different diets have been researched for larval rearing. Larvae have been raised on an artificial diet made up of enzymatic casein hydrolysate, enzymatic soy hydrolysate, fructose or saccharose, mineral salts, soy bean lecithin and oil, cholesterol, B vitamins, choline, inositol, and water (Vanderzant 1969). Feeding larvae a combination of dried, milled Angoumois grain moth adults, honey, brewer’s yeast, milk casein, and a vitamin mixture are examples of other meals. In earlier research, the feeding of larvae with lyophilised female honeybee brood powder was also examined. It is now not possible to totally substitute the natural diet of larvae with artificial diets. There have been numerous succeeding generations fed artificial diets, but there have also been some negative outcomes. In other instances, the length of the larvae’s developmental period, their rate of growth, and the proportions of pupal production and emergence have all decreased. As a result, artificial diets cannot be used in large-scale, ongoing mass rearing of larvae (Tulisalo 1984) Because all three instars of the Chrysoperla species are predatory, larval rearing is now the most expensive part of mass manufacturing. Most insectaria feed on massproduced insect prey, which is more expensive than artificial diets (often lepidopteran eggs: Sitotroga, Anagasta, or Corcyra). When the larvae were raised on these diets, all of the Chrysoperla spp. tested fared well. The creation of an artificial diet ought to remain a top goal. Either a liquid diet or a solid diet will allow lacewing larvae to grow and feed (see Cohen and Smith 1998a, b). Liquid diet production and encapsulation can be somewhat automated, but the price has remained high (Wang and Nordlund 1994). A wholly artificial, solid or semisolid food that appears to give considerable advantages over existing diets has been developed recently as a consequence of research that concentrated on in-depth observations of predator eating behaviour. The new diet does not require

558

15

Mass Production of Insect Predators

encapsulation, is reasonably affordable, and does not spoil soon (Cohen 1998; Cohen and Smith 1998a, b). It takes a lot of space and effort to mass-raise insects, especially cannibalistic predators; at the moment, automated, space-efficient massraising systems for Chrysoperla are being developed (Nordlund and Correa 1995). Quality Control Both the use of biological control and users’ view of biological control as a dependable pest management strategy depend on the standardisation of high-quality natural enemy production (e.g., Bigler 1992). However, because there are no stringent quality control regulations in the United States, the quality of naturally occurring enemies that are sold commercially can vary. For instance, a recent analysis of shipments from insectaries revealed that growers’ orders for Chrysoperla carnea were not regularly completed with the right species, and cannibalism dramatically decreased the survival of lacewings while in transit (O’Neil et al. 1998). Increased attention to maintaining correctly recognised, pure colonies and improved mass production and packing techniques can help to solve these issues (Bigler 1992). We advise the upcoming. At the beginning of the culture, the species of the Chrysoperla stock needs to be confirmed. To check for contamination, species identity should be reviewed occasionally throughout rearing (see Systematics section above). Cultures’ performance and survival through mass manufacturing should be periodically assessed (e.g., van Lenteren 1998). Recent research suggests that (1) the timing of stock collection in the field and (2) the periodic intervention of diapause or cold storage may avoid Chrysoperla stock degradation during continuous rearing (Jones et al. 1978). For instance, adults caught earlier in the season than those caught later in the season generated better progeny (Chang et al. 1996). Additionally, the induction of diapause improved the ability of offspring descended from late season cohorts to reproduce. These ecophysiological findings could be used productively to standardise the calibre of commercial stock with some carefully targeted research. Handling and shipping. To avoid hatching and cannibalism in transit, eggs that are ready to be sold should be delivered in insulated containers with cold packs as soon as possible after oviposition. Packaging with suitable packing material and food, such as Sitotroga or Anagasta eggs, can lower cannibalism and mortality when prefeed larvae are ordered. Here, more in-depth investigation is required. It is crucial for the insectary sector to provide standards that support the dependability of commercially generated natural enemies when looking at the quality control issue as a whole. In this aspect, it appears that the insectary business and scientists in Europe work together more effectively and cooperatively than they do in the United States (van Lenteren 1998). Chrysopa pallens was mass-reared using an artificial diet of shrimp, beef, beef liver, and egg yolk in order to employ it as a biological control agent in sustainable pest management. Additionally, a risk assessment protocol based on an artificial diet was created to look into how Cry1Ac, Cry1Fa, and Cry2Ab affect adult Chrysopa pallens survival and ability to reproduce. Chrysopa pallens was fed on diets containing cryoproteins and those without them (control). The boric acid-containing

15.5

Coleoptera

559

diet served as the positive control. Using double-antibody sandwiches, enzymelinked immunosorbent tests, and bioactivity verification bioassays, the temporal stability, bioactivity, and consumption of Cry proteins by Chrysopa pallens were validated. Pre-oviposition period, daily fecundity, total fecundity, and 30-day-old adult dry weights of Chrysopa pallens showed no statistically significant differences for the diets containing Cry1Ac, Cry1Fa, and Cry2Ab (50 g/g) compared to control. Positive control, however, showed a substantial decline in both survival and reproductive efficiency ( p 0.05). Our research shows that artificial diets are an excellent source of nutrition, improve Chrysopa pallens survival and reproduction, and can be used to raise predators in large numbers when natural diets are scarce. Cry proteins are also safe for Chrysopa pallens adults, and the cultivation of Bt crops aids in the conservation of predators in sustainable agriculture (Ali et al. 2018).

15.5

Coleoptera

The scientific literature, according to Riddick and Chen (2014), demonstrates modest technological advancements that have led to the production of coleopteran predators, particularly lady beetles. For polyphagous species, a variety of fictitious prey/foods and insect-free artificial diets are offered. The development of artificial diet-based production systems for polyphagous as opposed to oligophagous animals may be more efficient in terms of time and money. To lessen the detrimental impacts of crowding and cannibalism in colonies, more research is required to ascertain how to adjust rearing (population) densities—relative to food quality/quantity, cage size, oviposition, and mating. The metabolic rate of predators can be controlled by temperature to change the size of the colony throughout times of high or low demand. More investigation is required to develop quality control procedures for other coleopteran predators, such as routinely checking items for appropriate fitness and undesired infections prior to or after distribution to clients. Utilising the eggs of the Angoumois grain moth, Sitotroga cerealella, an unique mass production method for raising the mealybug predator Brumoides suturalis (Coleoptera: Coccinellidae) was created (Lepidoptera: Gelechiidae). This method was created with the aim of eliminating issues related to the wintertime raising of mealybugs, which are the typical host insects for Brumoides suturalis. On Sitotroga cerealella eggs, the Brumoides suturalis larval and pupal periods were 14.64 and 6.03 days, respectively. When compared to mealybug eggs, the adult lifetime of Sitotroga cerealella was much less (38.6 days) (49.0 days). However, compared to mealybugs (160.09 Sitotroga cerealella eggs), Brumoides suturalis’ fecundity was much higher on Sitotroga cerealella eggs (204.6 eggs per female), with values of 18.4, 17.5, and 18.18. As a result, we can anticipate getting about 55 adults if we start a culture of Brumoides suturalis with 100 eggs and raise them on Sitotroga cerealella eggs (Sunil Joshi et al. 2013). Riddick and Wu (2015) created methods for rearing Coleomegilla maculata, a ladybird beetle, and assessed how rearing density affected the beetle’s survival, growth, and development. The idea that a low to moderate rearing density has little

560

15

Mass Production of Insect Predators

to no impact on development and survival was put to the test. First instar Coleomegilla maculata larvae were fed powdered brine shrimp (Artemia franciscana) eggs and reared to pupae at a density of 1, 5, 10, 15, or 20 per arena (2.5 cm high, 9.0 cm diameter, and 159 cm3 volume). At densities of 1 and 5, more larvae survived, but no changes were seen at concentrations of 10, 15, or 20. At the 10, 15, and 20 densities, the median survival rate for larvae was at least 90%, and for pupae, it was 100%. The rearing density had no impact on the length of time for development, body weight, or sex ratio. Overall, our study demonstrates that C. maculata larvae can be successfully raised in containers supplied with powdered Artemia franciscana eggs at a density of 20 larvae/159 cm3 (0.126 larvae/cm3). It should be able to scale and increase container size and Coleomegilla maculata density in these containers (Riddick and Wu 2015). Later, container size optimisation for Oenopia conglobata bulk rearing (Coleoptera: Coccinellidae). The experiment used four different Oenopia conglobata rearing densities (1, 5, 10 and 20 larvae/container, regarded as low, moderate, medium, and high rearing density, respectively) and three different types of containers of varying sizes (106, 310, and 785 ml, small, medium, and large, respectively). Throughout the course of the investigation, Oenopia conglobata was fed Ephestia kuehniella (Lepidoptera: Pyralidae) egg yolks. All containers with low raising density had the best larval survival rate (100%), but medium containers with medium and high rearing density had the lowest (31.5–35.0%). The shortest pupal development duration (4.3–4.4 days) occurred in medium and big containers at medium and moderate rearing density, respectively. The shortest larval development period (8.6 days) was in medium containers at low rearing density. Small and medium containers with low raising density incurred the highest rearing costs, whereas big containers with high rearing densities incurred the lowest rearing costs. For the economically and quickly mass rearing of Oenopia conglobata, large containers with high rearing density are advised due to the survival rates, development times, and financial costs involved (Mamay and Mutlu 2019). The most prevalent and extensively researched aphid predators are coccinellids. For instance, Coccinella septempunctata was fed on dried aphids, but the experiment was unsuccessful because only 35% of the larvae matured into adults (Hodek and Honêk 2013). Although fecundity decreased marginally, quick-frozen aphids seem to be a good substitute for live aphids for Coccinella septempunctata larvae and adults (Shand et al. 1966). It is discussed how to raise Cryptolaemus montrouzieri. A pumpkin was covered by at least 50 Maconellicoccus hirsutus gravid females. On colonies of Maconellicoccus hirsutus that were 20 days old, ten adults of Cryptolaemus montrouzieri who were about to oviposit were released for 10 days. In 50–55 days following the initial pumpkin infestation, each of these colonies produced an average of 250 Cryptolaemus montrouzieri adults (Kishore et al. 1993). Menochilus sexmaculatus was fed in four different ways during its five developmental stages, and the natural and artificial diets were evaluated on each of them: (a) live aphids (La), (b) dried aphids (Da), (c) frozen aphids (Fa), and (d) chicken liver (Cl) (Khan and Khan 2002).

15.5

Coleoptera

561

Table 15.11 Menochilus sexmaculatus’s pre-oviposition, oviposition, egg-laying, and lifespan days were all fed live, frozen, and dried aphids, chicken live, and sucrose (Source: Khan and Khan 2002) Treatments Live aphids (La) Frozen aphids (Fa) Dried aphids (Da) Chicken live (Cl) Sucrose (Su)

Pre-oviposition days 4.7 18.5

Oviposition days 28.5 34

Egg laying/ day 30.5 6.4

Longevity days 166 63

21.6 0 0

30 0 0

2.5 0 0

46.5 42.6 36.5

Aphids from stock cultures were collected and killed by being kept at a low temperature for 24 h. The dead aphids spent 4 h in a 55 °C oven. In a plastic container, the dried prey was maintained in a 2 °C refrigerator. The laboratory stock culture of live aphids was used to prepare frozen aphids, which were then placed in a tight container measuring (18 × 6 cm). For 24 h, these containers were stored in the freezer. The dead aphids were separated into little portions when they had completely frozen, placed in small containers (9 × 3 cm), and preserved at the same temperature throughout the duration of the experiment (Khan and Khan 2002). The artificial diet was made by drying 5 kg of chicken liver completely at 90 °C for 2 h before blending it into a powder. The best form of chicken liver to use to create a homogenised powder was completely dry. The dried material was divided into 500 g of blended powder, which was then combined with 200 mL of sucrose solution (5% wet volume) and blended again for another 5–10 min to create a paste. The resulting paste was packaged in aluminium foil and frozen for storage. The cylindrical shape of the diet material was wrapped in a single sheet of parafilm, an insulating material. Preservatives and antibiotics were not used. Every day, 50 g of chicken liver material that had been preserved was taken and heated in a water bath to melt in a solution of sucrose. Using the tiny capillary tubes, the molten substance was transformed into small, uniform droplets. The droplets were evenly spaced apart on a parafilm sheet. Every day, the drops were prepared to prevent contamination. In order to provide them with a separate eating area, little pieces of parafilm were inserted in petri plates (Khan and Khan 2002). Results reveal that Menochilus sexmaculatus responded much better to feeding on live aphids (M. persicae) compared to artificial diets made up of dried aphids, frozen aphids of the same species, and chicken liver diet during all four immature stages (first four instars). The average weight increase of larvae in the first through fourth instars was significantly impacted by the natural diet (La) compared to the other treatments. When a predator fed on Da, Fa, and Cl, the survival rate was lower (25–30%). With the exception of the control, which is a normal diet, egg production stays lower in all treatments. Pre-oviposition intervals, oviposition intervals, egg-laying capacity, and adult longevity all changed (Table 15.11) (Khan and Khan 2002).

562

15

Mass Production of Insect Predators

As a polyphagous predator, the multicoloured Asian ladybeetle Harmonia axyridis loves aphids but also consumes coccids, lepidopteran eggs, and other insects. The juvenile larvae (second and third instars), which are used for releases, and the adults, which are used to maintain stock cultures, are two developmental phases that are essential for the mass production and usage of Harmonia axyridis as a biocontrol agent. Recently, certain artificial diets for adults and larvae in the second through fourth instars were produced. By contrasting the biochemical makeup of the ladybeetle larvae reared on the fake meals with those fed on E. kuehniella, these diets were tested and improved using an analytical technique and subsequent interactions. For insects that began as second-instar larvae, adult emergence was close to 80% on the best diets (Sighinolfi et al. 2008). One of the most significant predatory coccinellids in Egypt is the seven-spotted ladybird, Coccinella septempunctata (Coleoptera: Coccinellidae). The biology and mass rearing of this predator in Egypt are poorly characterised. The Bean Aphid (Aphis fabae) was the primary food source for the larvae and adults of Coccinella septempunctata. The adults laid 1 to 25 clusters of eggs per female. 1:1 is the sex ratio. The lifespan of female predators was between 21 and 26 days, whereas that of male predators was between 24 and 29 days. The incubation period lasted between 2 and 3 days. Coccinella septempunctata larva goes through four instars. Under laboratory conditions, the average number of aphids devoured by each of the four instar larvae is 35, 63, 96, and 290. Under consistent settings of 23 °C and 60 RH%, the predator’s whole developmental period occupied 16–21 days (Mahyoub et al. 2013). A novel technique for manipulating Coccinella septempunctata eggs was created, and it entails placing a plastic cylinder within the cages for egg laying, then moving the egg masses out of the cage and onto a different machine so they can be ready to be stuck on a card for release (Mahyoub et al. 2013). The major insect prey, Leptinotarsa decemlineata eggs, have nutritional qualities that are similar to those of the diets based on chicken liver and tuna fish. In people raised on artificial diets, developmental time and preoviposition duration were much longer, while egg viability, survival from egg to adult, and fecundity were significantly shorter. These modifications might represent some degree of diet adaptation on the part of the predators following numerous generations of in vitro upbringing. After 11 generations of in vitro rearing, there were no modifications in egg viability. When raised on an artificial diet, adult female weight was unchanged; however, after 11 generations of in vitro rearing, a considerable rise was seen. Under all treatments, females typically weighed 11.6 mg more than males (Rojas et al. 2000). Aphidophagous ladybirds (species that feed on aphids) also have a greater metabolic rate, grow and move more swiftly, and age more quickly than coccidophagous species (Dixon et al. 2016). Ephestia kuehniella (Lepidoptera: Pyralidae) eggs, a commercial mixture of Ephestia kuehniella and Artemia sp. (Anostraca: Artemiidae) cysts, a liver-based artificial diet, and the natural prey, Aphis gossypii, are used to rear Harmonia axyridis (Coleoptera: Coccinell (Hemiptera: Aphididae). The diet regimens had a variety of effects on the Harmonia axyridis life metrics that were being studied. On moth eggs, larvae developed quickly and successfully, while Harmonia axyridis fed

15.5

Coleoptera

563

on an artificial diet did not experience a quick and successful development. The rearing of Harmonia axyridis was made possible by the combination of moth eggs and brine shrimp cysts (Ricupero et al. 2020). Ricupero et al. (2020) advise against using it unless there are no other suitable feedstuffs available. The life table method utilised in this study could enhance the standard comparison between dietary regimens, enhancing the species’ rearing procedures (Ricupero et al. 2020). An introduced coccinellid from Japan named Sasajiscymnus tsugae is a promising biological HWA control agent. The major objectives of Sasajiscymnus tsugae mass rearing are to increase adult beetle output, decrease the need for human labour, and lower manufacturing costs. A modified rearing box produced much more adult S. tsugae and had a greater rate of egg-to-adult survival than a standard rearing box. 4400 more beetles were produced by the 30 modified boxes than by the 30 standard boxes. As a result of larval cannibalism at densities above 1650 larvae per rearing box, the survival rate of eggs to adults declined with time. The quantity of bug and spider predators accidentally introduced by hemlock twigs infected with Adelges tsugae—HWA also increased as the temperature rose. The amount of time needed to prepare modified and conventional boxes for the initial introduction of eggs was similar, but the amount of time needed to add water, honey, and twigs that had been infected with the Adelges tsugae—HWA was substantially less in the modified boxes. The more adult Sasajiscymnus tsugae produced and the time saved over the 35 days of planned maintenance as eggs hatched and larvae matured into adults made up for the greater expense and effort needed to modify each box (Conway et al. 2005). Twenty-eight of the aphidophagous species are members of the Coccinellini tribe, while six are members of the Coccidulini tribe. The shape of their last abdominal sternite was used in the lab to determine their sex (Hodek et al. 1973). They were then divided into pairs, each comprising a male and a female, and housed in a 9-cm Petri dish with a piece of filter paper that had been accordion folded to maximise the surface area for oviposition for 2 weeks at a temperature of 20 °C and a photo phase of 16 h. The ladybirds were moved to a fresh Petri dish each day and fed an excess of Acyrthosiphon pisum, a pea aphid raised on Vicia faba. Daily collections were made of the ladybird eggs found on the folded filter paper. For 10 of the species employed in this study, the pea aphid supports normal reproduction and is referred to as “essential food” (Hodek et al. 2012). It is possible that this aphid is not the best prey for the other species. Although prey quality influences clutch size, a research by Rana et al. (2002) indicates that it has little impact on ladybird egg size. Eight different species of coccidophagous ladybirds (species that feed on coccids) were either gathered from the wild or obtained from stock cultures kept in a lab. Two of them are from the Chilocorini Tribe, five are from the Coccidulini Tribe, and one is from the Noviini Tribe. Because coccidophagous ladybirds are far more specialised in their diet than aphidophagous species, these species were sexed, paired, and raised in the same manner as those described above, but fed one of a wider variety of food. Exochomus quadripustulatus, Cryptolaemus montrouzieri, Nephus reunioni, N. bisignatus, and N. includens were fed Planococcus citri, which was raised in the dark on potato sprouts. Icerya purchasi that had been raised on Pittosporum

564

15

Mass Production of Insect Predators

tobira was fed to Rodolia cardinalis. Aspidiotus nerii, a species raised on potato tubers, was given to Rhizobius lophantae and Chilocorus bipustulatus. The ovisacs that were still in each Petri dish at the end of the day were dissected under a binocular microscope and examined for ladybird eggs, as these ladybirds deposit their eggs inside or below ovisacs of their victim (Hemptinne et al. 2022).

15.5.1 Mass of Adults and Eggs and Ovariole Number Prior to being weighed precisely 0.1 mg, female ladybirds were given 10 days to get used to the lab environment. Depending on how well we were able to gather and raise each species, the number of females ranged from 3 to 16. Eggs under 24 h old were collected, taken out of the substrate they were put into or connected to, and individually weighed on the same microbalance to an accuracy of 0.1 mg. We planned to weigh five eggs from five consecutive ovipositions for each female. Some of the field-collected ladybirds, however, did not live long enough in the lab to oviposit five times; hence some samples are smaller than 25 eggs. The females were terminated gently after being weighed. Their abdominal tergites were removed, and their elytra were trimmed. By grasping the oviduct with forceps and yanking it out of the abdominal cavity, their ovaries were extracted. On a microscope slide, ethylene blue was used to stain them. The number of ovarioles in each ovaries was then counted using a binocular microscope to determine the amount of investment required for reproduction (Hemptinne et al. 2022). The interaction between log (adult mass) and diet type was not significant, much like in the linear mixed models (LMM). It was left out of the finished model. The fact that log (adult mass) and log (egg mass) are not correlated suggests that huge females do not generally produce larger eggs. However, because coccidophagous ladybirds lay noticeably smaller eggs than aphidophagous species, food type clearly has an impact. The interactions between adult mass and food type were never significant for the phylogenetic generalised least squares (PGLS) analyses, hence they were also removed. Depending on the phylogenetic tree employed in the investigation, the models without interaction showed that log (egg mass) significantly scales with log (adult mass), with an exponent ranging from 0.646 to 0.737. Compared to aphidophagous species, coccidophagous ladybirds lay eggs that are substantially smaller. Although the phylogenetic tree used in the PGLS study has an impact on the value of the Pagels’s coefficient, investigations of deviance show that the results are usually statistically different from 1 but not from 0 (no phylogenetic signal) (Hemptinne et al. 2022). This suggests that, rather from being a result of phylogenetic relatedness, the variation in egg mass is related to the type of prey that ladybirds hunt. The interaction between log (adult mass) and diet type was not significant for the generalised linear mixed-effects model (GLMM), hence it was left out of the final model. Coccidophagous ladybirds have a much lower reproductive investment than those that eat aphids, with log (reproductive investment) being slightly correlated with log (adult mass). The slopes between log (adult mass) and log (ovariole

15.5

Coleoptera

565

number), however, are identical in the two groups of ladybirds. The interactions between adult mass and diet type were likewise not significant for the PGLS analysis; hence they were left out of the calculations. The models without interactions demonstrate a significant scaling relationship between log (reproductive investment) and log (adult mass), with an exponent varying depending on how the phylogenetic tree was built from 0.452 to 0.465. However, coccidophagous ladybirds invest just as much in reproduction as aphidophagous species do (Hemptinne et al. 2022). Artificial Diet Two species of Coccinellids have been raised on such diets. Coleomegilla maculata can be raised on a variety of meridic diets; however, Harmonia oxivridis can grow on a different diet that includes lyophilised powder made from drone honeybee larvae. Additionally, this diet seems promising for other species of coccinellids (Hodek and Honêk 2013). A portion of fresh pork liver sold as human food (35.8 g), Isio 4 oil (Lesieur®) (2.4 g), olive oil (3.0 g), sucrose (12.0 g), glycerine (3.0 g), aqueous amino acid solution (40.0 g), yeast extract (3.5 g), and Vanderzant’s vitamin mixture (Sigma®) were combined to create the artificial diet used in the current study (0.3 g). Tyr (0.25 mg/g), His (0.62 mg/g), Arg (0.94 mg/g), and Ethanolamine (0.19 mg/g) were all present in the aqueous amino acid solution. Until it was utilised, the diet was kept at -20 °C. The food was given to the larvae or adults in the form of droplets with a diameter of 0.3–0.5 cm that were placed on Bristol paper, dried at room temperature (22–24 °C) for 18–24 h, and then kept in storage at 4 °C until needed. The drops were separated from the substrate before usage. To deliver water, an Eppendorf tube (1.5 ml) with cotton plugs was filled with distilled water (Sighinolfi et al. 2008). The results of Ghanim et al. (2021) demonstrated that the two predators were successfully reared on these diets; the average duration of the larval instars for Hypodamia tridecimpunctata and Chilomenus propinquaisis when reared on the artificial diet (AD1) was 24.6 and 21.7 days, respectively, while this duration was recorded for these predators when they were reared on the artificial diets AD2 was 19.9 and 17.6 days. When these two coccinellid predators were raised on artificial diet (AD1), the pupal stage lasted an average of 6.2 and 5.4 days, respectively. Meanwhile, the data showed that when raised on the artificial meals AD2 and 4.9 days, respectively, the pupal stage persisted on average (Ghanim et al. 2021). Harmonia oxivridis was able to mature from initial instar stages to adults thanks to an artificial food. However, the food source had a big impact on all the pre-imaginal factors. The larvae produced on the artificial diet displayed considerably longer developmental durations and lower adult emergence rates (27.6% emerged adults on the artificial diet vs. 91.5% on Ephestia kuehniella eggs, 2 = 134) than those grown on Ephestia kuehniella eggs. In comparison to first instar larvae, the majority of those raised on the artificial diet perished either at emergence or as third-instar larvae (38.0 and 16.7%, respectively). When compared to the artificial diet group, the weights of the newly emerged adults were substantially higher for the control coccinellids. When compared to the control insects, the daily

566

15

Mass Production of Insect Predators

weight gains and the quantity of eggs deposited by females (E10) raised as larvae on the artificial diet were significantly lower. In contrast, the length of the pre-oviposition phase, the quantity of eggs laid per mg of female weight, and adult longevity were unaffected by the larval feeding sources. In the larval stages, females fed the artificial diet had better fertility (% of hatched eggs) (Sighinolfi et al. 2008). Tetranychid mites have a significant natural opponent in the ladybird beetle Stethorus gilvifrons, which also serves as a biological control agent for these plant pests. When given three artificial diets, Stethorus gilvifrons growth, survivability, and ability to reproduce were examined. The artificial food components that Stethorus gilvifrons could be effectively grown on for one generation without the use of tetranychid mites were investigated. Methods: Sucrose, honey, royal jelly, agar, yeast, and date palm pollen made up artificial meals that were supplemented with hen’s egg yolk (AD1, the basic diet), Ephestia kuehniella eggs (AD2), or Ephestia kuehniella eggs and 2,4-dihydroxybenzoic acid (AD3) (AD3). The preoviposition and immature development periods of adults and larvae of Stethorus gilvifrons fed on AD1 were shorter than those of those fed on AD2 and AD3. Females fed AD3 had considerably more eggs overall deposited than those fed the other diets. On AD3, AD2, and AD1, respectively, Stethorus gilvifrons’ intrinsic rate of increase (r) was highest. The fact that Stethorus gilvifrons functioned best on AD3 suggests that this artificial diet has the potential to be used for the mass raising of this significant predatory ladybird beetle (Ebrahimifar et al. 2020). Under continuous laboratory settings (26.1 °C and 65.5% RH), the impact of various meals on the biological traits of Coccinella undecimpunctata was examined. The biological characteristics of Coccinella undecimpunctata were assessed using three diets: AD1 (as a basic artificial diet), AD2 (as an improvement artificial diet), aphid, Aphis craccivora frozen as well as live aphid, and Aphis craccivora as a control. When larvae fed on control, the results showed that the mean larval duration was 10.64 days. For frozen aphid, AD2 and AD1, it increased to 14.3, 14.6, and 21.3 days, respectively. The percentages of larvae that survived after being raised on control, AD2, AD1, and frozen aphid, respectively, were 93.33, 83.33, 73.33, and 60.00%. The control group had a pupation rate of 92.86%, but the frozen aphid had a pupation rate of only 66.7%. 92.31% of adults emerged from the control group, with a sex ratio of 54.17%. When larvae fed on AD2, it was 80.0% with a 62.5% sex ratio. On control, a mated female’s ovipositional period lasted 56.63 days and she produced 992.3 eggs throughout the course of her life, whereas AD1 produced the fewest (189.5 eggs). For egg fertility (95.62 and 75.9%), as well as egg hatching (95.0 and 79.2%), treatment control and AD2 produced the greatest results. The control group had the highest growth index (4.75), which was nearly identical to the artificial diet (AD2) (3.39). Overall, AD2 was better on most of the aforementioned metrics and played a significant role in the entire life cycle of Coccinella undecimpunctata for mass production (Youssif and Helaly 2021). Using the pea aphid Aphis craccivora as a control, the ladybeetle Coccinella septempunctata is mass-raised on four artificial diets (A, B, C, and D) (CK). According to the findings, the developmental period until emergence was as follows, going from short to long: CK (12.3 days), D (16.7 days), A (17.4 days), C (17.5

15.6

Quality Control

Table 15.12 Composition of artificial feed (Liu et al. 2022)

567

Ingredients Apis mellifera pupae (g) Bombyx mori pupae (g) Tenebrio molitor pupae Sucrose Pork liver Rapeseed pollen Aphid Beta carotene Olive oil Banana Agar Purified water Total

A 30 0 0 115 20 20 2 1 2 10 0 900 1100

B 0 30 0 115 20 20 2 1 2 10 0 900 1100

C 0 0 30 115 20 20 2 1 2 10 0 900 1100

D 30 0 0 115 20 20 2 1 2 10 0 900 1100

days), B (18.6 days). A new adult’s weight is rated from highest to lowest as follows: CK (339.5 g) > A (205.3 g) > D (197.7 g) > B (174.9 g) > C. Larval eclosion rates are ranked from highest to lowest as follows: CK (90.0%) > C (87.5%) = D (87.5%) > B (80.0%) > A (57.5%) (169.5 g). Fecundity ratios between the experimental group and the control group were 80.46% (A), 39.2% (B), 45.3% (C), and 53.0% (D). The percentages of hatching were 59.5% (A), 46.0% (B), 57.6% (C), 54.5% (D), and 53.9%. (CK). Adult Coccinella septempunctatas fed a combined artificial food had a greater death rate than those fed a control diet. When given artificial diet A, Coccinella septempunctata did not significantly lessen oviposition in comparison to the control diet. In order to raise Coccinella septempunctata in large numbers, diet A can be utilised (Liu et al. 2022) (Table 15.12).

15.6

Quality Control

The quality monitoring of commercially generated insects hasn’t gotten much attention up until lately. There are two likely causes for this. Because mass-raised insects are typically produced and used by the same institution, there is no market competitor driving ongoing product innovation. Second, it can be challenging to define, monitor, and manipulate quality in mass-produced insects. This is in large part due to a widespread lack of understanding of the traits that allow an insect to fulfil its intended function in pest management (Boller and Chambers 1977). Insect raising requires the same level of quality control attention as any other sector since it is crucial to survive in today’s cutthroat marketplace. Additionally, the ideas behind quality control are the same. According to Chambers and Ashley (1984), they are 1. A management practise known as "quality control" produces, maintains, and enhances quality.

568

15

Mass Production of Insect Predators

2. This process calls for the establishment of standards, evaluation of adherence to those standards, action on that information, and product improvement planning. 3. Product control, process control, and production control are the three components of quality control. Because they all provide information, create feedback loops, and regulate quality, these three sectors are strongly tied to one another. Evaluation of the product’s conformance to requirements and quality standards is the goal of product control. Products that don’t adhere to specified specifications can either be corrected or removed. Process control provides information on the efficiency of industrial processes. It regulates these processes to ensure that variations in the procedures do not lead to deviations from the product standards. Production output consistency, the volume of goods produced, and the timeliness of those goods’ production are all governed by production control. The usefulness of animal foods including beef liver, chicken eggs, chicken liver, and potted meat for growing different predators was examined by Eric and Riddick (2009) in his brief review. The results are shown in the table below, along with citations (Table 15.13).

15.7

Future Recommendation

• More study is required to create quality control standards for predators that are acceptable in terms of fitness. • It is crucial to consider the genetic component while breeding and widely dispersing entomophages. • In order to provide a suitable artificial diet for predators, the chemical makeup of various hosts from nature and laboratories will be examined. • It is important to analyse and compare the primary nutrients in artificial foods with those in natural foods and laboratory hosts. A new diet could be created soon based on the findings. • Future research on the compositions of minerals and amino acids is necessary. Identification of info-chemicals that attract reduviid for feeding the food is also necessary.

Geocoris punctipes (adults) Dicyphus tamaninii (nymphs adult)

Carcinops pumilio (nymphs adult) Harmonia axyridis (nymphs adult) Harmonia axyridis (nymphs adult) Orius laevigatus (nymphs adult)

Predator Calosoma sycophanta (larvae) Thanasimus dubius [larvae]

Chicken egg yolk, chicken liver, casein hydrolysate

Chicken whole egg, chicken liver, casein hydrolysate

Beef liver, ground beef—in parafilmpackets

Beef liver, ground beef—in parafilm

Beef liver, fatty ground beef, hen’s egg yolk—in Parafilm

Coccinellidae

Anthocoridae

Lygaeidae

Miridae

Veal, veal gravy, chicken egg, potted meat, infant formula, casein hydrolysate—in Parafilmcapsules PRO-PLEXTM protein additive

Artificial diet—animal protein base Beef liver, chicken meat

Coccinellidae

Histeridae

Cleridae

Family Carabidae

Survival (increase), longevity (increase), predation (increase)

Positive effects Survival (increase) Size (increase)

Development time (incrased), size (decrease)

Size (decrease)

Development time (increase), survival (decrease)

Development time (increase), oviposition (decrease)

Development time (increase), size (decrease), oviposition (decrease)

Development time (increase), size (decrease), oviposition (decrease)

Negative impacts Development time (increase)

Future Recommendation (continued)

Survival (H) by 4th– 5th generation

Size (H), oviposition (H) Predation (H)

Survival (H)

Survival (H)





Significant effect –

Table 15.13 Examples of experiments that examined how predatory beetles, real bugs, and a lacewing’s life parameters were affected by eating artificial diets versus fake or natural prey

15.7 569

Beef liver, ground beef, whole hen’s egg—in parafilm

Beef liver, fatty ground beef, hen’s egg yolk, casein—in parafilm Hen’s egg yolk—in parafilmdomes

Miridae

Miridae

Beef, whole hen’s egg—in mylar-parafilm

Pentatomidae

Pentatomidae

Pork liver, fatty ground beef—in parafilm domes Pork liver—in parafilmdomes

Pentatomidae



No significant effect in predation

Survival and longevity increase

Survival, longevity and oviposition increases

Positive effects





Development time increase; survival, total oviposition and longevity decreases

Oviposition (decrease)

Survival (H)

Development time (increase), size (decrease), oviposition (decrease), oogenesis (decrease) Oogenesis decrease

Longevity (H) –



– Development time, survival and size decreases



Size (H), oviposition (H) –

Oviposition (H) Predation (H)

Significant effect

Development time (increase) Survival (increase)

Negative impacts

15

Perillus bioculatus adults Perillus bioculatus adults Perillus bioculatus nymphs, adults

Miridae

Beef liver, fatty ground beef, hen’s egg yolk, casein— reformulated diet Pork liver, whey powder—in parafilm

Miridae

Hyaliodes vitripennis (nymphs, adults) Hyaliodes vitripennis (nymphs, adults) Macrolophus caliginosus (nymphs, adults) M. caliginosus (nymphs, adults)

Beef liver, fatty ground beef, hen’s egg yolk—in Parafilm

Miridae

Dicyphus tamaninii (nymphs adult) Dicyphus tamaninii (N, A)

Artificial diet—animal protein base

Family

Predator

Table 15.13 (continued)

570 Mass Production of Insect Predators

Chicken liver, tuna fish—in parafilm capsules

Beef liver, fatty ground beef, hen’s egg yolk—in parafilm

Beef liver, fatty ground beef, hen’s egg yolk—in parafilm

Pentatomidae

Pentatomidae

Pentatomidae

Development time (increase), size (decrease), oviposition (decrease)

Development time (increase), survival (decrease), oviposition (decrease) Development time (increase), survival (decrease), size (decrease) Survival (H) Longevity (H)





Source: Then Eric and Riddick (2009) – indicates no significant effects recorded (H) No significant effect (H) on a given life to a control, for designated predator; Podisus maculiventris adults Beef beef, hen’s egg yolk—in parafilmsize (;), predation (H) Chocorosqui and De Clercq (1999) Rearin Podisus maculiventris nymphs, adults beef liver, whole hen’s egg—in Mylar-parafilm development, oviposition (;) Podisus maculiventris nymphs, adults beef liver, whole hen’s egg—in Mylar-parafilmbody size (;), oviposition (;), oogenesis (;), vitellogenesis (;) Wittmeyer et al. (2001) P. maculiventris nymphs, adults beef liver, whole hen’s egg—in Mylar-parafilm development time (:), size (;), survival (H), oviposition (;) Coudron et al. (2002) nymphs, adults (Pentatomidae) beef liver, ground beef development time (:), size (;), ovarian weight (;) P. nigrispinus adults beef liver, ground beef ovarian development (;), oogenesis (;) Chrysoperla rufilabris larvae (Chrysopidae) Beef liver, fatty ground beef, whole hen’s eggs—

Perillus bioculatus (nymphs, adults) Picromerus bidens (nymphs, adults) Podisus maculiventris (nymphs, adults)

15.7 Future Recommendation 571

572

15

Mass Production of Insect Predators

References Ahmadi Z, Madadi H (2021) Life table, consumption rate and cold storage of Nabis pseudoferus (Hemiptera: Nabidae), the predator of agricultural crop pests. J Entomol Soc 41(1):73–100 Ali I, Shuai Z, Sajjad A, Jin-Jie C (2018) Artificial diet-based investigation on the impact of purified Cry1Ac, Cry1Fa and Cry2Ab on the survival and reproductive performance of adult green lacewing, Chrysopa pallens (Rambur) (Neuroptera: Chrysopidae). Phytoparasitica 46(1): 127–135 Arijs Y, De Clercq P (2004) Liver-based artificial diets to produce Orius laevigatus. BioControl 49 (5):505–516 Arvaniti KA, Kordas NA, Fantinou AA, Perdikis DC (2021) Impact of prey supply levels on growth performance and optimization of the mass rearing of an omnivorous mirid predator. J Pest Sci 94(3):947–958 Ashfaq M, Abida N, Gulam MC (2002) A new technique for mass rearing of green lacewing on commercial scale. Pak J Appl Sci 2:925–926 Ashfaq M, Nasreen A, Cheema GM (2004) Advances in mass rearing of Chrysoperla carnea (Stephen) (Neuroptera: Chrysopidae). South Pac Stud 24:47–53 Baskaran RK, Sasikumar S, Rajavel DS, Suresh K (2009a) Influence of semi-synthetic diet on Fecundity of Paragus serratus. Ann Plant Prot Sci 17(1):235–236 Baskaran RK, Sasikumar S, Rajavel DS, Suresh K (2009b) Studies on Population fluctuation of guava aphid, Aphis gossypii Glover (Hemiptera: Aphididae) and its predators. J Entomol Res 33(3):247–255 Bigler F (1992) Quality control of mass reared arthropods. In: Proceedings, 5th workshop of the IOBC Global Working Group. Swiss Federal Research Station for Agronomy, Zurich Biondi A, Guedes RNC, Wan FH, Desneux N (2018) Ecology, worldwide spread, and management of the invasive South American tomato pinworm, Tuta absoluta: past, present, and future. Annu Rev Entomol 63:239–258. https://doi.org/10.1146/annurev-ento-031616-034933 Boller EF, Chambers DL (1977) Quality aspects of mass-reared insects. In: Biological control by augmentation of natural enemies. Springer, Boston, pp 219–235 Bonte M, De Clercq P (2008) Developmental and reproductive fitness of Orius laevigatus (Hemiptera: Anthocoridae) reared on factitious and artificial diets. J Econ Entomol 101 (4):1127–1133 Calvo FJ, Lorente MJ, Stansly PA, Belda JE (2012) Pre-plant release of Nesidiocoris tenuis and supplementary tactics for control of Tuta absoluta and Bemisia tabaci in reenhouse tomato. Entomol Exp Appl 143:111–119 Campoy A, Egea-Casas O, Pérez-Bañón C, Rojo S (2022) Effect of cold storage on the pupal development of two pollinators, Eristalinus aeneus and Eristalis tenax. Entomol Exp Appl 170(2):110–121 Castañé C, Zapata R (2005) Rearing the predatory bug Macrolophus caliginosus on a meat-based diet. Biol Control 34(1):66–72 Chambers DL, Ashley TR (1984) Putting the control in quality control in insect rearing. In: King EG, Leppla NC (eds) Advances and challenges in insect rearing agricultural research service. United States department of agriculture, New Orleans, LA, pp 256–260 Chang YF, Tauber MJ, Tauber CA (1996) Reproduction and quality of F1 offspring in Chrysoperla camea: differential influence of quiescence, artificially induced diapause, and natural diapause. J Insect Physiol 42:521–528 Chocorosqui VR, De Clercq P (1999) Developmental and predatory performance of Podisus maculiventris (Say) (Heteroptera: Pentatomidae) reared on a meat-based artificial diet. In Proceedings of the 51st International Symposium on Crop Protection. Med Fac Landbouww Univ Gent 64(3a), 1999, pp 229–234) Cohen AC (1985) Simple method for rearing the insect predator Geocoris punctipes (Heteroptera: Lygaeidae) on a meat diet. J Econ Entomol 78(5):1173–1175

References

573

Cohen AC (1998) Solid-to-liquid feeding: the inside(s) story of extra-oral digestion in predaceous Arthropoda. Am Entomol 44:103–117 Cohen AC, Crittenden P (2004) Deliberately added and “cryptic” antioxidants in three artificial diets for insects. J Econ Entomol 97(2):265–272 Cohen AC, Smith LK (1998a) A new concept in artificial diets for Chrysoperla mfilabris: the efficacy of solid diets. Biol Control 13:49–54 Cohen AC, Smith LK (1998b) A new concept in artificial diets for Chrysoperla rufilabris: the efficacy of solid diets 1. Biol Control 13(1):49–54 Conway HE, Burton KB, Hendrix CA, Burgess LW, Culin JD (2005) Comparison of two different box styles for mass rearing of Sasajiscymnus tsugae (Coleoptera: Coccinellidae), a biological control agent of hemlock woolly adelgid (Hemiptera: Adelgidae) 1. Can Entomol 137(5): 622–629 Coudron TA, Wittmeyer J, Kim Y (2002) Life history and cost analysis for continuous rearing of Podisus maculiventris (Say) (Heteroptera: Pentatomidae) on a zoophytophagous artificial diet. J Eco Entomol 95(6):1159–1168 Crocker RL, Whitcomb WH (1980) Feeding niches of the big-eyed bugs Geocoris bullatus, G. punctipes, and G. uliginosus (Hemiptera: Lygaeidae: Geocorinae). Environ Entomol 9(5): 508–513 Daane KM, Hagen KS, Mills NJ (1998) Predaceous insects for insect and mite control. In: Ridgway RL, Hoffmann MP, Inscoe MN, Glenister CS (eds) Mass-reared natural enemies: application, regulation, and needs. Thomas Say Publications, Lanham, pp 61–115 De Clercq P, Degheele D (1992) A meat-based diet for rearing the predatory stinkbugs Podisus maculiventris and Podisus sagitta (Heteroptera, Pentatomidae). Entomophaga 37:149–157 Dixon AFG, Sato S, Kindlmann P (2016) Evolution of slow and fast development in predatory ladybirds. J Appl Entomol 140(1-2):103–114 Drescher W, Tornier I (1990) Development of a method for mass production of the pollinator and aphid predator Episyrphus balteatus Deg. (diptera, syrphidae). Int Symp Pollinat 288:464–468 Ebrahimifar J, Shishehbor P, Rasekh A, Hemmati SA, Riddick EW (2020) Effects of three artificial diets on life history parameters of the ladybird beetle Stethorus gilvifrons, a predator of tetranychid mites. Insects 11(9):579 Edney EB (2012) Water balance in land arthropods. Springer, Cham El-Arnaouty SA, Abdel-KHalek S, Hassan H, Shahata M, Game M, Mahmoud N (1998) Mass rearing of aphidophagous predators, Chrysoperla carnea and Harmonia axyridis on Ephestia kuehniella eggs. Regional symposium for appl bio-control in Mediterranean countries, Cairo 9:25–29 Eric W, Riddick (2009) Benefits and limitations of factitious prey and artificial diets on life parameters of predatory beetles, bugs, and lacewings: a mini-review. BioControl 54:325–339 Ferkovich SM, Shapiro JP (2004) Increased egg-laying in Orius insidiosus (Hemiptera: Anthocoridae) fed artificial diet supplemented with an embryonic cell line. Biol Control 31 (1):11–15 Finney GL (1948) Culturing Chrysopa californica and obtaining eggs for field distribution. J Econ Entomol 41(5):719–721 George PE (2000) Nymphal cannibalistic behavior of Rhynocoris kumarii Ambrose and Livingstone (Heteropreta: Reduviidae) to prey deprivation and mass rearing conditions. J Biol Control 14(1):1–3 George PE, Ambrose DP (2000) Nymphal cannibalism in reduviids-a constraint in mass rearing. In: Biotechnological applications for IPM. IBH Publication, New Delhi, pp 119–123 Ghanim AA, El-Serafy HA, Hassan HA, Mohamed NE, Abu El Matti FB (2021) Biological aspects of two coccinellid predatory insects reared on artificial diets and natural preys under constant temperature. J Plant Prot Pathol 12(1):19–22 Grenier S, Clercq PD (2003) Comparison of artificially vs. naturally reared natural enemies and their potential for use in biological control. In: Quality control and production of biological control agents: theory and testing procedures. CABI Publishing, Wallingford, pp 115–131

574

15

Mass Production of Insect Predators

Grundy PR, Maelzer DA, Bruce A, Hassan E (2000) A mass-rearing method for the assassin bug Pristhesancus plagipennis (Hemiptera: Reduviidae). Biol Control 18(3):243–250 Gupta T, Chandish R, Ballal A (2013) A mass rearing protocol for the anthocorid predator, Blaptostethus pallescens Poppius. In: 13th IOBC-MRQA Workshop, Bangalore, India, Mass Production & Quality Assurance of Invertebrates S2.O1, pp 32–33 Hemptinne JL, Lecompte E, Sentis A, Dixon AF, Magro A (2022) Prey life-history influences the evolution of egg mass and indirectly reproductive investment in a group of free-living insect predators. Ecol Evol 12(1):e8438 Hill CJ (1989) The effect of adult diet on the biological of butterflies (the common crow butterfly). Oecologia 81(2):258–266 Hodek I, Honêk A (2013) Ecology of coccinellidae, vol 54. Springer, Cham Hodek I, Růžička Z, Sehnal F (1973) Termination of diapause by juvenoids in two species of ladybirds (Coccinellidae). Experientia 29(9):1146–1147 Hodek I, Honek A, Van Emden HF (eds) (2012) Ecology and behaviour of the ladybird beetles (Coccinellidae). Wiley, Hoboken Holling CS (1966) The functional response of invertebrate predators to prey density. Memoirs Entomol Soc Canada 98(48):5–86 Hou RF, Hsiao JH (1979) Antimicrobial agents in semi-synthetic diet for the diamondback moth, Plutella xylostella (L.) (Lepidoptera:Yponomeutidae). Plant Prot Bull 21:251–256 Hung YT, Yang CC, Saska P, Tuan SJ (2021) Comparison of artificial diets and natural prey for mass rearing of Orius strigicollis (Hemiptera: Anthocoridae) using demographic characteristics to optimize cost-efficiency. J Econ Entomol 114(4):1523–1532 Igarashi K, Nomura M (2013) Development and reproduction of Geocoris varius (Hemiptera: Geocoridae) on two types of artificial diet. Appl Entomol Zool 48(3):403–407 Igarashi K, Nomura M, Narita A (2013) Application of a powdered artificial diet to promote the establishment of the predatory bug Geocoris varius (Hemiptera: Geocoridae) on strawberry plants. Appl Entomol Zool 48(2):165–169 Iwai H, Niijima K, Matsuka M (2007) An artificial diet for aphidophagous syrphids, Episyrphus balteatus (de Geer) and Eupeodes bucculatus (Rondani) (Diptera: Syrphidae) using drone honeybee brood powder. Appl Entomol Zool 42(2):167–172 Iwai H, Niijima K, Matsuka M (2009) Improvement of artificial diet for aphidophagous syrphids, Episyrphus balteatus (de Geer) and Eupeodes bucculatus (Rondani) (Diptera: Syrphidae)— additional effects of fatty acids and preservatives. Appl Entomol Zool 44(3):439–446 Jones SL, Kinzer RE, Bull DL, Ables JR, Ridgway RL (1978) Deterioration of Chrysopa carnea in mass culture. Ann Entomol Soc Am 71:160–162 Juan A, Ramos M, Guadalupe Rojas M, Coudron TA, Huynh MP, Zou D, Shelby KS (2023) Artificial diet development for entomophagous arthropods. https://doi.org/10.1016/B978-0-12822106-8.00011-7 Jun HJ, Kim KS, Ham EH (2022) Basic studies aiming at Orius minutus (Hemiptera: Anthocoridae) mass-rearing. Insects 13(1):77 Khan MR, Khan MR (2002) Mass rearing of Menochilus sexmaculatus Fabricius (Coleoptera: Coccinellidae) on natural and artificial diets. Int J Agric Biol 4:107–109 Khanzada MS, Wang S, Huang NX, Pang H, Tan X, Khanzada SR (2019) Optimization of microencapsulated artificial diets for mass rearing of the predacious big-eyed bug, Geocoris pallidipennis. Entomol Generalis 2019:353–363 Kishore R, Manjunath D, Kumar P, Datta RK (1993) Mass production of Australian lady bird beetle, Cryptolaemus montrouzieri Mulsant (Coccinellidae: Coleoptera). Indian J Seric 32(2): 184–188 Kubota T, Shiga M (1995) Successive mass rearing of Chrysopids (Neuroptera: Chrysopidae) on eggs of Tribolium castaneum (Coleoptera: Tenebrionidae). Jpn J Appl Entomol Zool 39(1): 51–58 Kumar SM, Sahayaraj K, Cohen A (2012) Gross morphology and histology of head and salivary apparatus of the predatory bug, Rhynocoris marginatus. J Insect Sci 12(1):1–12

References

575

Leroy PD, Verheggen FJ, Capella Q, Francis F, Haubruge E (2010) An introduction device for the aphidophagous hoverfly Episyrphus balteatus (De Geer) (Diptera: Syrphidae). Biol Control 54(3):181–188 Liu C, Liu H, Wu X, Xiao K, He H, Huang Q, Pu D (2022) Effects of artificial diet on the biological characteristics of the ladybeetle Coccinella septempunctata. Entomol Res 52:223–231 Madadi H, Ahmadi Z and Heshmati A (2016) Biology and life table parameters of predatory bugs Nabis pseduferus (Hemiptera: Nabidae). In: Proceeding of 22nd Iranian Plant Protection Congress, p 584 Mahyoub JA, Mangoud AA, Ghamdi KH, Ghramh H (2013) Method for mass production the seven spotted lady beetle, Coccinella Septempunctata (Coleoptera: Coccinellidae) and suitable manipulation of egg picking technique. Egypt Acad J Biol Sci 6(3):31–38 Mamay M, Mutlu C (2019) Optimizing container size and rearing density for rapid and economic mass rearing of Oenopia conglobata (Linnaeus, 1758) (Coleoptera: Coccinellidae). Turk J Entomol 43(4):395–408 McEwen PK, Kidd NA (1995) The effects of different components of an artificial food on adult green lacewing (Chrysoperla carnea) fecundity and longevity. Entomol Exp Appl 77(3):343– 346 Montoro M, De Clercq P, Overgaard J, Sigsgaard L (2020) Fitness consequences of artificial diets with different macronutrient composition for the predatory bug Orius majusculus. Entomol Exp Appl 168(6-7):492–501 Morales-Ramos JA, Rojas MG, Shapiro-Ilan DI (2013) Mass production of beneficial organisms: invertebrates and entomopathogens. Academic, Cambridge, pp 203–240 Morrison RK, King EG (1977) Mass production of natural enemies. In: Biological control by augmentation of natural enemies. Springer, Boston, pp 183–217 Mukhopadhyay A, Sannigrahi S (1993) Rearing success of a polyphagous predator Geocoris ochropterus (Lygaeidae) on preserved ant pupae of Oecophylla smaragdina. Entomophaga 38(2):219–224 Nordlund DA, Correa JA (1995) Improvements in the production systems for green lacewings: an adult feeding and oviposition unit and hot wire egg harvesting system. Biol Control 5:179–188 Norioka N, Okada T, Hamazume Y, Mega T, Ikenaka T (1984) Comparison of nutritive value of egg yolk and egg white and whole egg. J Biochem 97:19–28 O’Neil RJ, Giles V, Obrycki JJ, Mahr DL, Legaspi JC, Katovich K (1998) Evaluation of the quality of four commercially available natural enemies. Biol Control 11:1–8 Patana R (1982) Disposable diet packet for feeding and oviposition of Lygus hesperus (Hemiptera: Miridae). J Eco Entomol 75(4):668–669 Pehlivan S (2021) Influence of the eggs of Ephestia kuehniella (Lepidoptera: Pyralidae) reared on different diets on the performance of the predatory bug Orius laevigatus (Hemiptera: Anthocoridae). Eur J Entomol 118:51–56 Rana JS, Dixon AFG, Jarošík V (2002) Costs and benefits of prey specialization in a generalist insect predator. J Anim Ecol 71(1):15–22 Ricupero M, Dai C, Siscaro G, Russo A, Biondi A, Zappalà L (2020) Potential diet regimens for laboratory rearing of the harlequin ladybird. BioControl 65(5):583–592 Riddick EW, Chen H (2014) Production of coleopteran predators. In: Mass production of beneficial organisms. Academic, Cambridge, pp 17–55 Riddick EW, Wu Z (2015) Effects of rearing density on survival, growth, and development of the ladybird Coleomegilla maculata in culture. Insects 6(4):858–868 Ridgway RL, Morrison RK, Badgley M (1970) Mass rearing a green lacewing. J Econ Entomol 63(3):834–836 Rojas MG, Morales-Ramos JA, King EG (2000) Two meridic diets for Perillus bioculatus (Heteroptera: Pentatomidae), a predator of Leptinotarsa decemlineata (Coleoptera: Chrysomelidae). Biol Control 17:92–99 Sadat Mahdavi T and Madadi H (2016) Predation and fecundity rates of Nabis pseudoferus Remane fed on cotton aphid. In: Proceeding of 22nd Iranian Plant Protection Congress, p 642

576

15

Mass Production of Insect Predators

Sahayaraj K (2002) Small scale laboratory rearing of a reduviid predator, Rhynocoris marginatus Fab.(Hemiptera: Reduviidae) on Corcyra cephalonica Stainton larvae by larval card method. J Cent Eur Agric 3(2):137–148 Sahayaraj K (2012) Artificial rearing on the nymphal developmental time and survival of three reduviid predators of Western Ghats, Tamil Nadu. J Bio Pesticides 5:218–221 Sahayaraj K, Balasubramanian R (2016) Artificial rearing of reduviid predators for pest management. Springer, Singapore, p 180 Sahayaraj K, Muthukumar S (2011) Zootoxic effects of reduviid Rhynocoris marginatus (Fab.) (Hemiptera: Reduviidae) venomous saliva on Spodoptera litura (Fab.). Toxicon 58(5):415–425 Sahayaraj K, Sujatha S (2011) Temperature-dependent biology & physiology reduviids. Hauppauge, Nova Science, p 146 Sahayaraj K, Martin P, Karthikraja S (2003) Suitable sex ratio for the mass rearing of reduviid predator Rhynocoris marginatus (Fab.). J Appl Zool Res 14(1):34–37 Sahayaraj K, Martin P, Selvaraj P, Raju G (2006) Feeding behaviour of reduviid predators on meat and insect-based artificial diets. Belgian J Entomol 8:55–65 Sattar M, Abro GH (2011) Mass rearing of Chrysoperla carnea (Chrysopidae) adults for integrated pest management programs. Pak J Zool 43:483–487 Shand WA, Shands MK, Simpson GW (1966) Techniques for mass producing Coccinella septempunctata. J Eco Entomol 59(4):1023–1024. https://doi.org/10.1093/jee/59.4.1022 Sighinolfi L, Febvay G, Dindo ML, Rey M, Pageaux J, Baronio P, Grenier S (2008) Biological and biochemical characteristics for quality control of Harmonia axyridis (Pallas) (Coleoptera, Coccinellidae) reared on a liver-based diet. Arch Insect Biochem Physiol 68(1):26–39 Sunil Joshi CR, Ballal, and Lakshmi BL (2013) Development of a novel mass production technique for Brumoides suturalis (Fabricius) (Coleoptera: Coccinellidae), a predator of mealybugs. In: 13th IOBC-MRQA Workshop, Bangalore, India, Mass Production & Quality Assurance of Invertebrates S1.P5, p 29 Tan XL, Wang S, Zhang F (2013) Optimization an optimal artificial diet for the predatory bug Orius sauteri (Hemiptera: Anthocoridae). PLoS One 8(4):e61129 Taylor JR, Schmidt JM (1996) Factors regulating predation by first-instar spined assassin bugs [Sinea diadema (Fabricius)] (Hemiptera: Reduviidae). J Insect Behav 9(1):23–35 Tomson M, Sahayaraj K, Kumar V, Avery PB, McKenzie CL, Osborne LS (2017) Mass rearing and augmentative biological control evaluation of Rhynocoris fuscipes (Hemiptera: Reduviidae) against multiple pests of cotton. Pest Manag Sci 73(8):1743–1752 Tulisalo U (1984) Mass rearing techniques. In: Canard M, Semeria Y, New TR (eds) Biology of chrysopidae. Junk, Boston, pp 213–220 van Lenteren JC (1998) Designing and implementing quality control of beneficial insects: towards more reliable biological pest control. Sting 18(July):2–32 Van Lenteren JC, Bueno VHP, Burgio G, Lanzoni A, Montes FC, Silva DB, Hemerik L (2019) Pest kill rate as aggregate evaluation criterion to rank biological control agents: a case study with Neotropical predators of Tuta absoluta on tomato. Bull Entomol Res 109(6):812–820 Vanderzant ES (1969) An artificial diet for larvae and adults of Chrysopa carnea, an insect predator of crop pests. J Econ Entomol 62(1):256–257 Vieira NF, Truzi CC, Veiga ACP, Sipriano-Nascimento TP, Vacari AM, De Bortoli SA (2018) Life table of Xylocoris afer (Hemiptera: Anthocoridae) feeding on eggs of Corcyra cephalonica (Lepidoptera: Pyralidae) and Plutella xylostella (Lepidoptera: Plutellidae). J Asia Pac Entomol 21(4):1379–1383 Wang R, Nordlund DA (1994) Use of Chrysoperla spp. (Neuroptera: Chrysopidae) in augmentative release programmers for control of arthropod pests. Biocontrol News Inf 15:51–57 Wittmeyer JL, Coudron TA (2001) Life table parameters, reproductive rate, intrinsic rate of increase, and estimated cost of rearing Podisus maculiventris (Heteroptera: Pentatomidae) on an artificial diet. J Econ Entomol 94(6):1344–1352

References

577

Wittmeyer JL, Coudron TA, Adams T (2001) Reproductive development in the predacious spiner soldier bug, Podisus maculiventris (Say) (Heteroptera: Pentatomidae) on artificial diet. Invertebr Reprod Dev 39:9–20 Xie ZN, Wu ZX, Nettles WC Jr, Saldaña G, Nordlund DA (1997) In vitro culture of Trichogramma spp. on artificial diets containing yeast extract and ultracentrifuged chicken egg yolk but devoid of insect components. Biol Control 8(2):107–110 Youssif MAI, Helaly WM (2021) Evaluation of different diets on biological parameters of ladybird beetle, Coccinella undecimpunctata L. (Coleoptera: Coccinellidae). J Plant Prot Pathol 12(5): 357–363 Zapata R, Specty O, Grenier S, Febvay G, Pageaux JF, Delobel B, Castane C (2005) Carcass analysis to improve a meat-based diet for the artificial rearing of the predatory mirid bug Dicyphus tamaninii. Arch Insect Biochem Physiol 60:84–92 Zhang L, Zhao H, Yang Y, Ge Y, Shi W (2022) Effects of nutritional composition of different prey eggs on development and reproduction of the predatory bug, Orius sauteri (Hemiptera: Anthocoridae). J Econ Entomol 115(6):1869–1876. https://doi.org/10.1093/jee/toac159

Bioefficacy of Insect Predators Under Laboratory

16

Contents 16.1 16.2 16.3

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Predators in General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hemipteran Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.1 Anthecoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.2 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.3 Lygaeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.4 Nabidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.5 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.3.6 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.4 Lacewings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.6 Trichoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.7 Storage Pests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.8 Factors Influencing Functional Responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.8.1 Plant Species and Their Morphology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.8.2 Pesticides and Biopesticides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.9 Interaction Multiple Natural Enemies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.10 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

16.1

579 581 582 582 587 590 591 592 593 600 604 608 608 611 612 614 618 619 619

Introduction

Any predator’s ability to exert biological control or bioefficacy under laboratory settings must be evaluated using a crucial technique called functional or numerical response. Because of delays in the predator’s reproductive response, functional and/or numerical interactions between predators and their prey are usually cyclical in nature. In essence, as prey becomes more plentiful, each predator is able to catch more food, and after a while, this food is transformed into additional predators. This primary predator feeding response is known as a functional response, although it is

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_16

579

580

16

Bioefficacy of Insect Predators Under Laboratory

Fig. 16.1 Functional response curves of various sorts, as proposed by Hassell et al. (1976, 1977)

actually understood to be an integral component of the prey’s density-dependent regulatory system. The number of prey consumed by each predator per unit of time rises with prey density until predators are satiated. There are three fundamental types of functional responses. By analysing the coefficients of attack rate and handling time, the functional response curves may be described and distinguished (time spent by the predator in subduing and eating the prey). The handling time and coefficient of attack rate provide estimates of the satiation threshold and the steepness of the increase in predation with increasing prey density, respectively (Cabral et al. 2009). There are various functional response types that might result in a constant (I), decreasing (II), or increasing (III) rate of prey killing, such as a linear increase (Type I), an increase that decelerates to a plateau (Type II), or a sigmoid increase (Type III) (Hassell et al. 1976, 1977) (Fig. 16.1). Given that these curves show how quickly a predator kills its prey at various prey concentrations, they may also be used to gauge how well a predator manages prey populations and predict how effective the predator will be as a biological control agent. According to the predator’s rate of capture and attack (a), a type I response is characterised by a linear rise in consumption (Ne) as prey density (N0) rises, followed by a steady increase in consumption above a threshold density as a result of predator satiation. With the exception of passive feeders like spiders that build webs and species that feed through filters, type I reactions are rare and result in density-independent per capita mortality for prey. A type II functional response results in saturating prey consumption, which causes prey to experience high per capita death at low concentrations and reducing mortality at higher densities. The type II function, which is thought to be the most typical version, takes into consideration processing and ingestion time (h) in addition to a search/attack rate (a)

16.2

Predators in General

581

parameter. A low-density sanctuary from predators, a mid-density peak in per capita mortality, and then dropping mortality due to predator satiation are all characteristics of type III responses, which are sigmoidal in shape. The parameters b and c, which are unique to the type III response, mathematically describe this pattern, which appears when a predator’s capacity to find or consume prey is not constant across prey density (type II and III responses can also be mathematically described using the generalised form, where the magnitude of the scaling exponent q determines the degree to which the curve shifts from decelerating hyperbolic to sigmoidal) (Robert and Kevin 2020). The number of prey assaulted can then be used to forecast the growth, survival, and reproduction of the predator.

16.2

Predators in General

In a lab setting, a functional response research of the eight most frequent arthropod cotton bollworm, Helicoverpa zea, egg predators was carried out. Single predators were subjected to various prey density treatments while being deprived for 24 h. At 6, 12, and 24 h after the start of the feeding trials, the predator response was seen. Hippodamia convergens and Collops quadrimaculatus adults and Chrysopa oculata larvae demonstrated the highest consumption rates (116, 85, and 119 eggs/24 h, respectively) at the highest prey density (150 eggs per predator), followed by Hippodamia convergens larvae (51 eggs/24 h), adult Geocoris punctipes (Say) (45 eggs/24 h), adult Scymnus loe While adult Collops quadrimaculatus, Geocoris punctipes, Hippodamia convergens, and larval Hippodamia convergens and Chrysopa oculata had type 2 functional response, adult Notoxus spp., N. capsiformis, and O. insidiosus did not. Bollworm eggs were devoured by all predators most frequently at 35 °C and least frequently at 15 °C; the predation rate at 35 °C was four times higher than that at 15 °C (Parajulee et al. 2006). Each predator was tested against four densities of Halyomorpha halys (Hemiptera: Pentatomidae) eggs: 26, 52, 78, and 104 eggs. The predators were Acheta domesticus (Orthoptera: Gryllidae), Melanoplus femurrubrum (Orthoptera: Acrididae), Orius insidiosus (Hemiptera: Anthocoridae), and when Halyomorpha halys egg densities are low, only Acheta domesticus showed a Type II functional response with non-negative estimates of handling time and attack rate, indicating the possibility of a density-dependent mortality factor. For female Acheta domesticus, the predicted maximum predation rates were 189 eggs (95), or nearly seven egg masses. The largest number of eggs for males was 116 (35), or 4.5 egg masses. The remaining predators showed a Type I functional response and are unlikely to act as a stabilising factor in the population dynamics of Halyomorpha halys (Poley et al. 2018). Prey density is the primary factor influencing the functional response of predators; however, there are also other parameters that correlate with prey density, including as temperature, humidity, and photoperiods. For instance, the first, second, and third instars of the prey Spodoptera litura were examined by larvae and adults of Harmonia axyridis in a variety of temperatures (15, 20, 25, 30, and 35 °C). Using

582

16

Bioefficacy of Insect Predators Under Laboratory

logistic regression, the kind and parameters of the functional responses were identified and fitted to the Roger’s random predator equation. Predation intensity varied depending on the predator and prey stages, but it got worse as it got warmer and the predator got older. The female and fourth instar of the predator Harmonia axyridis preyed more on the first instar of the prey, followed by the second and third instars of the prey Spodoptera litura. For the first, second, and third instars of Harmonia axyridis, there was no predation on the larger prey. Across all temperatures and the three different species of prey the larvae and adult H. axyridis devoured, a type II (hyperbolic) functional response curve was produced. The maximum predation rates, handling times, and space clearance rates of H. axyridis varied with temperature and prey size, rising with increasing temperature and falling with decreasing prey size, indicating greater predation will occur on younger prey. This work suggests that temperature and prey/predator size combine to shape functional responses, which may make it more difficult to devise efficient biocontrol techniques to control this dangerous insect (Islam et al. 2022).

16.3

Hemipteran Predators

16.3.1 Anthecoridae As a potential biocontrol agent for Megalurothrips sjostedti, a significant pest of Kenyan French beans, Phaseolus vulgaris, the functional response of Orius albidipennis, was assessed. At temperatures of 15, 25, and 28 °C and concentrations of 5, 10, 20, and 30 larval and adult Megalurothrips sjostedti per cage, over a 24-h period, the functional response of adult Orius albidipennis to these pests was examined. At larger densities and higher temperatures, more Megalurothrips sjostedti adults and larvae were killed. Both the type I and type II functional response models were well fitted by the data. Both the second instar larva and the adult Megalurothrips sjostedti attacked at higher rates as the temperature rose. The handling times for the adults increased as the temperature rose, but those for the larvae dropped as the temperature rose. The functional responses of various Orius sauteri stages to Thrips palmi densities were investigated. The majority of Orius sauteri stages had the Holling Type-II reaction. The amount of Thrips palmi that Orius sauteri nymphs ate grew as the stage of the species progressed; however, late fifth-stage nymphs devoured a lot fewer Thrips palmi larvae than early fifth-stage nymphs. Thrips palmi larvae were slightly less frequently ingested by Orius sauteri adult females than by early fifthstage nymphs. Male adult Orius sauteri devoured less Thrips palmi larvae than female adult Orius sauteri did. More Thrips palmi larvae were obliterated by Orius sauteri nymphs than by Thrips palmi adults. The two stages of Thrips palmi that adult females of Orius sauteri killed had no discernible differences from one another. With rising temperatures, adult female Orius sauteri Thrips palmi consumption increased. The preference for Thrips palmi adults and larvae varied as Orius

16.3

Hemipteran Predators

583

sauteri developed, according to research on selective predation on various stages of Thrips palmi (Nagai and Yano 2000). Three-day-old adult females of Megalurothrips usitatus and Orius sauteri, both Hemiptera: Anthocoridae, were pitted against one another (10, 20, 40, 60, 80, 100, and 120 prey per predator, respectively). The maximum predation rate was 45.3 over a 24-h period, and the functional response of Orius sauteri to increasing Megalurothrips usitatus density was described by Holling’s disc equation. Orius sauteri’s intraspecific interference became more pronounced as its density rose. Orius sauteri, with an intrinsic rate of increase (r) of 0.16 and fecundity of 95.4 eggs per female, was able to complete its life cycle by eating on Megalurothrips usitatus. Over the course of their lifetimes, female and male minute pirate bugs each ingested an average of 304.7 and 104.0 thrips. According to these findings, Orius sauteri may be a useful biological control agent for Megalurothrips usitatus in integrated pest management. In China, cowpea is severely harmed by the bean flower thrips, Megalurothrips usitatus (Thysanoptera: Thripidae). In this work, the tiny pirate bug Orius sauteri (Hemiptera: Anthocoridae), which feeds on Megalurothrips usitatus, had its predation functional response and life table parameters examined in a lab. The maximum predation rate was 45.3 over a 24-h period, and the functional response of Orius sauteri to rising Megalurothrips usitatus density was characterised by Holling’s disc equation. Orius sauteri’s intraspecific interference became more pronounced as its density rose. Orius sauteri, with an intrinsic rate of increase (r) of 0.16 and fecundity of 95.4 eggs per female, was able to complete its life cycle by eating on Megalurothrips usitatus. Over the course of their lifetimes, female and male minute pirate bugs each ingested an average of 304.7 and 104.0 thrips. Based on these findings, Orius sauteri may be a useful biological control agent for Megalurothrips usitatus in integrated pest management (Liu et al. 2018). Megalurothrips sjostedti was examined as a candidate for biocontrol by Orius albidipennis. In Kenya, trybom is a significant pest of Phaseolus vulgaris, or French beans. At temperatures of 15, 25, and 28 °C and densities of 5, 10, 20, and 30 larval and adult M. sjostedti per cage, over a 24-h period, the functional response of adult Orius albidipennis to these organisms was examined. At larger densities and higher temperatures, more Megalurothrips sjostedti adults and larvae were killed. Both the type I and type II functional response models were well fitted by the data. Both the second instar larva and the adult Megalurothrips sjostedti attacked at higher rates as the temperature rose. The handling times for the adults increased as the temperature rose, but those for the larvae dropped as the temperature rose. These results’ consequences are examined (Gitonga et al. 2002). The two-spotted spider mite, Tetranychus urticae (Acari: Tetranychidae), is primarily controlled by Orius niger and Orius minutus, while the onion thrips, Thrips tabaci (Thysanoptera: Thripidae), is an important field pest of potatoes in the Ardabil region of Iran. In a controlled experiment, Fathi and Nouri-Ganbalani (2010) compared the functional responses of Orius niger and Orius minutus to female mites and second-instar thrips larvae at various prey densities (5, 10, 20, and 40 prey/arena). The resulting data were correctly fitted to Type II functional

584

16

Bioefficacy of Insect Predators Under Laboratory

Table 16.1 Functional response of the predator Anthecoride to various pests, with references Predator Orius albidipennis Orius albidipennis Orius albidipennis Orius albidipennis Orius sauteri Orius sauteri

Pest Megalurothrips sjostedti Bemisia tabaci egg

Response Type I and II Type II

References Gitonga et al. (2002) Shahpouri et al. (2019)

Bemisia tabaci nymphs

Type III

Shahpouri et al. (2019)

Type I and II Type II Type II

Gitonga et al. (2002)

Orius sauteri Orius niger

Megalurothrips sjostedti Thrips palmi Megalurothrips usitatus Spodoptera frugiperda Tetranychus urticae

Type II Type II

Orius niger

Thrips tabaci

Type II

Orius minutus

Tetranychus urticae

Type II

Orius minutus

Thrips tabaci

Type II

Orius laevigatus Orius laevigatus Orius vicinus Orius vicinus

Bemisia tabaci Tetranychus urticae Bemisia tabaci Tetranychus urticae

Type II Type II Type II Type II

Di et al. (2021) Fathi and Nouri-Ganbalani (2010) Fathi and Nouri-Ganbalani (2010) Fathi and Nouri-Ganbalani (2010) Fathi and Nouri-Ganbalani (2010) Pehlivan et al. (2020) Pehlivan et al. (2020) Pehlivan et al. (2020) Pehlivan et al. (2020)

Nagai and Yano (2000) Liu et al. (2018)

response models for four predator–prey interactions, including Orius niger to second instar thrips larvae (a = 0.009 h1; and Th = 1.62 h), Orius niger to female mites (a = 0.006 h1 and Th = 1.28 h), Orius minutus to second instar thrips larvae (a = 0.008 h1 and The number of female mites attacked by Orius minutus was higher than that by Orius niger. However, the number of second-instar thrips larvae attacked by Orius niger was high. These findings demonstrate that Orius niger and Orius minutus can both contribute significantly to a biological control programme against onion thrips and the two-spotted spider mites that are infesting potato farms in this area (Fathi and Nouri-Ganbalani 2010). Few anthecrids exhibited type I or type III curves, but the majority displayed type II functional response (Table 16.1). Under laboratory conditions, the predatory abilities of Blaptostethus pallescens and Anthocoris muraleedharani towards the papaya mealybug, Paracoccus marginatus, and the cotton mealybug, Phenacoccus solenopsis, were investigated in India (Ballal et al. 2012). A. muraleedharani could eat 66 Phenacoccus solenopsis crawlers in total during the nymphal stage, and 141 crawlers during the adult stage. Young B. pallescens nymphs (3–4 days old) could not eat P. solenopsis crawlers, while mature nymphs could eat 35 Phenacoccus solenopsis crawlers and adult nymphs could eat 23 crawlers. The feeding potential for Blaptostethus pallescens on Paracoccus marginatus was determined to be 18, 29, and 31 crawlers for young

16.3

Hemipteran Predators

585

nymphs, mature nymphs, and adults, respectively. However, feeding on Paracoccus marginatus was observed to severely shorten the lifespan of Blaptostethus pallescens. Paracoccus marginatus could not be predated upon by Anthocoris muraleedharani. Anthocoris muraleedharani appeared to be a more ferocious predator of Phenacoccus solenopsis based on its better predatory potential, higher adult longevity, and shorter nymphal length despite the fact that B. pallescens could feed on both mealybug species (Ballal et al. 2012). Orius albidipennis versus Bemisia tabaci (5, 8, 10, 15, 20, 25, 30, and 35 eggs and/or third instar nymphs) (Hemiptera: Aleyrodidae). The findings demonstrated that, when fed on third instar nymphs and whitefly eggs, respectively, Orius albidipennis exhibited type II and III functional responses. The handling times (Th) of O. albidipennis, when fed on eggs and third-instar nymphs of Bemisia tabaci, were calculated to be 0.35 and 1.01 h. The bug’s assault rate (a) on eggs was calculated to be 0.0723 h-1, and its constant (b) when feeding on third-instar whitefly nymphs was 0.0069. The calculation yielded 68.39 eggs and 23.20 third instar nymphs as the maximum attack rate (T/Th). Conclusion: Orius albidipennis was successful in containing Bemisia tabaci (Shahpouri et al. 2019). In a lab setting, the predatory bugs Orius laevigatus and Orius vicinus were tested for their functional reactions to various egg densities of the whitefly, Bemisia tabaci (Hemiptera: Aleyrodidae), and red spider mite, Tetranychus urticae (Acari: Tetranychidae). Females of the aforementioned predators were fed various amounts of eggs (2, 4, 6, 8, 16, 32, 64, and 128) from both species of prey for 24 h in a controlled setting. Utilising Holling’s Disc Equation, the functional response parameters were evaluated. To both prey, both predators displayed a Type II response. Orius laevigatus and Orius vicinus were used to calculate the attack rates (a) and handling times (Th) of the predators for spider mite eggs, while Orius laevigatus and Orius vicinus were used for whitefly eggs (a: 0.772 Th: 0.006). In addition, female Orius laevigatus devoured an average of more Bemisia tabaci eggs than did male Orius laevigatus. Orius laevigatus, on the other hand, was a more effective predator of Tetranychus urticae eggs. As a result, these findings suggest that these predators could work as a team to control populations of T. urticae and Bemisia tabaci in agricultural settings (Pehlivan et al. 2020). In another study, by feeding them with various densities of thrips larvae of Frankliniella schultzei, three anthocorid bugs, namely Blaptostethus pallescens, Cardiastethus affinis, and Montandoniola indica (Heteroptera: Anthocoridae), functional response was evaluated. All three species’ predation rates increased as prey numbers rose. Montandoniola indica displayed type-III reaction, while Blaptostethus pallescens and Cardiastethus affinis both demonstrated type-II response. Montandoniola indica, which was followed by B. pallescens in terms of effectiveness among the predators, had the highest predation rate, highest 1/Th, highest a/Th, and shortest handling time (Varshney et al. 2020). Regarding ingestion of Spodoptera frugiperda, similar experiments were conducted on Orius sauteri (Hemiptera: Anthocoridae) and Harmonia axyridis (Coleoptera: Coccinellidae). According to experimental findings, the predatory activities of Orius sauteri and Harmonia axyridis nymphs towards Spodoptera

586

16

Bioefficacy of Insect Predators Under Laboratory

frugiperda eggs and larvae suit Holling’s Type II functional response model. Importantly, the parameters on Spodoptera frugiperda first instar larvae were 700.24, 0.5602, and 0.0008 days, respectively, and the theoretical maximum number of prey consumed per day (Na-max), the instantaneous attack rate (a′), and the handling time (Th) of Orius sauteri nymphs on Spodoptera frugiperda eggs were 15.19, 0.7444, and 0.049 days, respectively. Another potential biocontrol agent for this pest is Orius sauteri. Our research offers a theoretical foundation for using natural enemies to manage Spodoptera frugiperda in the field. More study is needed to develop field-based conservation biological control strategies that will boost predator populations and Spodoptera frugiperda reduction (Di et al. 2021). Plant phenology is a significant additional component that controls the functional response. Various plant species from different groups have phenological characteristics that control insect populations. The predators that lived in agroecosystems had high bioefficacy. For instantance, two common wheat cultivars (Falat and Pishtaz) with different leaf morphological features were used to examine the impact of physical plant features on life table characteristics and functional response of Orius albidipennis females feeding on Schizaphis graminum (Rondani). Numerous characteristics relating to Orius albidipennis performance were significantly influenced by the trichome density. Despite exhibiting type III functional responses on both cultivars, Pishtaz had a lower searching efficiency and a longer handling time than Falat. On Falat, the maximum attack rate was higher than on Pishtaz. The Orius predators’ reduced maximum predation and longer handling times on Pishtaz leaves may be attributed to the plant’s leaves having much more surface trichomes than Falat leaves, which physically impeded bug mobility and reduced prey encounter rate. Moreover, the predatory bug’s foraging behaviour was adversely affected by an increase in trichome density. Additionally, it was discovered that females favour leaves with less trichomes as oviposition hosts. Trichomes appeared to be a barrier for the Orius bug. Compared to Pishtaz cultivar, Falat cultivar has a much higher intrinsic rate of rise (rm). Similar to this, Falat had a higher net reproduction rate and finite rate of rise than Pishtaz. To summarise, it may be said that cultivars with lower trichome density may benefit more from Orius albidipennis ability to suppress S. graminum on wheat (Gholami et al. 2022). The results of the current study make it clear that anthercorid predators could be utilised in thrips biological control programmes as a supplement.

16.3.1.1 Numerical Response Barman and Ghosh (2022) provide the RM (Rosenzweig-MacArthur) model, the BD model (RM type model with Beddington-DeAngelis functional response), the RMI model (RM model with intraspecific rivalry among predators), and the BDI model as our four proposed predator–prey models (BD model with intraspecific competition among predators). Each model includes a time delay in the numerical reaction of the predators. First, we examine each model’s delay-induced stability. We demonstrate that a coexisting stable equilibrium in RM and BD models is always unstable as the delay increases. In RMI and BDI models, a stable equilibrium is not always made unstable by an increase in latency, though. In fact, in the latter two models, the stable

16.3

Hemipteran Predators

Table 16.2 Numerical response of Rhynocoris kumarii on Mylabris indica under laboratory condtion

587

No. of predtor One Two Four Eight

Predtor rate (number per day) 24 h 48 h 72 h 0.23 0.17 0.08 0.40 0.27 0.28 0.81 0.43 0.67 0.88 0.85 0.72

96 h 0.15 0.22 0.48 0.77

equilibrium may also continue to be stable due to variable delay. One of the main findings is that intraspecific competition has an impact on the local stability’s invariance property in the RMI and BDI models. We demonstrate analytically that stability switching cannot happen in any of the models. Later, we employ separate prey and predator harvesting, which could lead to stability switching. If populations fluctuate in an unharvested system, significant effort may be necessary to maintain equilibrium. Prey harvesting and predator harvesting may result in opposing dynamic modes under the same natural circumstance (an unharvested environment). Research on the numerical response of three harpactorine assassin bugs, Rhynocoris kumarii, Rhynocoris marginatus, and Rhynocoris fuscipes, revealed that they had a positive numerical response by killing more prey in terms of the population of accessible prey per predator at a given time (Ambrose 1999, 2003). Predator survival was put to the test when prey density increased, decreasing searching ability and increasing discovery time, attack rate, and fecundity (Table 16.2).

16.3.2 Pentatomidae On potato plants, newborn Leptinotarsa decemlineata larvae, were being attacked by fifth instars of the hemipteran predator Perillus bioculatus. One to 100 preys per plant were observed, and predation rates were calculated by fitting conventional functional response models to the data. The functional response curves of type II and type III did not adequately fit the data. The observed attack rate did not increase with prey density, as predicted by the type II model, nor was it constant across all prey densities (as in the type III model). Instead, it decreased over a prey density range of 1–25, then significantly rose at a prey density of 50. An explanation for the observed pattern could be that additional successful feeding events occurred at high concentrations but not at low densities as a result of area-restricted search after prey consumption. Prey density had no discernible effect on per-prey handling times, and the mean observed handling time was essentially equivalent to an estimate based on the type III model (Heimpel and Hough-Goldstein 1994). In a lab setting and a potato field, researchers assessed the predator Podisus maculiventris’ functional reaction and search tactics as it attacked third-instar Leptinotarsa decemlineata, a Colorado potato beetle. Predators in petri dish arenas in the lab attacked a maximum of 7.43 victims in a 24-h period as prey density grew.

588

16

Bioefficacy of Insect Predators Under Laboratory

As prey density rose, estimates of the area searched decreased. The cumulative amount of time predators spent handling prey may be to blame for the drop in search. The predator’s functional response in the wild differed from that observed in the lab. No matter how many prey were available, predators in the wild attacked a roughly consistent number of them. Similar to the laboratory, when prey density rose, the predator’s estimated search area shrank. However, because predators in the outdoors attacked so few prey, as opposed to in the laboratory, the drop in search could not be explained by the length of time that predators had spent handling prey. The relationship between predator search and prey density was explained using a model of predator search in the wild. Results are contrasted with comparable measurements of Podisus maculiventris and other predator species’ assault rates and search tactics in soybeans, a different crop. The same search pattern and intensity of attack were seen regardless of the predator, victim, or crop. With relation to the use of predators in biological control, it is described how attack rates and search patterns are similar across different predator–prey systems as well as the differences between measurements made in a lab and in the field of a predator’s functional response (see O’Neil 1997). Previously, in 1991, Wiedenmann and O’Neil (1991) carried out laboratory studies to gauge the Podisus maculiventris’ functional response in a condensed arena. In 9-cm Petri dishes with 1, 2, 4, 8, 12, or 16 Mexican bean beetle thirdinstar larvae, predators were kept apart for 24 h (Epilachna varivestis). When 16 preys were offered, the results revealed a typical type II response, with up to nine preys being attacked. In order to determine handling time and the maximum number of victims that may be assaulted in a 24-h period, the data were fitted with Holling’s disc equation. The findings diverged significantly from earlier field measurements of Podisus maculiventris predation, showing that attack rates in a straightforward laboratory setting are constrained by distinct behaviours than are significant in the field. In order to use the data to anticipate the effect of the predator on field populations of prey, we therefore propose that assessing functional response must be done in an appropriate test arena (Wiedenmann and O’Neil 1991). In a lab setting, the predatory stinkbugs Podisus maculiventris and Podisus nigrispinus were observed preying on the tomato looper Chrysodeixis chalcites. Both Podisus species’ nymphs readily attacked the prey’s eggs, larvae in their second, fourth, and sixth instars, as well as pupae. On fourth- and sixth-instar caterpillars, predator nymph development times were, however, often shorter. Although they did not eat eggs, adult females showed high predation rates against Chrysodeixis chalcites larvae and pupae. Podisus maculiventris predation rates at 23 °C were often comparable to those of the smaller P. nigrispinus. Podisus maculiventris and Podisus nigrispinus adult females’ functional reactions to the abundance of fifth-instar Chrysodeixis chalcites caterpillars on sweet pepper plants closely resembled Holling’s type II response. For Podisus maculiventris, the estimated attack rate and handling time were 0.057 h-1 and 4.71 h, and for Podisus nigrispinus, they were 0.046 h-1 and 4.37 h. Additionally, in two greenhouse tests, Podisus maculiventris nymphs were introduced to battle tomato looper caterpillars on sweet pepper plants. Fourth-instar caterpillar populations were reduced by 40% in

16.3

Hemipteran Predators

589

Table 16.3 Pentatomidae predator’s response against various pests with references Predator Eocanthecona furcellata Perillus bioculatus Podisus maculiventris Podisus maculiventris Podisus nigrispinus Podisus maculiventris Podisus nigrispinus Podisus nigrispinus

Pest Latoia lepida Leptinotarsa decemlineata Leptinotarsa decemlineata Chrysodeixis chalcites Chrysodeixis chalcites Epilachna varivestis Spodoptera frugiperda Chrysomya putoria

Maximum prey attacked 91 3 prey/predator/h

References Senrayan (1988)

7.43

Heimpel and HoughGoldstein (1994) O’Neil (1997)

40%

De Clercq et al. (1998)

65%

De Clercq et al. (1998)

9.0 3.8

Wiedenmann and O’Neil (1991) Zanuncio et al. (2008)

86%

Botteon et al. (2017)

48 h and leaf-feeding damage after 1 week when fourth-instar nymphs of the predator were introduced at a predator: prey ratio of 1:3.3 (De Clercq et al. 1998). When offered with caterpillars of Latoia lepida, a sporadic pest of numerous agricultural crop plants, Eocanthecona furcellata showed similar outcomes. The findings (91, 68, 6, 51, 6, 42, 2, and 344 from the first to the fifth instar prey) show the significance of prey density and defence in affecting the assault reaction of the predator (Senrayan 1988). Female Podisus nigrispinus predator predation rates were examined in relation to the defence of the prey, Spodoptera frugiperda (Lepidoptera: Noctuidae). With densities of one, two, four, six, and eight larvae, respectively, Podisus nigrispinus preyed on third instars of Spodoptera frugiperda that were 1.0, 1.4, 1.2, 3.8, and 3.0 and 0.4, 0.8, 1.6, 2.8, and 3.2. The attack rate and manipulation time for Podisus nigrispinus females fed with larvae with defence were 0.67 0.39 and 6.72 2.88 h, respectively, and 2.51 0.16 and 0.51 0.77 h for those without protection. Predator Podisus nigrispinus’s functional response varies with prey Spodoptera frugiperda defence and density, with more eating of prey without defence at higher densities (Zanuncio et al. 2008). According to reviewed data, Pentatomidae predators responded to a variety of pests by reducing at least 68% of their prey (Table 16.3). As an alternate food source, Chrysomya putoria larvae (Diptera: Calliphoridae) were consumed by Podisus nigrispinus. Using ecological modelling and the Leslie matrix population model, the demographic factors of fecundity and survival were examined in the various life stages of the Podisus nigrispinus, resulting in histograms of the various life stages across time. Additionally, the functional response of Podisus nigrispinus was examined at 24 and 48 h on seven densities of third-instar Chrysomya putoria larvae. The development period from egg to adult for predators was 23.15 days, and their adult survival rate was 65%. At 24 and 48 h,

590

16

Bioefficacy of Insect Predators Under Laboratory

the predator had a type III functional response to Chrysomya putoria ingestion. The prey proved practical for use in the laboratory rearing and maintenance of Podisus nigrispinus thanks to the Leslie-matrix modelling of the age structure that allowed perpetuation of the predator population over time steps (Botteon et al. 2017).

16.3.3 Lygaeidae Geocoris punctipes, a predaceous lygaeid that had not previously been identified as a whitefly predator, was examined in the lab as a potential natural adversary of Bemisia tabaci, the sweet potato whitefly. With whiteflies as prey, it exhibited stalking behaviour akin to that seen with aphids. A hitherto unrecognised behaviour was discovered in which the predators attached the prey’s wings to various surfaces using salivary secretions, enabling labial probing and feeding. The pattern of the Holling type II functional response appeared to be followed by prey consumption as a function of prey number. The handling period per prey item was between 180 and 240 s. The handling time spent on early versus later catches did not differ. The nutritional value of whiteflies was assessed using the following parameters: crude protein, lipids, and carbs. Geocoris punctipes, a predaceous lygaeid not previously known to be a whitefly predator, was tested in the lab as a promising candidate for biological control of sweet potato whiteflies. Performance in terms of predator behaviour, total daily handling time, functional response, energy budget, and nutritional quality all support the hypothesis that Geocoris punctipes is a promising candidate for biological control of whiteflies (Cohen and Byrne 1992). Approximately 100 corn earworm eggs and 125 beet armyworm neonates, on average, were devoured by Geocoris floridanu nymphs in the corresponding prey treatments. Predation rates on Heliothis zea eggs and Spodoptera exigua neonates were comparable from the first to the third instar of Geocoris floridanu nymphs; however, during the fourth and fifth instars, predation on Spodoptera exigua larvae was stronger than on maize earworm eggs. There were no changes in the amount of prey that male and female predators ingested within the prey treatments. Geocoris floridanu female and male bugs ingested more Spodoptera exigua neonate larvae than Heliothis zea eggs over the course of their preimaginal development (Torres et al. 2004). The outcomes of Rehman et al. (2020) demonstrated a type II functional response for every stage of O. strigicollis’ life when fed each species of whitefly. When given T. vaporariorum rather than B. tabaci nymphs, O. strigicollis of various life stages required less time to handle their meal. When fed B. tabaci, O. strigicollis nymphal growth was noticeably slower than that of T. vaporariorum nymphs. Additionally, feeding adult females B. tabaci nymphs as opposed to T. vaporariorum nymphs resulted in a statistically shorter overall pre-oviposition period. Additionally, when given B. tabaci rather than T. vaporariorum, O. strigicollis had greater survival rates and overall fecundity. When fed either whitefly species, there were no appreciable variations in any of the population metrics of O. strigicollis. Geocoris punctipes, a generalist predator, is a useful natural foe of the Lygus bug. In the current study, Gómez-Domínguez et al. (2021) used laboratory tests to

16.3

Hemipteran Predators

591

ascertain if Geocoris punctipes engaged in intra-guild predation of P. relictus while growing in Lygus lineolaris Palisot de Beauvois nymphs. Geocoris punctipes preyed more on first and second instar Lygus lineolaris nymphs than third and fourth instar nymphs in choice and no-choice studies. Geocoris punctipes consumed Lygus lineolaris nymphs with P. relictus eggs more frequently than nymphs with P. relictus larvae in both choice and no-choice studies, but no nymphs with P. relictus pupae. In experimental arenas, Geocoris punctipes and P. relictus released simultaneously had an additive effect on Lygus lineolaris mortality, killing 90%, 80%, and 70% of first, second, and third instar nymphs, respectively. However, more investigation into the interactions between these two natural enemies in commercial strawberry fields is required before any concrete pest control suggestions can be provided (Gómez-Domínguez et al. 2021). Geocoris punctipes, a generalist predator, preyed on P. relictus inside its own guild when it was still a nymph of Lygus lineolaris Palisot de Beauvois. Geocoris punctipes preyed on more first and second instar nymphs of L. lineolaris than third and fourth instar nymphs in choice and no-choice studies. Geocoris punctipes consumed Lygus lineolaris nymphs with P. relictus eggs more frequently than nymphs with P. relictus larvae in both choice and no-choice studies, but no nymphs with P. relictus pupae. In experimental arenas, Geocoris punctipes and P. relictus released simultaneously had an additive effect on Lygus lineolaris mortality, killing 90, 80, and 70% of first, second, and third instar nymphs, respectively. However, more investigation into the interactions between these two natural enemies in commercial strawberry fields is required before any concrete pest control suggestions can be provided (Gómez-Domínguez et al. 2021). Based on these findings, Geocoris specie could potentially function as a biological control agent in integrated pest management because it could live and maintain populations on both species of whiteflies (IPM).

16.3.4 Nabidae The predator Nabis pseudoferus exhibits sit-and-wait behaviour and is a generalist. Although Nabidae family insects are widespread generalist predators in agroecosystems, little is known about how often they prey on fodder crop pests. Agallia constricta (Hemiptera: Cicadellidae) and Ceratagallia agricola (Hamilton), two common leafhopper pests of red clover (Trifolium pratense Fabaceae), were studied by Stasek et al. (2018) to establish the functional response and prey preference of Nabis Latreille species (Hemiptera: Cicadellidae). We also discovered how long A. constricta survived after being attacked by Nabis species over the course of 5 days. With a preference for A. constricta, the Nabis species showed a Type III functional response to both leafhopper species. After 5 days, the 10-A. constricta/ cage treatments and the 20-A. constricta/cage treatments had the highest survival probabilities (0.19 and 0.23, respectively). These findings suggest that Nabis species may aid in reducing leafhopper populations in systems involving forage and crops (Stasek et al. 2018).

592

16

Bioefficacy of Insect Predators Under Laboratory

16.3.5 Miridae A Holling’s type-II model provided a good description of the functional responses of the mirid Cyrtorhinus lividipennis to both the egg and the first-instar nymph of BPH (Nilaparvata lugens). The greatest number of BPH eggs consumed daily by the female, male, and third-instar nymph of Cyrtorhinus lividipennis were 22, 18, and 6 eggs, respectively. Depending on the mirid age, the handling time h and the instantaneous search rate an altered. The mirid Cyrtorhinus lividipennis preferred BPH eggs to nymphs (Sivapragasam and Asma 1985). In greenhouse tomato without the use of pesticides, the predator Tupiocoris cucurbitaceus is commonly observed feeding on whiteflies in Argentina. This study assessed the nymphs of Bemisia tabaci, eggs and larvae of Tuta absoluta, nymphs of Myzus persicae, and nymphs of Tetranychus urticae for a period of 24 h. It also measured the consumption of males, females, large and tiny nymphs, and nymphs. Whitefly intake ranged from 15.7 to 38.2 nymphs per predator, with the females of Tupiocoris cucurbitaceus consuming the most eggs (147.4) and nymphs of Myzus persicae and Tetranychus urticae (19.8). Individual Tetranychus urticae preyed on 3.3 tiny larvae, but all predator stages had insignificant predation on large larvae. Finally, 27.8 Tetranychus urticae females were eaten by each predator. This knowledge expands our understanding of the prey range and feeding preferences of this predator (López et al. 2019). In horticultural crops, dicyphine mirids are crucial biological control agents (BCAs). In Portugal, protected tomato crops are home to Dicyphus cerastii Wagner, which has been seen feeding on a variety of tomato pests. The ability of this species to prey is little understood, though. We assessed the functional response (FR) and predation rate of female predators on various densities of four prey species—Myzus persicae first instar nymphs (large mobile prey), Bemisia tabaci fourth instar nymphs, Ephestia kuehniella eggs (large immobile prey), and Tuta absoluta eggs—in order to investigate the predation capacity of Dicyphus cerastii and (small immobile prey). Tomato leaflets were the subject of experiments that lasted for 24 h in Petri dish arenas. All of the studied prey showed type II FR in Dicyphus cerastii. By ingesting an average of 88.8 B. tabaci nymphs, 134.4 E. kuehniella eggs, 37.3 M. persicae nymphs, and 172.3 T. absoluta eggs, the predator successfully preyed upon all prey. Prey size and mobility may have an impact on a predator’s ability to catch prey, according to differences in the FR parameters for attack rate and handling time. Dicyphus cerastii emerged as an intriguing potential BCA for tomato crops in light of the extremely high predation rates discovered for all prey species (Abraços-Duarte et al. 2021). Both female adults and fifth instar Frankliniella schultzei, a polyphagous pest that affects numerous crops in polyhouses in India, was the target of a type III functional response in Dortus primarius (Hemiptera: Miridae). Both the fifth instar and female adults ingested 21.6 and 28.6 thrips larvae per day, respectively, at the greatest prey density (40). Attack rate, handling time, a/Th, and T/Th functional response metrics demonstrated this predator’s effectiveness against Frankliniella schultzei (Varshney and Budhlakoti 2022).

16.3

Hemipteran Predators

593

16.3.6 Reduviidae Females of two species of reduviids, Sinea confusa and Zelus renardii, were observed as they fed, and the effect of relative predator–prey weight ratios on eating behaviour was examined. We measured handling time and extracted biomass of prey throughout a wide range of predator–prey weight ratios using 10–140 mg moth larvae, Heliothis virescens (provided in 7 weight groups) as prey (0.3–4.5:1). As the ratio of predators to prey rose, handling time dropped exponentially and extracted biomass increased linearly. Handling times for both species were around 100 min at predator–prey ratios of about 1:1. The amount of biomass that could be extracted rose linearly as prey weight increased, however these increases weren’t proportionate to rising prey weights or rising handling durations. For both species, the average rate of consumption was 110 g/min. S. confusa greatly outperformed Zelus renardii in terms of relative consumption rate. We put out the idea of the major investor strategy for predators that can consume reasonably large prey through extra-oral digesting. As can be observed, major investors make significant time and material investments in each huge prey item. They then have to harvest a sizable nutrient return from each prey before moving on to assault a new prey. Major investors would not follow functional response kinetics with large prey as a result (Cohen and Tang 1997). Acanthaspis pedestris, a reduviine assassin bug, is a predator of several insect pests. The preference of Acanthaspis pedestris for five different species of grasshoppers (Orthoptera: Acrididae), including Tylotropidus variecornis, Orthacris maindroni, two unidentified slant-faced grasshoppers, and one unidentified coneheaded grasshopper, as well as its stage preference for Orthacris maindroni, was examined in the current study. Among the five different varieties of acridid grasshoppers, the adults of A. pedestris strongly favoured the Orthacris maindroni (39%) followed by the slant-faced sp.1 (25%), Tylotropidus variecornis (16.38%), and cone-headed grasshopper (16.24%). Rarely (14.8%) did Acanthaspis pedestris favour the slant-faced grasshopper sp.2. The adults of Acanthaspis pedestris strongly liked the size groups of (1.0–1.7 cm) (32%) followed by (0.5–1.0 cm) (26.34%) and (1.7–2.2 cm) (22.46%) of their most chosen prey Orthacris maindroni. (0–0.5 cm) had the lowest preference rate (19.04%). Acanthaspis pedestris responded functionally to Orthacris maindroni in a type II manner. The number of prey killed and the prey density were shown to be positively correlated. As the prey density grew, the attack ratio fell. The maximal predation indicated by the “K” value (4.16), which was always limited to denser prey, The amount of time spent hunting and the density of the prey were found to be negatively correlated. The outcome shown that Acanthaspis pedestris favoured 1.0–1.7 cm size groups of Orthacris maindroni and displayed type II functional response to it, suggesting the potential for biocontrol of Acanthaspis pedestris on Orthacris maindroni (Table 16.4) (Balakrishnan et al. 2011). The positive connection found between the prey density and the number of prey killed by the predator suggests that Sphedanolestes variabilis exhibited Holling’s type II functional response. As the prey density grew, the attack ratio fell. The

594

16

Bioefficacy of Insect Predators Under Laboratory

Table 16.4 Reduviid predators’ posture in relation to various prey densities, with references Predator Sinea confusa

Pests Heliothis virescens

Order Lepidoptera

Response Positive

Zelus renardii

Heliothis virescens

Lepidoptera

Positive

Acanthaspis pedestris Sphedanolestes variabilis Rihirbus trochantericus Epidaus bicolor

Orthacris maindroni Lepidopteran pests

Orthoptera

Type II

Lepidoptera

Type II

Helopeltis spp.

Hemiptera

Type II

Helopeltis theivora

Hemiptera

Type II

Helopeltis theivora

Hemiptera

Type II

Aphis gossypii

Hemiptera

Type II

Phenacoccus solenopsis Dysdercus cingulatus Helopeltis spp.

Hemiptera

Type II

Hemiptera

Type II

Hemiptera

Positive

Picture-winged flies Pterophorus lienigianus Diaphania indicus

Diptera

Type II

Srikumar et al. (2017) Srikumar et al. (2017) Sahayaraj et al. (2012) Sahayaraj et al. (2012) Sahayaraj et al. (2012) Srikumar et al. (2014) Kalsi et al. (2014)

Lepidoptera

Positive

Chandral et al. (2009)

Lepidoptera

Type II

Dysdercus koenigii Phenacoccus solenopsis Nilaparvata lugens Dysdercus cingulatus Clavigralla gibbos

Heteroptera Heteroptera

Type II Type II

Nagarajan et al. (2010) Tomson (2021) Tomson (2021)

Hemiptera

Type II

Sunil et al. (2018)

Hemiptera

Type II

Hemiptera

Type II

Ambrose et al. (2009a, b) Ambrose et al. (2000)

Hieroglyphus banian Eutectona machaeralis Aphis craccivora

Hemiptera

Type II

Ambrose et al. (2000)

Lepidoptera

Type II

Ambrose et al. (2013)

Hemiptera

Type II

Spodoptera litura

Lepidoptera

Type II

Lepidoptera

Type II

Sahayaraj and Asha (2010) Muniyandi et al. (2011) Kumar et al. (2019)

Sycanus collaris sinensis Rhynocoris longifrons Rhynocoris longifrons Rhynocoris longifrons Cydnocoris gilvus Zelus longipes Rhynocoris fuscipes Rhynocoris fuscipes Rhynocoris fuscipes Rhynocoris fuscipes Rhynocoris fuscipes Rhynocoris marginatus Rhynocoris marginatus Rhynocoris marginatus Rhynocoris marginatus Rhynocoris kumarii Rhynocoris kumarii Rhynocoris kumarii

References Cohen and Tang (1997) Cohen and Tang (1997) Balakrishnan et al. (2011) Ambrose et al. (2009a, b) Bhat et al. (2013)

(continued)

16.3

Hemipteran Predators

595

Table 16.4 (continued) Predator

Coranus fuscipennis

Pests Odontotermes brunneus Corcyra cephalonica

Order

Response

References

Lepidoptera

Type II

Lam et al. (2020)

relationship between the searching time and the prey density turned out to be negative (Ambrose et al. 2009a, b). As one of the most significant commercially significant pests of cashew, Helopeltis spp., Rihirbus trochantericus shown Holling’s type II functional response. Rihirbus trochantericus’ molecular characterisation will be very helpful in establishing the species’ identification at any point of its life cycle (Bhat et al. 2013). Epidaus bicolor and Sycanus collaris’ functional responses to Helopeltis theivora (Hemiptera: Miridae) have been investigated, and Camellia sinensis has become a significant tea pest in southern India. Type II functional responses were present in both species of reduviids (Srikumar et al. 2017). Rhynocoris species are widespread, ferocious carnivores. In Petri dish arenas with cotton leaves, Rhynocoris longifrons feeding on five different densities of the cotton aphids Aphis gossypii, Phenacoccus solenopsis, and Dysdercus cingulatus were seen. When data were fit to Holling’s disc equation, the reduviid predator displayed a Type II functional response to all hemipteran pests examined. As the predator grew older and adults ingested the most Dysdercus cingulatus and Rhynocoris, the predator’s rate of predation steadily increased. When Aphis gossypii was given to the reduviid, a different pattern was seen. Rhynocoris was attacked at a fairly low but steady rate, with different stages of the predator’s life being generally more effective. Rhynocoris longifrons’ ability to kill adult stages of all evaluated prey species calls for further study of its biological control potential against cotton pests in pot and controlled fields. These findings showed that Rhynocoris longifrons could consume more Aphis gossypii at high prey densities, but predators also significantly decreased other cotton pests, making it a potential option for application as a commercial biological control agent for cotton hemipteran pests in India (Sahayaraj et al. 2012). Previously, Rhynocoris fuscipes showed a decrease in handling time and an increase in seeking capacity as prey density increased. The predator’s large daily prey intake at greater prey densities than at lower densities demonstrated the predator’s capacity for biocontrol. The predator’s IV stadium did not dramatically shrink with increasing prey density, but its V stadium did significantly shorten. The prey density has no effect on the survival or longevity of the predators. The pre-oviposition time lengthened as a result of low prey numbers. Prey density was correlated with enhanced fecundity and hatchability (Ambrose and Claver 1997). According to a study on the functional behaviour of Rhynocoris fuscipes in connection to Pterophorus lienigianus caterpillars that infest Solanum melongena, the predator may be able to control an expanding pest population by killing more pests. Rhynocoris fuscipes increased the amount of prey it killed at

596

16

Bioefficacy of Insect Predators Under Laboratory

lower densities in response to the rising pest population density. This was further supported by the finding that there was a link between prey density and prey killed. Additionally, a negative association was found between the prey density and attack ratio due to the predator Rhynocoris fuscipes’ inverse relationship between attack ratio and prey level. At all prey densities, a negative connection between prey density and predator seeking time was found (Chandral et al. 2009). Holling’s type II curvilinear functional response was demonstrated by Rhynocoris fuscipes in reaction to the cucumber leaf folder Diaphania indicus. As the prey density rose, the predator’s attack frequency rose as well. At greater prey densities, the maximum predation represented by ‘k’ value (2.30) was discovered. However, the lowest attack ratio (0.52) and the highest attack ratio (y/x) (0.99) were found at 1 and 8 prey/predator densities, respectively. At all prey densities, there was a link between prey density and prey killed and a correlation between prey density and predator seeking time. Predator spent less time looking and more time handling at high prey density, whereas at low prey density, searching time was always observed to be longer than handling time. Handling times vary depending on the predator’s chase speed, the prey’s escape rate, and the success of the prey capture (Nagarajan et al. 2010). In field cages in Tamil Nadu, India, in 1998, the functional reaction of fourthinstar nymphs of Rhynocoris marginatus to adults of Clavigralla gibbos and Hieroglyphus banian on pigeon pea (Cajanus cajan) was examined. Both prey species displayed a type II functional response. Prey density increased along with searching time, with maximum predation being seen at the highest prey densities (Ambrose et al. 2000). Rhynocoris marginatus responded to Eutectona machaeralis teak skeletoniser with a type II Holling’s curvilinear functional response. The predator’s assault frequency rose as the density of its prey did. The observed value of “k” for the highest level of predation was 4.16. However, at densities of 1 and 8 prey/predator, respectively, the highest and lowest attack ratios (y/x) (0.69 and 0.52) were recorded. Prey density and prey killed were shown to be positively correlated, however prey density and predator seeking time were found to be negatively correlated. Predator spent less time looking and more time handling at high prey density, whereas at low prey density, predator spent more time searching and less time handling. The rate at which the predator pursued the prey and the prey escaped or the success of the prey catch, however, affected the handling time (Ambrose et al. 2013). In India, the main sucking pest of cashew (Anacardium occidentale) is Helopeltis spp. (Hemiptera: Miridae). It has been noted that Cydnocoris gilvus (Reduviidae: Harpactorinae) may prey on Helopeltis species. When prey density increased, the predator killed more prey than it did when prey density was lower (Srikumar et al. 2014). Rhynocoris kumarii fourth nymphal instars were examined on various Aphis craccivora densities by measuring the functional response that was visible. The outcomes showed that Rhynocoris kumarii demonstrated Holling’s type II functional response and responded favourably to the rising aphid density. Similar to that, searching time also shrank as prey density rose. However, prey density was a factor that affected the first instar reduviid’s searching time. Rhynocoris kumarii in its fifth instar and as an adult avoided aphid. The fourth nymphal instar studied demonstrated

16.3

Hemipteran Predators

597

a larger potential for predation than the other three instars. Rhynocoris kumarii fourth instar nymphs can be utilised in a groundnut ecosystem in a ratio of 1:75.0 to control aphids (Sahayaraj and Asha 2010). Rhynocoris kumarii favoured Spodoptera litura larvae of the tobacco cutworm (35.29%), followed by Achaea janata larvae of the castor semilooper (29.41%), and Tyloprobidus variecornis larvae (21.57%). The blister beetles Mylabris indica (13.45%) and Mylabris pustulata (8.03%) were given a low preference. Mylabris pustulata was the least liked of the blister beetles. Rhynocoris kumarii favoured the 5–10 mm size group (44.64%) out of the several size groups of its most chosen prey, Spodoptera litura, followed by the smaller size groups (0–5 mm, 26.95%) and (10–15 mm, 18.16%). The 15–20 mm size range received the least amount of favour (10.23%). In response to Spodoptera litura larvae, Rhynocoris kumarii also demonstrated type II functional response. Prey density and prey killed were found to be positively correlated (y = 0.598 + 0.279x; r = 0.926). As the prey density grew, the attack ratio fell (y = 0.496–0.011x; r = -0.732). The highest prey density was consistently determined to be the limit of the maximal predation represented by the “k” value (4.53). The amount of time spent looking for prey and its density were shown to be negatively correlated (y = 4.383–0.311x; r = -0.975) (Muniyandi et al. 2011). Investigations were done on Rhynocoris kumarii and Odontotermes brunneus. There were used termite densities of 1, 2, 4, 8, 16, and 32. The fourth and fifth nymphal instars, as well as adult males and females of Rhynocoris kumarii, were fasted for 24 h in test cages that measured 14 by 22 by 10 cm. This experiment was conducted in a lab setting with 28–34 °C temperature, a 12-h photoperiod, and 65–70 relative humidity. When Odontotermes brunneus’ prey density increased, all of Rhynocoris kumarii experimental life stages responded by killing more prey than they did at lower prey densities. Predators displayed the type II curvilinear functional response as a result. High prey density was required for the maximum predation (K ) (24.889, 24.722, 22.889 and 24.194 for IV, V nymph and adult male and female respectively). Rhynocoris kumarii seeking ability rose with termite population, which demonstrated its predatory potential (Kumar et al. 2019). Zelus longipes, a generalist predator and potential biological control agent of picture-winged flies (Diptera: Ulidiidae), which significantly reduce sweet corn yields in Florida, is a milkweed assassin bug (Hemiptera: Reduviidae). We investigated Zelus longipes capacity to act as a biocontrol agent against the four ulidiid pests Euxesta stigmatias, Euxesta eluta, Euxesta annonae, and Chaetopsis massyla in maize fields. The distributions of Zelus longipes and ulidiids within a plant and within a field, as well as the functional responses of Zelus longipes to ulidiid prey, were identified. In the R1, R2, and R3 corn stages, the basal or middle leaves at 9:00 a.m. EST, the ears at 13:00 a.m. EST, and the top and tassel at 17: 00 a.m. EST typically had the highest concentrations of Zelus longipes and ulidiids. As a result, Zelus longipes and ulidiids appeared to move in unison during the day from the lowest to the tallest sections of the corn plant. Based on Taylor’s power law, Iwao’s patchiness regression, index of dispersion, and Lloyd’s patchiness indices of dispersion, aggregated (clumped) distributions for Zelus longipes and ulidiids were most prevalent within the corn field, especially in the later R2 and R3 stages. While

598

16

Bioefficacy of Insect Predators Under Laboratory

there were contradictory results for dispersion indices between times of day and between predators and prey, predator and prey populations were lower in the R1 stage. At 13:30 EST, ulidid distributions in R1 were primarily uniform and regular, although they aggregated at 9:00 and 17:00 EST. Although generally aggregated at 13:00 EST, Zelus longipes R1 distributions were random or regular between 9:00 EST and 17:00 EST. Zelus longipes had handling times of 1.0–1.39 h for males and 0.67–0.97 h for females, respectively. They also displayed type II functional responses to Euxesta stigmatias, Euxesta eluta, Euxesta annonae and consumed about five flies daily. Zelus longipes looks to be a viable biocontrol agent of ulidiid flies in corn, despite the fact that its population abundance can vary between seasons (Kalsi et al. 2014). Studies were done to find out how Rhynocoris marginatus’s functional response to Dysdercus cingulatus was affected by the predator’s age, sex, and prey size (Hemiptera: Pyrrhocoridae). Holling’s type II functional response was present in Rhynocoris marginatus in response to Dysdercus cingulatus life phases. The positive association found verified that more prey ( y) were killed at the highest prey density (x) in a given amount of time (Tt). Lower prey densities were where the highest attack ratio (y/x) was discovered. Predators who were older took longer to catch their prey than those who were younger. When handling their prey, males were slower than females of the same age group (b). Higher prey densities were the only ones that allowed for the maximal predations denoted by k values. Younger predators to younger prey were reported to have higher k/Tt values (Ambrose et al. 2009a, b). Rhynocoris fuscipes has a 98.513.63 predatory rate. The daily predation rates for nymphs in the first, second, third, fourth, and fifth instars were respectively 0.84, 1.14, 1.33, 1.13, and 1.41. The positive connection found between the density of the prey Nilaparvata lugens and the quantity of prey killed by the predator suggests that Rhynocoris fuscipes exhibited Holling’s type II functional response. As the prey density grew, the attack ratio fell. The relationship between the searching time and the prey density turned out to be negative (Sunil et al. 2018). In order to ascertain the impact of host species on biology, host stage preference, and biological control effectiveness in a laboratory setting, the relationships between a predatory reduviid, Rhynocoris fuscipes, and three prey species, namely, the red cotton bug Dysdercus koenigii (Heteroptera: Pyrrhocoridae), cotton mealybug Phenacoccus solenopsis (Hemiptera: Pseudococcidae) When feeding on Corcyra cephalonica, Dysdercus koenigii, or Phenacoccus solenopsis, Rhynocoris fuscipes finished the nymphal stage in 41 days, 45 days, and 50 days, respectively. When Corcyra cephalonica was fed, adult longevity, fecundity, and egg viability were higher; they were lowest when Phenacoccus solenopsis was fed. Corcyra cephalonica benefited from favourable life table conditions. The predator’s third and fourth instars preferred Dysdercus koenigii third instars. The predator had selected the fourth and fifth instars of Dysdercus koenigii, respectively, as well as adults. Phenacoccus solenopsis adults were preyed upon by all predator instars. It has been noted that the reduviid displayed type II functional response in response to increasing Dysdercus koenigii and Phenacoccus solenopsis density. Positive relationships unmistakably imply that the pest will also benefit (Tomson 2021).

Hemipteran Predators

No. of Helopeltis antonii fed/day

16.3

599

16 14 12 10 8 6 4 2 0

Reduviid Predator

Fig. 16.2 Helopeltis antonii predation by various reduviid species: a 5-day observation

In Vietnam, the reduviid Coranus fuscipennis (Heteroptera: Reduviidae) plays a crucial role in the biological control of vegetable pests. Under controlled laboratory conditions, the predator Coranus fuscipennis’ functional reaction to the larvae of the rice meal moth Corcyra cephalonica (Lepidoptera: Pyralidae) was assessed. The findings demonstrate that the Coranus fuscipennis nymph and adult responded to varying prey densities. They killed more prey at denser prey levels and less prey at denser prey levels, resulting in a linear type II functional response curve (Holling 1959). When raising the Coranus fuscipennis at high prey concentrations, the maximum consumption was always found to be constrained. Although there is a positive link between the quantity of prey that are killed and the prey density, there is a negative association between the Coranus fuscipennis predation rate and prey densities (R ranges from 0.70 to 0.98). The negative connection demonstrated that the C. fuscipennis’s search time for prey decreased as prey densities rose (R between 0.85 and 0.98). The Corcyra cephalonica developed the reduviid Coranus fuscipennis in a lab, and they can utilise it to biologically control several vegetable pests (P. rapae, S. litura, and P. xylostella) in Vietnam (Lam et al. 2020). When given a choice in a laboratory setting, the reduviid predator Sycanus indagator preferred the larvae of the greater wax moth (Galleria mellonella) to the larvae of the fall armyworm (Spodoptera frugiperda) (Bass and Shepard 1974). Under natural laboratory circumstances, the numerical responses of Alloeocranum biannulipes to various densities (1, 2, 4, 6, and 8) of D. porcellus larvae were assessed. A count of the predator’s devoured prey was made. The findings showed that Alloeocranum biannulipes females preying on Dinoderus porcellus (8) larvae with the highest larval density had the highest predation rate (4.3 larvae/predator/day) (Fig. 16.2) (Loko et al. 2022).

600

16.4

16

Bioefficacy of Insect Predators Under Laboratory

Lacewings

The green lacewing, Chrysoperla carnea (Neuroptera: Chrysopidae), is a significant predator of spider mites and other soft-bodied pests in Iran. In many regions of the world, the two-spotted spider mite, Tetranychus urticae (Acari: Tetranychidae), is a significant pest in various crop systems, including cotton. In this work, 6-h laboratory tests were used to examine the functional response of the three larval instars of Chrysoperla carnea on adult females of Tetranychus urticae. As a substrate for cotton leaf discs, the studies were carried out in Petri dishes with a 6 cm diameter that were lined with a layer of solidified agar solution. The predator’s first and second larval instars displayed Type II functional reactions against the prey. However, the predator’s third instar larvae displayed a Type III functional response. The first and second larval instars of the predator were predicted to have attack rates (a) and handling times (Th) of 1.758 h-1, 0.995 h and 0.163 h-1, 0.038 h, respectively. Attack coefficient (b) and handling time for the predator’s third instar larvae were 0.014 and 0.032 h, respectively. The predator’s last instar larvae were found to be subject to the highest theoretical maximum predation, which was assessed to be 187.5, 157.89, and 36.81, respectively. According to the study’s findings, Chrysoperla carnea larvae, particularly those in their last instar, have a good chance of catching adult female two-spotted spider mites. Therefore, the use of pesticides against this pest will probably be decreased by incorporating this predator into control plans. However, additional field-based investigations are required for a thorough assessment of Chrysoperla carnea biocontrol capabilities toward acariane preys (Hassanpour et al. 2009). In 2010, researchers in Japan examined the functional behaviour of the native green lacewing Chrysoperla nipponensis and the imported green lacewing Chrysoperla carnea when they fed on the cotton aphid Aphis gossypii at seven different densities (Homoptera: Aphididae). Based on logistic regression analysis, Chrysoperla nipponensis and Chrysoperla carnea demonstrated a Type II functional response. Chrysoperla carnea had a higher maximum prey intake than Chrysoperla nipponensis. Both species’ handling times reduced after 24 h; however, Chrysoperla nipponensis handling time was longer at 12 and 24 h than it was for Chrysoperla carnea. In the second and third instars, Chrysoperla nipponensis had a somewhat greater attack coefficient than Chrysoperla carnea. These findings suggest that Chrysoperla nipponensis may be a potential candidate for application as a commercial biological control agent for aphids in Japan, even if Chrysoperla carnea may consume more aphids at high prey densities. As far as environmental safety is concerned, Chrysoperla nipponensis will be more significant than Chrysoperla carnea because it is a native species that has adapted to the Japanese environment and non-target effects may be avoided (Montoya-Alvarez et al. 2010). Programs for biological control must be based on an understanding of predator– prey interactions. Given the theory, Hassanpour et al. (2011) conducted a study to assess the functional response of three larval stages of the green lacewing, Chrysoperla carnea, which feeds on eggs and first-instar cotton bollworm larvae, Helicoverpa armigera. Chrysoperla carnea larvae in their first and second instars

16.4

Lacewings

601

showed type II functional responses to both prey stages. Chrysoperla carnea larvae of the third instar, however, displayed a type II functional response to H. armigera larvae of the first instar, but a type III functional response to the eggs. Chrysoperla carnea larvae of the first instar attacked Helicoverpa armigera eggs at a rate that was significantly greater than that of the larvae, but the second instar Chrysoperla carnea larvae attacked Helicoverpa armigera larvae at a rate that was much higher than that of the eggs. Chrysoperla carnea third instar larvae were attacked at a rate of 1.015 h-1, and the attack coefficient on the eggs was 0.036. The third instar larvae handling times on eggs and larvae were 0.087 and 0.071 h, respectively. The Chrysoperla carnea larvae in their third instar were observed to prey most heavily on the eggs of Helicoverpa armigera. According to the study’s findings, Chrysoperla carnea larvae, particularly those in the third instar, exhibited a good capacity for controlling Helicoverpa armigera eggs and larvae. However, additional field-based research is required for a thorough assessment of Chrysoperla carnea’s capacity to suppress Helicoverpa armigera (Hassanpour et al. 2011). The lifespan of the prey is regarded as a desirable aspect of the predators’ functions (Table 16.5). The impact of prey density on the biology and functional response of the green lacewing, Chrysoperla carnea, was investigated in Pakistan in 2014 by Batool et al. In 9 cm petri dishes, freshly laid eggs from the Sitotroga cerealella (Lepidoptera: Gelechiidae) species were fed to newly emerging Chrysoperla carnea larvae. The prey density was found to have a substantial impact on Chrysoperla carnea’s favourable consumption rate, development, and fertility. In general, when prey density increased, we saw maximum consumption with the shortest developmental period, highest fecundity, and longest adult longevity. Predatory potential was high in all treatments when the prey density was increased. The daily predation rate of Chrysoperla carnea gradually rose throughout the first two larval instars and peaked in the third instar (Batool et al. 2014). The functional response research of thirdinstar Chrysoperla nipponensis (Neuroptera: Chrysopidae) larvae reared on artificial diet and Corcyra cephalonica eggs was undertaken, according to Memon et al. (2015) in 2015. Aphid, Aphis craccivora, papaya mealybug, Paracoccus marginatus, and whitefly, Bemisia tabaci, were used as prey at varying densities. Chrysoperla nipponensis larvae had type II functional responses to all kinds of prey. On mealybug and whitefly, respectively, the greatest search rates (á) of larvae raised on artificial food and C. cephalonica eggs were 0.688 and 0.404. The handling times for aphid (2.133 and 2.051) and mealybug (0.459 and 0.410) were greatest and shortest for both Corcyra cephalonica eggs and artificial diet raised larvae, respectively. Aphids and mealybugs, respectively, were the prey on which artificial diet and larvae raised from Corcyra cephalonica eggs preyed at the highest and lowest rates (0.487 and 0.468 and 2.433 and 2.177, respectively) (Memon et al. 2015). In a laboratory setting, Chrysoperla carnea larval instars were observed feeding on Brevicoryne brassicae, a cabbage aphid. The prey densities employed were 10, 15, 20, 25, and 30 (second to third) nymphal instar aphids for the first instar larvae, 10, 20, 30, 40, and 50 aphids for the second instar larvae, and 15, 30, 45, 60, and 75 aphids for the third instar larvae. Results indicated that in the first, second, and third larval instar stages of Chrysoperla carnea, an increase in prey density led

602

16

Bioefficacy of Insect Predators Under Laboratory

Table 16.5 Lacewings and their prey’s interactions with references Predator Chrysoperla carnea Chrysoperla carnea Chrysoperla carnea Chrysoperla carnea Chrysoperla carnea Chrysoperla carnea Chrysoperla carnea Chrysoperla carnea Chrysoperla carnea Chrysoperla nipponensis Chrysoperla nipponensis Chrysoperla nipponensis Chrysoperla nipponensis Chrysopa pallens Chrysoperla externa

Pest Tetranychus urticae-II instar Tetranychus urticae-III instar Brevicoryne brassicae Aphis gossypii

Order Acari

Response Type II

References Hassanpour et al. (2009) Hassanpour et al. (2009) Zada et al. (2016)

Acari

Type III

Homoptera Homoptera

Positive response Type II

Helicoverpa armigera-I instar Helicoverpa armigera-eggs Sitotroga cerealella

Lepidoptera

Type II

Lepidoptera

Type III

Lepidoptera

Type II

Choristoneura rosaceana Saissetia oleae

Lepidoptera

Type II

Hemiptera

Type-II

Aphis craccivora

Homoptera

Type II

Rios-Velasco et al. (2017) Mahzoum et al. (2020) Memon et al. (2015)

Paracoccus marginatus Bemisia tabaci

Homoptera

Type II

Memon et al. (2015)

Homoptera

Type II

Memon et al. (2015)

Aphis gossypii

Homoptera

Type II

Frankliniella occidentalis Hypothenemus hampei

Thysanoptera

Type III

Montoya-Alvarez et al. (2010) Sarkar et al. (2019)

Lepidoptera

Type II

Botti et al. (2022)

Montoya-Alvarez et al. (2010) Hassanpour et al. (2011) Hassanpour et al. (2011) Batool et al. (2014)

to an increase in prey consumption up to a specific limit. In the first, second, and third larval instar stages of Chrysoperla carnea, the maximum recorded prey numbers devoured were 22, 24, and 45 for the highest prey densities of 30, 50, and 75. In all three of Chrysoperla carnea larval instar stages, the same pattern was seen for the times spent searching, handling, and resting. The third instar larvae of the Chrysoperla carnea were discovered to have a better propensity for eating than the first and second ones. Additionally, the Chrysoperla carnea first instar larvae displayed shorter searching, handling, and resting times, which increased in the second and third instar larval stages. The findings of this study also demonstrated that Brevicoryne brassicae had a high predation capability for Chrysoperla carnea larvae, particularly the third instar (Zada et al. 2016). In 2017, a laboratory experiment compared Chrysoperla carnea to four densities of single-instar larvae of the five instars of the obliquebanded leafroller Choristoneura rosaceana (Lepidoptera: Tortricidae). The objectives were to identify

16.4

Lacewings

603

the type and other characteristics of the green lacewing’s functional response, including its capacity as a predator, in order to determine whether it could be used to supplement biological control in apple (Malus domestica Borkh; Rosales: Rosaceae) orchards where the obliquebanded leafroller has recently been introduced in Mexico. The bug quickly multiplied, resulting in substantial foliar damage and some fruit deformity. Based on logistic regression research, third-instar green lacewing larvae responded Type II functionally to four densities of single-instar obliquebanded leafroller larvae. The average number of second-instar larvae ingested by a green lacewing predator was 4, at a rate of 1.93 per 24 h, which was the highest average quantity among the 5 instars of prey. Two third-instar prey larvae were devoured on average by each predator, and each third-instar green lacewing consumed 1.5 third-instar obliquebanded leafrollers in a 24-h period. The handling periods for the remaining instars ranged from 23.48 min for the fifth instar to 31.56 min for the fourth instar. The third-instar predator collected and consumed third-instar prey larvae in the least handling time (h), i.e., only 6.46 min. However, third- and fourth-instar prey larvae had slightly higher attack coefficients (a) of green lacewings, with a respective 0.19 h (11.4 min) and 0.15 h (9.0 min) until the first attack, as opposed to the other instars’ 0.09–0.11 h (5.4–6.6 min). According to the findings, lepidopteran leafrollers could potentially be controlled biologically by using the green lacewing (Rios-Velasco et al. 2017). Chrysoperla carnea tested for black scale in 2020. The significant pest Saissetia oleae (Hemiptera: Coccidae) affects a number of crops, including the olive tree. Chrysoperla carnea’s effectiveness as a predator on Saissetia oleae has not yet been investigated. In the current study, the functional response of Chrysoperla carnea larvae fed on S. oleae nymphs was examined. Increasing densities of S. oleae second and third nymph stages were provided to freshly emerged specimens of the three larval instars of Chrysoperla carnea in a controlled laboratory setting. After 24 h, the amount of S. oleae that had died was counted, and Chrysoperla carnea’s functional response was evaluated. The slain prey increased with larger Saissetia oleae densities up to a maximum that was constrained by the handling duration in the three larval stages of Chrysoperla carnea. While the maximum attack rate was much higher for the third instar, the attack rate did not significantly differ between the three instars. Chrysoperla carnea’s first larval instar required more handling time than the third instar. Our findings suggested that Saissetia oleae could serve as a food supply for all Chrysoperla carnea larval stages. Additionally, the predator’s third larval stage proved the most effective at lowering Saissetia oleae numbers (Mahzoum et al. 2020). These findings imply that Chrysoperla carnea larvae may aid in the management of cabbage aphids, Helicoverpa armigera, Saissetia oleae, Choristoneura rosaceana, Sitotroga cerealella, and Brevicoryne brassicae in sustainable agriculture. Due to the diversity of its prey, wide distribution, and superior predatory abilities, Chrysopa pallens is one of the most efficient entomophagous predators. The average quantity of Frankliniella occidentalis (Thysanoptera: Thripidae) larvae eaten by Chrysopa pallens in its first, second, and third larval instars. In the first, second, and third instars of Chrysopa pallens, the linear coefficients in the logistic regression

604

16

Bioefficacy of Insect Predators Under Laboratory

analysis were 0.0725, 0.1687, and 0.0789, respectively. This suggests that all three instars of Chrysopa pallens exhibited a type III functional response to larval Frankliniella occidentalis. Western flock thrips were preyed upon at maximum observed rates (SE) by Chrysopa pallens in its first, second, and third instars, correspondingly, of 33.6, 40.5, and 38. Based on 95% CI, the second instar lacewings had a significantly greater attack rate than the other two instars. Chrysopa pallens’ three instars’ handling times did not differ considerably from one another (Sarkar et al. 2019). For the first time, Botti et al. (2022) document the predation of the coffee berry borer Hypothenemus hampei by a green lacewing Chrysoperla externa in a laboratory setting, demonstrating the predator’s capacity to enter pest galleries, remove immature stages, and feed on them. We also noticed third instar larvae preying on adult Hypothenemus hampei. With this note, we expand the reported list of species that Hypothenemus hampei control still hasn’t thoroughly studied (Botti et al. 2022).On all cucumber varieties, the predator Chrysoperla carnea displayed functional response type III. On various cucumber cultivars infested by the melon aphid, Aphis gossypii, the attack constant (b) for the third instar Chrysoperla carnea ranged from 0.00121 to 0.00168 and its handling time (Th) from 0.3145 to 0.4596 h (Asasi et al. 2022).

16.5

Coleoptera

Aphids, mites, beetles, and butterfly eggs are just a few of the many prey items that ladybird beetles eat. In India, a lab experiment was conducted to examine how varying predation intervals and prey concentrations affected Coccinella septempunctata’s functional response (Lipaphis erysimi). It was discovered that when predation period and prey density increased, so did the rate of prey eating. Over half of the prey animals were eaten in the first 3 h. Coccinella septempunctata’s predation behaviour complied with Holling’s Type II equation. While predation duration increased, the search rate of larvae dropped linearly as prey density rose. The appetite of predators that had been deprived for 24 h was greater than that of unstarved predators (Shukla et al. 1990). Using tiny containers, researchers in a lab examined the consumption of diamondback moth, Plutella xylostella larvae, and 24 species of adult carabids and 2 species of carabid larvae. Using tiny, confined arenas that replicated field circumstances, 13 species of adult carabids were tested for their ability to suppress diamondback moth larvae. Adult carabids consumed between 0 and 23 fourth instars each day of each sex, ranging from Amara simplicidens Morawitz’s female to Chlaeniusposticalis motschulsky’s male. Through the entire larval stage, the larvae of Chlaeniusmicans and Chlaeniusmicans posticalis ingested 191 and 92 early fourth instar diamondback moths per individual, respectively. These larvae were frequently found in the crowns of cabbage. C. micans, C. posticalis, and Dolichus halensis produced >95% mortality for diamondback moth larvae over 4-day trials in the enclosure experiment. The other 9 species in 6 genera, however, contributed to

16.5

Coleoptera

605

Table 16.6 Coleopteran predator’s response against various storage pests with references Predator Coccinella septempunctata Pterostichus planicollis Coleomegilla maculate Eriopis connexa Eriopis connexa Coccinella undecimpunctata Propylea quatuordecimpunctata Hyperaspis polita Hyperaspis polita Delphastus catalinae Delphastus pallidus Harmonia axyridis Oenopia kirbyi

Pest Lipaphis erysimi

Order

Response Type II

References Shukla et al. (1990)

Plutella xylostella Leptinotarsa decemlineata Macrosiphum euphorbiae Tetranychus evansi Myzus persicae

Lepidoptera

Type II

Coleoptera

Type II

Suenaga and Hamamura (1998) Munyaneza and Obrycki (1997) Sarmento et al. (2007) Sarmento et al. (2007) Cabral et al. (2009)

Type III Acari

Type II Type II

Aphis fabae

Hemiptera

Type II

Phenacoccus solenopsis Hyperaspis polita Bemisia tabaci Bemisia tabaci Spodoptera frugiperda Macrosiphum rosae

Hemiptera

Type III Type II

Hemiptera Hemiptera Lepidoptera

Type II Type II Type II

Hemiptera

Type II

Papanikolaou et al. (2014) Seyfollahi et al. (2019) Seyfollahi et al. (2019) Kumar et al. (2020) Di et al. (2021)

the 71–77% mortality that Pterostichus planicollis caused (Table 16.6) (Suenaga and Hamamura 1998). A comparison of the functional response of Coleomegilla maculate (Coleoptera: Coccinellidae) fourth instars was undertaken in 1997 by Munyaneza and Obrycki in lab, greenhouse, and field settings. Individual larvae of the Colorado potato beetle, Leptinotarsa decemlineata (Coleoptera: Chrysomelidae), were placed in 9-cm petri dishes for 24 h. The typical egg mass was 15 eggs per mass. The amount of potato leaf used to feed Coleomegilla maculate larvae in the greenhouse and field ranged from 0.5 to 35 Leptinotarsa decemlineata egg masses/m2. Under laboratory, greenhouse, and field circumstances, fourth instars of Coleomegilla maculate displayed a type II functional response to Leptinotarsa decemlineata eggs. Prey density and predator search effectiveness had an inverse relationship (Munyaneza and Obrycki 1997). The sort of prey also influences the predator’s voracity. For instance, the concept is demonstrated by the functional responses of adult females of Eriopis connexa to various densities of Macrosiphum euphorbiae and Tetranychus evansi. Eriopis connexa displayed a Type III sigmoidal functional response curve when consuming aphids. When Eriopis connexa was given mites (Tetranychus evansi) instead of aphids, this behaviour abruptly changed to an exponential (Type II) functional

606

16

Bioefficacy of Insect Predators Under Laboratory

response. Such disparate patterns demonstrated the requirement for this coccinellid to adopt various strategies depending on the type of available prey. Because it was thought that predators could only control prey numbers when they used a Type III functional response. Eriopis connexa would work well as a biological control agent against Macrosiphum euphorbiae, but it would not be able to stop the spread of Tetranychus evansi (Sarmento et al. 2007). Under controlled laboratory conditions, the voracity and functional responsiveness of Coccinella undecimpunctata adults (males and females) and larvae in the fourth instar on Myzus persicae were assessed. This investigation was conducted in little plastic boxes with various aphid concentrations. When fourth instar larvae received 130 aphids and adult males or females received 90 aphids, satisfaction was attained. The fourth instar larvae of the Coccinella undecimpunctata showed a shorter handling time than adults, and attack rates across females and males as well as between larvae and adults were comparable. Under controlled circumstances, Coccinella undecimpunctata proved to be a successful predator for the biological control of Myzus persicae. It is hypothesised that the presence of both the larvae and adults of this ladybird beetle’s fourth instar could increase the effectiveness on field pest suppression (Cabral et al. 2009). In both long- and short-term functional response trials, the 14-spotted ladybird beetle Propylea quatuordecimpunctata (Coccinellidae) and its prey Aphis fabae (Hemiptera: Aphididae) were employed. According to the findings, adults of the Propylea quatuordecimpunctata display a type II functional response. Between long- and short-term studies, the predicted attack rates and handling times varied significantly, showing that the digestive process inhibits the predatory potential of Propylea quatuordecimpunctata (Papanikolaou et al. 2014). According to recent research, the functional response is also influenced by taxa, characteristics, temperature, and arena size in addition to the previously listed parameters (life stages, temperatures, or sexes). The findings of Uiterwaal and DeLong (2018) demonstrate that temperature, arena size, prey type, predator stage, and predator mass all have an impact on functional response parameters (handling time and space clearance rate). Although complicated by interaction terms, handling time generally decreased with predator size, temperature, and stage while space clearance rate generally rose with each of these factors. The handling periods were longest while handling coleopteran prey. Additionally, they emphasised how experimental arena size has a significant and consistent impact on space clearance rate. All other factors, such as predator mass and temperature, are less significant in predicting foraging rates at low prey densities than arena size. The selection of the arena size therefore complicates attempts to employ functional response trials in laboratories to assess the effectiveness of biocontrol predators (Uiterwaal and DeLong 2018). Phenacoccus solenopsis’ main predator in Iran is Hyperaspis polita. When the ladybird Hyperaspis polita preyed upon different stages of the mealybug Phenacoccus solenopsis, the functional response of the fourth instar larva, adult female, and male was investigated. With the exception of the adult female stage of Hyperaspis polita, which had a type III functional response to Phenacoccus

16.5

Coleoptera

607

solenopsis nymphs in the first instar, all stages of Hyperaspis polita displayed type II functional responses to all stages of Phenacoccus solenopsis. The attack rates did not significantly differ when the prey and predator stages changed. However, handling times varied depending on the prey and predator stages. This value, which was the shortest for any predator stage that consumed first-instar Phenacoccus solenopsis nymphs, was calculated to be 0.1100, 0.1868, and 0.2939 h for adult females and males, fourth-instar larvae, and fourth-instar larvae, respectively. The fourth instar larval stage had the shortest handling time, and then adult females and males to various prey stages. As Phenacoccus solenopsis evolved from one life stage to the next, the maximum predation rate (T/Th) declined. The adult female mealybug provided the lowest estimated value, and the fourth instar larval stage was the most predatory. The fourth larval instar stage of Hyperaspis polita is thought to be the most effective predatory stage for use in biological control programmes, indicating that the species has notable feeding capability (Seyfollahi et al. 2019). However, Seyfollahi et al. (2019) noted that more field-based research is required before a thorough assessment of Hyperaspis polita’s biocontrol capabilities toward Phenacoccus solenopsis can be formed. The ability of two predatory beetles, Delphastus catalinae and Delphastus pallidus, to suppress two of the most infamous members of the Bemisia tabaci species complex, the Middle Eastern Asia Minor 1 (MEAM1 or biotype B) and Mediterranean (MED or biotype Q) whiteflies. Under laboratory conditions, the functional responses of two predatory beetles at prey densities ranging from 20 to 120 whitefly eggs/leaf disc and prey-stage preferences with different ratios of whitefly eggs and nymphs were examined. The logistic regression model found that both species of beetles responded Type II functionally to MEAM1 and MED eggs. D. catalinae appeared to be a more suitable biocontrol candidate for whitefly pests based on the observed predation parameters (handling time, attack rate, and the amount of eggs devoured). The findings showed that both beetle species could eat more than 50 eggs each day and favoured whitefly eggs over the earliest nymphal instars. Although these results are encouraging, more research must be done on their feeding capability in greenhouse and semi-field settings to validate their effectiveness against MEAM1 and MED whiteflies (Kumar et al. 2020). Similar research on the intake of Spodoptera frugiperda was done on Harmonia axyridis (Coleoptera: Coccinellidae). The findings of the experiments showed that Holling’s Type II functional response model best described how Harmonia axyridis nymphs preyed on Spodoptera frugiperda eggs and larvae. In Harmonia axyridis, the Na-max, a′, and Th values for adults on Phenacoccus solenopsis eggs were 130.73, 1.1112, and 0.085 days, respectively, while on Spodoptera frugiperda first instar larvae were 1401.1, 0.8407, and 0.0006 days, respectively. These findings demonstrated that Harmonia axyridis is an extremely ferocious predator of Spodoptera frugiperda eggs and early larvae. Our research offers a theoretical foundation for the use of natural enemies to manage Spodoptera frugiperda in the field. More study is needed to develop field-based conservation biological control strategies that will boost predator populations and Spodoptera frugiperda reduction (Di et al. 2021). The Oenopia kirbyi (Coccinellidae: Coleoptera) battles the

608

16

Bioefficacy of Insect Predators Under Laboratory

Macrosiphum rosae rose aphid in 2022 (Aphididae: Hemiptera). The newly emerged coccinellid larvae were raised separately in test tubes harbouring several populations of Macrosiphum rosae nymphs in the second and third instars. First, second, third, and fourth instar coccinellid larvae had a respective predatory potential of 8.70, 20.70, 28.10, and 31.40 aphids. Mature males eaten 491 aphids throughout the course of their lives, whereas adult females ingested 560.16 aphids over the course of 544–580 aphids. The assault rate (a) and handling time characteristics of Oenopia kirbyi against Macrosiphum rosae show how the predator’s predation efficiency increased with its developmental stage (Th). First, second, third, fourth instars, and adults all experienced predation at rates of 19.17, 27.69, 30.63, 32.49, and 34.27 aphids per 24 h, respectively. Oenopia kirbyi fourth instar coccinellid larvae and adults (male and female) have been observed to be voracious eaters and can be successfully used in integrated pest management (IPM) programmes against rose aphids in conjunction with other natural enemies as a biocontrol agent of M. rosae (Gaikwad et al. 2022).

16.6

Trichoptera

For two of its naturally occurring prey species, experiments were done to see how the substrate type affected Plectrocnemia conspersa Curtis’s functional response. Predation rates are lower for both prey species on substrates that are similar to those that the prey prefers in the wild. Due to the later types’ supply of safe havens for prey, the predation rate on simple and certain more complex substrates differs. Because of how the two prey species behave differently when captured in a predator’s net, there are differences in the predation rates on the two species. The stability of predator– prey interactions and the availability of prey to predators are explored in relation to the effect of substrate on P. conspersa’s functional response (Hildrew and Townsend 1977).

16.7

Storage Pests

In storage environments, Xylocoris flavipes and a number of other insects from the subfamily Lyctocorinae typically act as predators. Xylocoris flavipes has the potential to be used in biological control programmes due to its capacity to suppress populations of storage pests. It feeds on more than 13 species of insects from three orders. There have been few studies on the functional response in Xylocoris flavipes (Hemiptera: Anthocoridae). Xylocoris flavipes, a generalist predator of insects found in stored goods. Both the prey species and the surrounding environment had an impact on the suppression of the Indianmeal moth, Plodia interpunctella, and the almond moth, Cadra cautella (Lepidoptera: Pyralidae). Xylocoris flavipes release reduced the numbers of Indianmeal and almond moths by as much as 71.4% and 78.8%, respectively, before a severe freeze and cold snap in January wiped out the almond moth population. Throughout the 7-month test period, the Indianmeal moth

16.7

Storage Pests

609

Table 16.7 Anthecoride and reduviid predators’ response against various storage pests with references Predator Xylocoris flavipes Xylocoris flavipes

Pest Oryzaephilus surinamensis P. interpunctella

Type II

Xylocoris flavipes

Rhyzopertha dominica

Type II

Xylocoris flavipes

Tribolium castaneum

Type III

Amphibolus venator Alloeocranum biannulipes Alloeocranum biannulipes Alloeocranum biannulipes

Tribolium confusum Teretrius nigrescens

Type II Type II

Prostephanus truncatus

Type I, III Type II

Dinoderus porcellus

Response Type II

References Donnelly and Phillips (2001) Donnelly and Phillips (2001) Donnelly and Phillips (2001) Donnelly and Phillips (2001) Nishi et al. (2004) Loko et al. (2017) Loko et al. (2020) Loko et al. (2022)

was suppressed. One element of a comprehensive peanut management campaign based on the discharge of biological agents, Xylocoris flavipes, may be useful (Brower and Mullen 1990). There is significant literature on the biological control abilities of virus predators. Without naming the predators, the biocontrol effectiveness of Xylocoris flavipes (Abdel-Rahman et al. 1983), seed beetles (Berger et al. 2017), wheat grains (Ferdous et al. 2010), and peanut storages (Brower and Mullen 1990; Arbogast 1984). On the stored-product insect pest Tribolium confusum, the predatory prowess of Amphibolus venator was investigated. After being starved for 3 days, Amphibolus venator adults (5–15 days old) were then individually placed into plastic containers holding various prey densities (3, 5, 10, 15, or 20) of larvae, pupae, or adults of Tribolium confusum. The experiment was conducted for 1 day at 25 and 30 °C. Also looked at was the amount of Tribolium confusum killed by Amphibolus venator over the course of 10 days with a prey density of 10. All tested life stages—mature larvae, pupae, and adults—were attacked by Amphibolus venator. Compared to the other stages, Amphibolus venator preferred to target adult larvae. High prey density led to an increase in the quantity of prey that Amphibolus venator killed, and all prey stages showed Holling’s Type II response in the functional response of Amphibolus venator. In general, females killed more prey than males. Amphibolus venator preyed more successfully at 30 °C than it did at 25 °C (Table 16.7) (Nishi et al. 2004). LeCato and Arbogast (1979) investigated Xylocoris flavipes predation on the eggs and larvae of the Angoumois grain moth, Sitotroga cereallela (Lepidoptera: Gelichiidae), at various prey densities, while Sing (1997) investigated the functional response of Xylocoris flavipes to several different bruchid species infesting legumes. In relation to Oryzaephilus surinamensis (Coleoptera: Cucujidae) (small and large

610

16

Bioefficacy of Insect Predators Under Laboratory

larvae), Tribolium castaneum (Coleoptera: Tenebrionidae) (small larvae), P. interpunctella (egg and small larvae), and Rhyzopertha dominica (Coleoptera: Bostrichidae), Donnelly and Phillips (2001) studied the functional response (egg and larvae). Holling’s type II model best captures the functional response of Xylocoris flavipes; however, prey that were more challenging to subdue, like Tribolium castaneum larvae, elicited a type III reaction. There were more female-attacked prey than male-attacked prey in both settings for particular prey species and life stages. The following were the greatest attack rates for the various prey species in empty glass jars over a 24-h period: There were 27.3 small larvae and 1.6 large larvae in Tribolium castaneum, 24.3 small larvae and 17.4 large larvae in Oryzaephilus surinamensis, 27.2 eggs and 23.7 small larvae in P. interpunctella, and 16.6 internally feeding larvae in Rhyzopertha dominica. Following are the greatest numbers of each prey species that were killed in glass jars of wheat over a 48-h period: The Oryzaephilus surinamensis P. interpunctella; 13.7 small larvae, 12.8 large larvae Rhyzopertha dominica has 22.0 eggs and 12.1 internally fed larvae, whereas there are 41.4 eggs and 14.7 tiny larvae. A single Xylocoris flavipes was able to find and kill 27.1 out of 50 P. interpunctella eggs and 8.1 out of 10 Rhyzopertha dominica larvae inside kernels blended into 18,000 kernels of wheat in 48 h, according to experiments with wheat-filled jars (Donnelly and Phillips 2001). To better understand their roles in the biological control of this significant pest of stored yam chips, Alloeocranum biannulipes (Hemiptera: Reduviidea) and Teretrius nigrescens (Coleoptera: Histeridae) underwent functional evaluations at five different densities of larvae and pupae of Dinoderus porcellus (Coleoptera: Bostrichidae). Based on a logistic regression model’s analysis of their functional responses to larvae, both predators displayed Type II responses. In contrast to Alloeocranum biannulipes, Teretrius nigrescens considerably killed more Dinoderus porcellus larvae. When both predators were offered pupae of Dinoderus porcellus, however, this behaviour shifted to a linear functional response (Type I), probably as a result of the pupae’s immobility. In terms of the dead pupae, there was also no discernible difference between Teretrius nigrescens and Alloeocranum biannulipes. Both predators’ Holling disc equation parameters were computed. Teretrius nigrescens and Alloeocranum biannulipes had handling times for Dinoderus porcellus larvae of 0.254 and 0.677 h, respectively, and their rates of searching efficiency were 0.289 and 0.348 h1. According to the findings, Teretrius nigrescens is a better choice for augmentative release for controlling Dinoderus porcellus than Alloeocranum biannulipes. To draw definitive conclusions, however, semifield researches are necessary (Loko et al. 2017). In 2020, it was discovered that mature females of the predatory bug Alloeocranum biannulipes (Hemiptera: Reduviidae) responded both numerically and functionally to variations in the density of the larger grain borer Prostephanus truncatus (Coleoptera: Bostrichidae), which was infesting cassava chips. Predators that were starving were subjected to six repetitions of treatments that varied in prey density. Daily records were kept of the prey taken, eggs placed, adults that emerged, and hatching rates. Consumption rose linearly (R2 = 0.858) with the increase in

16.8

Factors Influencing Functional Responses

611

larval density when feeding on Prostephanus truncatus larvae. The adult Alloeocranum biannulipes displayed the highest consumption rates when there were six Prostephanus truncatus pupae per predator, the maximum pupal density (1.67 pupae). When Prostephanus truncatus larvae and pupae were devoured by Alloeocranum biannulipes, type I and type III functional responses were noted. Predator attack estimated at 0.027 hours per hour for larvae and 0.125 hours per hour for pupae. The adult predator handled the pupae for 0.352 h, with a predicted daily maximum predation (T/Th) of 68.18 P. truncatus pupae. When exposed to larger prey densities, Alloeocranum biannulipes responded numerically in a way that was positively correlated with pupal density, laying more eggs per female and having a higher hatching rate. The efficiency of Alloeocranum biannulipes’ food conversion into eggs rose with falling larval densities and was steady with rising pupal densities. Alloeocranum biannulipes responded in both functional and numerical ways, which suggests that it can effectively regulate P. truncatus pupae at both low and high densities as well as larvae at both low and high densities. So, Alloeocranum biannulipes would provide a good option for P. truncates’ biological control (Loko et al. 2020). Females of Alloeocranum biannulipes showed the highest predation rate (4.3 larvae/predator/day) when feeding on Dinoderus porcellus larvae with the maximum density (number 8). (Loko et al. 2022).

16.8

Factors Influencing Functional Responses

Various behaviours and environmental factors affect functional reactions. Theoretical population and community ecologists paid little attention to the effects of temperature on predator–prey dynamics during the twentieth century, and the majority of published models that included temperature dependence of consumption rates were complex and system specific. The functional response is also influenced by the host plant’s presence or absence. In petri dish arenas and on plants with different leaf structures, the lady beetle Propylea quatuordecimpunctata’s reaction to the abundance of the Russian wheat aphid, Diuraphis noxia, was observed. Both adult and larval beetle behaviour in dishes closely matched a type II response. Estimates of “a” and “Th” on whole plants differed greatly from those made from dishes, and they also varied depending on the type of plant. At each aphid density, beetle adults ate more aphids on the narrow-leaved Indian ricegrass (Oryzopsis hymenoides) than on the broad-leaved crested wheatgrass (Agropyron desertorum). Additionally, logistic regression revealed that Indian ricegrass would respond in a type II manner as opposed to crested wheatgrass would respond in a type III (sigmoidal) manner. The complex response on crested wheatgrass may have resulted from variations in the proportion of aphids in refuges that are depending on densities (such as rolled leaves). Plant characteristics can affect the stability of predator–prey dynamics and the efficacy of biological control by altering the forms and dimensions of functional responses (Messina and Hanks 1998). The functional response of female adults of the two-spot ladybird, Adalia bipunctata (Coleoptera: Coccinellidae), was studied in petri dish arenas containing sweet pepper leaves infested with various densities of

612

16

Bioefficacy of Insect Predators Under Laboratory

the green peach aphid, Myzus persicae. Temperature is another significant factor that affects the functional response of many predators (Hemiptera: Aphididae). At three different test temperatures, ranging from 19 to 27 °C, the predator had a type II functional response. The maximum number of prey that a predator could theoretically capture increased with temperature. Based on the random predator equation, the projected attack rates on a leaf area of 20–25 cm2 ranged from 0.13 h1 at 19 °C to 0.35 h1 at 27 °C. The predator’s assault rates did not alter much between 23 and 27 ° C. As temperature increased from 19 °C (0.39 h) to 27 °C, handling time considerably decreased (0.24 h). Adalia bipunctata has significant rates of predation on Myzus persicae throughout a wide range of temperatures, according to this study, suggesting the potential for augmentative releases against this aphid pest (Jalali et al. 2010). Moreover, the relative energy worth of prey and the predator’s ability to locate and capture it are all influenced by the predator’s body size. According to scaling theory, as predator body size increases, space clearance rate should rise and handling time should fall. Additionally, some studies show that the size of the arena may be relevant in determining the rate at which space is cleared, while others say it is not (Uiterwaal and DeLong 2018).

16.8.1 Plant Species and Their Morphology Five-day-old Orius insidiosus females ingested considerably more thrips on bean plants than on tomato plants across all prey densities (5.75 and 3.39 thrips per predator per day, respectively). On pepper, a moderate number of thrips were consumed (4.16 thrips per predator per day). This was especially noticeable at a thrips density of 30 per plant, where female Orius insidiosus consumed 11.3, 7.0, and 3.8 thrips on bean, pepper, and tomato, respectively. Similar to how more thrips were consumed on bean plants than tomato plants, The quantity and percentage of prey that Orius insidiosus consumed were both influenced by thrips density. At a density of 5 (0.51), substantially more thrips were preyed upon than at the other densities (0.33, 0.25, and 0.24 at densities of 10, 20, and 30 thrips per plant, respectively). There were no discernible interactions between plant and prey density; hence, the effect of prey density on predation rate was equal across the three plants. Nevertheless, variations in Orius insidiosus’ functional response curves on the three plants had an impact on their shape. Both type-I and type-II curves offer good fits to the data for beans and peppers based on r2, while only type I fits the data for tomatoes (Table 16.8). It appears that the inclusion of larger thrips densities in the experiment would result in a better fit of the bean and pepper data because a higher maximal consumption rate occurred on bean and pepper than on tomato (to a type-II rather than a type-I response). Last but not least, tomato plants had a lower attack coefficient (0.251) than bean and pepper plants (0.432 and 0.434, respectively). Results indicate that Orius insidiosus hunted for thrips on tomato plants less effectively than it did on bean and pepper plants (Coll and Ridgway 1995). In a different investigation, bean, tomato, and pepper plants in a greenhouse were used to gauge Orius insidiosus’ functional reaction to one of its principal victims, the

16.8

Factors Influencing Functional Responses

613

Table 16.8 Different predator’s response against various storage pests of different plants/cultivers with references Predator Orius insidiosus

Pest Thrips

Order Thysanur

Plants Tomato

Response Type-I and type II

Orius insidiosus

Thrips, aphids, leafhoppers, and whiteflies Spodoptera exigua

Lepidoptera

Bean, tomato, pepper, and corn plots Type II

Positive response on pepper and tomato Sweet pepper and eggplant

Podisus nigrispinus Podisus nigrispinus

Spodoptera exigua

Lepidoptera

Type III

Tomato

Orius albidipennis

Schizaphis graminum

Type III

Falat (variety) Pishtaz (variety)

References Coll and Ridgway (1995) Coll and Ridgway (1995) De Clercq et al. (2000) De Clercq et al. (2000) Gholami et al. (2022)

western flower thrips, Frankliniella occidentalis. The predator captured fewer thrips on tomato plants than on bean plants, and tomato plants had a lower attack coefficient than bean and pepper plants. The form of the functional reaction of Orius insidiosus to its prey was influenced by variations in the predator’s searching behaviour on the three different plants. At two field sites, bean, tomato, pepper, and corn plots were observed to determine the association between populations of Orius insidiosus and its prey (thrips, aphids, leafhoppers, and whiteflies). Orius insidiosus density was noticeably higher on bean and corn than on pepper at both sites (densities in tomato were intermediate). Although the phenology of the predator varied across crops, it generally followed its prey’s population dynamics. In contrast to bean and com, pepper and tomato showed a delayed (reproductive) response to prey, which was density-dependent and behavioural. Furthermore, compared to tomato and pepper, bean and corn had a larger ratio between the densities of Orius insidiosus nymphs and adults, a sign of population expansion (Coll and Ridgway 1995). Using potted sweet pepper, eggplant, and tomato plants, the impact of the host plant on the functional response of Podisus nigrispinus females to concentrations of fourth instar Spodoptera exigua was examined. Sweet pepper and eggplant showed a type II response in the analysis of logistic regression; however, tomato showed a type III response. Both the random predator and disc equations were used to examine the data. The disc equation and the random predator equation predicted the attack rates (a) on sweet pepper and eggplant to be 0.048 and 0.093 h-1, respectively. On tomatoes, attack frequency was a function of prey density (a = bN), with b averaging 0.006 for the disc predator equation and 0.008 for the random predator equation. The estimated handling times (Th) for eggplant and sweet pepper were 2.5–2.8 h for the disc equation and 3.0–3.2 h for the random predator equation, respectively. With an

614

16

Bioefficacy of Insect Predators Under Laboratory

estimated value of 5.6 h in both models, predators demonstrated noticeably longer handling times on tomato. Another experiment evaluated the stay of nymphal predators on various plants in the absence of prey. On eggplant and sweet pepper plants, predators were discovered in over 60% of cases, but only in 25% of cases on tomato plants. Allelochemicals and glandular trichomes on the plant’s surface are thought to affect how well the predator functions on tomato, reducing its effectiveness at searching and its capacity for predation (De Clercq et al. 2000). The same impact was previously noticed and reported by Anderson et al. (2001) in Notonecta kirbyi, Giroux et al. (1995) in Coleomegilla maculata (Coleoptera, Coccinellidae), Gresens et al. (1982) in Celithemis fasciata (Odonata: Libellulidae), and Mack and Smilowitz (1982a, b) in Cole (Coleoptera: Coccinellidae). The functional response is also influenced by host plant and plant structure (Carter et al. 1984; Messina and Hanks 1998; De Clercq et al. 2000). Different plant morphological characteristics have an impact on the phytophagous insects’ biocontrol agents. In the present study, two cultivars of common wheat (Falat and Pishtaz) with different leaf morphological characteristics were used to examine the effects of physical plant features on life table characteristics and functional response of Orius albidipennis females feeding on Schizaphis graminum. A wide range of characteristics relating to Orius albidipennis’ performance were significantly influenced by the trichome density. Despite exhibiting type III functional responses on both cultivars, Pishtaz had a lower searching efficiency and a longer handling time than Falat. On Falat, the maximum attack rate was higher than on Pishtaz. The Orius predators’ reduced maximum predation and longer handling times on Pishtaz leaves may be attributed to the plant’s leaves having much more surface trichomes than Falat leaves, which physically impeded bug mobility and reduced prey encounter rate. Moreover, the predatory bug’s foraging behaviour was adversely affected by an increase in trichome density. Additionally, it was discovered that females favour leaves with less trichomes as oviposition hosts. Trichomes appeared to be a barrier for the Orius bug. Compared to Pishtaz cultivar, Falat cultivar has a much higher intrinsic rate of rise (rm). Similar to this, Falat had a higher net reproduction rate and finite rate of rise than Pishtaz. To summarise, it may be said that cultivars with lower trichome density may benefit more from O. albidipennis’ effectiveness at preventing S. graminum on wheat (Gholami et al. 2022).

16.8.2 Pesticides and Biopesticides Pesticides have been shown to have both deadly and sublethal effects on populations of both targeted and untargeted arthropods. Acanthaspis pedestris, a non-target potential reduviid biological control agent, was used to study the effects of the pesticide cypermethrin on functional response, predatory behaviour, and mating behaviour. As the cypermethrin concentration was raised, the severity of the anomalous behaviour grew. The pesticide had a detrimental impact on the operational reaction events, including attack ratio, handling time, and rate of discovery. In

16.8

Factors Influencing Functional Responses

615

Acanthaspis pedestris, cypermethrin also decreased the predatory effectiveness and lengthened the mating processes. Cypermethrin changed the type II (decelerating curve) of the functional response into a type IV (dome-shaped curve) (Claver et al. 2003). Neonicotinoid pesticides are frequently employed to combat sucking pests, and helpful arthropods like natural enemies might expect sublethal effects. In many Chinese agricultural systems, Serangium japonicum serves as a significant predator and potential biological control agent against Bemisia tabaci. We investigated the effects of imidacloprid’s toxicity on the functional response to B. tabaci eggs in Serangium japonicum. Adult Serangium japonicum plants subjected for 24 h to dried residues of imidacloprid at the recommended field rate on cotton against Bemisia tabaci revealed substantial mortality rates. The lowest rate, 5 ppm, was determined to be sublethal because it did not significantly increase mortality compared to the control group. There is a risk for S. japonicum in treated fields, as indicated by the fatal rate 50 and hazard quotient (HQ) estimates of 11.54 ppm and 3.47, respectively. Serangium japonicum functional response to Bemisia tabaci eggs was impacted by exposure to dried imidacloprid residues at the sublethal rate (5 ppm) on cotton leaves. This resulted in longer handling times and lower egg consumption peaks. The amount of Bemisia tabaci eggs devoured on treated leaves was substantially less than on untreated leaves, indicating that imidacloprid residues also disrupted predator voracity. Imidacloprid systemically administered at the recommended field rate (for cotton) showed no toxicity to S. japonicum, nor impacted the functional response of the predator, and all effects vanished within a few hours after transfer to untreated cotton leaves. Imidacloprid’s sublethal effects on Serangium japonicum in our study are likely to have a negative impact on their ability to develop and reproduce, which could ultimately lead to a decline in the number of predators (He et al. 2012). These findings point to the significance of investigating imidacloprid’s possible impacts on Serangium japonicum in order to create effective integrated pest management strategies for Bemisia tabaci in China. The third-instar larvae of Tuta absoluta treated to LC30 (2.03104 conidia/mL) values of Metarhizium anisopliae Sorokin isolate DEMI 001 were inspected by Nabis pseudoferus females. Following inoculation, the predator was exposed to six densities of the prey (1, 2, 4, 8, 10, and 16) at 0, 24, 48, and 72 h. The type II functional response to prey density was present in Nabis pseudoferus in all treatments, suggesting that predation rises asymptotically to a satiation level. The attack rates (a) for treatments administered 48 and 72 h after infection were 0.1052 0.0440 and 0.0509 0.0133 h-1, respectively, respectively. The control group’s maximum theoretical predation rate (T/Th) was determined to be 10.96 (Alikhani et al. 2019). Tobacco leaf miner pest control may benefit from combining Metarhizium anisopliae with Nabis pseudoferus, according to Alikhani et al. (2019), although this needs to be verified in actual field situations. Pesticides have an impact on most predators’ rate of predation, including Podisus nigrispinus. One of the most prevalent asopine species in the Neotropics, it has been found in various South and Central American nations and is a crucial biological

616

16

Bioefficacy of Insect Predators Under Laboratory

control agent for a variety of crops. Podisus nigrispinus, a predator that feeds on Spodoptera frugiperda (Lepidoptera: Noctuidae) strains resistant to lambdacyhalothrin, was fed on Bt cotton expressing Cry1Ac (Bollgard®) during research conducted by Malaquias et al. in 2014. The following conditions were employed with Spodoptera frugiperda larvae: resistant strains to lambda-cyhalothrin strains 1 and 2 fed on Bollgard® cotton leaves (DP 404 BG); and resistant strains to lambdacyhalothrin strains 3 and 4 fed on non-genetically modified cotton leaves (cultivar DP4049). Imidacloprid had an impact on the predatory behaviour of Podisus nigrispinus, and the type II asymptotic curve best captured the functional response data. In the presence of imidacloprid, the handling time (Th) of predator females did not change between treatments. However, a rise in the density of larvae supplied resulted in a decrease in the attack rate. The predation of Podisus nigrispinus females on Spodoptera frugiperda larvae was dramatically reduced when exposed to imidacloprid, notably at a density of 16 larvae/predator, regardless of the treatment (S. frugiperda strain or cotton cultivar). The insecticide imidacloprid affects P. nigrispinus’ predation behaviour on Spodoptera frugiperda larvae, indicating that cotton crops should use it sparingly (Malaquias et al. 2014). On nymphs in the fifth instar, diazinon, fenitrothion, and chlorpyrifos were examined for their sublethal effects on the functional response of the predatory insect Andrallus spinidens (Hem.: Pentatomidae), a potential biological control agent. The experiment was carried out using final instar larvae of Chilo suppressalis (Lepidoptera: Pyralidae) at different densities (2, 4, 8, 16, 32, and 64) as prey at 25 2 °C, 60% 10% relative humidity (RH), and a photoperiod of 16:8 h (L: D). The outcomes of logistic regressions showed type II functional responses in all pesticide treatments and the control. The mean number of preys devoured by A. spinidens significantly decreased after testing pesticides, according to a comparison of functional response curves. In comparison to the other treatments, A. spinidens functional response curve to chlorpyrifos treatment was noticeably lower. When compared to the control, the spraying of insecticides in this study resulted in a drop in the attack rate and an increase in the handling time of exposed bugs. The fenitrothion and chlorpyrifos treatments, respectively, had the lowest attack rate (0.023 0.007) and the longest handling time (3.97 0.62). The findings indicated that integrated pest management (IPM) programmes should take A. spinidens negative effects of these insecticides into account (Gholamzadeh et al. 2014). Table 16.9 shows the significant effects of several pesticides and biopesticides on various natural enemies. In many agricultural systems in North America, the spined soldier insect, Podisus maculiventris, is a significant biocontrol agent. It has lately been found in South America as well. Many biological control businesses produce and market Podisus maculiventris. Many economically significant insect pests, including leafhoppers, aphids, and whiteflies, are preyed upon by this generalist predator. When exposed to logistic regression of the number of prey consumed as a function of the initial prey density, female predators on uncontaminated plants demonstrated a response that best fit a quadratic model with a negative first-order term. This outcome shows that a type II functional reaction occurred. Roger’s equation for type II functional response thus anticipated the curve of the number of larvae devoured by the predator based on

16.8

Factors Influencing Functional Responses

617

Table 16.9 References for the disposition of various pesticides, biopesticides, and fungicides by predators on varying densities of their prey Predator Acanthaspis pedestris Andrallus spinidens

Order/family Reduviide

Pesticides Cypermethrin

Response Type II–IV

Pentatomidae

Pentatomidae

Diazinon, Fenitrothion, Chlorpyrifos Imidacloprid

Coleoptera

Imidacloprid

Scolothrips takahashii

Thrips

Macrolophus pygmaeus

Miridae

Abamectin, fenpropathrin, fungicidesmancozeb and carbendazim Thiacloprid chlorantraniliprole

Type II, longest handling time, lowest attack rate Altered predatory rate Not affected functional response; negatively affect development and reproductive capacity; and may ultimately reduce predator population growth Altered Holling-II to Holling-III

Podisus maculiventris Serangium japonicum

Nesidiocoris tenuis

Andrallus spinidens

Pentatomidae

Diazinon Fenitrothion

Coccinella septempunctata

Coleoptera

Chlorpyrifos Thiamethoxam, lambdacyhalothrin cypermethrin, imidacloprid, profenophos chlorpyrifos

Thiacloprid reduce the attack rate and increase handling time. Chlorantraniliprole increased handling time, but not the attack rate Decreased preys consumption Altered handling time Altered Holling-II to Holling-III

References Claver et al. (2003) Gholamzadeh et al. (2014)

He et al. (2012)

Li et al. (2006)

Martinou and Stavrinides (2015)

Gholamzadeh et al. (2014)

Afza et al. (2021)

the initial prey density, estimates of female predator attack frequency, and handling duration on plants without imidacloprid drench treatment. However, neither the quadratic nor the cubic models were significant for the first-order term for predators

618

16

Bioefficacy of Insect Predators Under Laboratory

exposed to cabbage plants treated with 5.04 and 10.08 mg of imidacloprid. Therefore, there was no evidence of a functional response to these treatments. Regardless of the insecticide concentration (average of 7.25, 1.45, and 1.19 preys consumed/ predator for 0.0, 5.04, and 10.08 mg a.i./plant, respectively), predation was always higher (>twofold) without imidacloprid exposure. Spined soldier bug females responded to increasing diamondback moth populations with increased predation and increased prey density, plateauing at high densities. This is a typical type II functional response. The predatory species’ efficiency in biocontrol programmes may be restricted by this documented functional response to decreased levels of prey density. The functional response parameters, such as the predatory rate, attack rate, scrounging ability, handling time, and switching time, are generally altered by pesticide applications. It was knowledgeable about a variety of predators, including Podisus nigrispinus, (Malaquias et al. 2014). According to Cloyd and Bethke (2011), a variety of plant-feeding insects, such as aphids, mealybugs, and whiteflies, are commonly managed in greenhouses and/or interiorscapes (plant interiorscapes and conservatories) using the neonicotinoid insecticides imidacloprid, acetamiprid, dinotefuran, thiamethoxam, and clothianidin. These systemic pesticides, nevertheless, might also be damaging predators and other natural enemies. When predatory insects and mites: 1. Feed on pollen, nectar, or plant tissue contaminated with the active ingredient; 2. Consume the active ingredient of neonicotinoid insecticides while ingesting plant fluids; or, 3. Feed on hosts (prey) that have consumed leaves contaminated with the active ingredient, neonicotinoid systemic insecticides may have a negative impact on them.

16.9

Interaction Multiple Natural Enemies

Invertebrate predator interactions may have an impact on pest control. One of the most notable herbivores on creeping thistle, Cirsium arvense is the shield beetle Cassida rubiginosa (Coleoptera, Chrysomelidae), which is regarded as a biological control agent against this weed. On Cassida rubiginosa larvae, Polistes dominulus (Hymenoptera, Vespidae) was responsible for 99.4% of the predation. Polistes dominulus could therefore negate the beetle’s increased release as a biocontrol agent (Schenk and Bacher 2002). The hemipteran species Macrolophus pygmaeus and Nesidiocoris tenuis (Hemiptera: Miridae) cohabit in tomato crops in Mediterranean nations and are natural enemies of a number of pests in agroecosystems. The multiple predator effects (MPEs) on prey suppression were calculated using the multiplicative risk model (MRM) and the substitutive model when two individuals of the predators foraged at the same densities on South American tomato pinworm, Tuta absoluta (Lepidoptera: Gelechiidae), eggs. Egg consumption increased with increasing egg density and the two predators exhibited a type III functional response.

References

619

Predation rates were strongly affected by prey density. Using the MRM, we found risk reduction at intraspecific treatments at high prey density. Applying the substitutive model, we detect risk enhancement at interspecific treatments at high egg density (Michaelides et al. 2018). The effects of age-dependent parasitism of host eggs on intraguild predation (IGP) between these two species were examined in a lab setting. IGP occurs when predators and parasitoids interact at multiple trophic levels, including intraguild predation (IGP). The predatory bug was exposed to 40 parasitised and nonparasitised eggs of various ages in no-choice and choice preference tests (24, 48, and 72 h old). Different combinations of tomato leafminer eggs (30:90, 45:75, 60:60, 75:45, and 90:30 nonparasitised): parasitised eggs were used, employing eggs of various ages, to investigate switching behaviour (24, 48, and 72 h old). In experiments with no options, the predatory bug’s highest eating rate was 39.21 eggs on nonparasitised eggs that were 24 h old, and its lowest feeding rate was 1.4 eggs on parasitised eggs that were 72 h old. The predatory bug preferred to eat nonparasitised eggs with 48and 72-h-old eggs in choice tests, but there was no discernible preference for the 24-h-old eggs, according to a comparison of the Manly’s indices. The switching test results revealed that in 72-h-old eggs, there was no significant linear regression between Manly’s index and various ratios of nonparasitised eggs to parasitised and nonparasitised eggs. However, with 24- and 48-h-old eggs, this regression was substantial, and the predator’s preference depended on the proportion of nonparasitised to parasitised tomato leafminer eggs. The current study’s findings demonstrated that the intensity of IGP between N. pseudoferus and T. brassicae decreased with the age of the parasitised egg (Mohammadpour et al. 2019).

16.10 Recommendations • Storage pests and their predators have very little job to do. It is advised to take steps to learn more. Predators’ bioefficacy was seen at large pests. • The majority of research had time limits of 24–96 h. However, for accurate assessments, we advised exposing the predators for only a brief (up to 96 h) and long period of time (stadium for young ones and 15–30 days for adults). • Functional response and CO2 elevation may be of utility in the near future. • Researchers could examine the combined impact of artificial pesticides and predator density on functional responsiveness.

References Abdel-Rahman HA, Shaumrar NF, Soliman ZA, El-Agoze MM (1983) Efficiency of the anthocorid predator ylocoris flavipes (Reut.) in biological control of stored grain insects. Bull Entomol Soc Egypt 11:27–34 Abraços-Duarte G, Ramos S, Valente F et al (2021) Functional response and predation rate of Dicyphus cerastii Wagner (Hemiptera: Miridae). Insects 12(6):530

620

16

Bioefficacy of Insect Predators Under Laboratory

Afza R, Riaz MA, Afzal M et al (2021) Adverse effect of sublethal concentrations of insecticides on the biological parameters and functional response of predatory beetle Coccinella septempunctata (Coleoptera: Coccinellidae) of brassica aphid. Sarhad J Agric 37(1):226–234 Alikhani M, Safavi SA, Iranipour S (2019) Predation response of Nabis pseudoferus (Hemiptera: Nabidae) on untreated and Metarhizium anisopliae-treated larvae of Tuta absoluta (Lepidoptera: Gelechidae). J Crop Prot 8(3):311–322 Ambrose DP (1999) Assassin Bugs Oxford IBH Publ Co Pvt Ltd, New Delhi, India and Science Publishers, Inc, New Hampshire, USA, p 337 Ambrose DP (2003) Biocontrol potential of assassin bugs (Hemiptera, Reduviidae). Indian J Expe Zool 6(1):1–44 Ambrose DP, Claver MA (1997) Functional and numerical responses of the reduviid predator, Rhynocoris fuscipes F. (Het., Reduviidae) to cotton leafworm Spodoptera litura F. (Lep., Noctuidae). J Appl Entomol 121(1–5):331–336 Ambrose DP, Claver MA, Mariappan P (2000) Functional response of Rhynocoris marginatus (Heteroptera: Reduviidae) to two pests of pigeonpea (Cajanus cajan). Indian J Agric Sci 70 (9):630–632 Ambrose DP, Kumaraswami NS, Nagarajan K (2009a) Influence of predator’s age, sex and prey size on the functional response of Rhynocoris marginatus (Fabricius) (Hemiptera: Reduviidae) to Dysdercus cingulatus Fabricius (Hemiptera: Pyrrhocoridae). Hexapoda 16(1):18–24 Ambrose DP, Sebasti Rajan XJ, Nagarajan K et al (2009b) Biology, behaviour and functional response of Sphedanolestes variabilis Distant (Insecta: Hemiptera: Reduviidae: Harpactorinae), a potential predator of lepidopteran pests. Entomol Croat 13(2):33–44 Ambrose DP, Nagarajan K, Kumar AG (2013) Interaction of reduviid predator, Rhynocoris marginatus (Fabricius) (Hemiptera: Reduviidae) with its prey teak skeletonizer, Eutectona machaeralis Walker (Lepidoptera: Pyralidae) as revealed through functional response. J Entomol Res 37(1):55–60 Anderson MT, Kiesecker JM, Chivers DP et al (2001) The direct and indirect effects of temperature on a predator–prey relationship. Can J Zool 79:1834–1841 Arbogast RT (1984) Biological control of stored-product insects: status and prospects. In: Insect management for food storage and processing. AACC, Washington, DC, pp 226–233 Asasi R, Hassanpour M, Golizadeh A et al (2022) Effect of some cucumber cultivars on biological and population growth parameters of Aphis gossypii (Glover) and functional response of Chrysoperla carnea (Stephens). Int J Veg Sci 6(1):17–32 Balakrishnan P, Nagarajan K, Kumar AG et al (2011) Host preference, stage preference and functional response of Acanthaspis pedestris Stål (Hemiptera: Reduviidae) to its most preferred prey the acridid grasshopper, Orthacris maindroni Boliver. Insect Pest Management, a current scenario. Entomology Research Unit, St. Xavier’s College, Palayamkottai, pp 210–217 Ballal CR, Gupta T, Joshi S (2012) Predatory potential of two indigenous anthocorid predators on Phenacoccus solenopsis Tinsley and Paracoccus marginatus Williams and Granara de Willink. J Biol Control 26(1):18–22 Barman B, Ghosh B (2022) Role of time delay and harvesting in some predator–prey communities with different functional responses and intra-species competition. Int J Simul Model 42(6): 883–901 Bass JA, Shepard M (1974) Predation by Sycanus Indagator on larvae of Galleria mellonella and Spodoptera frugiperda. Entomol Exp Appl 17(2):143–148 Batool A, Abdullah K, Mamoon-ur-Rashid M et al (2014) Effect of prey density on biology and functional response of Chrysoperla carnea (Stephens) (Neuroptera: Chrysopidae). Pak J Zool 46(1):129–137 Berger A, Degenkolb T, Vilcinskas A et al (2017) Evaluating the combination of a parasitoid and a predator for biological control of seed beetles (Chrysomelidae: Bruchinae) in stored beans. J Stored Prod Res 74:22–26 Bhat PS, Srikumar KK, Raviprasad TN (2013) Biology, behavior, functional response and molecular characterization of Rihirbus Trochantericus Stal Var. Luteous (Hemiptera: Reduviidae:

References

621

Harpactorinae) a potential predator of Helopeltis Spp. (Hemiptera: Miridae). Entomol News 123 (4):264–277 Botteon VW, Neves JA, Godoy WAC (2017) Functional response and matrix population model of Podisus nigrispinus (Dallas,1851) (Hemiptera: Pentatomidae) fed on Chrysomya putoria (Wiedemann, 1818) (Diptera: Calliphoridae) as alternative prey. Neotrop Entomol 46(2):137– 143 Botti JMC, Martins EF, Franzin ML et al (2022) Predation of coffee berry borer by a green lacewing. Neotrop Entomol 51(1):160–163 Brower JH, Mullen MA (1990) Effects of Xylocoris flavipes (Hemiptera: Anthocoridae) releases on moth populations in experimental peanut storages. J Entomol Sci 25(2):268–276 Cabral S, Soares AO, Garcia P (2009) Predation by Coccinella undecimpunctata L. (Coleoptera: Coccinellidae) on Myzus persicae Sulzer (Homoptera: Aphididae): effect of prey density. Biol Control 50(1):25–29 Carter MC, Sutherland D, Dixon AFG (1984) Plant structure and the searching efficiency of coccinellid larvae. Oecologia 63:394–397 Chandral S, Sinazer RL, Daniel S et al (2009) Evaluation of biocontrol potential of Rhynocoris fuscipes Fabricius (Heteroptera: Reduviidae: Harpactorinae) on caterpillars of Pterophorus lienigianus Z., infesting Solanum melongena Linn., using functional response. Hexapoda 16 (1):71–75 Claver MA, Ravichandran B, Khan MM et al (2003) Impact of cypermethrin on the functional response, predatory and mating behaviour of a non-target potential biological control agent Acanthaspis pedestris (Stål) (Het., Reduviidae). J Appl Entomol 127(1):18–22 Cloyd RA, Bethke JA (2011) Impact of neonicotinoid insecticides on natural enemies in greenhouse and interiorscape environments. Pest Manag Sci 67(1):3–9 Cohen AC, Byrne DN (1992) Geocoris punctipes as a predator of Bemisia tabaci: a laboratory evaluation. Entomol Exp Appl 64(2):195–202 Cohen AC, Tang R (1997) Relative prey weight influences handling time and biomass extraction in Sinea confusa and Zelus renardii (Heteroptera: Reduviidae). Environ Entomol 26(3):559–565 Coll M, Ridgway RL (1995) Functional and numerical responses of Orius insidiosus (Heteroptera: Anthocoridae) to its prey in different vegetable crops. Ann Entomol Soc Am 88(6):732–738 De Clercq P, Merlevede F, Mestdagh I et al (1998). Predation on the tomato looper Chrysodeixis chalcites (Esper) (Lep., Noctuidae) by Podisus maculiventris (Say) and Podisus nigrispinus (Dallas) (Het., Pentatomidae)) J Appl Entomol 122(1–5):93–98 De Clercq P, Mohaghegh J, Tirry L (2000) Effect of host plant on the functional response of the predator Podisus nigrispinus (Heteroptera: Pentatomidae). Biol Control 18(1):65–70 Di N, Zhang K, Xu Q et al (2021) Predatory ability of Harmonia axyridis (Coleoptera: Coccinellidae) and Orius sauteri (Hemiptera: Anthocoridae) for Suppression of Fall Armyworm Spodoptera frugiperda (Lepidoptera: Noctuidae). Insects 12(12):1063 Donnelly BE, Phillips TW (2001) Functional response of Xylocoris flavipes (Hemiptera: Anthocoridae)-effects of prey species and habitat. Environ Entomol 30(3):617–624 Fathi SAA, Nouri-Ganbalani G (2010) Assessing the potential for biological control of potato field pests in Ardabil, Iran: functional responses of Orius Niger (wolf.) and O. minutus (L.) (Hemiptera: Anthocoridae). J Pest Sci 83(1):47–52 Ferdous J, Islam W, Parween S (2010) Suppression of a population of Tribolium castaneum by IGRS in presence of a predator, Xylocoris flavipes. Trop Agric Res Ext 13(2):25–28 Gaikwad MB, Verma SC, Sharma PL et al (2022) Functional response and predatory potential of coccinellid predator, Oenopia kirbyi Mulsant (Coccinellidae: Coleoptera) on rose aphid, Macrosiphum rosae L. (Aphididae: Hemiptera). Int J Trop Insect Sci 42:2233–2239 Gholami N, Aleosfoor M, Hosseini M et al (2022) Predation functional response and demographic parameters of Orius albidipennis (Hemiptera: Anthocoridae) on Schizaphis graminum (Hemiptera: Aphididae): effect of host plant morphological attributes. Biocontrol Sci Tech 32 (3):362–380

622

16

Bioefficacy of Insect Predators Under Laboratory

Gholamzadeh CM, Hajizadeh J, Ghadamyari M et al (2014) Sublethal effects of diazinon, fenitrothion and chlorpyrifos on the functional response of predatory bug, Andrallus spinidens Fabricius (Hem.: Pentatomidae) in the laboratory conditions. J King Saud Univ Sci 26(2):113– 118 Giroux S, Duchesne RM, Coderre D (1995) Predation of Leptinotarsa decemlineata (Coleoptera, Chrysomelidae) by Coleomegilla maculata (Coleoptera, Coccinellidae)—comparative effectiveness of predator developmental stages and effect of temperature. Environ Entomol 24:748– 754 Gitonga LM, Overholt WA, Löhr B et al (2002) Functional response of Orius albidipennis (Hemiptera: Anthocoridae) to Megalurothrips sjostedti (Thysanoptera: Thripidae). Biol Control 24(1):1–6 Gómez-Domínguez NS, Rodríguez-Leyva E, Lomeli-Flores JR et al (2021) Examining potential intra-guild predation of Peristenus relictus by Geocoris punctipes, and its effects on the biological control of Lygus lineolaris. Biol Control 156:104557 Gresens SE, Cothran ML, Thorp JH (1982) The influence of temperature on the functional response of the dragonfly Celithemis fasciata (Odonata: Libellulidae). Oecologia 53:281–284 Hassanpour M, Nouri-Ganbalani G, Mohaghegh J et al (2009) Functional response of different larval instars of the green lacewing, Chrysoperla carnea (Neuroptera: Chrysopidae), to the twospotted spider mite, Tetranychus urticae (Acari: Tetranychidae). J Food Agric Environ 7 (2):424–428 Hassanpour M, Mohaghegh J, Iranipour S et al (2011) Functional response of Chrysoperla carnea (Neuroptera: Chrysopidae) to Helicoverpa armigera (Lepidoptera: Noctuidae): effect of prey and predator stages. Insect Sci 18(2):217–224 Hassell MP, Lawton JH, Beddington JR (1976) The components of arthropod predation: the prey death-rate. J Anim Ecol 45:135–164 Hassell MP, Lawton JH, Beddinger JR et al (1977) Sigmoid functional responses by invertebrate predators and parasitoids. J Anim Ecol 46:249–262 He Y, Zhao J, Zheng Y et al (2012) Lethal effect of imidacloprid on the coccinellid predator Serangium japonicum and sublethal effects on predator voracity and on functional response to the whitefly Bemisia tabaci. Ecotoxicology 21(5):1291–1300 Heimpel GE, Hough-Goldstein JA (1994) Components of the functional response of Perillus bioculatus (Hemiptera: Pentatomidae). Environ Entomol 23(4):855–859 Hildrew AG, Townsend CR (1977) The influence of substrate on the functional response of Plectrocnemia conspersa (Curtis) larvae (Trichoptera: Polycentropodidae). Oecologia 31:21–26 Holling CS (1959) Some characteristics of simple type of predation and parasitism. Canadian Entomol 91:385–395 Islam Y, Shah FM, Güncan A et al (2022) Functional response of Harmonia axyridis to the larvae of Spodoptera litura: the combined effect of temperatures and prey instars. Front. Plant Sci 13 Jalali MA, Tirry L, De Clercq P (2010) Effect of temperature on the functional response of Adalia bipunctata to Myzus persicae. BioControl 55(2):261–269 Kalsi M, Seal DR, Nuessly GS et al (2014) Distribution of Zelus longipes (Hemiptera: Reduviidae) in South Florida corn fields and its functional response to corn-infesting picture-winged flies (Diptera: Ulidiidae). Environ Entomol 43(5):1223–1234 Kumar AG, Ramar M, Murugan K (2019) Functional response of an assassin bug Rhynocoris kumarii Ambrose and Livingstone (Insecta: Hemiptera: Reduviidae) to Odontotermes Brunneus (Hagen) (Insecta: Isoptera: Termitidae). Int J Environ Sci Nat Resour 19(1):19–24 Kumar V, Mehra L, McKenzie CL et al (2020) Functional response and prey stage preference of Delphastus catalinae and D. pallidus (Coleoptera: Coccinellidae) on Bemisia tabaci (Hemiptera: Aleyrodidae). Biocontrol Sci Technol 30(6):581–591 Lam TX, Lien NTP, Cuong NQ (2020) The functional response and predatory ability of the reduviid Coranus fuscipennis Reuter (Heteroptera: Reduviidae) fed on the rice meal moth Corcyra cephalonica (Stainton). Vietnam J Biotechnol 18(2):265–272

References

623

LeCato GL, Arbogast RT (1979) Functional response of Xylocoris flavipes to Angoumois grain moth and influence of predation on regulation of laboratory populations. J Econ Entomol 72 (6):847–849 Li DX, Tian J, Shen ZR (2006) Effects of pesticides on the functional response of predatory thrips, Scolothrips takahashii to Tetranychus viennensis. J Appl Entomol 130(5):314–322 Liu P, Jia W, Zheng X et al (2018) Predation functional response and life table parameters of Orius sauteri (Hemiptera: Anthocoridae) feeding on Megalurothrips usitatus (Thysanoptera: Thripidae). Fla Entomol 101(2):254–259 Loko YL, Djagoun AD, Dannon EA et al (2017) Functional response of the predators Alloeocranum biannulipes (Hemiptera: Reduviidae) and Teretrius nigrescens (Coleoptera: Histeridae) feeding on Dinoderus porcellus (Coleoptera: Bostrichidae) infesting yam chips. Environ Entomol 46(1):84–91 Loko YLE, Onzo A, Datinon B et al (2020) Population dynamics of the predator Alloeocranum biannulipes Montrouzier and Signoret (Hemiptera: Reduviidae) feeding on the larger grain borer, Prostephanus truncatus (Horn) (Coleoptera: Bostrichidae), infesting cassava chips. Egypt J Biol Pest Control 30(1):1–12 Loko YLE, Toffa J, Gavoedo DM et al (2022) Effect of population density on oviposition, development, and survival of Alloeocranum biannulipes (Hemiptera: Reduviidae) preying on Dinoderus porcellus (Coleoptera: Bostrichidae). J Basic Appl Zool 83(1):1–8 López SN, Orozco Muñoz A, Andorno A et al (2019) Predatory capacity of Tupiocoris cucurbitaceus (Hemiptera Miridae) on several pests of tomato. Bull Insectol 72(2):201–205 Mack TP, Smilowitz Z (1982a) Using temperature-mediated functional response models to predict the impact of Coleomegilla maculata (DeGeer) adults and third instar larvae on green peach aphids. Environ Entomol 11:46–52 Mack TP, Smilowitz Z (1982b) CMACSIM, a temperature dependent predator–prey model simulating the impact of Coleomegilla maculata (DeGeer) on green peach aphids on potato plants. Environ Entomol 11:1193–1201 Mahzoum AM, Villa M, Benhadi-Marín J et al (2020) Functional response of Chrysoperla carnea (Neuroptera: Chrysopidae) larvae on Saissetia oleae (Olivier) (Hemiptera: Coccidae): implications for biological control. Agronomy 10(10):1511 Malaquias JB, Ramalho FS, Omoto C et al (2014) Imidacloprid affects the functional response of predator Podisus nigrispinus (Dallas) (Heteroptera: Pentatomidae) to strains of Spodoptera frugiperda (JE Smith) on Bt cotton. Ecotoxicology 23(2):192–200 Martinou AF, Stavrinides MC (2015) Effects of sublethal concentrations of insecticides on the functional response of two mirid generalist predators. PLoS One 10(12):e0144413 Memon SA, Omar D, Muhamad R et al (2015) Functional responses of green lacewing Chrysoperla nipponensis (Neuroptera: Chrysopidae) reared on natural and herb based artificial diet. J Entomol Zool Stud 3:80–83 Messina FJ, Hanks JB (1998) Host plant alters the shape of the functional response of an aphid predator (Coleoptera: Coccinellidae). Environ Entomol 27(5):1196–1202 Michaelides G, Sfenthourakis S, Pitsillou M et al (2018) Functional response and multiple predator effects of two generalist predators preying on Tuta absoluta eggs. Pest Manag Sci 74(2): 332–339 Mohammadpour M, Hosseini M, Karimi J et al (2019) Effect of age-dependent parasitism in eggs of Tuta absoluta (Lepidoptera: Gelechiidae) on intraguild predation between Nabis pseudoferus (Hemiptera: Nabidae) and Trichogramma brassicae (Hymenoptera: Trichogrammatidae). J Insect Sci 19(3):27 Montoya-Alvarez AF, Ito K, Nakahira K et al (2010) Functional response of Chrysoperla nipponensis and C. carnea (Neuroptera: Chrysopidae) to the cotton aphid Aphis gossypii Glover (Homoptera: Aphididae) under laboratory conditions. Appl Entomol Zool 45(1):201–206 Muniyandi J, Kumar AG, Nagarajan K (2011) Host preference, stage preference and functional response of assassin bug, Rhynocoris kumarii Ambrose and Livingstone (Hemiptera: Reduviidae) to its most preferred prey tobacco cutworm, Spodoptera litura (F.). Insect Pest

624

16

Bioefficacy of Insect Predators Under Laboratory

Management, a current scenario. Entomology Research Unit, St. Xavier’s College, Palayamkottai, pp 240–248 Munyaneza J, Obrycki JJ (1997) Functional response of Coleomegilla maculata (Coleoptera: Coccinellidae) to Colorado potato beetle eggs (Coleoptera: Chrysomelidae). Biol Control 8 (3):215–224 Nagai K, Yano E (2000) Predation by Orius sauteri (Poppius) (Heteroptera: Anthocoridae) on Thrips palmi Karny (Thysanoptera: Thripidae): functional response and selective predation. Appl Entomol Zool 35(4):565–574 Nagarajan K, Rajan K, Ambrose DP (2010) Functional response of assassin bug, Rhynocoris fuscipes (Fabricius) (Hemiptera: Reduviidae) to cucumber leaf folder, Diaphania indicus Saunders (Lepidoptera: Pyraustidae). Entomon 35(1):1–7 Nishi A, Imamura T, Miyanoshita A et al (2004) Predatory abilities of Amphibolus venator (Klug) (Hemiptera: Reduviidae), a predator of stored-product insect pests. Appl Entomol Zool 39 (2):321–326 O’Neil RJ (1997) Functional response and search strategy of Podisus maculiventris (Heteroptera: Pentatomidae) attacking Colorado potato beetle (Coleoptera: Chrysomelidae). Environ Entomol 26(6):1183–1190 Papanikolaou NE, Milonas PG, Demiris N et al (2014) Digestion limits the functional response of an aphidophagous coccinellid (Coleoptera: Coccinellidae). Ann Entomol Soc Am 107(2): 468–474 Parajulee MN, Shrestha RB, Leser JF et al (2006) Evaluation of the functional response of selected arthropod predators on bollworm eggs in the laboratory and effect of temperature on their predation efficiency. Environ Entomol 35(2):379–386 Pehlivan S, Alinç T, Achiri TD et al (2020) Functional responses of two predatory bugs (Hemiptera: Anthocoridae) to changes in the abundance of Tetranychus urticae (Acari: Tetranychidae) and Bemisia tabaci (Hemiptera: Aleyrodidae). Eur J Entomol 117:49 Poley K, Bahlai C, Grieshop M (2018) Functional response of generalist predators to Halyomorpha halys (Hemiptera: Pentatomidae) eggs. Environ Entomol 47(5):1117–1127 Rehman SU, Zhou X, Ali S et al (2020) Predatory functional response and fitness parameters of Orius strigicollis Poppius when fed Bemisia tabaci and Trialeurodes vaporariorum as determined by age-stage, two-sex life table. Peer J 8:e9540 Rios-Velasco C, Nájera-Miramontes D, Jacobo-Cuellar JL et al (2017) Predation capability and functional response of Chrysoperla carnea1 to Choristoneura rosaceana2 under laboratory conditions. Southwest Entomol 42(3):677–690 Robert PD, Kevin AH (2020) Predator type influences the frequency of functional responses to prey in marine habitats. Biol Lett 16:20190758. https://doi.org/10.1098/rsbl.2019.0758 Sahayaraj K, Asha A (2010) Biological control potential evaluation of Rhynocoris kumarii Ambrose and Livingstone (Hemiptera: Reduviidae) on Aphis craccivora (Koch.) (Hemiptera: Aphididae). Indian J Agric Res 44(4):281–287 Sahayaraj K, Kalidas S, Tomson M (2012) Stage preference and functional response of Rhynocoris longifrons (Stål) (Hemiptera: Reduviidae) on three hemipteran cotton pests. Braz Arch Biol Technol 55:733–740 Sarkar SC, Wang E, Zhang Z et al (2019) Laboratory and glasshouse evaluation of the green lacewing, Chrysopa pallens (Neuroptera: Chrysopidae) against the western flower thrips, Frankliniella occidentalis (Thysanoptera: Thripidae). Appl Entomol Zool 54(1):115–121 Sarmento RA, Pallini A, Venzon M et al (2007) Functional response of the predator Eriopis connexa (Coleoptera: Coccinellidae) to different prey types. Braz Arch Biol Technol 50:121– 126 Schenk D, Bacher S (2002) Functional response of a generalist insect predator to one of its prey species in the field. J Anim Ecol 71(3):524–531 Senrayan R (1988) Functional response of Eocanthecona furcellata (Heteroptera: Pentatomidae) in relation to prey density and defence with reference to its prey Latoia lepida (Cramer) (Lepidoptera: Lemacodidae). Proc Anim Sci 97(4):339–345

References

625

Seyfollahi F, Esfandiari M, Mossadegh MS et al (2019) Functional response of Hyperaspis polita (Coleoptera, Coccinellidae) to the recently invaded mealybug Phenacoccus solenopsis (Hemiptera, Pseudococcidae). Neotrop Entomol 48(3):484–495 Shahpouri A, Yarahmadi F, Zandi Sohani N (2019) Functional response of the predatory species Orius albidipennis Reuter (Hemiptera: Anthocoridae) to two life stages of Bemisia tabaci (Genn.) (Hemiptera: Aleyrodidae). Egypt J Biol. Pest Control 29(1):1–6 Shukla AN, Singh R, Tripathi CPM (1990) Effect of predation period on the functional response of Coccinella septempunctata Linn. (Coleoptera: Coccinellidae), a predator of Lipaphis erysimi Kalt. (Hemiptera: Aphididae). J Adv Zool 11(1):27–32 Sing SE (1997) Suppression of bruchids infesting stored grain legumes with the predatory bug Xylocoris flavipes (Reuter)(Hemiptera: Anthocoridae). Master thesis, McGill University, PP 134 Sivapragasam A, Asma A (1985) Development and reproduction of the mirid bug, Cyrtorhinus lividipennis (Heteroptera: Miridae) and its functional response to the brown planthopper. Appl Entomol Zool 20(4):373–379 Srikumar KK, Bhat PS, Raviprasad TN et al (2014) Biology, behaviour and functional response of Cydnocoris gilvus Brum. (Heteroptera: Reduviidae: Harpactorinae) a predator of Tea Mosquito Bug (Helopeltis antonii Sign.) on cashew in India. J Threat Taxa 6(6):5864–5870. https://doi. org/10.11609/JoTT.o3815.5864-70 Srikumar KK, Sukumaran S, Kumar BS et al (2017) Life history and functional response of two species of Reduviids (Hemiptera: Reduviidae: Harpactorinae) in tea. J Agric Urban Entomol 33(1):44–56 Stasek DJ, Radl JN, Crist TO (2018) The functional response and prey preference of generalist Nabis (Hemiptera: Nabidae) predators to leafhopper prey (Hemiptera: Cicadellidae). Can Entomol 150(2):190–200 Suenaga H, Hamamura T (1998) Laboratory evaluation of carabid beetles (Coleoptera: Carabidae) as predators of diamondback moth (Lepidoptera: Plutellidae) larvae. Environmental Entomol 27 (3):767–772 Sunil V, Sampathkumar M, Lydia C et al (2018) Biology, predatory potential and functional response of Rhynocoris fuscipes (Fabricius) (Hemiptera: Reduviidae) on rice brown planthopper, Nilaparvata lugens (Stal) (Homoptera: Delphacidae). J Exp Zool 21(1):259–263 Tomson M (2021) Functional response, host stage preference and development of Rhynocoris fuscipes (Fab.) (Heteroptera: Reduviidae) for two cotton pests. J Biopest 14(1):12–21 Torres JB, Silva-Torres CS, Ruberson JR (2004) Effect of two prey types on life-history characteristics and predation rate of Geocoris floridanus (Heteroptera: Geocoridae). Environ Entomol 33(4):964–974 Uiterwaal SF, DeLong JP (2018) Multiple factors, including arena size, shape the functional responses of ladybird beetles. J Appl Ecol 55(5):2429–2438 Varshney R, Budhlakoti N (2022) Biology and functional response of the predator, Dortus primarius (Distant) (Hemiptera: Miridae) preying on Frankliniella schultzei (Trybom) (Thysanoptera: Thripidae). Egypt J Biol Pest Control 32(1):1–7 Varshney R, Budhlakoti N, Ballal CR (2020) Functional response of three species of predatory pirate bugs to different densities of blossom thrips, Frankliniella schultzei Trybom (Thysanoptera: Thripidae). Curr Sci 118(5):00113891 Wiedenmann RN, O’Neil RJ (1991) Laboratory measurement of the functional response of Podisus maculiventris (Say) (Heteroptera: Pentatomidae). Environ Entomol 20(2):610–614 Zada H, Ahmad B, Nadeem M et al (2016) Functional response of Chrysoperla carnea Stephen (Neuroptera: Chrysopidae) fed on cabbage aphid, Brevicoryne brassicae (Linnaeus) under laboratory conditions. Pak J Zool 48:165–169 Zanuncio JC, Silva CADD, Lima ERD et al (2008) Predation rate of Spodoptera frugiperda (Lepidoptera: Noctuidae) larvae with and without defense by Podisus nigrispinus (Heteroptera: Pentatomidae). Braz Arch Biol Technol 51:121–125

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen House/Polyphagous

17

Contents 17.1 17.2

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hemiptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2.1 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2.2 Reduviids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2.3 Pentatomide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2.4 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.2.5 Entatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.3 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.4 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.5 Multiple Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.5.1 Predation by a Complex of Predators Made Up of Adults of the Lygaeid . . . 17.6 Multiple Controlled Conditions in Field Cage for Life Traits and Bioefficacy . . . . . . . . 17.7 Recommendation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

17.1

627 629 630 634 636 637 640 641 645 648 650 657 659 659

Introduction

Researchers have researched a number of arthropod predators by enclosing entire rows of crop plants in field cages. Instead of focussing on specific insects, these investigations mostly estimated average predation levels for representative populations. Although these data can be modified for systems modelling, it is frequently challenging to remove the mistake brought on by predators from entering the field. Therefore, it is advised to first conduct bioefficacy tests under controller field cage, polyhouse, or screen house conditions in order to determine the predator’s suitability. Additionally, it is a requirement for all biological control programmes. Cages alter the microenvironment, particularly the temperature, which is regarded to be crucial in deciding how encounters between predators and prey turn out. Interestingly, there was no difference between the number of grain aphids in control plots and those where the number of ground predators was reduced when only polythene # The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_17

627

628

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

Fig. 17.1 One of the authors is introducing reduviid predators near a cotton plant while it is in a screen house environment

enclosures, 60 cm high, buried to a depth of 30 cm, and not cages, were used. These do not alter the microenvironment of the manipulated plots; they allow aphids to emigrate, but exclude ground predators (says Kindlmann et al. 2015). It might be challenging to experimentally evaluate predation in biological control, but it is essential to understand how predators affect the dynamics of pest population growth. The effectiveness of current predators is documented, the effects of earlier releases are quantified, and pest life phases or seasonal times that could be targeted for subsequent releases are also identified (Obrycki John 1992). Ecologists frequently conduct field studies to learn more about the structure and function of ecosystems. The comprehensive causal understanding required for prediction, however, is frequently not provided by trials carried out at a scale representative of a whole ecosystem. Instead, ecologists attempt to grasp such causal relationships by conducting experiments in tiny enclosures or cages. In order to isolate certain species combinations and gain a thorough grasp of the interspecies interaction mechanisms, enclosure cages provide the fine-scale precision and control required. However, meeting a few crucial design requirements is crucial for the relevance of such insights to predicting the performance of an entire ecosystem. These requirements include ensuring that enclosure experiments are carried out in actual field settings rather than in artificial laboratory settings, that the behaviour of mobile species is not significantly hindered, and that the experiment is carried out over time scales that correspond to the life cycles of the species (Schmitz 2008). To prevent the introduction of any macroorganism, the cage, screen house, or polyhouse should be completely enclosed or coated with standard materials (Fig. 17.1). Given the orientation, it ought to be built from east to west. Light, temperature, humidity, CO2, and/or O2 levels should all be under control for a very advanced setup. The sort of crops kept inside the cage should be maintained at the specified size.

17.2

17.2

Hemiptera

629

Hemiptera

In terrestrial and aquatic habitats, heteroopteran predators play a significant role in predatory guilds. The majority of heteropteran species have omnivorous diets, and most heteropteran families have records of intraguild predation. Zoophytophagous species commonly interact with one another within their own guilds. By offering an abundance of alternate prey, an increase in extraguild prey density is frequently projected to decrease intraguild predation between guild members. However, a rise in the density of intraguild predators may also be linked to an increase in the density of extraguild prey, strengthening intraguild predation. It is challenging to assess the cumulative impact of these potentially conflicting impacts on intraguild predation. The majority of research has been done in the lab, with communities of predators and prey that have been artificially simplified as well as spatial and temporal scales that may not accurately represent field settings. Before reporting an observational case study examining the impact of extraguild prey density on the intensity of intraguild predation at larger spatial and temporal scales in unmanipulated cotton fields, we review experimental studies examining how extraguild prey density influences the intensity of intraguild predation. Aphid and mite abundance in fields were significantly correlated with enhanced survival of intraguild prey and not with higher densities of intraguild predators (lacewing larvae). In this system, the broader spatial and temporal scales of commercial agriculture are likewise affected by extraguild prey’s capacity to reduce the intensity of intraguild predation, as previously observed in small-scale field trials (Lucas and Rosenheim 2011). Due to its voracity and affinity for whitefly prey, Serangiurn parcesetosum is a promising biological control agent against whiteflies in the Bemisia complex. It has attracted little study attention despite its promise as a biological control agent against Alcyrodid whiteflies. Legaspi et al. (2001) investigated Serangiurn parcesetosum adults as predators of citrus blackfly eggs in a field cage trial to measure the biological control effectiveness; the investigation was carried out in a commercial grapefruit orchard near Hargill, Texas. Adult Serangiurn parcesetosum were collected from a colony that was raised in Mission, Texas. Individual terminals with 6 to 10 leaves each were encircled by screen cages. The Serangiurn parcesetosum were starved for about 12 h before being released (one beetle per cage in the treatment cages), while there were no predators in the control cages. After allowing the predators to feast on the citrus blackfly eggs for 12 days, the leaf terminals were taken to the lab to count how many eggs hatchet. Because Serangiurn parcesetosum punctured eggs in earlier studies without fully devouring them, egg hatch rather than predation was detected. There were no discernible changes between the control and treatment groups in the pre-release egg counts. The average number of eggs per cage in the predator treatment was 383.4 as opposed to 367 in the control. In comparison to the control cages, considerably fewer eggs hatched after predation in the treated cages. Only 42.8% of citrus blackfly eggs hatch, even in the absence of predators. Egg hatching was reduced to 12.5% after Serangiurn parcesetosum preyed on them for 12 days. These findings suggest that Serangiurn parcesetosum may significantly

630

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

reduce citrus blackfly populations in the field while not being as voracious on eggs as it is on silver leaf whitefly nymphs (Legaspi et al. 2001). Predator Serangiurn parcesetosum

pests Citrus blackfly

Plant Citrus

Observation Not voracious on citrus blackfly eggs

References Legaspi et al. (2001)

17.2.1 Anthocoridae Strawberries are a crucial crop for the organic farming industry in Portugal. The western flower thrips, Frankliniella occidentalis (Thysanoptera: Thripidae), posed a significant threat to strawberries in the Algarve in 1990 and 1991. The majority of anthecoride predators were used to control thrips in naturally occurring agroecosystems and/or plants. However, certain intercrops or trap crops were also planted alongside the solo crops because adult predators require pollen. Orius spp. initially proved to be an efficient biocontrol agent for the western flower thrips (WFT) in commercial sweet peppers. By the end of the growing season, the imported Orius insidiosus had been replaced by the newly introduced native Orius niger. A common naturally occurring species in Portugal, Orius laevigatus (Heteroptera: Anthocoridae), was experimentally dispersed in strawberry plantings in 1993. In cages where the Orius were released, there were fewer thrips per bloom. There was a noticeable increase in the composite plant Chamaemelum mixtum in one of the cages where Orius laevigatus was released. This weed produced a lot of pollen, and after the weed flowered, there was a lower thrips population and a greater Orius population in this plot. It is unclear whether the decrease in thrips was brought on by Orius laevigatus, which multiplied as a result of feeding on weed pollen, or by a similar transfer of thrips to the Compositae, which were also drawn to their pollen. It would be very helpful to understand how compost weeds affect the number of thrips populations on strawberries (Frescata and Mexia 1996). Spanish greenhouse plantations are well known for their employment of Orius laevigatus as a pest control method. Frankliniella occidentalis, a western flower thrips, has been effectively and widely controlled in several polyhouse crops by using Orius spp. In Germany, good results were obtained with two releases of Orius laevigatus at 2-week intervals (0.5–1 insect/m2) in cucumbers. In the United Kingdom, releases of one to two Orius laevigatus per plant controlled Frankliniella occidentalis on pepper. Frankliniella occidentalis could be effectively controlled on greenhouse cucumbers in France by Orius majusculus. In the United Kingdom, Frankliniella occidentalis infestations of chrysanthemums in a greenhouse were successfully managed by Orius majusculus, and in the Netherlands, Orius insidiosus was released on Saintpaulia, roses, and other greenhouse plants. Chrysanthemum and Saintpaulia had less pests, while roses did not. Frankliniella occidentalis caused 90% damage to Pink Pompon and 30–50% damage to D-Mark in untreated chrysanthemum plots. In contrast, damage to D-Mark (a resistant variety) did not surpass 5%

17.2

Hemiptera

631

during the course of the growth season when Orius insidiosus was discharged, whereas damage to Pink Pompon did not reach 20% in the last 8 weeks. In France, Frankliniella occidentalis was discovered to be controlled by Orius laevigatus in greenhouse strawberries (refer the review of Ballal and Yamada 2016). The predatory behaviour of adult female Orius niger on Frankliniella occidentalis and Thrips tabaci (Thysan., Thripidae) larval I–II and adults was studied in controlled environment chambers in June 1997. This investigation was done on Capsicum annuum leaves with predator/number of thrips ratios of 1/5, 1/10, 1/20, and 1/30 for larval I–II and adults separately for each species of thrips. The larval stages (larvae I–II) and adults of the two thrips species were successfully preyed upon by Orius niger. The researcher Deligeorgidis (2002) came to the conclusion that Orius niger might be successfully used for the biological control of thrips in greenhouse crops. The population growth rate and biological control effectiveness of Orius strigicollis against Thrips palmi with various initial densities in plastic houses were evaluated in the field. Three times after 0, 5, 30, and 100 thrips were injected per 30 cucumber plants, Orius strigicollis was released. Each plot received a release of 360 predators altogether. Orius strigicollis was unable to control Thrips palmi population because Orius strigicollis population did not rapidly rise after inoculation of 5 Thrips palmi per 30 plants in each plot in plastic houses. Thrips palmi population rose swiftly with 16.3 individuals/plant in the fourth week. The density of Orius strigicollis rose in the plot with 30 thrips inoculation, reaching 2.1 individuals per plant in the third week and 9.8 individuals per plant in the sixth. In the sixth week, Thrips palmi density increased to 33.8 individuals per plant. The density of Orius strigicollis grew to 9 individuals per plant in the sixth week in the 100 thrips inoculation plot, which was comparable to the density in the 30 thrips inoculation plot. However, in the sixth week, Thrips palmi density quickly increased from 2.9 to 384.7 individuals per plant. Orius strigicollis’ ability to manipulate Thrips palmi did not manifest. Orius strigicollis density in the five thrips inoculation plots in autumn culture was so low that it was unable to control thrips density. The Thrips palmi density in the 100 thrips inoculation plot was so high that Orius strigicollis was unable to control it. Orius strigicollis had a density of 4.8 individuals per plant in the sixth week of the 30 thrips inoculation plot, and the thrips population had been controlled. However, due to the low temperature and brief photoperiod after the sixth week, Orius strigicollis was unable to control the thrips population. These findings suggest that Orius strigicollis may be able to suppress Thrips palmi on cucumber in the spring culture (Kim et al. 2006). Frankliniella occidentalis and Thrips tabaci (Thysanoptera: Thripidae) are the main pests of sweet pepper in Northwest Italy in 2002–2003. Research was done to evaluate the anthocorid Orius laevigatus’s existence naturally on pepper and to contrast its capacity for colonisation and predation with that of artificially introduced species. Twelve sweet pepper greenhouses were used for the experiments, and Orius laevigatus was released in six of them. Thrips and anthocorids were sampled on pepper by taking flowers, and Orius spp. was also taken from nearby natural flora over the period of late May to early October. Orius organisms were discovered in all

632

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

investigated greenhouses regardless of the releases; however, Orius niger—which was also discovered on a variety of wild plants—was the most common species. If not disturbed by pesticide treatment, it spontaneously colonised crops starting in late June and proved to be the most effective sweet pepper predator in the examined area. Orius laevigatus, on the other hand, was only ever discovered in the greenhouses where it had been discharged. However, before the native species naturally colonised the area as a result of its introduction, thrips were controlled (Bosco et al. 2008). Orius insidiosus was found to be a more effective predator of adult Scirtothrips dorsalis in laboratory experiments when compared to Amblyseius swirskii, although this trend was not seen with thrips larvae. In comparison to up to 13 Scirtothrips dorsalis and > 40% damage on control plants, both predator species maintained 0.5 thrips per leaf and 1% foliar damage on all pepper kinds after 5 weeks at a rate of 20 predators per infested pepper plant. In a second trial, a reduced rate of predators (10 per plant) led to about 20% foliar damage, compared to >90% on control plants, indicating slightly less effective control. In comparison to Amblyseius swirskii alone, plants treated with Orius insidiosus consistently had less adult Scirtothrips dorsalis and plant damage. In addition, we found that some pepper species were more susceptible to thrips than others, with Trinidad perfume and Brigade hybrid suffering the least damage in comparison to Large Red Cherry and Serrano. Our findings imply that both species may be utilised in conjunction without suffering from diminished efficacy due to intraguild predation (Doğramaci et al. 2011). These results demonstrate that both predators were efficient predators of the chilli thrips Scirtothrips dorsalis on pepper. On greenhouse pepper, the effectiveness of releasing Orius sauteri to manage thrips and aphids was tested. To create the population of Orius sauteri, flowering marigolds (Tagetes eracta) were utilised as the banker plant. According to the findings, Orius sauteri could manage pests on pepper plants. The control efficiencies of releasing Orius sauteri at rates of 0.5 or 1 per m2 on Tetranychus cinnabarinus and thrips varied from 97.20% to 99.95% 5 to 7 weeks after release. Orius sauteri had a control efficiency of 96.39% when released at a rate of 1 per square metre against aphids, but only 22.78% when released at a rate of 0.5 per square metre (Jiang et al. 2011). Gynaikothrips uzeli was found to be a highly effective prey item for Montandoniola confusa (=moraguesi) in greenhouse studies on extensively infested Ficus benjamina. Montandoniola confusa (=moraguesi) reproduced all year long, reducing thrips populations by 95% and leaf galls by up to 77% in just 5 weeks, according to tests with three Ficus benjamina cultivars. Orius insidiosus, in contrast, had no effect on Gynaikothrips uzeli populations inside leaf galls (Arthurs et al. 2011). Orius majusculus (Heteroptera: Anthocoridae) can be found in numerous Mediterranean crops throughout central and southern Europe during the growing season, as well as in the adjacent wild flora. Frankliniella occidentalis (Thysanoptera: Thripidae), a western flower thrips, has been successfully controlled in the field by this polyphagous predator, which preys on a variety of pests. Orius majusculus appears to be a viable biological control agent in greenhouses of that region. Having a diverse diet is particularly significant in areas where prey species change, like in ephemeral and disturbed agroecosystems of the Mediterranean area

17.2

Hemiptera

633

Table 17.1 Collection of anthecorid predators that have been tested against various pests of particular crops, along with references and recommendations Predator Orius laevigatus

Pests Frankliniella occidentalis

Plant Strawberries

Orius laevigatus

Frankliniella occidentalis

Various polyhouse

Observation Compost weeds should be maintained along with sole crop Release number is a deciding factor

Orius niger

Frankliniella occidentalis Thrips tabaci Thrips palmi

Sweet pepper

Successful predator

Cucumber Sweet pepper

Kim et al. (2006) Bosco et al. (2008)

Pepper

Successful predator

Marigolds

Successful predator

Montandoniola confusa Orius insidiosus Orius majusculus

Frankliniella occidentalis Thrips tabaci Scirtothrips dorsalis Thrips and aphids Gynaikothrips uzeli Gynaikothrips uzeli Frankliniella occidentalis

Bioefficcy depends on cultivation season Orius niger is a natural predator

Ficus benjamina Ficus benjamina Mediterranean crops

Established and successful predator Not a successful predator Promising predator

Blaptostethus pallescens

Tetranychus urticae

Cucumber

Promising biocontrol agent

Doğramaci et al. (2011) Jiang et al. (2011) Arthurs et al. (2011) Arthurs et al. (2011) Pumariño and Alomar (2012) Ghongade et al. (2022)

Orius strigicollis Orius laevigatus Orius insidiosus Orius sauteri

References Frescata and Mexia (1996) Ballal and Yamada (2016) Deligeorgidis (2002)

(Pumariño and Alomar 2012). Even while not all experiments are successful, the data show that these attempts can still be used to forecast the predator’s success (Table 17.1). The predation effectiveness of the anthocorid insect Blaptostethus pallescens on the eggs and active stages of the two-spotted spider mite Tetranychus urticae (Trombidiformes: Tetranychidae) on parthenocarpic cucumber cultivated under a polyhouse was examined by Ghongade et al. in 2022. In comparison to the use of Blaptostethus pallescens at 40 or 30 nymphs per plant (68.9% in egg and 69.4% in active stage of mites; 66.4% in egg and 68.2% in active stage of mites, respectively), the use of 50 nymphs per plant resulted in a significantly higher level of predation on the two-spotted spider mite population (76.4% in egg and 81.7% in Cucumber yields were noticeably higher in the plot where 50 nymphs of Blaptostethus pallescens (2554.8 g/plant) were introduced. Results indicate that releasing this anthocorid bug can successfully manage two-spotted spider mite (Ghongade et al. 2022). Additionally, they found that Tetranychus urticae was managed on parthenocarpic

634

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

cucumbers produced in polyhouses using Blaptostethus pallescens as part of an IPM system (Ghongade et al. 2022).

17.2.2 Reduviids In cotton (SVPR-2 var.) and castor (Tirunelveli local var.) field cages, Ravichandran (2004) made an effort to assess the biocontrol potential of Acanthaspis pedestris against Spodoptera litura, Helicoverpa armigera, and Achaea janata. In cages with Helicoverpa armigera infestations, Acanthaspis pedestris greatly reduced plant damage. Additionally, Acanthaspis pedestris reduced the harm in cages with Spodoptera litura and Achaea janata infestations. Compared to such control plots, yield was much higher in predator-released Spodoptera litura and Helicoverpa armigera-infected patches (Ravichandran 2004). Ables (1978) and Richman et al. (1980) showed similar reductions in artificial lepidopteran larval infestation by reduviid Zelus renardii and Arilus cristatus in cotton and soybean field cage plots (1980). Reduviids studied under field cages and screen houses demonstrated their efficacy against numerous pests of diverse crops (Table 17.2).

Table 17.2 Reduviids’ ability to prey on field cage crops including cotton, soybeans, and okra was investigated in a number of different nations Reduviid Acanthaspis pedestris

Zelus renardii Arilus cristatus Rhynocoris longifrons

Plants Cotton (SVPR2 var.) Castor (Tirunelveli local var.) Cotton and soybean

Pests Spodoptera litura, Helicoverpa armigera Achaea janata

References Ravichandran (2004)

Lepidopteran larvae

Ables (1978)

Cotton

Dysdercus cingulatus Phenacoccus solenopsis Helicoverpa armigera aphis gossypii Spodoptera litura

Kalidas (2012)

Rhynocoris longifrons Rhynocoris marginatus

Cotton

Rhynocoris kumarii Ecotomocoris tibialis

Okra Cotton

Spodoptera litura, Mylabris pustulata and Dysdercus cingulatus Helicoverpa armigera Anomis flava Dysdercus cingulatus

Catamiarus brevipennis Acanthaspis pedestris

Cotton

Dysdercus cingulatus

Okra

Helicoverpa armigera

Cotton

Sahayaraj et al. (2020) Ambrose and Claver (1999) Ambrose (2000) Sahayaraj and Ambrose (1997) Sahayaraj (1991) Sahayaraj (1991)

17.2

Hemiptera

635

Sahayaraj (1991) carried out two tests in India to test the bioefficacy of Ecotomocoris tibialis and Catamiarus brevipennis in a field cage (6 × 3 × 4.52″) lined with cotton. Ecotomocoris tibialis and Catamiarus brevipennis were intentionally introduced into the cage following Dysdercus cingulatus, and their predatory effectiveness was regularly monitored for 10 days. According to the findings, the predatory rates for Catamiarus brevipennis and Ecotomocoris tibialis, respectively, varied from 0.007 preys per predator per day to 0.064 preys per predator per day. On order to assess the bioefficacy of Acanthaspis pedestris against Helicoverpa armigera under the field cage (5 × 6″), Sahayaraj (1991) carried out another experiment in okra at the same time period. The prey rate varied from 0.0130 to 0.056 prey per predator per day. Under pot conditions, the reduviid Rhynocoris longifrons was tested for its ability to suppress the cotton pests Dysdercus cingulatus (second and third instar nymphs), Phenacoccus solenopsis (adult), Helicoverpa armigera (second and third instar larvae), and Aphis gossypii (all stages). In comparison to Dysdercus cingulatus and Phenacoccus solenopsis, the adult Rhynocoris longifrons demonstrated a greater predatory rate against Helicoverpa armigera when it was released either in the morning or the evening. Rhynocoris longifrons favoured hiding under pebbles during its life stages over leaves and cotton plants, which shows that the predator has adapted to its surroundings. The effectiveness of Rhynocoris longifrons’ augmentative release against a number of cotton pests in irrigated and rain-fed farmer fields was assessed. The findings showed that the predator significantly decreased the populations of Dysdercus cingulatus (53.8%), Aphis gossypii (11.8%), and Phenacoccus solenopsis (26.0%), all of which were rain-fed species. In the fields tested, Rhynocoris longifrons released plots (837.0 kg h″) had a higher cotton production and cost-benefit ratio than the control plots (715.5 kg h″), both of which were grown under irrigation (Kalidas 2012). Rhynocoriss kumarii was released in cotton and pigeon pea field cages by Ambrose and Claver (1999) and Claver and Ambrose (2001), who reported that the predator successfully controlled the pest. Spodoptera litura, Mylabris pustulata, and Dysdercus cingulatus were three cotton pests that were targeted by the release of Rhynocoris marginatus into enormous cotton field cages. In the absence of other pests, parasitoids, and predators, different pest species were put into separate cages. Throughout the evaluation period, Rhynocoris marginatus-free control experimental cages were kept for each prey setup. Spodoptera litura (57.5%), Mylabris pustulata (52.3%), and Spodoptera litura (45.8%) infestations were significantly decreased by Rhynocoris marginatus. Rhynocoris marginatus reduced the leaf, flower, and boll damages (32%, 35%, and 28% by S. litura, Mylabris pustulata, and Dysdercus cingulatus, respectively) and seed-cotton yield loss (1.4, 1.6, and 1.25 times in Spodoptera litura, Mylabris pustulata, and Dysdercus cingulatus infested cages, respectively) The effectiveness of this predator in preventing harm and controlling pests is suggested by field observations (Ambrose and Claver 1999). Rhynocoris longifrons (Reduviidae), a generalist predator of numerous cotton insect pests, is found in India. Under screen house settings, the predator’s

636

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

concealment behaviour, which is one of the major contributors to predation success, was studied. The screen house testing revealed that Rhynocoris longifrons life stages preferred to bury themselves beneath small pebbles. Under screen house circumstances, all Rhynocoris longifrons life stages demonstrated biocontrol potential against the four insect pests. Their ability to manage populations, however, remained constant throughout the day and night (Sahayaraj et al. 2020).

17.2.3 Pentatomide According to field cage tests conducted in 1987, Colorado potato beetle populations at high densities (=450 per plant) might be reduced by 50% by releasing 5–10 Perillus bioculatus and Podisus maculiventris per plant. More leaf protection was offered by Perillus bioculatus than by Podisus maculiventris. Perillus bioculatus reduced Colorado potato beetle densities of 100 per plant by 8.5% in 1988 in identical cage experiments when it was released at rates of 2 to 8 per plant. In 1988, field plot testing with Perillus bioculatus verified cage results, and they showed that 1 and 3 predators per plant reduced Colorado potato beetle numbers by 30 and 62%, respectively. In comparison to the untreated control, three Perillus bioculatus per plant considerably reduced defoliation and improved yield by 65%. Our findings suggest that releasing predatory stinkbugs by inoculation to control Colorado potato beetle populations may be a key element of an integrated management strategy (Biever and Chauvin 1992). In field cages containing larvae of the yellow-margined leaf beetle, Microtheca ochroloma (Coleoptera: Chrysomelidae), a pest of organic crucifer (Brassicaceae) crops in the USA, three release rates of the spined soldier bug, Podisus maculiventris (Hemiptera: Pentatomidae), were tested in February–March 2009 and February– March 2010. Four, ten, or sixteen first instars of Podisus maculiventris were distributed equally among six caged turnip (Brassica rapa) plants on the same day that 132 first instars of Microtheca ochroloma were put into the cages. No predators were placed in the control treatment cages. During the first four sample days in 2009, the mean number of Microtheca ochroloma larvae gradually decreased. The mean total Microtheca ochroloma counts in the two treatments with higher release rates were considerably lower on the seventh and final sample day than in the low releaserate and control treatments. 39.1% of Podisus maculiventris survivors overall received high release-rate therapy. During the first four sample days in 2010, the mean number of Microtheca ochroloma larvae gradually decreased. There were no appreciable variations in the predator release treatments on the fourth sample date or the final (ninth) sampling date. In the high-release rate therapy, the overall survival rate of Podisus maculiventris was the lowest (3.1%). This study produced two possible recommendations for growers: If the plants are anticipated to have no more than seven leaves per plant, release 10 Podisus maculiventris first instars per six plants; otherwise, release four Podisus maculiventris first instars per six plants (Montemayor and Cave 2012).

17.2

Hemiptera

637

Rhynocoris longifrons (Reduviidae), a generalist predator of numerous cotton insect pests, is found in India. Under screen house settings, the predator’s concealment behaviour, which is one of the major contributors to predation success, was studied. The screen house testing revealed that Rhynocoris longifrons life stages preferred to bury themselves beneath small pebbles. Under screen house circumstances, all Rhynocoris longifrons life stages demonstrated biocontrol potential against the four insect pests. Their ability to manage populations, however, remained constant throughout the day and night (Sahayaraj et al. 2020).

17.2.4 Miridae Predatory mirid bugs (Hemiptera: Miridae) have become increasingly common in horticulture crops in recent years. Mirid insects are zoophytophagous predators, meaning they exhibit omnivorous behaviour and consume both plants and arthropods as food. Whiteflies, lepidopteran eggs, and mites are just a few examples of the many prey that mirid bugs can feed on successfully. In various cropping systems, zoophytophagous mirid predators are utilised in conjunction with targeted insecticides. The primary generalist predators of tomato pests in Europe are Macrolophus pygmaeus, Nesidiocoris tenuis, Dicyphus bolivari, and Dicyphus errans (Hemiptera: Miridae), though only Macrolophus pygmaeus and Nesidiocoris tenuis are readily available and utilised commercially. According to Pérez-Hedo et al. (2020), the following steps should be taken to improve the performance of mirids as biocontrol agents and aid in their global expansion: 1. highlight the numerous species and biotypes that have not yet been described and examine their applicability; 2. demonstrate how it is possible to use the plant defences induced by mirids to improve pest management; 3. make the case that genetic selection of improved mirid strains is feasible; 4. examine the use of companion plants and the use of alternative foods to improve the management of mirid bugs; and, 5. discuss strategies for the expansion of miri. Nesidiocoris tenuis is employed in regions with warmer climates, but Macrolophus pygmaeus is primarily used in production areas with temperate climates. The natural species Dicyphus hesperus is commercially accessible in North America and is employed to manage numerous pests of greenhouse-grown tomatoes in temperate regions of Canada, the northern United States, and most recently Mexico (Pérez-Hedo et al. 2020; Shipp and Wang 2006; Calvo et al. 2016). However, plant bugs can potentially harm tomato plants, particularly if their preferred prey is in short supply and they are confined to greenhouses, which limits their ability to spread (Castañé et al. 2011; Urbaneja-Bernat et al. 2019; Sanchez 2008).

638

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

Both the sweetpotato whitefly, Bemisia tabaci, and the potato psyllid, Bactericera cockerelli (Hemiptera: Psyllidae), are significant pests of tomato in the United States, Mexico, and Central America. Since the late 1990s, British Columbia has studied the North American native predatory mirid Dicyphus hesperus for use in managing greenhouse tomato pests. Previous studies showed that the mirid insect Dicyphus hesperus can successfully manage both pests on tomato in cage trials and has favourable rates of development and reproduction when grown on Bemisia tabaci and Bactericera cockerelli. As a result, it might have application as a biological pest control agent. In two further tests, Dicyphus hesperus was tested for its ability to control Bemisia tabaci and B. cockerelli in tomato plants in huge cages that replicated commercial greenhouse settings during the two separate cropping seasons (fall-winter and summer). Three replicates were used in each season’s randomised complete block design, which had two treatments: (1) No Dicyphus hesperus, receiving Bemisia tabaci and Bactericera cockerelli, and (2) Dicyphus hesperus, receiving Bemisia tabaci and Bactericera cockerelli as no Dicyphus hesperus plus and Dicyphus hesperus. The predator successfully established itself in the crop, multiplied there, and significantly decreased the populations of whiteflies and psyllids during both cropping seasons. Additionally, neither leaves nor flowers showed signs of plant damage; however, a little amount of fruit damage that is likely of negligible economic importance was noted. Our findings show that the application of augmentative Dicyphus hesperus releases would enhance biologically based management tactics in tomato and likely contribute to a rise in the acceptance of such programmes in tomato in North America Calvo et al. (2016). Because tomato and other solanaceous crops produce sticky exudates that harm other natural enemies, mirids are well-adapted to these crops. Once more in North America, three Dicyphus hesperus release rates were examined on greenhouse tomatoes to control the sweetpotato whitefly. In cages containing four tomato plants and one mullein banker plant, adult Dicyphus hesperus were released at rates of one, two, or three per tomato plant each week for 3 weeks. One week after the third release, there were fewer whitefly eggs in cages getting predators than in untreated cages. Two weeks after the third release, there were fewer whitefly nymphs in cages receiving predators. Between the predator release treatments, there were no statistically significant variations in whitefly eggs or nymphs. Whitefly nymphs were reduced by 60% at the maximum release rate. There were no changes in the number of Dicyphus hesperus among release regimens 42 days following the first predator releases (Smith and Krey 2019). According to Smith and Krey (2019) findings, Dicyphus hesperus can help manage Bemisia tabaci on greenhouse tomatoes, but it might not be enough as a stand-alone tactic (Smith and Krey 2019). The majority of the controlled field cage trials were done to manage tomato pests (Table 17.3). A significant tomato pest called Tuta absoluta (Lepidoptera: Gelechiidae) is rapidly expanding over the globe. On tomato plants infested with Tuta absoluta in large cages in an experimental greenhouse, we conducted an experiment to assess the control potential and risk for plant damage of three Neotropical mirid species, Campyloneuropsis infumatus (Hemiptera: Miridae), Engytatus varians (Hemiptera: Miridae), and Macrolophus basicornis (Hemiptera: Miridae). We monitored the

17.2

Hemiptera

639

Table 17.3 The following is a list of predatory mirid bugs that have been tested against tomato pests, along with references and recommendations Predator Dicyphus hesperus Dicyphus hesperus Campyloneuropsis infumatus Engytatus varians Macrolophus basicornis Engytatus varians

Engytatus nicotianae

Pests Dicyphus Hesperus Bemisia tabaci Bactericera cockerelli Bemisia tabaci Tuta absoluta Bactericera cockerelli Bactericera cockerelli

Plant Tomato

Observation Successful agent

Tomato

Successful agent

Tomato

Successful agent

Tomato

Predtor cause necrotic rings Able to establish and control pest Potential biocontrol tool

Tomato

References Calvo et al. (2016) Smith and Krey (2019) van Lenteren et al. (2018)

Pérez-Aguilar et al. (2019)

Veronesi et al. (2022)

population growth of the three mirids exposed to Tuta absoluta as well as Tuta absoluta alone in separate cages in the greenhouse across three intervals of 9 weeks each. The amount of Tuta absoluta eggs and larvae per leaf, the number of mirid predators per leaf, the percentage of Tuta absoluta-damaged leaves and fruits, and the weight of fruits were all determined on a weekly basis. Infested tomato plants by Tuta absoluta were successfully colonised by two mirid predators, Campyloneuropsis infumatus and Macrolophus basicornis, which greatly decreased Tuta absoluta populations and subsequently boosted yield. Due to plant feeding, these two mirid species hardly ever caused harm to tomato plants or fruits. Surprisingly, the species Engytatus varians did not establish and control pest populations in any of the tests despite having high predation rates in laboratory testing (van Lenteren et al. 2018). For the first time ever, the polyphagous predator Engytatus varians (Hemiptera: Miridae), which is extensively spread, has been observed feeding continuously on the nymphs of the highly destructive pest of numerous solanaceous crops, Bactericera cockerelli, in Mexico (Hemiptera: Triozidae). For 12 weeks, the predation of this mirid on tomato plants (Solanum lycopersicum) was examined in cages measuring 7 square metres in a greenhouse. Two release rates of Engytatus varians (Ev) adults—1 and 4 adults per plant—were investigated and contrasted with the control rate of 0 Ev per plant. Regardless of the pest life stage, 1 or 4 adults of Ev/plant introduced caused an 80 to 90% reduction in both nymphs and adults of Bactericera cockerelli. In the treatments (30), there were considerably more Engytatus varians nymphs and adults present each day than there were in the control (3). However, Engytatus varians can feed on tomato plants and leave necrotic rings on the leaves because of its zoophytophagous tendencies; as a result, this kind of damage was also noted. Although there were no discernible variations between the treatments, the amount of necrotic rings/leaves was inversely linked to the presence

640

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

of prey (says Pérez-Aguilar et al. 2019). Additionally, their findings suggest that Engytatus varians can establish and manage Bactericera cockerelli populations in greenhouse settings without seriously harming tomato plants (Pérez-Aguilar et al. 2019). The B biotype of Bemisia tabaci, sometimes referred to as the sweetpotato whitefly, is a pest of numerous agronomic, ornamental, and horticultural crops, including tomato. The Tomato Yellow Leaf Curl Virus (TYLCV), which is mostly linked to nymphal feeding, is transmitted by the Sweetpotato Whitefly, which is also the cause of irregular tomato ripening. In greenhouse production, the predatory mirids Macrolophus pygmaeus and Nesidiocoris tenuis (Heteroptera: Miridae) are utilised successfully to control whiteflies. Bactericera cockerelli, sometimes known as the tomato-potato psyllid (TPP), is a significant pest of solanaceous crops. We investigated the potential of the mirid bug Engytatus nicotianae (adults) as a biocontrol agent of Bactericera cockerelli on greenhouse tomato plants as an alternative to the use of pesticides. The experiment consisted of two parts: (a) a choice feeding assay in the lab to examine the preferences of Engytatus nicotianae when presented with Bactericera cockerelli eggs and nymphs; and (b) a greenhouse experiment to evaluate E. nicotianae’s potential as a biocontrol agent under conditions that are similar to those found in the marketplace. The laboratory choice experiment revealed that Engytatus nicotianae preferred the first two instars of Bactericera cockerelli, then the third instar, and last the eggs. Only after all preceding life stages had been digested were the fourth instars consumed. Even though Bactericera cockerelli feeds on tomato leaves and stems, in the cage experiment, the Engytatus nicotianae -only treatment produced the same amount of fruits and flowers as the control (no insects). The Bactericera cockerelli-only treatment, in comparison, produced less fruits and flowers. In comparison to the Bactericera cockerelli-only treatment, the simultaneous introduction of Engytatus nicotianae and Bactericera cockerelli significantly reduced Bactericera cockerelli build-up and prevented the pest from establishing at all in four of the seven duplicates of the experiment. These results imply that Engytatus nicotianae should be investigated as a possible biocontrol agent for Bactericera cockerelli, and additional study is required to establish the ideal release configuration (Veronesi et al. 2022).

17.2.5 Entatomidae The predatory bug Andrallus spinidens is well-known in Iran for its use in the biocontrol of rice defoliator caterpillars. In field and semi-field settings, investigations were conducted to ascertain the predator’s capacity for predation in 2001–2002. On a rice plant, 20 Naranga aenescens larvae in their fifth instar were released in separate cages with fifth instar nymphs and adults. Adults and nymphs of the second to fifth instars were introduced to individual pots of rice plants in the glasshouse, where 20 preys were released. The findings showed that, under field settings, every adult and larva of the fifth instar killed, respectively, 1.44 and 1.87 larvae every day. Predation rates for the second through fifth instars and adults in the

17.3

Coleoptera

641

glasshouse were 0.43, 0.77, 1.28, 1.40, and 1.92, and 0.41, 0.76, 1.20, 1.23, and 2.20 prey/predator/day, respectively, for the first and second years. Typically, under glasshouse conditions, each predator in its second instar needs 11 to 13 prey animals to mature within 12 to 15 days. The average number of prey per male and female predator per day for greenhouse and field tests was 4.06 and 3.55, respectively. For the 5’h instars of the greenhouse and field, the sex ratios (M/F) were 1 and 1.08, respectively. The need for predator-foraging investigations under more realistic conditions is indicated by the differences between the findings of this study and those obtained from laboratory experiments (Mohaghegh and Najafi 2003).

17.3

Coleoptera

Knowledge of the causes that lead to herbivorous insect mortality is crucial to creating more effective management techniques for their abundance in crops and a better understanding of their population dynamics. Predaceous coccinellids are more frequently utilised than any other taxon as a metaphor for biological management. Can local natural enemies thwart the actions of herbivores introduced for biological control of weeds, and if so, can their level of activity be predicted from tests that employ resident herbivores as hosts? This was a crucial subject that needed to be addressed. Exclusion tests were carried out at three locations in central New York State with a focus on the leaf beetle Galerucella nymphaeain stands of the introduced weed purple loosestrife in order to investigate these topics (Lythrum salicaria). Galerucella calmariensis and Galerucella pusilla, two European species that are imported and dispersed in North America for the biological control of purple loosestrife, are congeneric with this beetle. Galerucella nymphaeae eggs were preyed upon by general predators, such as the common lady beetle Coleomegilla maculata, from late spring until the end of summer. From one-third of Galerucella nymphaeae’s egg masses were attacked at this time, and the percentage of eggs inside each egg mass that were destroyed or injured rose from about 50 to 90%. Galerucella nymphaeae larvae and pupae at all sites survived less in open than in closed cages in the middle and end of the summer, but not earlier. The absence of parasitized or diseased beetles and the presence of arthropod predators show that predators were mostly to blame for the decreased survival in open cages (Nechols et al. 1996). Nechols et al. (1996)’s findings led them to expect that local species of general predators would occasionally prevent the European Galerucella calmariensis and Galerucella pusilla from colonising new areas or functioning as well. Therefore, it is preferable to continue using protective cages both during the introduction and subsequent dispersal of these natural enemies. Moreover, future assessments of biological control initiatives including Galerucella spp. should take general predator activity into account (Nechols et al. 1996). Schmaedick and Shelton (1999) evaluated the impacts of arthropod predators on Pieris rapae eggs and larvae on cabbages (Brassica oleracea variety capitata) in New York State using two methods of predator exclusion. Pieris rapae survival on

642

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

cabbage plants confined to keep predators out was compared to survival on plants caged with the bottom uncovered to let in arthropod predators but not larger predators like birds. Two cohorts were monitored in each of two unsprayed cabbage plots for a total of eight cohorts over the course of 2 years. The estimated arthropod mortality of eggs and larvae ranged from 23 to 80%, averaged 53% for all 8 cohorts, and primarily affected the eggs and first instars. Exclusion tests were also performed to compare the mortality of individual Pieris rapae eggs protected by Tanglefoot rings against that of eggs that were left unprotected by the rings. For the whole egg stage, mortality attributable to arthropod predators in six cohorts divided into two fields ranged from 0 to 44%. Our research shows that arthropod predators in cabbage fields frequently cause fairly significant mortality rates for Pieris rapae eggs and first instars. Understanding these predators’ crucial function is the first step in finding ways to optimise their activity in commercial crops (Schmaedick and Shelton 1999). Hoogendoorn and Heimpel (2004) performed field-cage investigations on the direct interactions between Coleomegilla maculata, a coccinellid species endemic to North America, and Harmonia axyridis, a species imported from Asia (Coleoptera: Coccinellidae). There was no significant effect of Harmonia axyridis on the survival or weight gain of Coleomegilla maculata, but Harmonia axyridis larvae weighed more when kept with Coleomegilla maculata for 5 days than when kept with equal numbers of conspecifics. The researchers compared the mortality and weight gain of larvae of both species in field cages that enclosed one or both species with corn plants containing high or low aphid numbers. This shows that Harmonia axyridis faced more intraspecific competition than interspecific competition with Coleomegilla maculata. According to Hoogendoorn and Heimpel, Coleomegilla maculata’s geographical distribution over the plants varied between single-species and two-species treatments, suggesting that this species avoided interactions with Harmonia axyridis (2004). The hemlock woolly adelgid (HWA), Adelges tsugae, is being controlled biologically by Laricobius nigrinus (Coleoptera: Derodontidae) (Homoptera: Adelgidae). In Virginia, US researchers looked into the effects of predator exclusion studies on survival, reproduction, and HWA populations during a 2-year period. One of three treatments—caged hemlock branches with predators, caged hemlock branches without predators, or uncaged hemlock branches—was chosen for each branch during the first year. From February to April, Laricobius nigrinus adults persisted and gave birth to up to 41 offspring per female. When compared to densities on branches not exposed to predators, Laricobius nigrinusexposed branches showed a much higher rate of reduction in Adelges tsugae densities. Additionally, compared to caged and uncaged branches without predators, the final density of sistens and progrediens was considerably lower on caged branches harbouring Laricobius nigrinus. Over two 10-week sampling periods in year 2, Laricobius nigrinus survival and predation were assessed: (November– January and February–April). Laricobius nigrinus endured the entire 6-month experiment, with a survival rate of 89% in January and 55% in April. Per beetle, 38 young were born between February and April. On caged branches with Laricobius nigrinus than on those without predators, the decline in adelgids was

17.3

Coleoptera

643

much greater measured in terms of both numbers of adelgids and % reduction per branch (Lamb et al. 2005). In Michigan, a complex guild of insect predators, including the native lacewing Chrysoperla carnea, the native gall midge Aphidoletes aphidimyza, and the foreign coccinellid Harmonia axyridis, feed on the soybean aphid Aphis glycines. Some members of this guild may also prey on other guild members in addition to eating Aphis glycines. These interactions could have a favourable, negative, or no effect on the biological regulation of Aphis glycines. We compared aphid populations in microcosms with either Aphidoletes aphidimyza larvae or Chrysoperla carnea larvae alone, with both a Harmonia axyridis adult and either Aphidoletes aphidimyza or Chrysoperla carnea larvae, and without predators to examine the effects of intraguild predation on soybean aphid population dynamics. The lady beetle behaved as an intraguild predator when Chrysoperla carnea or Aphidoletes aphidimyza larvae were also present. Contrary to microcosms with only the intraguild and aphid prey, intraguild feeding did not cause a release of aphid populations. In field cages, a similar outcome was discovered. When compared to cages that excluded predators, cages that allowed large predators had lower populations of Chrysoperla carnea and Aphidoletes aphidimyza larvae but also much fewer aphids. Thus, Harmonia axyridis’ positive effect on Aphis glycines outweighed its detrimental effect as an intraguild predator in both laboratory and field investigations (Gardiner and Landis 2007). These results also show that the uncommon Harmonia axyridis, while acting as an intraguild predator and maybe contributing to local losses in Chrysoperla carnea and Aphidoletes aphidimyza, is currently crucial for the biological control of Aphis glycines as a whole (Gardiner and Landis 2007). In the western United States, the hemlock woolly adelgid, Adelges tsugae, is preyed upon by the lady beetle Scymnus (Pullus) coniferarum (Coleoptera: Coccinellidae). Scymnus coniferarum may play a significant role as a predator in preventing Adelges tsugae populations from increasing to harmful levels in this area. Darr et al. (2016) conducted this study to evaluate the potential of this predator as a biological control agent for Adelges tsugae in the eastern United States and to confirm Scymnus coniferarum as a biocontrol agent. At two locations in southwestern Virginia, Scymnus coniferarum predation, reproductive potential, and survival were assessed in field cages on Tsuga canadensis that was adelgid-infested. Between December 2012 and June 2014, sampling was done to determine how Scymnus coniferarum affected both generations of Adelges tsugae (sistens and progrediens). During both outdoor trials, adult Scymnus coniferarum consumed Adelges tsugae in all life stages and at rates comparable to other adelgid-specific predators. The lack of Scymnus coniferarum oviposition evidence may be linked to the cold temperatures and the scarcity of prey. The 2013 winter sample period at the higher elevation site saw the highest Scymnus coniferarum mortality, whereas the 2014 spring sample period saw the lowest mortality. At locations in southwest Virginia, Scymnus coniferarum showed a high predation rate on v and survived for prolonged periods of time, showing that this species could be a successful predator of hemlock woolly adelgid in similar temperatures (Darr et al. 2016). Different coleopteran (particularly

644

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

Table 17.4 Below is a collection of coleopteran predators that have been tested against various pests, along with references and recommendations Predator Coleomegilla maculata Coleomegilla maculate Harmonia axyridis Laricobius nigrinus

Pests Galerucella nymphaeaeeggs Aphid

Plant Purple loosestrife

Observation Predator colonized Need further evaluation Interactions between the predators

References Nechols et al. (1996) Hoogendoorn and Heimpel (2004)

Adelges tsugae

Hemlock

Lamb et al. (2005)

Harmonia axyridis Aphidoletes aphidimyza Chrysoperla carnea Scymnus coniferarum Arthropod predators

Aphis glycines

Soybean

Seasons-based progeny establishment and control the pest Intraguild predator

Adelges tsugae

Tsuga canadensis Cabbages

Established in the cage, effective predator Biocontrol agent

Harmonia axyridis

Danaus plexippu Aphids Aphis gossypii

Darr et al. (2016) Schmaedick and Shelton (1999) Koch et al. (2003)

Hippodamia convergens

Pieris rapae

Potential predator

Potential predator

Gardiner and Landis (2007)

Shrestha and Parajulee (2013)

Coccinellids) were tested with various crops as preys, and the results demonstrated that predators could quickly establish themselves in the test environment and effectively manage the target pests thanks to their intergrained behaviour (Table 17.4). Using no-choice predation tests in both lab and caged field settings, the potential for a traditional biological control agent, Harmonia axyridis, to have non-target impacts on populations of the monarch butterfly, Danaus plexippu, was investigated. In three different functional response trials, immature Danaus plexippus were preyed upon by adult and larval Harmonia axyridis. At roughly 25 and 15, respectively, nonlinear functional responses for third instar Harmonia axyridis preying on Danaus plexippu eggs and first instars attained plateaus. Adult Harmonia axyridis preying on Danaus plexippueggs showed a linear response. Large field cages with potted Asclepias syriaca L. were infested by Danaus plexippu, a first instar, and Harmonia axyridis, a third instar. First instar Danaus plexippu survival in cages containing Harmonia axyridis larvae was considerably lower than in cages lacking Harmonia axyridis larvae. According to the current research, Harmonia axyridis may be a stressor for populations of Danaus plexippu. To determine the chance of Danaus

17.4

Neuroptera

645

plexippu exposure to Harmonia axyridis in the wild and to calculate the possibility that Harmonia axyridis may negatively affect Danaus plexippu in the presence of other prey, particularly aphids, more research is required. These facts might be utilised to create an ecological risk evaluation (Koch et al. 2003). One Hippodamia convergens adult per plant, released at a prey density of one aphid per leaf, held the aphid Aphis gossypii population below economic threshold for the duration of the growing season, according to a field cage research (Shrestha and Parajulee 2013). To determine the extent to which Pterostichus melanarius, a frequent carabid species in Québec soybeans, affects Aphis glycines populations through laboratory and field cage studies. Between 16.8% (during low aphid density) and 33.7% (during high aphid density) of Pterostichus melanarius tested positive for aphid throughout the growth season. Furthermore, despite the fact that Pterostichus melanarius feeds on Aphis glycines, laboratory feeding tests did not show a significant decline in Aphis glycines populations caused by carabid beetles. When pest concentrations are high, these findings point to a very modest interaction between Pterostichus melanarius and Aphis glycines, but the high predation rate when aphid densities are particularly low shows that these natural enemies may serve as significant early-season predators (Firlej et al. 2013).

17.4

Neuroptera

The green lacewing, Chrysopa carnea, reduced larval populations from 73.8 to 99.5% (266 larvae/acre) when inundative releases of its larvae and eggs were tested in field cages to assess their impacts on the bollworm, Heliothis zea, and the tobacco budworm, Heliothis virescens. The variations in alternate prey availability, the quantity of released lacewings, and the population of Hetiothis spp. were likely the causes of the variations in percentage reduction. When the predator’s eggs and larvae were released in areas with lots of prey, 23.5 and 25.0% of them, respectively, were later found to be pupae or adults. The pupae and adults were only partially recovered, perhaps due to escapes and ineffective collecting. Chrysopa carnea colonies of various sizes were created by dispersing various amounts (Ridgway and Jones 1968). Chrysoperla rufilabris release rates on the management of Bemisia argentifolii inside field cages in organically produced watermelon and Lablab purpureus, a leguminous forage crop, and silverleaf whitefly in organically cultivated watermelon. Chrysoperla rufilabris larvae of the second instar were released in cages at rates of 0 (control), 10, 25, and 50. (0.37 m2 area). Prior to release, whitefly counts were performed, and those results were compared to those obtained 48 h later. Comparing the control in the watermelon field to the predator treatment that had the greatest whitefly counts throughout the season, the control had about 35% more whiteflies (25 lacewings per plant). The second half of the season was when the consequences of predator releases were most noticeable. Increased Chrysoperla rufilabris release rates did not, however, lead to greater pest suppression. Because prey densities were already too high before the experiment began, it’s possible that

646

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

there were no discernible patterns of prey suppression in the legume field (Legaspi et al. 1996). Additionally, they stated that rather than predation, crop phenology and environmental factors appeared to be what influenced Bemisia argentifolii population dynamics in the lab. To best control the whitefly populations, causes for observed variations in the levels of pest suppression provided by lacewing larvae are discussed (Legaspi et al. 1996). Rhopalosiphum padi population growth was inhibited by Chrysopa carnea larvae in trials conducted in lab and outdoor cages. The number of aphids per shoot on each day of the experiment dropped by 10% compared to aphid populations when no larvae were present when the initial predator-prey ratio was 2 1-day-old larvae per 100 aphids (= 50 aphids for one larva). The index dropped by 50% when the predator-prey ratio was 20 larvae per 100 aphids (= 5 aphids per one larva). In field cages, 0.5 chrysopid eggs per one aphid, 1 egg, 21%, 3 eggs, 51%, and 10 eggs reduced the aphid index by 11%, 21%, 51%, and 91%, respectively. For the same reduction in aphid populations, roughly 15–20 times more eggs than larvae were required (Rautapää 1977). On caged cotton, the functional response of Chrysopa carnea larvae in their third instar feeding on four densities of Heliothis virescens eggs was ascertained (Gossypium hirsutum). The mean search rate for Chrysopa carnea larvae was 1.08 × 105 ha per predator day, or 0.11 × row per predator day (Stark and Whitford 1987). For the two most prevalent predators of Hyalopterus pruni (Hemiptera: Aphididae) in prune orchards in California’s Central Valley, Harmonia axyridis (Coleoptera: Coccinellidae) and Chrysopa nigricornis (Neuroptera: Chrysopidae), field observation, field cages, and laboratory arenas were compared as methods to estimate daily per capita The field observation method produced the highest estimates and the laboratory arena method the lowest for both predator species. Daily per capita consumption rose with larval size. The necessity to assess both biomasses killed and biomass consumed is one of the most significant possible explanations for variations between estimates for each approach, as Harmonia axyridis consumed nearly all of each prey item while Chrysopa nigricornis always killed more biomass than they consumed. This study implies that when prey are colonial and predators are relatively sedentary, field cages may be more effective for measuring daily intake than the laboratory arena method, which can result in an underestimate. The field observation method, along with observations of the duration and pattern of feeding activity throughout the day, is the best choice for calculating daily per capita intake for highly mobile predators or predators of dispersed prey (Latham and Mills 2009). In laboratory trials conducted at 25 °C, the predatory ability and preferred prey of larvae of Chrysoperla carnea (Neuroptera: Chrysopidae) on eggs or larvae of Pieris brassicae (Lepidoptera: Pieridae) were assessed in the absence and presence of cabbage aphids as an alternative prey. Both instars fed on cabbage aphids and butterfly eggs and larvae, with the third instar being the most ravenous. The caterpillars were a clear favourite of the lacewings over the butterfly eggs. Third instar lacewings either maintained their predation on Pieris brassicae eggs or larvae or reduced it by about 80% in the presence of aphids, whereas second instar

17.4

Neuroptera

647

lacewings either entirely stopped or reduced it by around 70%, respectively. Thus, aphids were clearly preferred by both instars over Pieris brassicae eggs by both. However, third instar lacewings preferred caterpillars over aphids, whereas second instar lacewings favoured the opposite. The findings suggest that Pieris brassicae may be resistant to third instar Chrysoperla carnea as a biocontrol agent (Huang and Enkegaard 2010). An significant predator of the red spider mite, Oligonychus coffeae, which infests Indian tea is the green lacewing, Mallada desjardinsi. In a lab setting, the best predator-prey ratios were 1:50 and 1:33; however, in a greenhouse setting, 1:33 and 1:25 ratios worked well (says Vasanthakumar and Babu 2013). They said that the findings might be seen as a first step toward using this predator in an IPM programme to control red spider mite infestations in tea (Vasanthakumar and Babu 2013). Pest infestations in net houses can occasionally be similar to those in open fields, necessitating the use of insecticides. Chrysoperla zastrowi sillemi (Neuroptera: Chrysopidae), a green lace wing, was tested for its ability to prey on tomatosucking pests in a screen-house environment to test the idea that bioagents may be more effective in constrained plant growing conditions. A total of three releases were made at intervals of 7 days, with the first release occurring during the emergence of the aphid, Myzus persicae, and the whitefly, Bemisia tabaci. The release rates were set at 4, 5, and 6 s instar grubs plant. The pooled data for the years 2018 and 2019 showed that there was no statistically significant difference in the effectiveness of the release rates of 4, 5, and 6 grubs plant in lowering the population of these pests. The release rate and timing of these releases were the basis for a factorial analysis that revealed a single release of C. zastrowi sillemi at 4 grub plants was effective against sucking pests in tomato grown in a screen-house (Nair et al. 2020). The adults of Chrysoperla carnea were captured in Pakistan using an aerial net, kept in plastic jars (151,525 cm), and then transferred to a plastic cage (233,838 cm) with ventilation holes on both sides. This is done in a lab setting to determine the predator’s capacity for biological control. A significant predator of many aphid species, including Lipaphis erysimi and Diuraphis noxia, is the green lacewing, Chrysoperla carnea. Chrysoperla carnea is the most potent predator for aphid species management in biological control. Chrysoperla carnea larval stages’ functional responses are calculated over a period of 24 h at various prey densities (10, 20, 30, 40, and 50). Using a logistic regression model, the functional response of three larval instars was estimated. Roger’s random predator equation was used to quantify handling time (Th) and attack rate (a). The findings showed that the first, second, and third larval instars of Chrysoperla carnea on both Lipaphis erysimi and Diuraphis noxia display a type II functional response and consumption rate increased with the rise in prey density. It is determined that the attack rate reduces as the handling time grows. The handling time and attack rate of the Chrysoperla carnea third instar feeding on Lipaphis erysimi were 0.029 and 1.55 h, respectively, while those feeding on Diuraphis noxia were 0.14 and 0.72 h, respectively. This research showed that the Chrysoperla carnea third larval instar had an excellent predatory capability for Lipaphis erysimi and Diuraphis noxia (Hameed et al. 2022). According to the study

648

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

Table 17.5 Lacewings that have been tested to repel certain pests, along with references and recommendations Predator Chrysopa carnea Chrysoperla rufilabris Chrysopa carnea Chrysopa carnea Harmonia axyridis Chrysopa nigricornis Chrysoperla carnea Chrysoperla zastrowi sillemi Chrysoperla carnea

Pests Heliothis zea Heliothis virescens Bemisia argentifolii Rhopalosiphum padi Heliothis virescens

Plant

Watermelon Cotton

Observation Predators populatin— Based controle efficiency Pest reduction is not related to predator Potential predator

Cotton

Recorded good searching ebility

Hyalopterus pruni

Orchards

Best methodology to estimate predatory potential

Pieris brassicae

Cabbage

Potential biocontrol agent

Myzus persicae Bemisia tabaci

Tomato

Effective predator

Lipaphis erysimi Diuraphis noxia

Efficient predator

References Ridgway and Jones (1968) Legaspi et al. (1996) Rautapää (1977) Stark and Whitford (1987) Latham and Mills (2009)

Huang and Enkegaard (2010) Nair et al. (2020) Hameed et al. (2022)

cited above, the majority of Chrysopa species are capable of lowering the tested pest species in field cage tests (Table 17.5).

17.5

Multiple Predators

In the majority of agricultural cropping systems, generalist predators are prevalent. The overall impact of the predator complex can be significantly influenced by interactions between predators. However, because the degree of predation is typically unknown and challenging to quantify, pest control from these predators is frequently disregarded as a component of integrated pest management (IPM). In field cage tests with cotton, Lingren et al. (1968), VanDen Bosch et al. (1969), and López Jr et al. (1976) also noted a similar type of reduction in the population of bollworms by nabid, Nabis arnericoferus, pentatomid (Podisus inaculiventris), and lygacid (Geocoris punctipes) predators. Three field trials evaluated the effect of arthropod predators on the mortality of the first instars of the range caterpillar, Hemileuca oliviae. First instars of H. oliviae were placed into experimental units in the first two studies; half of the units had cages and insecticides to keep off predators. The findings indicated that mortality was not affected by density and that the absence of possible predators had a substantial impact on larval survival from the initial level of larvae. Generalist predators, like the ant Crematogaster punctulata Emery, may

17.5

Multiple Predators

649

Table 17.6 List of predators that have been tested against various pests, along with references and recommendations Predator Nabis arnericoferus, Podisus inaculiventris Geocoris punctipes

Pests Bollworm

Plant Cotton

Observation Good predators

Orius strigicollis Amblyseius cucumeris

Thrips palmi

Green pepper, Sweet pepper cucumber

Potential predators

Geocoris pallens, Nabis americoferus, Chrysopa carnea

Lygus bugs

53–76% control

References Lingren et al. (1968) VanDen Bosch et al. (1969) López Jr et al. (1976) Kim et al. (2006)

Leigh and Gonzalez (1976)

play a significant role in ecosystems, according to evidence from sweep-net samples, pitfall traps, and general observations. Consideration is given to the potential effects of spiders, cicindelids, mantids, and gryllacridids. The third experiment demonstrated that C. punctulata is a predator of H. oliviae (Shaw et al. 1987). In a different study, greenhouse green pepper, sweet pepper, and cucumber were used to assess the efficacy of Orius strigicollis and Amblyseius cucumeris as natural enemies against thrips. Formula (Dcontrol Dtreatment)/Dcontrol100 (Dcontrol Dtreatment)/Dcontrol100 was used to determine control efficacy. Dcontrol is the average density of thrips on plots where no natural enemies were released, and D-Treatment is the level of treatment. The average thrips density on the plots where natural enemies were unleashed is known as the “D-treatment.” As a result, Frankliniella occidentalis on green pepper and sweet pepper is controlled by Orius strigicollis. Additionally, it has a 61.2–74.4% control effectiveness against Thrips palmi on cucumber. Amblyseius cucumeris had control efficiencies of 12.938.312.938.3 and 17.187.017.187.0 against Frankliniella occidentalis on green pepper and sweet pepper, respectively. It had a control effectiveness of 90.497.490.497.4 against T. palmi on cucumber. Field testing revealed that the early application of natural enemies was successful in lowering thrips density. In order to effectively manage Frankliniella occidentalis on green pepper and sweet pepper in the spring, three to six consecutive releases of five to six Orius strigicollis individuals per crop are required. Amblyseius cucumeris should be discharged more than 100 times each crop in order to successfully manage Thrips palmi on cucumber in the autumn (Kim et al. 2006). Table 17.6 shows the list of predators that have been tested against various pests, along with references and recommendations. Lygus hesperus effectiveness as a predator was assessed using Geocoris pallens, Nabis americoferus, and Chrysopa carnea. Lygus insect populations of egg-nymphs or nymphs were greatly reduced by Geocoris pallens. When kept in sleeve cages, Nabis americoferus was successful as a predator in the egg–nymph stages, but was

650

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

useless in huge (183 cm × 183 cm × 366 cm) field cages. Chrysopa carnea did not work. 53–76% of lygus bugs were controlled by the natural predator complex when they were in the egg-nymph stage (Leigh and Gonzalez 1976). To ascertain daily ingestion rates of the soybean looper’s eggs and larvae, Pseudoplusia inc/udens, about 200 predator specimens from 16 species were examined in field cage testing. The nabids Reduviolus roseipennis, Tropiconabis capsiformis, and Hoplistoscelis deceptivus, the green lacewing Chrysopa rufilabris, and the Iygaeids Geocoris punctipes and Geocoris uliginosus showed the highest rates of predation. The nabids Tropiconabis capsiformis and Hoplistoscelis deceptivus, the Iygaeids Geocoris punctipes and Geocoris uliginosus, and the carabid beetle Calleida decora all consumed the tiniest larvae. The most medium-sized larvae were devoured by the reduviid Arilus cristatus and the pentatomid Stiretrus anchoago (Richman et al. 1980). The percentage of Pectinophora gossypiella egg masses attacked, egg consumption within egg masses attacked, and overall egg consumption by the predators was examined in field cages in cotton during the summer of 1970 in the Coachella Valley, California, for various densities of the same predator species, for various predators at the same density, and for various locations of egg masses on the plants. Pink bollworm egg predators with some promise included Chrysopa carnea, Geocorus pallens, and Nabis americoferus; those with less promise included Notoxus calcaratus Horn and Spanogonicus albofasciatus; and those in the middle were Geocorus punctipes and Orius tristicolor (Irwin et al. 1974).

17.5.1 Predation by a Complex of Predators Made Up of Adults of the Lygaeid In Mississippi, field cages were used to test the effects of Geocoris punctipes, Nabis roseipennis and/or Tropiconabis capsiformis, Hippodamia convergens, and Coleomegilla maculata on Heliothis virescens eggs, first-instar larvae, or both, on late pinhead square and early bloom stage cotton. Predator:prey (P:p) ratios of 1, 2, 3, 4, and 5 were utilised, with prey densities of 4 (11, 512/ha) or 8 per cage. To assess predator efficacy and phosphorus-32 or carbon-14 labelling of prey, respectively. For a 48-h period, the average percent egg predation varied from 2.1 to 12.1 depending on the P:P ratio. As a function of P:P ratio, the average percent predation on larvae varied from 5.3 to 22.0. The hemiptera consumed more larvae than the coleoptera, and Geocoris punctipes was the top predator of eggs. The average area of discovery for the predator complex when exposed to eggs and larvae was 6.98103 m/ day and 2.34102 m/day, respectively, given the range of predator densities utilised (Thead et al. 1987). In a laboratory setting, the relative effectiveness of four predators against three distinct stages of the tobacco bud worm, Reliothis virescens, or the cotton bollworm, Reliothis zea, was investigated. The species that were examined on cotton in field cages included Chrysopa carnea, Geocoris punctipes, Coleomegilla maculata, and Podisus maculiventris. Coleomegilla maculata was generally the most effective

17.5

Multiple Predators

651

predator of Heliothis eggs and first stage larvae when searching was restricted. On first instars, Chrysopa carnea was the most effective predator when the first stage predators were compared. The Podisus maculiventris third stage nymphs and adults were the most effective predators of the third stage larvae. The immature stages of predators on Heliothis eggs and first instars that were most effective under the enlarged searching criteria were Chrysopa carnea larvae. When they were adults, Coleomegilla maculata and Geocoris punctipes were the most effective predators of eggs and first-stage larvae, respectively. Compared to the other examined predators, Podisus maculiventris third instars and adults ingested more Heliothis third instars. There were significantly fewer tobacco budworm larvae when Chrysopa carnea and Podisus maculiventris were released at rates of 100,00/acre in field cages. While the reductions attributed to Podisus maculiventris primarily occurred when the tobacco budworm larvae were larger, those caused by Chrysopa carnea were more pronounced in the egg and small larva numbers (López Jr et al. 1976). We looked examined how generalist insect predators’ intraguild interactions affected the suppression of the aphid Aphis gossypii, an herbivore. In field enclosure/exclosure studies, we manipulated members of the predator community, including three hemipteran predators and larvae of the predatory green lacewing Chrysoperla carnea, in order to answer the following four questions: 1. Does Chrysoperla carnea get eaten by generalist hemipteran predators? 2. Is intraguild predation (IGP) a significant cause of death for Chrysoperla carnea? 3. Do predator species interact significantly or do they behave independently and additively? 4. Could a trophic cascade effect cause the experimental addition of some predators to boost aphid densities? The consumption of Chrysoperla carnea and other carnivorous arthropods by a number of generalist predators has been directly seen in the wild. When other predators were present, the survival of lacewing larvae was drastically reduced, indicating that IGP was a primary cause of mortality. Predation, not competition, was the primary cause of the lacewing larvae’s decreased survival. IGP significantly altered the interactions between lacewings’ effects on aphid population control and either Zelus renardii or Nabis predators. Although the trophic web was too complex to distinguish between different trophic levels within the predatory arthropod community, some trophic links were strong enough to produce cascades from higherorder carnivores to the level of herbivore population dynamics: in one experiment, the addition of either Z. renardii or Nabis predators generated enough lacewing larval mortality to free aphid populations from lacewing predator control. We draw the conclusion that intraguild predation is common in this system and may have significant effects on a critical herbivore’s population dynamics (Rosenheim et al. 1993). In a series of lab and field tests, Losey and Denno (1998) examined the interaction between foliar-foraging (Coccinella septempunctata) and groundforaging (Harpalus pennsylvanicus) predators of the pea aphid (Acyrthosiphon pisum). The interaction between the predators was demonstrated in both open and

652

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

closed field cages, in addition to laboratory experiments, where it was shown that the combined effect of the two predators on the growth of the aphid population was significantly greater than the sum of their individual effects. These findings suggest that models estimating the influence of various predator complexes must take into account predators’ positive interactions and that the significance of ground-foraging predators in agroecosystems may need to be reevaluated (Losey and Denno 1998). Through laboratory and field cage tests in Northern Switzerland, the ability of three aphidophagous predators, Adalia bipunctata, Aphidoletes aphidimyza, and Episyrphus balteatus, to suppress the pink apple aphid, Dysaphis plantaginea, a significant pest on apple in Europe, was evaluated. All three predators were effective in feeding on Dysaphis plantaginea on apple seedlings in a lab setting. The size of the branches of apple trees, which ranged in leaf surface area from 150 cm2 to 960 cm2, had no effect on the larvae of Aphidoletes aphidimyza success in looking for individual aphids. Six hours and 48 h after the discharge of a single second-instar Aphidoletes aphidimyza larva, respectively, 50 and 70% of individual aphids were discovered and destroyed. Aphidoletes aphidimyza and, to a lesser extent, Episyrphus balteatus were found to be reliable and effective predators of Dysaphis plantaginea in spring conditions in a first field cage experiment conducted in 1996. They were unaffected by cool temperatures and rainy weather. Aphidoletes aphidimyza and Episyrphus balteatus larvae were released alone and together on apple seedlings that were infested with aphids in a second field cage experiment in 1997 to examine interactions between these two promising control agents. Aphid population growth was significantly impacted negatively by both species. Since the two species did not significantly interact, an additive model best accounts for their combined effect. Aphid densities were lowered to 5% of the control by combined releases of the two predator species. This suggests the possibility of increasing the number of times these native aphid predators is released to manage Dysaphis plantaginea (Wyss et al. 1999). Researchers used a mirid C. lividipennis wolf spider for their lab experiments. Brown plant hopper (BPH, Nilaparvata lugens), Lycosa pseudoannulata, and rove beetle, Paederus fuscipes, in the mortality of nymphs. When the predators were placed in separate cages with BPH nymphs, Paederus fuscipes had the best nymphal survival (53.60%), whereas L. pseudoannulata had the lowest (32.60%), followed by C. lividipennis (43.0%), demonstrating the importance of each predator species. The nymphal survival rate was lowest (21.8%) in the treatment with L. pseudoannualata and Paederus fuscipes when two predator species were introduced together. This was equivalent to the treatments using C. lividipennis in combination with Lycosa pseudoannulata (25.4%) and Paederus fuscipes (29.6%), though. Except for the treatment including C. lividipennis, a combination of all three predators reduced nymphal survival to its lowest level (20.6) but did not statistically vary from other treatments (Geetha and Gopalan 1999). Using field cages, it was determined if larval Spodoptera frugiperda (Lepidoptera: Noctuidae) cannibalised maize under field settings. When maize plants were infected with two or four fourth-instar larval generations over the course of 3 days, cannibalism was found to be responsible for about 40% of the mortality. Trials in the

17.5

Multiple Predators

653

field looked at how Spodoptera frugiperda natural enemies fared in relation to larval density. On maize plants with higher degrees of larval feeding damage, predator abundance (earwigs, staphylinids, other predatory beetles, and Chrysoperla spp.) was significantly higher; however, the link between predator abundance and the number of Spodoptera frugiperda larvae per plant was less obvious. Predation risk will be higher for larvae living in big groups as larval damage is likely a more accurate predictor of past larval density than numbers gathered during an evaluation. In sorghum, parasitism caused 7.1% of the larval mortality and affected six species of Tachinidae and Hymenoptera. The likelihood of a parasitoid larval attack was unaffected by larval density or the placement of larvae within a plant (Chapman et al. 2000). When caged with varied densities of fourth or fifth instar azalea lace bug, Stephanitis pyrioides prey, azalea plant bug (Rhinocapsus vanduzeei) fifth instars and a commercially produced green lacewing (Chrysoperla rufilabris) first and second instars displayed a type II functional response. Attack coefficients for combined fourth and fifth instar prey for Rhinocapsus vanduzeei and Chrysoperla rufilabris were statistically comparable (0.052 and 0.057, respectively). Rhinocapsus vanduzeei required much more time to handle than Chrysoperla rufilabris (3.96 h) (2.41 h). As the initial azalea lace bug density increased, both predators’ search efficiency generally decreased. In comparison to Rhinocapsus vanduzeei, Chrysoperla rufilabris killed considerably more fourth and fifth instar prey (8.0 and 6.0, respectively) in a 24-h period. According to the findings, Chrysoperla rufilabris is a better option for augmentative release for controlling azalea lace bugs than Rhinocapsus vanduzeei. However, Rhinocapsus vanduzeei, a member of the guild of lace bug natural enemies, can reduce azalea lace insect numbers in the landscape and should be taken into account in conservation efforts (Stewart et al. 2002). In a microcosm system, under controlled laboratory conditions, intraguild predation between female erigonid spiders Erigone atra (Blackwall) and Oedothorax apicatus, Araneae, Erigonidae, and lacewing larvae (second instar larvae of Chrysoperla carnea, Neuropt., Chrysopidae) and interaction effects of predator combinations on cereal aphi The microcosm studies included 15 wheat seedlings, 15 Sitobion avenae (Aphididae) as the starting population, as well as a female spider, a lacewing larva, or a combination of a spider and a lacewing larva. They were conducted over the course of 7 days. In comparison to lacewing larvae housed alone, intraguild predation by female spiders of the species E. atra and Oedothorax apicatus dramatically raised the death rate of lacewing larvae by 44 and 31%. In compared to controls without predators, the final aphid counts in the microcosms were dramatically decreased by all single predator treatments (spiders, lacewing larvae), as well as the predator combos. Both spider species had a similar predation impact on aphid populations, and there was no statistically significant difference. For both spider species, an additive impact of the predator combinations—“spider plus surviving lacewing larva”—was discovered, leading to lower aphid populations than the single predator treatments. The effects on aphids were non-additive when the lacewing larva was killed by an E. atra female, however aphid numbers did not statistically increase in comparison to the lacewing larva treatment. The effects of the spider and C. carnea larva on the quantity of aphids

654

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

were additives when the lacewing larva was killed by an Oedothorax apicatus female. Intraguild predation was not observed when there were additional prey (fruit flies and Collembola), and E. atra females had no discernible impact on the survival of lacewing larvae. Additionally, in the presence of fruit flies and Collembola, E. atra females had no discernible impact on aphid populations. However, when combined with a lacewing larva that survived, a considerable drop in the aphid population was seen in comparison to the lacewing larva treatment. The presence or absence of an E. atra female had no statistically significant impact on the body mass of lacewing larvae at the end of the trial (Dinter 2002). The three main predators of Zea mays (sweet corn) in western New York are Orius insidiosus, Coleomegilla maculata, and Harmonia axyridis. The main insect pest is Ostrinia nubilalis, also known as the European corn borer. The goal of this study was to examine the rates of Ostrinia nubilalis egg predation for these three species and determine how the availability of alternative food affects egg predation by these predators. According to laboratory findings, all three predators consume the eggs of Orius nubilalis. More eggs were devoured by Coleomegilla maculata than by Harmonia axyridis or Orius insidiosus. On a diet of Orius nubilalis eggs, the larvae of Coleomegilla maculata and Orius insidiosus easily finished development, whereas Harmonia axyridis larvae were unable to do so. For some stage of all species, the presence of corn leaf aphids (Rhopalosiphum maidis) and corn pollen decreased egg predation per bug. In field cage investigations, the reduction in Orius nubilalis egg predation linked to the presence of aphids was verified. The findings were consistent throughout the coccinellid populations examined. When aphids and pollen are abundant, there is less biological control of Orius nubilalis because there are more options for food. Field studies comparing aphids, predator populations, and Orius nubilalis egg predation demonstrate that reduced egg predation per insect more than offsets the higher populations seen in these situations (Musser and Shelton 2003). In field cages containing larvae of the yellow margined leaf beetle, Microtheca ochroloma (Coleoptera: Chrysomelidae), a pest of organic crucifer (Brassica ceae) crops in the United States, three release rates of the spined soldier bug, Podisus maculiventris (Hemiptera: Pentatomidae), were tested to determine the bug’s capacity for predation In February and March of 2009 and in February and March of 2010, the experiment was carried out twice. Six turnip (Brassica rapa) plants in cages received an equal distribution of four (=low), ten (=medium), or sixteen (=high) Podisus maculiventris first instars on the same day that 132 Microtheca ochroloma first instars were put into the cages. No predators were placed in the control treatment cages. During the first four sample dates in 2009, there was a steady decline in the mean number of Microtheca ochroloma larvae. The mean total Microtheca ochroloma counts in the two treatments with higher release rates were considerably lower on the seventh and final sample day than in the low-release rate and control treatments. 39.1% of Podisus maculiventris survivors overall received the highrelease rate therapy. During the first four sample dates in 2010, there was a steady decline in the mean number of Microtheca ochroloma larvae. There were no appreciable variations in the predator release treatments on the fourth sample date

17.5

Multiple Predators

655

or the final (ninth) sampling date. In the high-release rate therapy, Podisus maculiventris overall survival rate (3.1%) was the lowest. This study produced two possible recommendations for growers: If the plants are anticipated to have seven leaves per plant, release 10 first instars of Podisus maculiventris per six plants; otherwise, release four first instars of Podisus maculiventris per six plants (Montemayor and Cave 2012). This section discusses the selective advantages of cannibalism in relation to the danger of predatory behaviour and parasitism. Enhancing predator species richness frequently strengthens prey suppression, according to studies that manipulate predator diversity and assess the impact on herbivore abundance. Greater prey suppression may be a result of mechanisms like complementarity or facilitation, which are regarded as “real” benefits of variety since it is an emergent characteristic of the multispecies predator community. Or it can be brought on by the identity effect, a “apparent” advantage of diversity that arises from the increased possibility of having one exceptionally ravenous predator species as the number of predator species increases (Long and Finke 2014). We simultaneously changed the species richness and species composition of predators attacking bird cherry-oat aphids (Rhopalosiphum padi) on wheat using different greenhouse and field studies (Triticum aestivum). We discovered that assemblages of predators with a diversity of species typically reduce aphid populations more effectively than monocultures of a single species. Individual predator species’ performance varied, nevertheless, and species-rich assemblages did not outperform all compositions of one species, which suggests an identity effect. Particularly well-performing throughout tests were single-species compositions of the lady beetle Coleomegilla maculata, and on average, predator assemblages with a lady beetle predator had lower overall aphid abundances than compositions when lady beetles were absent. Together, these findings show that lady beetles, especially Coleomegilla maculata, play a major role in natural pest control and imply that predator species composition and identity are crucial considerations in the fight to preserve this vital ecosystem service (Long and Finke 2014). Saleh et al. (2017) assess the impact of releasing Chrysoperla carnea larvae in their second instar at various predator-to-prey ratios to manage tetranychid mites on Phaseolus vulgaris plants in semi-field circumstances over the summer of 2016. When the P:P ratio was 1:10 or 1:120, the results showed that Tetranychus urticae could be effectively controlled after 6 days, whereas at higher ratios (1:40 and 1:50), the tetranychid mite population began to decline after 12 days. The findings demonstrated that, at 1:10 and 1:20 P:P ratios, the numbers of kidney bean tetranychid mites dropped after 6 days after release by 100 and 86.6%, respectively. Tetranychus urticae had percentage redactions of 100, 100, 93.3, 90.2%, and 84.7 days with P:P releases of 1:10, 1:20, 1:30, 1:40, and 1:50, as opposed to Chrysoperla carnea larvae, which released after 9 days. The acquired data demonstrated that for a period of 12 days following the release, there were no tetranychid mites. A substantial negative correlation between P:P ratios and the kidney beantetranychid mite reduction percentage was found using regression analysis (Saleh et al. 2017).

656

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

Oulema melanopus (Coleoptera: Chrysomelidae), sometimes known as the cereal leaf beetle (CLB), is an invasive pest that has recently been discovered in the Canadian Prairies. We conducted a field investigation to measure the predation of CLB eggs as well as a series of laboratory assays to identify probable predators. The most dependable predators of eggs and larvae in no-choice Petri dish testing were various species of common lady beetles (Coccinellidae), rove beetles (Staphylinidae), and ground beetles (Carabidae). Numerous larvae were devoured by Nabis spp. (Hemiptera: Nabidae) and wolf spiders (Araneae: Lycosidae), but no eggs. When Sitobion avenae, a substitute food source, was provided, Hippodamia spp., Coccinella septempunctata (Coleoptera: Coccinellidae), and Pterostichus melanarius (Coleoptera: Carabidae) also consumed CLB eggs on potted plants (Hemiptera: Aphididae). In our field investigation, we discovered that, on average, 24.5% of sentinel eggs vanished during the course of a day, most likely as a result of predation (Kheirodin et al. 2019). In order to study the interactions between experimentally simulated combinations of predator guilds (generalists and specialists) travelling to the crop field from the surrounding landscape, and three fertilisation treatments, Aguilera et al. (2021) set up 2 2 2 m mesocosms in an oat field (no fertilisation, organic manure and inorganic mineral fertilisation). The mesocosms in an ongoing, long-term agricultural field experiment simulated the spread of different predator types, including no predators, generalist predators (wolf spiders), and specialised predators, from the surrounding landscape to study the effects of local management of organic manure or inorganic mineral fertilisation (ladybirds). The authors investigated the top-down versus bottom-up processes that controlled aphid density under various fertilisation regimes and the impact of predator population composition on the level of pest suppression. Results show beneficial synergistic effects between predator spillover and manure fertilisation on the inhibition of aphid growth. Top-down suppression of aphids was more effective when manure was fertilised and when predators specialised were present (ladybirds). In plots using inorganic fertiliser, bottom-up impacts on the plant biomass growth predominated. The same yield was produced by both organic and inorganic fertilisation but by distinct processes. Given that inorganically fertilised plants are mostly driven by bottom-up effects, the quantity of locally emerging predators in the manure treatment increased top-down pest control and produced plant biomass levels comparable to those of organically fertilised plants (Aguilera et al. 2021). According to this study, organic fertilisation promoted the local establishment of predators, enhancing top-down pest control. Local predator communities, in contrast, were unable to control aphid populations when treated with inorganic materials or without fertilisation. Here, aphid population development has to be slowed down by predator infiltration from outside the crop field. It becomes clear that managing landscapes to support migratory predators is especially crucial for crop fields without manure additives (Aguilera et al. 2021). The Huanglongbing (HLB) disease, which is endangering global citrus production, is linked to Diaphorina citri (Hemiptera: Liviidae), a major citrus pest. Tamarixia radiata, a parasitoid, and chemical pesticides have been used to control Diaphorina citri in the past (Hymenoptera: Eulophidae). Predatory mites, syrphid

17.6

Multiple Controlled Conditions in Field Cage for Life Traits and Bioefficacy

657

flies, lacewings, and lady beetles, among other species, have been shown to feed on Diaphorina citri in citrus agroecosystems. Although Tamarixia radiata has had varying degrees of effectiveness, the use of pesticides is controversial owing to environmental and social issues, particularly given that Diaphorina citrii is primarily distributed in residential areas. Finding sustainable and supplementary methods to eradicate this pest is therefore urgently needed. In field surveys, generalist predators like green lacewings and coccinellids have frequently been linked to Diaphorina citri, but they have rarely been put to the test. Five predators were tested for their ability to prey on prey under laboratory, greenhouse, and semi-field conditions: the coccinellids Diomus pumilio and Rhyzobius lophanthae (Coleoptera: Coccinellidae), the green lacewings Chrysoperla comanche and Chrysoperla rufilabris (Neuroptera: Chrysopidae), and the brown lacewing Sympher (Neuroptera: Hemerobiidae). Under laboratory conditions, all of them could consume Diaphorina citrian and lower its populations, but important distinctions between greenhouse and field cages developed, with some predators (the green lacewings and D. pumilio) being more successful. We propose potential use for these predators as an alternative biological control technique to address present Diaphorina citri control limits in commercial orchards, in neighbourhoods, and at the urban-agricultural interface (Gomez-Marco et al. 2022).

17.6

Multiple Controlled Conditions in Field Cage for Life Traits and Bioefficacy

Leigh and Gonzalez (1976) studied a variety of predators inside a field cage.They checked Lygus hesperus Knight’s predator effectiveness was assessed using a series of 5 treatments: control or check, Geocoris pallens, Nabis americoferus, Chrysopa carnea, and the naturally occurring predator complex. Lygus insect populations of egg-nymphs or nymphs were greatly reduced by Geocoris pallens. When kept in sleeve cages, Nabis americoferus was effective as a predator on the egg-nymph stages, but proved ineffectual in large (183 cm 183 cm 366 cm) field cages. Chrysopa carnea did not work. In the egg-nymph stage, the natural predator complex controlled lygus bugs by 53–76%. In 1980, in field cage testing, 200 predator samples from 16 species were examined to assess the daily eating rates of the eggs and larvae of the soybean looper, Pseudoplusia includens. The nabids Reduviolus roseipennis, Tropiconabis capsiformis, and Hoplistoscelis deceptivus, as well as the lygaeids Geocoris punctipes, and Chrysopa rufilabris, had the highest predation rates. The lygaeids Geocoris punctipes and Geocoris uliginosus, the carabid beetle Calleida decora, and the nabids T. capsiformis and H. deceptivus consumed the most tiny larvae. The most medium-sized larvae were devoured by the pentatomid Stiretrus anchorago and the reduviid Arilus cristatus (Richman et al. 1980). Due to the enhanced favourable habitat, the green peach aphid (GPA), Myzus persicae, multiplied more quickly in cages with many plants than in cages with a single plant. In cages with a single plant and many plants, respectively, the growth

658

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

rates of the GPA populations were lowered by 6.8 and 16.7% with the addition of the predators Geocoris bullatus and Nabis alternatus Parshley. Based on predator efficacy and power, G. bullatus and N. alternatus consumption rates rose 1.9 to 2.7-fold in cages with >50 GPA compared with 50 GPA per cage. Both predator species’ intake rates increased 1.9 times when compared between cages with a single plant and numerous plants. The GPA populations in the higher temperature cages showed declining development rates in outdoor cages with three temperature ranges and no predators present. In low- and medium-temperature cages, the growth rates further fell with the addition of predators by 10.7 to 21% and 25% correspondingly compared to the control cages (Tamaki et al. 1981). Under field-cage conditions, the average daily predation rate and the impact of the prey on the growth and survival of the several nymphal stages of Podisus maculiventris were investigated. The prey consisted of third instar larvae of the huge white butterfly Pieris brassicae and the Colorado potato beetle Leptinotarsa decemlineata. The findings demonstrate that the fifth instar nymphs and adults of the predator Podisus maculiventris would be promising agents for the biological control of larvae of Leptinotarsa decemlineata and Pieris brassicae under climatic conditions similar to those occurring in the Thessaloniki region (N. Greece) (Stamopoulos and Chloridis 1994). Adelges tsugae Annand, sometimes known as the hemlock woolly adelgid, is a destructive non-native pest of hemlock in eastern North America. In Japan, where Laricobius osakensis is a major predator, A. tsugae is a natural species. In the mountains of southwest Virginia, USA, in plant hardiness zones 5b and 6b, the performance of adult and immature stages of Laricobius osakensis was assessed in sleeve cages on Tsuga canadensis infested with adelgid. Throughout all of the biweekly sample periods from December 2010 to May 2011, including the winter months when the temperature typically fell below 0 °C, adults consumed adelgids and lay eggs. Predation of A. tsugae/predator in cages with one adult pair—one male and one female—was 0.3–0.9/day in December and January, rose to 2.5/day in February, and then fell to 0.15/day in early May. The oviposition rate increased from 0.02 eggs per day from December to mid-January to a peak of 1.5 eggs per day in early April, then decreased to 0.4 eggs per day in late April, delaying the changes in feeding by 2–4 weeks. Despite temperatures as low as 18 °C, mortality was 20% in cages left unattended for 2 months throughout the winter (cages examined biweekly had higher mortality, likely due to disturbance). 34 and 37 offspring were found in cages that were left unattended for 2 months in the winter or the first few weeks of spring, respectively. A pair of adults and their offspring ingested 2.5 and 2.4 adult adelgids or ovisacs each day throughout each biweekly period. In the 28 days following the placement of the eggs in the cages, 48% and 95% of the retrieved larvae had matured and had each destroyed 43 and 39 ovisacs, respectively. According to this study, Laricobius osakensis may have the ability to be an efficient biological control agent of A. tsugae in the majority of the eastern US regions where it is a pest (Wallace and Hain 2000). Studies in the lab on the predator Cyrtorhinus lividipennis’s numerical response to Nilaparvata lugens, one of its natural prey, established the reproductive response to egg production, prey-dependent birth rate, prey-dependent mortality rate, and life

References

659

cycle duration to various egg densities. 20 replications of 1, 5, 10, 15, 20, and 30 eggs/predator per cage were used during the experiment. The findings showed that Cyrtorhinus lividipennis’ numerical response was positively correlated with prey density, with higher quantities of eggs laid, a longer lifespan, and a later adult age observed at a high prey density than a low prey density. The prey density had little impact on the adult longevity or sex ratio, though. Cyrtorhinus lividipennis growth required a minimum prey density of 5 eggs per day, whereas 10 eggs per day were required for egg production. When Cyrtorhinus lividipennis devoured one Nilaparvata lugens egg, the examination of the numerical reaction provided information to calculate the prey-dependent birth rate and prey-dependent mortality rate. According to the findings, prey density-dependent birth rates were 0.123 eggs per bug per day and prey density-dependent mortality rates were 0.027 bugs per day (Tangkawanit et al. 2018). The most prevalent species of mirid bug found in cotton fields in northern China is Apolygus lucorum. Apolygus lucorum showed positive prey consumption rates on cotton aphids throughout a 24-h period in a Petri plate. On cotton aphid, a similar effect was also observed under screen house conditions (9 and 18 days after the releases) (Li et al. 2020). In both studies, there was a considerable decrease in aphid abundance when mirid bugs were present. These findings also offer new proof that Apolygus lucorum is effective at reducing aphid populations (Li et al. 2020). According to the potential reduction of cotton aphids by mirid bugs, linear regression studies revealed a substantial correlation between rising abundances of mirid bugs and generalist predators and falling aphid abundance. Dortus primarius, a mirid predator, was able to mature from an egg to an adult on a tomato plant infested with Frankliniella schultzei larvae in 23–24 days with a death rate of 15 and 11.76% for fourth and fifth instar nymphs, respectively. Female adults and the fifth instar both displayed a class III functional response. Both the fifth instar and female adults ingested 21.6 and 28.6 thrips larvae per day, respectively, at the greatest prey density (40). Attack rate, handling time, a/Th, and T/Th functional response metrics demonstrated this predator’s effectiveness against Frankliniella schultzei (Varshney and Budhlakoti 2022).

17.7

Recommendation

It is mandatory to test any predator’s bioefficacy in a field cage, a screen room, or a greenhouse. Therefore, we advise you to do this study under a variety of climatic conditions, using crops of various cultivers (varieties), at various times of the year.

References Ables JR (1978) Feeding behaviour of an assassin bug, Zelus renardii. Ann Entomol Soc Am 71:476–478

660

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

Aguilera G, Riggi L, Miller K et al (2021) Organic fertilisation enhances generalist predators and suppresses aphid growth in the absence of specialist predators. J Appl Ecol 58(7):1455–1465 Ambrose DP (2000) Suppression of Helicoverpa armigera (Hubner) and Anomis flava (Fabricius) infesting okra by the reduviid predator Rhynocoris kumarii Ambrose and Livingstone in field cages. Pest Manag Hortic Ecosyst 6:32–35 Ambrose DP, Claver MA (1999) Suppression of cotton leafworm Spodoptera litura, flower beetle Mylabris pustulata and red cotton bug Dysdercus cingulatus by Rhynocoris marginatus (Fabr.) (Het., Reduviidae) in cotton field cages. J Appl Entomol 123(4):225–229 Arthurs S, Chen J, Dogramaci M et al (2011) Evaluation of Montandoniola confusa Streito and Matocq sp. nov. and Orius insidiosus say (Heteroptera: Anthocoridae), for control of Gynaikothrips uzeli Zimmerman (Thysanoptera: Phlaeothripidae) on Ficus benjamina. Biol Control 57(3):202–207 Ballal CR, Yamada K (2016) Anthocorid predators. In: Ecofriendly pest management for food security. Academic Press, London, pp 183–216 Biever KD, Chauvin RL (1992) Suppression of the Colorado potato beetle (Coleoptera: Chrysomelidae) with augmentative releases of predaceous stinkbugs (Hemiptera: Pentatomidae). J Econ Entomol 85(3):720–726 Bosco L, Giacometto E, Tavella L (2008) Colonization and predation of thrips (Thysanoptera: Thripidae) by Orius spp. (Heteroptera: Anthocoridae) in sweet pepper greenhouses in Northwest Italy. Biol Control 44(3):331–340 Calvo FJ, Torres-Ruiz A, Velázquez-González JC et al (2016) Evaluation of Dicyphus hesperus for biological control of sweetpotato whitefly and potato psyllid on greenhouse tomato. BioControl 61:415–424 Castañé C, Arnó J, Gabarra R et al (2011) Plant damage to vegetable crops by zoophytophagous mirid predators. Biol Control 59:22–29 Chapman JW, Williams T, Martínez AM et al (2000) Does cannibalism in Spodoptera frugiperda (Lepidoptera: Noctuidae) reduce the risk of predation? Behav Ecol Sociobiol 48(4):321–327 Claver MA, Ambrose DP (2001) Evaluation of Rhynocoris kumarii Ambrose & Livingstone (Hemiptera: Reduviidae) as a potential predator of some lepidopteran pests of cotton. J Biol Control:15–20 Darr MN, McAvoy TJ, Brewster CC et al (2016) Field-cage evaluation of survival, reproduction, and feeding behavior of adult Scymnus coniferarum (Coleoptera: Coccinellidae), a predator of Adelges tsugae (Hemiptera: Adelgidae). Environ Entomol 45(6):1527–1535 Deligeorgidis PN (2002) Predatory effect of Orius Niger (Wolff)(Hem., Anthocoridae) on Frankliniella occidentalis (Pergande) and Thrips tabaci Lindeman (Thysan., Thripidae). J Appl Entomol 126(2–3):82–85 Dinter A (2002) Microcosm studies on intraguild predation between female erigonid spiders and lacewing larvae and influence of single versus multiple predators on cereal aphids. J Appl Entomol 126(5):249–257 Doğramaci M, Arthurs SP, Chen J et al (2011) Management of chilli thrips Scirtothrips dorsalis (Thysanoptera: Thripidae) on peppers by Amblyseius swirskii (Acari: Phytoseiidae) and Orius insidiosus (Hemiptera: Anthocoridae). Biol Control 59(3):340–347 Firlej A, Doyon J, Harwood JD et al (2013) A multi-approach study to delineate interactions between carabid beetles and soybean aphids. Environ Entomol 42(1):89–96 Frescata C, Mexia A (1996) Biological control of thrips (Thysanoptera) by Orius laevigatus (Heteroptera: Anthocoridae) in organically-grown strawberries. Biol Agric Hortic 13(2):141– 148 Gardiner MM, Landis DA (2007) Impact of intraguild predation by adult Harmonia axyridis (Coleoptera: Coccinellidae) on Aphis glycines (Hemiptera: Aphididae) biological control in cage studies. Biol Control 40(3):386–395 Geetha N, Gopalan M (1999) Effect of interaction of predators on the mortality of nymphs of brown plant hopper, Nilaparvata lugens Stal. J Entomol Res 23(2):179–181

References

661

Ghongade DS, Sangha KS, Dhall RK et al (2022) Field evaluation of Blaptostethus pallescens Poppius (Hemiptera: Anthocoridae) in controlling Tetranychus urticae Koch on parthenocarpic cucumber under protected environments. Int J Acarol 48(2):139–144 Gomez-Marco F, Gebiola M, Simmons GS et al (2022) Native, naturalized and commercial predators evaluated for use against Diaphorina citri. Crop Prot 155:105907 Hameed Y, Rizwan M, Shahzaib M et al (2022) Functional response of Chrysoperla carnea (Stephens)(Neuroptera: Chrysopidae) to Lipaphis erysimi (Kaltenbach) and Diuraphis noxia (Kurdjumov). Pakistan J Zool 55:1–5 Hoogendoorn M, Heimpel GE (2004) Competitive interactions between an exotic and a native ladybeetle: a field cage study. Entomol Exp Appl 111(1):19–28 Huang N, Enkegaard A (2010) Predation capacity and prey preference of Chrysoperla carnea on Pieris brassicae. BioControl 55(3):379–385 Irwin ME, Gill RW, Gonzalez D (1974) Field-cage studies of native egg predators of the pink bollworm in southern California cotton. J Econ Entomol 67(2):193–196 Jiang Y, Wu Y, Duan Y et al (2011) Control efficiencies of releasing Orius sauteri (Heteroptera: Anthocoridae) on some pests in greenhouse pepper. Chin J Biol Control 27(3):414–417 Kalidas S (2012) Augmentative biological control potential evaluation of cotton pests using reduviid. Ph.D. Thesis. Manonmaniam Sundaranar University, Tirunelveli, Tamil Nadu, India Kheirodin A, Costamagna AC, Cárcamo HA (2019) Laboratory and field tests of predation on the cereal leaf beetle, Oulema melanopus (Coleoptera: Chrysomelidae). Biocontrol Sci Tech 29 (5):451–465 Kim JH, Byeon YW, Kim YH et al (2006) Biological control of thrips with Orius strigicollis (Poppius)(Hemiptera: Anthocoridae) and Amblyseius cucumeris (Oudemans)(Acari: Phytoseiidae) on greenhouse green pepper, sweet pepper and cucumber. Korean J Appl Entomol 45(1):1–7 Kindlmann P, Yasuda H, Kajita Y et al (2015) Predator efficiency reconsidered for a ladybird-aphid system. Front Ecol Evol 3:27 Koch RL, Hutchison WD, Venette RC et al (2003) Susceptibility of immature monarch butterfly, Danaus plexippus (Lepidoptera: Nymphalidae: Danainae), to predation by Harmonia axyridis (Coleoptera: Coccinellidae). Biol Control 28(2):265–270 Lamb AB, Salom SM, Kok LT (2005) Survival and reproduction of Laricobius nigrinus fender (Coleoptera: Derodontidae), a predator of hemlock woolly adelgid, Adelges tsugae Annand (Homoptera: Adelgidae) in field cages. Biol Control 32(2):200–207 Latham DR, Mills NJ (2009) Quantifying insect predation: a comparison of three methods for estimating daily per capita consumption of two aphidophagous predators. Environ Entomol 38(4):1117–1125 Legaspi JC, Correa JA, Carruthers RI et al (1996) Effect of short-term releases of Chrysoperla rufilabris (Neuroptera: Chrysopidae) against silverleaf whitefly (Homoptera: Aleyrodidae) in field cages. J Entomol Sci 31(1):102–111 Legaspi JC, Ciomperlik M, Legaspi BC Jr (2001) Field cage evaluation of Serangium parcesetosum as a predator of blackfly eggs. Southwest Entomol 26(2):171–172 Leigh TF, Gonzalez D (1976) Field cage evaluation of predators for control of Lygus hesperus knight on cotton. Environ Entomol 5(5):948–952 Li W, Wang L, Jaworski CC et al (2020) The outbreaks of nontarget mirid bugs promote arthropod pest suppression in Bt cotton agroecosystems. Plant Biotechnol J 18(2):322 Lingren PD, Ridgway RL, Jones SL (1968) Consumption by several common arthropod predators of eggs and larvae of two Heliothis species that attack cotton. Ann Entomol Soc Am 61(3):613– 618 Long EY, Finke DL (2014) Contribution of predator identity to the suppression of herbivores by a diverse predator assemblage. Environ Entomol 43(3):569–576 López JD Jr, Ridgway RL, Pinnell RE (1976) Comparative efficacy of four insect predators of the bollworm and tobacco budworm. Environ Entomol 5(6):1160–1164

662

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

Losey JE, Denno RF (1998) Positive predator–predator interactions: enhanced predation rates and synergistic suppression of aphid populations. Ecology 79(6):2143–2152 Lucas É, Rosenheim JA (2011) Influence of extraguild prey density on intraguild predation by heteropteran predators: a review of the evidence and a case study. Biol Control 59(1):61–67 Mohaghegh J, Najafi I (2003) Predation capacity of Andrallus spinidens (F.)(Het.: Entatomidae) on Naranga aenescens Moore (Lep.: Noctuidae) under semi-field and field conditions. Appl Entomol Phytopathol 71(1):57–68 Montemayor CO, Cave RD (2012) Evaluation of the predation capacity of Podisus maculiventris (Hemiptera: Pentatomidae) on Microtheca ochroloma (Coleoptera: Chrysomelidae) in field cages. J Econ Entomol 105(5):1719–1725 Musser FR, Shelton AM (2003) Predation of Ostrinia nubilalis (Lepidoptera: Crambidae) eggs in sweet corn by generalist predators and the impact of alternative foods. Environ Entomol 32 (5):1131–1138 Nair IJ, Sharma S, Kaur R (2020) Efficacy of the green lace wing, Chrysoperla zastrowi sillemi (Esben-Peterson)(Neuroptera: Chrysopidae), against sucking pests of tomato: an appraisal under protected conditions. Egypt J Biol Pest Control 30(1):1–6 Nechols JR, Obrycki JJ, Tauber CA et al (1996) Potential impact of native natural enemies on Galerucella spp. (Coleoptera: Chrysomelidae) imported for biological control of purple loosestrife: a field evaluation. Biol Control 7(1):60–66 Obrycki John J (1992) Techniques for evaluation of predators of Homoptera. Florida Entomologist 75(4):472–476. https://doi.org/10.2307/3496128; Accessed 26 May 2022 Pérez-Aguilar DA, Martínez AM, Viñuela E et al (2019) Impact of the zoophytophagous predator Engytatus varians (Hemiptera: Miridae) on Bactericera cockerelli (Hemiptera: Triozidae) control. Biol Control 132:29–35 Pérez-Hedo M, Riahi C, Urbaneja A (2020) Use of zoophytophagous mirid bugs in horticultural crops: current challenges and future perspectives. Pest Manag Sci 77:33 Pumariño L, Alomar O (2012) The role of omnivory in the conservation of predators: Orius majusculus (Heteroptera: Anthocoridae) on sweet alyssum. Biol Control 62(1):24–28 Rautapää J (1977) Evaluation of predator-prey ratio using Chrysopa carnea Steph. in control of Rhopalosiphum padi (L.). Ann Agric Fenniae 16:103–109 Ravichandran B (2004) Biocontrol potential evaluation of a Reduviid predator Acanthaspis pedestris Stål (Insecta: Heteroptera: Reduviidae) on three chosen Lepidopteran insect pests. Ph.D. thesis. Manonmaniam Sundaranar University, Tirunelveli, Tamil Nadu, India, p 102 Richman DB, Hemenway RC Jr, Whitcomb WH (1980) Field cage evaluation of predators of the soybean looper, Pseudoplusia includens (Lepidoptera: Noctuidae). Environ Entomol 9(3):315– 317 Ridgway RL, Jones SL (1968) Field-cage releases of Chrysopa carnea for suppression of populations of the bollworm and the tobacco budworm on cotton. J Econ Entomol 61(4):892– 898 Rosenheim JA, Wilhoit LR, Armer CA (1993) Influence of intraguild predation among generalist insect predators on the suppression of an herbivore population. Oecologia 96(3):439–449 Sahayaraj K (1991) Bioecology, ecophysiology and ethology of chosen predatory hemipterans and their potential in biological control (Insecta: Heteroptera: Reduviidae). Ph.D. thesis, Madurai Kamaraj University, India. p 233 Sahayaraj K, Ambrose DP (1997) Field cage evaluation of predator, Ecotomocoris tibialis dist. (Heteroptera: reduviidae) to control Dysdercus cingulatus fab. J Insect Sci 10(1):65–66 Sahayaraj K, Kalidas S, Estelle LYL (2020) Bioefficacy of Rhynocoris longifrons (Stål) (Heteroptera: Reduviidae) against multiple cotton pests under screen house and field conditions. Sci Rep 10(1):1–9 Saleh FM, Tawfik AA, Elsherbeni MK (2017) Release of Chrysoperla carnea (Stephens) (Neruoptera: Chrysopidae) as a biological control agent against Tetranychid mites, Tetranychus urticae (Koch), on Phaseolus vulgaris L. Plantations under semi field conditions. J Plant Prot Pathol 8(12):615–617

References

663

Sanchez JA (2008) Zoophytophagy in the plantbug Nesidiocoris tenuis. Agric For Entomol 10:75– 80 Schmaedick MA, Shelton AM (1999) Experimental evaluation of arthropod predation on Pieris rapae (Lepidoptera: Pieridae) eggs and larvae in cabbage. Environ Entomol 28(3):439–444 Schmitz OJ (2008) From mesocosms to the field: the role and value of cage experiments in understanding top-down effects in ecosystems. In: Insects and ecosystem function. Springer, Berlin, Heidelberg, pp 277–302 Shaw PB, Owens JC, Huddleston EW et al (1987) Role of arthropod predators in mortality of early instars of the range caterpillar, Hemileuca oliviae (Lepidoptera: Saturniidae). Environ Entomol 16(3):814–820 Shipp JL, Wang K (2006) Evaluation of Dicyphus hesperus (Heteroptera: Miridae) for biological control of Frankliniella occidentalis (Thysanoptera: Thripidae) on greenhouse tomato. J Econ Entomol 99:414–420 Shrestha RB, Parajulee MN (2013) Potential cotton aphid, Aphis gossypii, population suppression by arthropod predators in upland cotton. Insect Sci 20(6):778–788 Smith HA, Krey KL (2019) Three release rates of Dicyphus hesperus (Hemiptera: Miridae) for management of Bemisia tabaci (Hemiptera: Aleyrodidae) on greenhouse tomato. Insects 10 (7):213 Stamopoulos DC, Chloridis A (1994) Predation rates, survivorship and development of Podisus maculiventris (Het.: Pentatomidae) on larvae of Leptinotarsa decemlineata (Col.: Chrysomelidae) and Pieris brassicae (Lep.: Pieridae), under field conditions. Entomophaga 39(1):3–9 Stark SB, Whitford F (1987) Functional response of Chrysopa carnea [Neur: Chrysopidae] larvae feeding on Heliothis virescens [Lep.: Noctuidae] eggs on cotton in field cages. Entomophaga 32 (5):521–527 Stewart CD, Braman SK, Pendley AF (2002) Functional response of the azalea plant bug (Heteroptera: Miridae) and a green lacewing Chrysoperla rufilabris (Neuroptera: Chrysopidae), two predators of the azalea lace bug (Heteroptera: Tingidae). Environ Entomol 31(6):1184– 1190 Tamaki G, Weiss MA, Long GE (1981) Evaluation of plant density and temperature in predatorprey interactions in field cages. Environ Entomol 10(5):716–720 Tangkawanit U, Hinmo N, Khlibsuwan W (2018) Numerical response of Cyrtorhinus lividipennis (Hemiptera: Miridae) to Nilaparvata lugens (Hemiptera: Delphacidae). J Entomol Sci 53 (2):171–179 Thead LG, Pitre HN, Kellogg TF (1987) Predation on eggs and larvae of Heliothis virescens [Lep.: Noctuidae] by an adult predator complex in cage studies on cotton. Entomophaga 32(2):197– 207 Urbaneja-Bernat P, Bru P, González-Cabrera J et al (2019) Reduced phytophagy in sugarprovisioned mirids. J Pest Sci 92:1139–1148 VanDen Bosch R, Leigh TF, Gonzalez D, Stinner RE (1969) Cage studies on predators of the bollworm in cotton. J Econ Entomol 62(6):1486–1489 van Lenteren JC, Bueno VHP, Calvo FJ et al (2018) Comparative effectiveness and injury to tomato plants of three Neotropical mirid predators of Tuta absoluta (Lepidoptera: Gelechiidae). J Econ Entomol 111(3):1080–1086 Varshney R, Budhlakoti N (2022) Biology and functional response of the predator, Dortus primarius (distant)(Hemiptera: Miridae) preying on Frankliniella schultzei (Trybom) (Thysanoptera: Thripidae). Egypt J Biol Pest Control 32(1):1–7 Vasanthakumar D, Babu A (2013) Life table and efficacy of Mallada desjardinsi (Chrysopidae: Neuroptera), an important predator of tea red spider mite, Oligonychus coffeae (Acari: Tetranychidae). Exp Appl Acarol 61(1):43–52 Veronesi ER, Saville DJ, van Koten C et al (2022) Potential of the mirid bug, Engytatus nicotianae, for the biological control of the tomato-potato psyllid in greenhouses. Crop Prot 156:105941

664

17

Bioefficacy Evaluation of Insect Predators Under Pot Condition/Screen. . .

Wallace MS, Hain FP (2000) Field surveys and evaluation of native and established predators of the hemlock woolly adelgid (Homoptera: Adelgidae) in the southeastern United States. Environ Entomol 29(3):638–644 Wyss E, Villiger M, Müller-Schärer H (1999) The potential of three native insect predators to control the rosy apple aphid, Dysaphis plantaginea. BioControl 44(2):171–182

Controlled Field Cage and Field Evolution

18

Contents 18.1 18.2 18.3

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Predators in General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hemipterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.3.1 Miridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.3.2 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.3.3 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.3.4 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.4 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.5 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.6 Dermaptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.7 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.7.1 Effectiveness of Syrphids as Field-Based Biocontrol Agents . . . . . . . . . . . . . . . 18.8 Predator Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.9 Comparison Between Organic and Conventional Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.10 Genetically Modified (GM) Crops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.11 Bio-Intensive Pest Management (BIPM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.12 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

18.1

665 666 671 671 672 676 679 681 684 687 687 687 688 688 688 693 695 695

Introduction

In cotton fields, predatory natural enemies play a significant role in pest control and may boost cotton growers’ profitability by lowering the need for chemicals. Integrated pest management (IPM) techniques that can result in a decrease in the usage of chemical pesticides depend heavily on biological control agents (BCAs), such as parasitoids and predators. The range, availability, and identity of alternative prey frequently have an impact on how pests are preyed upon by generalist predators. Occasionally, comparably desirable alternatives to the intended prey will divert the attacks of generalist predators, upsetting the balance of biological

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_18

665

666

18

Controlled Field Cage and Field Evolution

control. At the same time, over an extended period of time, alternative prey can aid in the development and maintenance of predator populations that can hunt pests. The same is true for predators that grow in other agricultural or natural settings; they can move into a field and attack the pest herbivores there. The optimum scenario might be for generalists to switch to targeting pests as non-pest prey naturally decline as herbivorous pests move in. However, data from straightforward laboratory arenas may not provide an insight into the net impacts under open-field conditions because apparent benefit or harm to biological control frequently reflects the spatiotemporal scale at which the indirect effects of alternative prey are investigated. The following groups have been involved in biological pest management: 1. 2. 3. 4. 5. 6. 7. 8. 9.

The Food and Agriculture Organization (FAO). Organisation for Economic Co-operation and Development (OECD). The North American Plant Protection Organization (NAPPO). The European and Mediterranean Plant Protection Organization (EPPO). The Ministry of Agriculture and Farmers’ Welfare Directorate of Plant Protection, Quarantine, and Storage. The International Plant Protection Convention (IPPC). Regional Plant Protection Organization (RPPO). International Plant Protection Australia. Thailand’s National Biological Control Research Centre.

18.2

Predators in General

For the purpose of eradicating native or non-native pests, “augmentation” in biological management refers to the practices that involve colonisation, periodic release (either by inoculation or inundation), and mass culture to enhance natural enemy populations. Chinese citrus growers utilised paper nests of ants (Oecophylla smaragdina) to defend their trees from other insects in the year 300 AD, which is when instances of predatory insects being used as agents of insect management methods were first recorded. For use in trials on the biological control of the balsam woolly aphid, 15 species of predators from Pakistan and India were brought to North Carolina between 1961 and 1965. Studies on the species’ recovery were conducted; however, there is currently no proof that any have succeeded in becoming established. This lack of establishment is generally due to significant climatic changes between the old and new environments as well as to predatory larvae and oviposition adults’ inability to accept their prey (Amman and Speers 1971). Adelges tsugae Annand, the hemlock woolly adelgid, has been the subject of little research regarding the effects of natural enemies. Using field surveys and cage exclusion trials, this 2-year study investigated the interaction between native predators and Adelges tsugae in south-eastern United States. In 1997 and 1998, predators were gathered in incredibly low quantities. A total of 81% of the predators gathered in 1998 comprised of Harmonia axyridis (Coleoptera: Coccinellidae), lacewings (Neuroptera: Chrysopidae and

18.2

Predators in General

667

Hemerobiidae), and gall gnats (Diptera: Cecidomyiidae). At all three sites, cage exclusion trials conducted in 1997 and 1998 found no discernible effects of predators. Due to the low densities of predators discovered while adelgids were in high abundance, it seems unlikely that native predators had any real control on adelgid populations. Therefore, releasing exotic predators into these areas under controlled conditions should be considered. Prior to widespread releases of alien predators, it is critical for scientists and resource managers to continue to understand the significance of pre-release assessments of native natural enemies (Wallace and Hain 1999). Bemisia tabaci (Hemiptera: Aleyrodidae), a pest of several crops, especially vegetables, is known as the sweet potato whitefly. For the control of the sweet potato whitefly, weekly inundated releases of the coccinellid predator Coccinella undecimpunctata, the common green lacewing predator Chrysoperla carnea (Neuroptera), and the mirid predator Macrolophus caliginosus (Miridae) were made in three different vegetable crops, including squash, cucumber, and cabbage. In the vegetable crops throughout the course of 20 weeks, each predator’s larvae or nymphs were released in numbers ranging from 1 million to 2.5 million. The majority of the season saw a reduction in whitefly populations of between 25% and 45% in each crop where each predator was deployed. Each predator had a comparable impact on the decline of the whitefly population. Release of the predators typically had no advantage because it was late in the season (October), when whitefly populations were low. Chrysoperla carnea had the highest number of predators retrieved during sampling in all crops; however, this was expected, given that more of this predator’s individuals were released during the experiment than any other. The effectiveness of these natural enemies in controlling Bemisia tabaci in vegetable crops is defined by these findings (Simmons and Abd-Rabou 2011). Only infrequently were either of the two natural enemies examined discovered in the crop (maximum 1/plot of 60 plants for Franklinothrips vespiformis and 4/plot of 60 plants for C. carnea). They were injected every 2 weeks, and the crop showed no signs of them becoming established. On the leaves, no green lacewing eggs could be seen. Only two F. vespiformis immature larvae were discovered once in a Thrips setosus colony. Both Euseius gallicus and Phytoseiulus persimilis failed to establish, most likely as a result of the usage of sulphur and the scarcity of prey. The predators exhibited no signs of reproductive activity even when T. setosus was present. When compared to the untreated control at the end of the trial, T. setosus was dramatically reduced by all biocontrol treatments. The findings also suggest that inundated releases of biological control agents may be able to at least match chemical applications in limiting T. setosus infestations on hydrangea. Between the approach utilising predatory thrips in conjunction with C. carnea and the one utilising green lacewing alone, there were significant changes in the development of the pest population. The combination of the two positives produced the best outcomes. The combined method that used predatory thrips exclusively as a curative measure when the initial T. setosus hotspots were discovered was the least expensive. Local releases of F. vespiformis would be less expensive and equally effective than ongoing full-field releases of the predatory thrips, but they would necessitate

668

18

Controlled Field Cage and Field Evolution

greater vigilance on the part of the growers. Since C. carnea is seven times less expensive than F. vespiformis, introducing it to an entire field early on in the crop cycle seemed the best course of action. It is simpler to incorporate it into a biological release programme because of its capacity to feed on a variety of pests. The green lacewing may eat spider mites in addition to being efficient against aphids. Both bugs frequently affect hydrangeas. Despite the fact that our research indicates that C. carnea and F. vespiformis can both be efficient predators for the management of Thrips setosus, the evaluated inundated release procedures are still too expensive when compared to the existing chemical methods. It is still uncertain how many predators need be introduced in order to effectively manage T. setosus, and this will undoubtedly depend on the state of the crops. Utilising outbreak point detection techniques or implementing conservation biocontrol technologies that allow the establishment of predators could reduce the cost of T. setosus biocontrol. However, early attempts to supplement Ephestia kuehniella eggs failed to increase the favourable effects’ durability. The establishment’s bottleneck might be the use of sulphur. It is necessary to determine whether sulphur affects the predators negatively in order to validate this. The planned biocontrol programme will need to average between 0.5 and 1 euros per square metre in order to be competitive with chemical control, which will help promote the use of biocontrol against Thrips setosus in hydrangea in the Netherlands. The growers’ mindsets will also need to alter. Beneficials can keep the pest at low densities but do not completely eradicate it. Nowadays, the majority of hydrangea producers use a zero-tolerance approach and spray for Thrips setosus. Others are aware that it is not possible to rely on insecticides for a prolonged period of time. Growers must also gain experience with various control methods. Currently, the majority of these growers solely employ predatory mites, although these are frequently ineffective at the field level. According to our research, predatory thrips and green lacewings are better alternatives (Pijnakker et al. 2019). Utilising native natural enemies as biological control agents for combating cassava mealybugs has been studied in Thailand. These agents include coccinellid predators Nephus sp., Brumoides sp., Cheilomenes sexmaculatus, and Micraspis discolor as well as predatory green lacewings Plesiochrysa ramburi and Mallada basalis. At the National Biological control Research Center (NBCRC)’s Central Regional Centre, Plesiochrysa ramburi and Mallada basalis are mass-reared, and, each month, 10,000,000 eggs of these green lacewings and 10,000,000 parasites are generated. These locally adapted natural enemies have been dispersed in mealybug infestation hotspots in the provinces of Kanchanaburi and Suphan Buri (Suasaard 2010). Additionally, the NBCRC has advocated using certain predatory insects to control certain pests (homopterans, orthopterans, hemipterans, and lepidopterans) in Thailand (Table 18.1).

18.2

Predators in General

669

Table 18.1 Crops, orthopteran, hemipteran, lepidopteran, and hemipteran pests, and suggested predatory insects for their control in Thailand Crop Sugarcane

Pests Sesamia inferens

Predator Anthicus ruficollis Formicimus braminus Proreus similans

Aleurolobus barodensis Aulacaspis tegalensis

Ceratovacuna lanigera

Catana parcesetosa Cheilomenes sexmaculatus Coccinella transversalis Cheilomenes sexmaculatus Chilocorus nigritus Chilocorus circumdatus Coccinella transversalis Ankylopteryx octopunctata Mallada basalis

Saccharicoccus sacchari

Cheilomenes sexmaculatus Coccinella transversalis Scymnus apiciflavus Synonycha grandis Proreus simulans Hyperaspis trilineata Gitona perspicax

Cassava

Corn

Pseudococcus jackbeardsleyi

Ostrinia furnacalis

Plesiochysa ramburi Cheilomenes sexmaculatus Micraspis discolor Proreus simulans Mallada basalis Anthicus ruficollis Ophionia indica Paederus fuscipes

Frankliniella williamsi Helicoverpa armigera

Orius persequens Wollastoniella rotunda Eocanthecona furcellata Eocanthecona robusta Sycanus collaris

Order Coleopter: Anthicidae Coleopter: Anthicidae Dermaptera: Chelisochidae Coleopter: Coccinellidae Coleopter: Coccinellidae

Neuroptera: Chrysopidae Neuroptera: Chrysopidae Coleoptera: Coccinellidae

Dermaptera: Chelisochidae Coleopter: Coccinellidae Diptera: Drosophilidae Neuroptera: Chrysopidae Coleopter: Coccinellidae Dermaptera: Chelisochidae Neuroptera: Chrysopidae Coleoptera: Anthicidae Coleoptera: Carabidae Coleoptera: Staphylinidae Hemiptera: Anthocoridae Hemiptera: Pentatomidae Hemiptera: Reduviidae (continued)

670

18

Controlled Field Cage and Field Evolution

Table 18.1 (continued) Crop

Pests

Predator Mallada basalis

Mythimna separata

Eocanthecona robusta Sycanus collaris

Patanga succincta

Soybean

Epicauta maclini Epicauta waterhousei Mylabris phalerata Sphex viduatus

Myzus persicae

Mallada basalis

Nezara viridula

Cheilomenes sexmaculatus Coccinella transversalis Scymnus apiciflavus Geocoris punctipes Sycanus collaris

Cotton

Helicoverpa assulta Amrasca biguttula Aphis gossypii

Eocanthecona furcellata Eocanthecona robusta Mallada basalis Mallada basalis Cheilomenes sexmaculatus Harmonia octomaculata Coccinella transversalis Micraspis discolor Syrphus balteatus

Thrips tabaci

Wollastoniella rotunda Wollastoniella parvicuneis Franklinothrips vespiformis Mallada basalis

Order Neuroptera: Chrysopidae Hemiptera: Pentatomidae Hemiptera: Reduviidae Coleoptera: Meloidae Hymenoptera: Sphecidae Neuroptera: Chrysopidae Coleoptera: Coccinellidae Hemiptera: Geocoridae Hemiptera: Reduviidae Hemiptera: Pentatomidae Neuroptera: Chrysopidae Neuroptera: Chrysopidae Coleoptera: Coccinellidae Diptera: Syrphidae Hemiptera: Anthocoridae Thysanoptera: Aeolothripidae Neuroptera: Chrysopidae

18.3

18.3

Hemipterans

671

Hemipterans

18.3.1 Miridae In a lab setting and a tomato greenhouse, the ability of Dicyphus hesperus to control Trialeurodes vaporariorum (Westwood), the greenhouse whitefly, and Tetranychus urticae Koch, the two-spotted spider mite, was evaluated. In the lab, Dicyphus hesperus adults readily consumed both pest species, whereas nymphs relied on either whitefly or mites to complete their development from an egg to an adult. However, compared to mites, whiteflies were a better food source for the growth and development of Dicyphus hesperus. Nymphs raised on whiteflies developed more quickly and had larger adult bodies as a result than did those raised on mites. Adults in a greenhouse release orientated to and laid their eggs on sentinel plants with whiteflies but did not do the same for plants with mites. On greenhouse tomatoes, Dicyphus hesperus adults laid their eggs, and the young finished growing in a greenhouse that also had whiteflies and mites. These findings are addressed in relation to the application of Dicyphus hesperus for biological pest control of greenhouse vegetables (McGregor et al. 1999). The establishment of the aphids Macrolophus pygmaeus, Dicyphus errans, Dicyphus tamaninii, and Deraeocoris pallens on sweet pepper plants with and without supplemental food (eggs of the flour moth Ephestia kuehniella and decapsulated cysts of the brine shrimp Artemia franciscana), as well as their effects on a However, the predators Macrolophus pygmaeus and Dicyphus tamaninii could successfully reduce aphid populations when released prior to an intentionally introduced aphid infestation. None of the predatory species could control an established population of aphids on sweet pepper plants. The combination of Macrolophus pygmaeus and a weekly application of extra food produced the best outcomes. Our findings thus show that the level and order of mirid predator and aphid plant penetration determine the effectiveness of biological control. Additional research is required to assess the effectiveness of mirid predatory bugs in sweet pepper crops grown in commercial greenhouses with a variety of pests and natural enemies. This research is particularly important to understand how the variety of food sources affects the behaviour and preferences of these insects when it comes to feeding (Messelink et al. 2015). Recently, the widespread polyphagous predator Engytatus varians (Ev) (Miridae) has been discovered preying on Bactericera cockerelli nymphs for the first time in Mexico (Hemiptera: Triozidae). For 12 weeks, tomato (Solanum lycopersicum) plants were caged in 7-m2 cages in a greenhouse to assess this mirid’s predation potential. Two release rates of Engytatus varians (Ev) adults—one and four adults per plant—were investigated and contrasted with the control rate of 0 Ev per plant. Regardless of the pest life stage, one or four adults of Ev/plant introduced caused an 80–90% reduction in both nymphs and adults of Bactericera cockerelli. In the treatments, there were considerably more Engytatus varians nymphs and adults present each day than in the control (30 and 3, respectively). However, Engytatus varians can feed on tomato plants and leave necrotic rings on the leaves because of

672

18

Controlled Field Cage and Field Evolution

its zoophytophagous tendencies; as a result, this kind of damage was also noted. Although the presence of prey was inversely correlated with the number of necrotic rings/leaf, there were no discernible changes across the treatments. According to our findings, Engytatus varians can establish and manage Bactericera cockerelli populations in greenhouse settings without seriously harming tomato plants (Pérez-Aguilar et al. 2019). Despite the presence of whiteflies and the fact that sesame was a superior plant resource than tomato when given the freedom to roam the field, Nesidiocoris tenuis did not stay on sesame. The mirid was equally prevalent on tomato and sesame plants, and it easily invaded treatment plots with only tomato plants. It has been discovered that mirids migrating into tomato crops focus on plants with high prey density, which may improve pest management on tomatoes. Nesidiocoris tenuis decreased the population of Bemisia tabaci in all 4 years to extremely low levels (10 whiteflies/leaf), which were comparable to those observed in the pots treated with pesticides frequently used by producers in the United States. In fact, as soon as the quantity of whiteflies started to rise, Nesidiocoris tenuis started to infiltrate all treatments, even the pesticide-treated pots. This suggests that whiteflies in open fields may be successfully controlled by Nesidiocoris tenuis. However, more research is required to see whether the mirids can offer pesticide-like control across vast fields, which are characteristic of production in the United States, Mexico, and China. Our study on reduced sesame planting showed that the minimum number of sesame plants required to maintain the mirid could be lowered while maintaining the mirid in the plots. To ensure that the mirids stay and successfully control Bemisia tabaci in big commercial tomato fields, more research is required to ascertain the quantity of sesame and planting distribution. Castillo et al. (2022) carried out six field trials over four consecutive years to create a method to preserve the mirids in order to demonstrate this. Due to the limited incidence of prey, pre-plant inoculation of tomato plants did not result in their establishment. We looked into the possibility of retaining the mirids using sesame (Sesamum indicum). Sesame intercropping kept N. tenuis populations stable for the course of the crop. Any of the open-field experiments never saw the establishment of Macrolophus praeclarus. Damage from mirid and Nesidiocoris tenuis was lessened, and Nesidiocoris tenuis damage was negligible (one necrotic ring/plant). Our findings suggest that intercropping sesame may offer a way to use mirids to manage the established pest B. tabaci and give tomato growers options in the event that Tuta absoluta invades the United States.

18.3.2 Reduviidae A key element of IPM is the augmentative release of predators, and, in particular, rcduviids have a significant impact on the control of several pests with economic relevance. The number of adult Dysdercus cingulatus population in the test and in the control field plots during the sampling dates was unaffected by the first release of Rhynocoris kumarii (September 9 and 14). On September 19, 24, and 29, 1996, the

18.3

Hemipterans

673

second release of R. kumarii eggs had no statistically significant impact on the population of Dysdercus cingulatus’s first instar. Again, the number of II and III nymphal instars of the Dysdercus cingulatus population in the test field plots was not different from that in the 65 control field plots following the release of adults and eggs. However, there was much more following the fourth release of the V instar of R. kumarii. After the introduction of the oligidic diet-reared reduviid predator Rhynocoris marginatus, populations of the insects Aphis craccivora, Helicoverpa armigera, Spodoptera litura, Mylabris indica, and grasshoppers significantly declined. It was evident that Rhynocoris marginatus significantly reduced the population of Pericallia ricini, Helicoverpa armigera, Aphis craccivora, and Spodoptera litura. Additionally, the production of peanuts was high in the predator raised on an artificial diet. The field-released oligidic diet-reared (OD) predator had the highest cost–benefit ratio (1:2), followed by the T1 (CC)-reared predator (1:1.8). Two pods on a groundnut plant were more than three pods among the treatments. The oligidic diet-reared predator plot produced the most groundnuts during the summer (1224 kg/ h), followed by T1 (936 kg/h), and the control plot (728 kg/h). The same pattern was seen in kharif as well. The field-released OD-reared predator experienced the highest percent preventable loss (PAL) during kharif (23.33%), followed by the released T1-reared predator (14.81%) (Sahayaraj and Balasubramanian 2016). In India, Rhynocoris marginatus, Ectomocoris tibialis, and Rhynocoris kumarii were released augumentatively, and their biological control capacity in various agroecosystems was assessed (Ambrose and Claver 1999; Sahayaraj 1999; Sahayaraj and Martin 2003; Sahayaraj and Ravi 2007). Rhynocoris longifrons, also known as the Indian assassin bug, is a natural enemy of the laboratory species D. cingulatus, Massonia pustulata, H. armigera, and S. litura (Ambrose et al. 2003; Kumar 2011). The authors hypothesised that R. longifrons could be used to supplement biological control against insect pests. To decrease Oryctes rhinoceros grubs and adults, the reduviid predator Platymeris longicollis was deployed in a coconut field (Antony et al. 1979). When Sahayaraj (1999) released Rhynocoris marginatus in a groundnut field, he saw lepidopteran pests being suppressed and noted a great yield of groundnuts. Grundy and Maelzer (2000) reported the management of different pests in Australia after releasing Pristhisancus plagipennis in a pigeon pea field. In Australia, once more, the authors carried out a similar experiment in 2002. To investigate the biological control capabilities of the Australian assassin bug, Pristhisancus plagipennis, third instar nymphs were released onto cotton plots at two release densities and two crop growth phases. Field populations of 0.51 and 1.38 nymphs per metre row were obtained with release rates of 2 and 5 nymphs per metre row, respectively, indicating that more than 70% of the nymphs died or fled within 2 weeks of release. For a 7-week period, Helicoverpa spp. larvae were reduced in the plots at effective release rates of 1.38 nymphs per metre row. The plots where P. plagipennis nymphs were released produced significantly higher crop yields, with an effective release rate of 1.38 nymphs per metre row, producing yields comparable to those of insecticide-treated plots. The findings indicate that Pristhisancus plagipennis, when enhanced by inundated discharge, has the ability

674

18

Controlled Field Cage and Field Evolution

to decrease Helicoverpa spp. larval densities in cotton crops (Grundy and Maelzer 2002). In a Rhynocoris marginatus-released field, Sahayaraj and Martin (2003) observed a good yield of groundnuts. At 5000/hectare, all life stages of the reduviid predator Rhynocoris marginatus were released into the groundnut field 30, 50, and 70 days after seeding (DAS). During the course of the investigation, it was possible to see Spodoptera litura, Helicoverpa armigera, Atractomorpha crenulata, Chrotogonus trachypterus, Aphis craccivora, Mylabris pustulata, and Mylabris indica. Spodoptera litura (85.89%) was the species that Rhynocoris marginatus greatly reduced, followed by Helicoverpa armigera (67.65%), Aphis craccivora (46.34%), A. crenulata, and C. trachypterus (42.86%). Rhynocoris marginatus made no difference on the populations of Mylabris spp., which were seen during the study period. Rhynocoris marginatus had no effect on the other predatory fauna present in the control and predator-released fields, including coccinellids (Menochilus sexmaculatus and Coccinella septumpunctata), praying mantises, wasps, damselflies (Agriocnemis femina), and spiders (Lycosa tista and Hippasa pisaurina). The field with released predators had a higher peanut production rate (1480 Kg/hectare) than did the field with only one crop (1104 Kg/hectare), and the results were statistically significant. The R. marginatus-released groundnut field likewise had the highest net gain and cost–benefit ratio. The following qualities are necessary for a natural enemy to be an effective biological control agent in such disturbed habitats: 1. Colonising ability, which allows it to keep up with the habitat’s spatial and temporal disruptions. 2. Temporal persistence, which enables it to maintain its population after colonisation even in the absence of the target pest species. 3. Opportunistic feeding habits, which allow it to take advantage of any opportunities that arise. In India, cotton was mulched to colonise under field conditions. Acanthaspis pedestris, a released reduviid, was discovered colonising mulched (waste, leaf, pot piece, and stone) plots. They were much more numerous in the stone-built plots, but the population of soil predatory arthropods was significantly larger in the plantain leaflet- and stone-built plots compared to the trash- and pot- and piece- and controlplanted plots (Ravichandran 2004). In a different study, four different kinds of mulch, including palmyra leaf, banana leaf, coconut spathe, and stone-laid plots, were selected as shelter. These materials offered more room for movement, surfaces, and hiding places as well as helped lower the temperature and preserve moisture in a cotton field of the Indian state of Tamil Nadu. To combat Spodoptera litura and Helicoverpa armigera, a total of 500 Rhynocoris fuscipes adults, fourth and fifth nymphal instars, and 300 eggs (ready to hatch the next day) of this species were released. The populations of Helicoverpa armigera and Spodoptera litura were lower in mulched plots than in control plots, according to the results. They were discovered below in stone-paved plots, coconut spathes, banana leaves, and palmyra leaves. The greatest populations of Spodoptera litura (2.80) and

18.3

Hemipterans

675

Helicoverpa armigera (3.38) were found in the control plots on 28 May 2010 and 11 April 2010, respectively. The lowest populations of S. litura on palmyra leaf (1.35), banana leaf (1.75), coconut spathe (1.48), and stone-laid (2.06) mulch plots as well as H. armigera on palmyra leaf (1.56), banana leaf (1.94), coconut spathe (0.30), and stone-laid (1.15) mulch plots were observed in mulched plots on April 25 and May 2, 2010, respectively. The freed reduviids were discovered colonising mulched plots (palmyra leaves: 5.21; banana leaves: 4.33; coconut spathes: 4.17; stone-laid: 2.77). When the R. fuscipes were released in the test plots, from 28 March to 9 May 2010, they sought refuge under the mulches. They were more prevalent in palmyra leaf plots than in coconut spathe, banana leaf, or stone-laid plots, and the population of soil and foliage predatory arthropods was also much higher in palmyra leaf plots. The coccinellids were more common than were the hemipteran predators among the foliage predators (Nagarajan 2010). The field colonisation of the assassin bug Rhynocoris kumarii and the biocontrol capability of predatory arthropods were researched in a cotton field in South India, following mulching with sorghum trash and coconut leaflets and with shelter provisioning with pieces of clay pots and stones. Rhynocoris kumarii’s third and fourth nymphal instars were published. When compared to control and other (mulched and shelter-provided) plots, plots with mulched cotton garbage had less Helicoverpa armigera larvae. Mulching had little impact on the quantity of adult Mylabris pustulata, though. When compared to other shelter-provided and control plots, the flower and boll damage was much reduced in garbage and leaf mulch plots. Additionally, more high-quality cotton was grown in the mulch plots compared to the control plots. Additionally, plots with coconut leaf mulch and trash mulch had significantly higher seed cotton yields than did control plots (Claver et al. 2003). The impact of different stages of the predator Coranus africana on the suppression of several populations of the whitefly Bemisia tabaci was assessed in Egypt. Different stages of the predator Coranus africana were distributed in tomato fields in the Qalyubia and Bani Sweif governorates throughout the 2 years 2007 and 2008 by one predator/plant to combat Bemisia tabaci infestation. According to the discharge of the predator in each treatment, the percentage of Bemisia tabaci infestation reduction varied. At Qalyubia during the first year, the reduction in the second week of release in the adult and immature stages was 93.4% and 100%, respectively. However, during the second year, this drop was 94.8% and 97.1%, respectively. However, over the course of 2 years, the percentage reduction at Bani Swaif was 97.4% and 92.7% and 99.2% and 91.2%, respectively (El-Sebaey and Abd El-Wahab 2010). Previously, in 2007, El-Sebaey and Abd El-Wahab (2010) researched the effectiveness of Coranus africana in eradicating several populations of the cotton aphid, Aphis gossypii, in cucumber and squash crops at Fayoum governorate in the following years, 2005 and 2006. In fields of cucumber and squash, C. africana was introduced in a ratio of one predator per plant against 3° of Aphis gossypii infestation. According to the release of the predator in each treatment, the percentage of Aphis gossypii infestation reduction varied. When there was a high infestation, the cucumber field experienced a reduction of 99.9 and 92.86% during the first and second years of the examination, respectively,

676

18

Controlled Field Cage and Field Evolution

whereas the squash field experienced a reduction of 88.6% and 90.1% throughout the course of the two subsequent years of the study, respectively. When the amount of infection was low or moderate in two crops, the insect population was completely eradicated. The basic yield metrics for the two crops during the 2 years of the experiment were expressed as the weight of the fruit and the number of fruits/plant. The number of Spodoptera litura and Helicoverpa armigera on the sampling date of 02 April 2001 in the control (3.12, 2.80) and test (2.24, 2.05) plots was unaffected by the adult Acanthaspis pedestris’ first release during the same period in 2001. A second release of Acanthaspis pedestris eggs on April 1, 2001 had a negligible impact on the populations of Spodoptera litura and Helicoverpa armigera on the day of the sampling (09 April 2001). The targeted pest species Spodoptera litura and Helicoverpa armigera on the sample date of 16 April 2001 were unaffected by the third release of nymphal instars (08 April 2001) after the release of adults and eggs. However, on the sample date of 14 May 2001, Acanthaspis pedestris had a considerable impact on the population of Helicoverpa armigera and Spodoptera litura following the fourth release (22 April 2001) of nymphal instars (III, IV, and V). Aphis craccivora (48.66/plant), Spodoptera litura (0.40/plant), Chrotogonus trachypterus (0.06/plant), A. crenulata (0.04/plant), Mylabris pustulata (0.70/ plant), and Helicoverpa armigera (0.15/plant) were all present in the highest concentrations. After the predators were released, their populations drastically decreased, especially that of Aphis craccivora (24.28). The populations of Helicoverpa armigera and Spodoptera litura both saw significant declines. However, despite the fact that it decreased the populations of C. trachypterus and A. crenulata, these changes were not statistically significant at the 5% level and had no effect on Mylabris pustulata (0.77 for the control- and reduviid-released fields, respectively). Rhynocoris kumarii decreased the populations of S. litura, C. trachypterus, A. crenulata, Mylabris pustulata, Helicoverpa armigera, and A. craccivora by 66.66, 40, 25, 0, 56.25, and 51.59%, respectively. Two pods were higher in both plots than one pod (10.32 in the control and 18.83 in the experimental), and the pod yield was higher in the predator-released plot (3.0 in the control and 1.94 in the experimental, respectively). Additionally, the experimental plot produced more groundnuts (1655.50 kgha′) than did the control plot (1041 kgha′), but the difference was statistically insignificant. The experimental plot likewise had a high cost–benefit ratio (1:2.09).

18.3.3 Anthocoridae Frankliniella occidentalis, a thrips, may be controlled by predatory bugs Orius insidiosus and Orius albidipennis. Orius albidipennis has a shorter life cycle than does Orius insidiosus; however, laboratory tests have shown that it lays much more eggs than the latter. Orius albidipennis is an anthocorid that may be regarded as the more effective biocontrol agent since it produces more eggs at the beginning of the oviposition cycle (van De Veire and Degheele 1995). When released simultaneously on four ornamental plants such Saintpaulia, Impatiens, Gerbera, and Brachyscome

18.3

Hemipterans

677

multifida infested with F. occidentalis in greenhouses, the predators Amblyseius (=Neoseiulus) cucumeris and O. insidiosus were particularly successful (Sörensson and Nedstam 1993). The ability of the two species of Chrysoperla carnea and Chrysoperla rufilabris to prey on Scirtothrips citri, a citrus thrips, was assessed. Eggs were taped to the tree branches, and larvae were deposited one at a time using a brush on the outside of citrus leaves. Citrus thrips considerably reduced the amount of fruit scarring in 6 of the 11 releases compared to the levels seen in the untreated control. By days 13–14 after release, thrips numbers (immature + adults) had decreased. The number of predatory mites on trees treated with various treatments showed no difference. Therefore, increasing Chrysoperla spp. releases in citrus orchards may help reduce citrus thrips populations (Khan and Morse 1999). In order to assess Orius strigicollis’ potential as a biological control agent, fieldrelease tests on eggplant and adzuki beans grown in open fields in central and southern Taiwan were carried out in 1998 and 1999. The densities of thrips gradually decreased in the regions where it was released. The number of thrips in the chemical control regions decreased right away after insecticides were applied, but they quickly rebounded. Four to six weeks after the first release, there were noticeable variations between the biocontrol regions and the chemical control areas. The biocontrol’s effects persisted for several weeks. When deployed at the right time, Orius strigicollis has proven to be a potent natural enemy for the management of thrips on specific crops in open fields (Wang et al. 2001a, b). Orius spp. (Heteroptera: Anthocoridae) augmentative releases are the main method of biological control of thrips (van Lenteren 2000; Reitz 2009). To reduce thrips management, Orius strigicollis is released in a supplemental manner (Wang et al. 2001a, b). In India, the black-headed caterpillar, Opisina arenosella, is a significant coconut pest. Opisina arenosella eggs and newly hatched larvae may be preyed upon by Cardiastethus exiguus. In two pest-infested locations of Kerala, India, during the summers of 2003–04 and 2004–05, Cardiastethus exiguus was released into the wild and evaluated there. In the crown portion of each tree, between 50 and 100 nymphs/ adults were released. There were three releases altogether. After the predator was released, the population of Opisina arenosella significantly decreased. Both release rates were shown to be equally effective at reducing the Opisina arenosella population (Lyla et al. 2006). In semi-field and field tests, Sigsgaard et al. (2006) examined the potential for control of the pear psyllid, Cacopsylla pyri, in pear orchards by Anthocoris nemoralis and Anthocoris nemorum. After 2 days of an experimental infestation, one adult Anthocoris nemorum or Anthocoris nemoralis female reduced the number of C. pyri eggs and young nymphs in sleeve cages by nearly a third and by half, respectively. Two A. nemorum, two Anthocoris nemoralis, or both together reduced Cacopsylla pyri eggs and nymphs by 72–90% over the course of 2 weeks in a natural infestation. In a modest field study, 5–6% of Cacopsylla pyri introduced stayed on branches where Anthocoris nemoralis eggs had been placed, compared to twice as many on untreated branches, indicating not only the strong impact of anthocorids but also their significant natural mortality. There was one exception: In the orchard with the highest initial psyllid infestation, the treatment with releases of only 10 Anthocoris nemoralis nymphs was not statistically different from the

678

18

Controlled Field Cage and Field Evolution

control. Field releases of 10 or 30 Anthocoris nemoralis, Cacopsylla pyri significantly decreased the number of pests. Field releases offer hope for creating a workable release technique (Sigsgaard et al. 2006). Murata et al. (2007) examined the effectiveness of the reduviid bug Amphibolus venator and the warehouse pirate bug Xylocoris flavipes against Tribolium confusum, which infests wheat flour. Tribolium confusum was suppressed by X. flavipes alone in 96.9% of cases, A. venator by itself in 76.2% of cases, and both bugs combined in 95.6% of cases. The fact that Amphibolus venator targeted X. flavipes adults but not nymphs suggests that simultaneous releases may be designed to control various storage pests. In their 2010 study, Hosseini et al. (2010) examined the impact of plant quality on the intraguild predation of greenhouse cucumber Aphidoletes aphidimyza (Diptera: Cecidomyiidae) and Orius laevigatus using A. gossypii as shared prey (with different nitrogen (N) fertilisation levels). Orius majusculus alone outperformed A. aphidimyza alone in terms of aphid control, regardless of N fertilisation levels. However, they came to the conclusion that the poor intraguild predation between the two predators did not have an impact on their predatory effectiveness. In France and Spain, the combined use of Neoseiulus spp. and Orius sp. effectively reduced Frankliniella occidentalis in protected strawberry crops (Sampson and Kirk 2016). The use of Orius majusculus in conjunction with predatory midges, predatory thrips, and parasitoids significantly improved the suppression of aphids and thrips infesting sweet pepper, supporting the idea that intraguild predation, which may be detrimental to biocontrol, may be offset by the beneficial effects of generalist predators for control of a variety of pests (Messelink et al. 2013). In greenhouse-grown sweet pepper, a combination of O. laevigatus and Macrolophus pygmaeus releases seemed to be the best method for thrips and aphid control (Messelink and Janssen 2014). Combining O. laevigatus and the predatory mite Amblyseius swirskii releases in Tunisia helped control the Frankliniella occidentalis infestation of greenhouse pepper (Elimem and Chermiti 2012). When compared to homes where no predators were released, Wong and Frank (2012) discovered that augmentative releases of Orius insidiosus significantly decreased thrips abundance in hoop houses. On cut roses, Rosa hybrida, Frankliniella occidentalis was tested for its role in Orius insidiosus. Female Orius insidiosus killed more prey than did males, although neither sex showed preference for A. swirskii over juvenile or adult Frankliniella occidentalis. We evaluated the control of Frankliniella occidentalis on roses with releases of Orius insidiosus in greenhouse studies that mimicked thrips infestations of cut rose crops during commercial production. Similar amounts of harvestable flowers were generated by pest-infested and predator-free roses, although the latter had, on average, two to three times as many thrips as the former. Because counts of thrips and predatory mites on flowers with both Orius insidiosus and A. swirskii were not statistically different from similar counts on flowers with only A. swirskii, concurrent releases of Orius insidiosus and A. swirskii did not appear to interfere with or enhance suppression of Frankliniella occidentalis on cut roses (Chow et al. 2010).

18.3

Hemipterans

679

In their review article in 2016, Ballal and Yamada noted the following. Many of these predators were commercially available in nations like the United Kingdom, Canada, the Netherlands, the United States, Israel, Italy, Germany, France, Poland, etc., which has led to their extensive utility against pests of field crops and polyhouses. Anthocorid predators are used in a variety of countries for pest management. In Japan, China, and Korea, Orius sauteri has been successful in controlling thrips palmi in eggplants. Orius insidiosus, Orius laevigatus, Orius majusculus, and Orius strigicollis have been introduced in Poland to fight greenhouse pests that affect vegetable crops. Psylla pyri has been effectively managed in pear orchards in Venice, Italy, using Anthocoris nemoralis, which has been released at 200–300 anthocorids/ha two to three times each year. The Philippines introduced Montandoniola moraguesi to Hawaii and Bermuda for the traditional biological control of the Ficus leaf gall thrips, Gynaikothrips ficorum. The species from Guadeloupe, the West Indies, Hawaii, Bermuda, Australia, and Florida appeared to be Montandoniola confusa, suggesting that the species originally introduced from the Philippines may have been Montandoniola confusa and not Montandoniola moraguesi. This anthocorid may only be found in the Mediterranean region and Africa. India sent Tetraphleps raoi to Kenya for release against Pineus pini on various Pinus spp. This led to the establishment of the Pineus pini population and its subsequent decline (Ballal and Yamada 2016). The potential to manage silverleaf whitefly (SLW) and Bemisia tabaci infestations and their consequences on tomato plant health and yield at both the early phases of the crop cycle and at high pest densities using two indigenous omnivorous biological agents, Dicyphus hesperus and Orius insidiosus, was studied. This study demonstrated the ability of Bemisia tabaci populations to expand quickly. The incidence of tomato irregular ripening (TIR) disorder and yield, however, was not significantly impacted by either the introduced Bemisia tabaci densities or the time of infestation. Our findings also demonstrate that while Orius insidiosus did not considerably lower SLW populations, the addition of three to five Dicyphus hesperus adults per cage reduces the quantity of Bemisia tabaci larvae and pseudonymphs. Our findings indicate that early in the crop cycle, when SLW levels are high, Dicyphus hesperus may be able to control them (Dumont et al. 2021).

18.3.4 Pentatomidae Native to North America, the spined soldier bug, Podisus maculiventris, is a generalist predator whose primary target includes Coleoptera and Lepidoptera larvae (McPherson 1982). Under controlled laboratory conditions, this predator has been successfully used to control pests in cotton (Gossypium hirsutum) and tomato (Solanum lycopersicum) plants as well as in greenhouses (Ables and McCommas Jr 1982; De Clercq et al. 1998). In a US greenhouse with cotton that was already naturally infested with Peridroma saucia, adults of Podisus maculiventris, a major predator of lepidopterous larvae on many crops, were released, and, within 48 h, 75% of the cutworm

680

18

Controlled Field Cage and Field Evolution

larvae had been devoured (Ables and McCommas Jr 1982). The field release in a castor plantation in the north of Thailand and the subsequent success in suppressing the castor semilooper, Achaea janata, were the key steps in the application of the pentatomid predator, E. furcellata, for augmentative biological control of insect pests. In agricultural ecosystems, predatory stink bugs (Hemiptera: Pentatomidae: Asopinae) are frequently utilised in augmentative releases to manage pests (HoughGoldstein and McPherson 1996; Tipping et al. 1999). In small plots of staked tomatoes planted in plastic or hairy vetch mulch, multiple releases of Podisus maculiventris (Heteroptera: Pentatomidae) led to higher predation rates of Colorado potato beetle (CPB) eggs, Leptinotarsa decemlineata. Over the course of the study’s years, no consistent impact of the mulch on CPB infestations was discovered. Edovum puttleri did not cause any parasitism in 1995, but, in 1997, it gave rise to parasitism in 0–30.6% per sample date. Podisus maculiventris nymphs were more numerous in release plots and attacked CPB egg masses with no hesitation. The cost of maintaining release rates that were high enough to achieve considerable control of the pest would be prohibitive on a wider scale, notwithstanding the ability of both beneficials to explore the structurally complicated plant and discover the pest (Tipping et al. 1999). According to Hough-Goldstein and McPherson (1996), augmentative releases of Podisus maculiventris in little field plots decreased the population of Leptinotarsa decemlineata, the Colorado potato beetle’s larvae. Spined soldier insect, Podisus maculiventris, consumed as least as many Colorado potato beetle, Leptinotarsa decemlineata, eggs, and larvae as did the two-spotted stink bug, Perillus bioculatus, in laboratory and microplot eating tests. Similar to Podisus bioculatus, Podisus maculiventris persisted in field microplots contaminated with Colorado potato beetle egg masses for a little while longer. In tiny field plots, the two predators were equally successful at devouring egg masses, lowering beetle larval populations and preventing potato defoliation. These outcomes are in contrast to earlier studies, in which Podisus bioculatus had demonstrated more efficacy. The relative efficiency of the two predators may rely on the prey life stages that are present at the time of the predator release since in laboratory experiments, larger Podisus maculiventris nymphs chose larger prey larvae but Podisus bioculatus nymphs did not (HoughGoldstein and McPherson 1996). Podisus maculiventris has been seen preying on Microtheca ochroloma in central Florida. Regarding its ability to prey on M. ochroloma field populations, however, little is known. Podisus maculiventris arrives in northern Florida in March and begins to overwinter there in October (Herrick and Reitz 2004). The predation ability of Podisus maculiventris was tested in the United States from February to March of 2009 and again in February to March of 2010 in field cages containing larvae of the yellow-margined leaf beetle, Microtheca ochroloma (Coleoptera: Chrysomelidae), a pest of organic crucifer (Brassicaceae) crops. Six turnip (Brassica rapa) plants in cages received an equal distribution of 4 (=low), 10 (=medium), or 16 (=high) P. maculiventris first instars on the same day that 132 M. ochroloma’s first instars were put into the cages. No predators were placed in the control treatment cages. During the first four sample dates in 2009, there was a

18.4

Neuroptera

681

steady decline in the mean number of M. ochroloma larvae. The mean total M. ochroloma counts in the two treatments with higher release rates were considerably lower on the seventh and final sample day than in the low release-rate and control treatments. Overall, 39.1% of Podisus maculiventris survivors received the high release-rate therapy. During the first four sample dates in 2010, there was a steady decline in the mean number of M. ochroloma larvae. There were no appreciable variations in the predator release treatments on the fourth sample date or on the final (ninth) sampling date. In the high release-rate therapy, the overall survival rate of Podisus maculiventris (3.1%) was the lowest. This study produced two possible recommendations for growers: If the plants are anticipated to have 7 leaves per plant, release 10 first instars of Podisus maculiventris per 6 plants; otherwise, release 4 first instars of Podisus maculiventris per 6 plants (Montemayor and Cave 2012).

18.4

Neuroptera

The responses of Chrysoperla spp. and biotypes to physical and biotic elements in their habitat, habitat preference, adult cryptic colouring, seasonal cycles, and presumably prey preferences all vary significantly. The releasing places for lacewings also differ significantly (they range from cotton fields in Texas and apple orchards in Washington to greenhouses in a variety of locations). Unfortunately, while formulating release strategies, the inherent heterogeneity among the Chrysoperla taxa and the variations in geographic regions, agroecosystems, or environmental circumstances have rarely been taken into account; as a result, the effectiveness of releases varies substantially (Daane et al. 1998). However, a few studies that took into account these problems produced suggestions for pairing species and biotypes with particular pest management circumstances, and some of these suggestions have improved the insectary business (Tauber and Tauber 1993). Many field investigations have demonstrated that lacewings can lower target population sizes. However, with regard to commercial releasing techniques, it is untrue. For instance, Ehler et al. (1997) hypothesised that low efficiency was caused by the poor quality of the insectary-reared stock or the introduction of the incorrect lacewing species when lacewings had no effect on the bean aphid, Aphis fabae, in sugar beets. This is similar to how different release methods and speeds can have a significant impact on lacewing effectiveness in field tests (Daane and Yokota 1997). For instance, Chrysoperla carnea has been distributed at rates ranging from 7500 to 2,00,000 eggs per hectare using a variety of techniques, including the careful release of larvae by hand and mechanical distribution of eggs (Daane et al. 1998). Although it can be anticipated that this volatility will affect field trial results, evaluation tests rarely adjust for or measure it. In small-plot and on-farm trials, the efficacy of the inundated release of the common green lacewing Chrysoperla carnea to control two vineyard pests, Erythroneura variabilis and the Western grape leafhopper Erythroneura elegantula, was investigated. In quarter-vine cages, Chrysoperla carnea larvae were manually

682

18

Controlled Field Cage and Field Evolution

Table 18.2 Erythroneura variabilis and Erythroneura elegantula decrease rates (%) in response to the release of Chrysoperla carnea at various densities Year 1900

Predator Chrysoperla carnea

Release rate / hectare 29,652

88,956

1991

Chrysoperla carnea

9884

19,768

Pests Erythroneura variabilis Erythroneura elegantula Erythroneura variabilis Erythroneura elegantula Erythroneura variabilis Erythroneura elegantula Erythroneura variabilis Erythroneura elegantula

Reduction (%) 23.5

30.3

33.6

31.4

released in 1900. In comparison to no-release experiments, leafhopper numbers were considerably decreased by 23.5 and 30.3% in plots that received 29,652 and 88,956 C. carnea larvae per hectare, respectively. In uncaged, three-vine plots in 1991, Chrysoperla carnea larvae were manually released onto the vines; the two release rates employed nearly matched industry recommendations. The Chrysoperla carnea-release and control plots did not vary significantly at the lower rate (9884 larvae per hectare). Leafhopper numbers were dramatically decreased by 33.6 and 31.4% in the first and second leafhopper generations, respectively, at an increased rate (19,768 larvae per hectare). The three commercially available lacewing species, Chrysoperla carnea, Chrysoperla comanche, and Chrysoperla rufilabris, were evaluated in 1992 using three-vine plots once more at a release rate of 19,768 larvae per hectare. The Chrysoperla rufilabris release treatment was the only discernible decrease in leafhopper numbers (Table 18.2). Large, replicated Chrysoperla carnea release and no-release plots were used in on-farm studies from 1990 to 1993 to evaluate the efficacy of commercial release programmes. Between 7413 and 37,065 eggs per hectare, every leafhopper generation of Chrysoperla carnea was released. Leafhopper densities were considerably lower in plots that had Chrysoperla carnea released than in unreleased plots in 9 out of 20 experiments. To find potential explanations for the variance in Chrysoperla carnea releases’ efficacy, data from all experiments were merged. Different release trials, rates, and techniques, as well as variations in prey density, are all possibilities. The average reduction in leafhoppers in Chrysoperla carnea-release plots was 29.5% in cages, 15.5% in three-vine plots, and 9.6% in commercial vineyards when compared to no-release plots. There was a strong, albeit sluggishly positive, association between release rate and efficiency. When lacewings were released as larvae rather than eggs, there was also a higher decrease in leafhopper nymphs. Combining data from all investigations, it was found that leafhopper density was correlated with the quantity and proportion of reduced leafhopper nymphs. The decrease in leafhopper population was frequently insufficient to bring leafhopper densities below the economic

18.4

Neuroptera

683

harm threshold when they were over the level that was proposed to cause economic damage (15–20 nymphs/leaf) (Daane et al. 1996). However, it is essential to evaluate the recommendations in actual use. Although Chrysoperla rufilabris does not thrive in dry environments, it is typically the best option for usage in greenhouses or other moist environments. Chrysoperla carnea, in comparison, thrives in dry environments and should be employed there. These suggestions were successfully included into the sales promotions of a few US insectaria, and the species are now properly marketed. To assess the costeffectiveness of the proposals, follow-up studies (comparative quantitative field tests of the two species) are currently required. The recommendations for release should also take into account current research on the death of Chrysoperla carnea eggs at high temperatures (>37 °C) (Daane and Yokota 1997), which should be tested in the field. In the second instance, research on the differences in seasonal responses and habitat preferences among the various Chrysoperla carnea species or populations resulted in provisional suggestions for pairing particular pest management contexts, like cropping systems, with particular biotypes. For instance, the light green Chrysoperla carnea from eastern United States is advised for use in field crops or vineyards, whereas the dark green Chrysoperla downesi is suggested for use in evergreen trees. These outstanding examples highlight the need to match the predator’s biological features to the crop and the environment’s physical parameters. They also demonstrate the need for comparative research in order to create lacewing usage guidelines that are specific to each species. In comparative quantitative investigations, these preferences should be clearly characterised, and the distinctions between species and biotypes should be made clearer. Additionally, generalisations about lacewing susceptibility may not be appropriate. At the University of California Lindcove Research and Extension Center near Visalia, California, in the spring of 1995, the predatory impact of two species of Chrysoperla (Chrysoperla Carnea and Chrysoperla rufilabris) against citrus thrips, Scirtothrips citri, populations was assessed. Eggs were taped to the tree branches, and larvae were deposited one at a time using a brush on the outside of citrus leaves. Citrus thrips considerably reduced the amount of fruit scarring in 6 of the 11 releases compared to the levels seen in the untreated control. By days 13–14 after release, thrips numbers (immature + adults) had decreased. The number of predatory mites on trees treated with various treatments showed no difference. Therefore, increasing Chrysoperla spp. releases in citrus orchards may help reduce citrus thrips populations (Khan and Morse 1999). In experimental field settings, Micromus igorotus (Neuroptera: Hemerobiidae), when introduced at a density of 800 adults/ha, failed to replace the dominant Dipha aphidivora (Lepidoptera: Pyralidae) after 30 days. Dipha aphidivora increased its natural population and decreased aphid intensity in a series of experiments, whereas Micromus sp. and unidentified syrphids showed little action. The study showed that Dipha aphidivora (Hemiptera: Aphididae) was the primary predator of Ceratovacuna lanigera and that its augmentative releases were effective in lowering the aphid population during the years of invasion at the study location (Srikanth et al. 2015).

684

18

Controlled Field Cage and Field Evolution

It is suggested by both biocontrol theory and biocontrol practise that generalist predators can function as BCAs in IPM. The green lacewing, Chrysopa pallens (Hemiptera: Chrysopidae), has been regarded as a valuable generalist predator for the biological management of pests in forestry and agriculture. It consumes several pests, including aphids, whiteflies, mites, and lepidopteran larvae, and is carnivorous both as an adult and as a larva. Throughout each of the experimental periods, changes in the population of adult and larval Frankliniella occidentalis that fed on glasshouse cucumber plants took place in various treatments. The initial densities of adult and larval Frankliniella occidentalis were established at the start of the experiment (12 October 2017) and did not significantly differ between replicates. This demonstrated that the population of Frankliniella occidentalis was widespread and well-established on cucumber crops. The population trend of larval Frankliniella occidentalis on the cucumber plants in the untreated control (CK) declined somewhat on the second sampling date, increased quickly until the fourth sampling date, and then decreased once again on the final sampling day. Frankliniella occidentalis adult population trends exhibited traits resembling those of the larvae in the untreated control. By the final sample date, the various densities of Chrysopa pallens larvae significantly decreased Frankliniella occidentalis densities in comparison to the control treatment. Generally speaking, the impact of Frankliniella occidentalis, both larval and adult, on glasshouse cucumbers rose according to the initial density of the predatory larvae. For beginning densities of 2, 4, 8, and 16 lacewings, the density reduction of the Frankliniella occidentalis population at the conclusion of the sampling was 11, 39, 59, and 68% for the larvae and 12, 43, 58, and 68% for the adults, respectively (Sarkar et al. 2019).

18.5

Coleoptera

The establishment rate for the aphidophagous species was 30.8–40.0% and that for the coccidophagous species was 17.6–74.1%, respectively, compared to the most commonly released species (with 10 or more introduction episodes). Among the ladybirds released, Rodolia cardinalis is regarded as the most successful due to its complete control over Icerya purchasi in a number of nations and over Icerya aegyptiaca (Hemiptera: Monophlebidae) in areas of Africa (reviewed by Cock et al. 2010). The Australian Cryptolaemus montrouzieri was initially employed in Italy in 1908 to manage the mealybug Planococcus citri (Hemiptera: Pseudococcidae), and it is now regarded as established in nearly all nations where it was purposefully imported to combat mealybugs (Kairo et al. 2013). According to Hodek and Evans (2012), Clitostethus oculatus specifically feeds on aleyrodids and significantly helped eradicate Aleurodicus dispersus (Hemiptera: Aleyrodidae) in American Samoa, Hawaii (USA), Guam, and Fiji. It is noted that significant control was boosted when it was combined with Encarsia spp. parasitoids (Tables 18.3). Both Rhyzobius lophanthae and the majority of its prey species were born in Australia (reviewed by Roy and Migeon (2010)). Rhyzobius lophanthae successfully established itself after being introduced to a new place, and it is currently widespread

18.5

Coleoptera

685

Table 18.3 The target family, the most often introduced species of ladybird, and a summary of success rates. If a BCA was introduced in 1 nation or region for 10 years to control a specific pest, then it is considered an introduction event

Ladybird species Cryptolaemus montrouzieri

Rodolia cardinalis Cryptognatha nodiceps Rhyzobius lophanthae Diomus hennesseyi Pseudoazya trinitatis Hippodamia convergens Chilocorus bipustulatus Chilocorus nigritus Chilocorus cacti Coccinella septempunctata Curinus coeruleus

Clitostethus oculatus Halmus chalybeus

Hyperaspis notata Chilocorus kuwanae Harmonia axyridis

Target family Coccidae, Diaspididae, Monophlebidae, Pseudococcidae Monophlebidae, Pseudococcidae Diaspididae

No. introductions 73

No. establishments (relative %) 46 (63)

No. partial, substantial, or complete successes reported (relative %) 26 (35.6)

58

43 (74.1)

35 (60.3)

23

6 (26.1)

3 (13.0) 4 (20.0)

Coccoidea, Diaspididae Pseudococcidae

20

11 (55.0)

17

3 (17.6)

0 (0)

Diaspididae

17

4 (23.5)

1 (5.9)

Aphididae, Monophlebidae Diaspididae

16

5 (31.3)

4 (25.0)

15

6 (40.0)

3 (20.0)

15

5 (33.3)

1 (6.7)

13

6 (46.2)

2 (15.4)

13

4 (30.8)

0 (0)

13

9 (69.2)

0 (0)

12

5 (41.7)

4 (33.3)

12

3 (25)

0 (0)

12

5 (41.7)

0 (0)

11

2 (18.2)

1 (9.1)

10

4 (40.0)

0 (0)

Coccoidea, Diaspididae Asterolecaniidae, Diaspididae Adelgidae, Aphididae Asterolecaniidae, Diaspididae, Pseudococcidae, Psyllidae Aleyrodidae Coccidae, Coccoidea, Diaspididae, Eriococcidae, Pseudococcidae Pseudococcidae Adelgidae, Coccoidea, Diaspididae

(continued)

686

18

Controlled Field Cage and Field Evolution

Table 18.3 (continued)

Ladybird species

Rodolia pumila Other species (eight or less releases) Total

Target family Aphididae, Monophlebidae, Psyllidae Monophlebidae

No. introductions

No. establishments (relative %)

No. partial, substantial, or complete successes reported (relative %)

10 462

7 (70.0) 96 (20.8)

5 (50.0) 31 (6.7)

822

270 (32.8)

120 (14.6)

Reference: Rondoni et al. (2021)

in several nations worldwide (Pellizzari and Porcelli 2014). Although the first releases were unsuccessful or only partially successful, the ladybird eventually provided complete control of Aulacaspis tegalensis on sugarcane in Tanzania and significant control of Aonidiella aurantii (all Hemiptera: Diaspididae) on citrus orchards in South Africa. It also significantly helped suppress Diaspidiotus perniciosus and Pseudaulacaspis pentagona in Europe (the introduction of Rodolia pumila Weise in the Pacific Islands has effectively reduced the prevalence of Icerya aegyptiaca on fruit trees; reviewed by Clausen (1978)). As a pest-predator of aphids, the ladybeetle Harmonia axyridis (Coleoptera: Coccinellidae) is dispersed throughout North America and North-west Europe (Koch and Galvan 2008). The field (vegetable fields planted with corn, eggplant, and pepper) evaluation for the release of the predator Coccinella undecimpunctata was the best in lowering the number of aphid’s insects (Aphis fabae) to 57 insects after 1 day from releasing predators (Ammar et al. 2013). Aphis gossypii was greatly reduced (>90%) in India following the inoculative release of 30, 40, or 50 adults of Harmonia sedecimnotata per 100 m2. As a result, it is advised to release 40 adults per 100 m2 to control the eggplant aphid population. Therefore, Harmonia sedecimnotata is among the most promising biological control agents for cotton aphids that can be used for immediate control by the release of adults for inoculation (Boopathi et al. 2020). Coccinella undecimpunctata was released in New Zealand in 1874 to manage aphids and mealybugs, marking the first known introduction of an alien ladybird. The cottony cushion scale, Icerya purchasi, was later successfully controlled by the introduction of a different species, Rodolia cardinalis, in the United States and in Europe (Hemiptera: Monophlebidae). In order to combat insect pests, including coccids, aphids, psyllids, and whiteflies since then, exotic ladybirds have been dispersed in new places (reviewed by Rondoni et al. 2021).

18.7

18.6

Syrphids

687

Dermaptera

The literature on dermapterans has limited field evaluation studies. Another reason why its population is not highly concentrated in any crop habitat is because mass production technology has not been adopted. In eastern Uganda, Aphis craccivora (Homoptera: Aphididae) is a commercially important pest of cowpea. In eastern and central Africa, farms are frequently visited by the predatory earwig Forficula auricularia (Dermaptera: Forficulidae). During the rainy seasons of 2004 and 2005, laboratory and field cage tests with different predator–prey ratios were set up in Uganda to test Forficula auricularia’s capacity to prey on Aphis craccivora. Forficula auricularia ingested 1.27–7.82 Aphis craccivora on average per day in the lab, compared to 3.66–7.24 in the field. Although this earwig did not exhibit high predation levels, their populations ought to be promoted because they help the environment regulate insect problems (Manyuli et al. 2008). According to Ramos et al. (2019), the initial damage rating of 2, which is equivalent to between 21% and 40% of the damage caused by Brontispa longissima, climbed to between 41% and 60% in the control treatment. Brontispa longissima was evaluated based on the average 1.5 in the treatment with Chelisoches morio (Dermaptera: Chelisochidae) predator release, which is a reduction of around 40% from the original harm. Damage in the control group, however, increased by around 20% of the initial damage.

18.7

Syrphids

18.7.1 Effectiveness of Syrphids as Field-Based Biocontrol Agents According to Chambers (1988), effectiveness is measured by how much aphid population growth is slowed down, maintained steady, or reversed by predatory activity (syrphids) as well as whether this keeps aphid population density below the threshold that causes economic damage. He described two approaches—qualitative and quantitative—that are typically employed for predator evaluation. While manipulating the initial prey and predator ratios and observing the outcome or estimating the abundance of the predator are the methods used in quantitative analysis, qualitative methods include checking aphid colonies for the presence of predators, excluding or including predators, and observational methods with graphical analysis. For instance, hoverfly oviposition on colonies of Diaphorina citri (Hemiptera: Liviidae) was examined in field studies, as were the phenology of flowering and its attractiveness to hoverflies Allograpta obliqua (Diptera: Syrphidae) and other natural enemies and the effect of potted alyssum and hoverfly predation on the mortality of Diaphorina citri nymphs (Irvin et al. 2021).

688

18.8

18

Controlled Field Cage and Field Evolution

Predator Density

The number of predatory arthropods detected in plots that were released and those that were not released was equal. For instance, 3.1 preying mantids (Mantis religiosa), 2.9 coccinellid predators (Menochilus sexmaculatus), 2.16 heteropteran predators (Onus sp., Geocoris sp., Cantheconidia sp., and Rhynocoris fuscipes), and 1.9 spiders were recorded. In all, 1.5 plants, 2.8 coccinellids, 2.24 heteropterans, 2.5 mantids, and 2.59 spiders were found in the control field! The test field plots contained five plants. In tagged plants of predator-released plots compared to those of predator-non-released plots, the yield of seed cotton was not statistically different (test X = 141.1 g and control X = 138.75 g/5 plants). Once more, the percentage of high-quality cotton in released plots was higher (63.1) than that in non-released control plots (61.0%) (Ravichandran 2004).

18.9

Comparison Between Organic and Conventional Fields

We discovered that for both Nabis and Geocoris, the probability of finding potato beetle DNA in organic compared to conventional fields was much higher in a straightforward comparison of the two farming methods without taking any arthropod community features into account. We observed no appreciable difference in detection in organic vs. conventional potato fields under more complicated models for Nabis predation on potato beetles. However, two models with strong support contained interactions between an agricultural system and a predator community attribute. Overall, Geocoris from organic fields were more likely than those from conventional fields to carry potato beetle DNA. Despite the fact that no wellsupported models included potato bug abundance, this is consistent with generally greater potato beetle densities under organic management (Lynch et al. 2021).

18.10 Genetically Modified (GM) Crops A significant predatory insect in cotton fields, the big-eyed bug Geocoris pallidipennis, can feed on a wide range of pests, including aphids, leafhoppers, cotton bollworm Helicoverpa armigera, and lepidopteran larvae. By preying on herbivorous insects that consume cotton, Geocoris pallidipennis, a significant natural enemy insect in cotton fields, is able to consume the Bacillus thuringiensis (Bt) protein generated in GM cotton. However, it is yet uncertain if GM cotton poses a threat to Geocoris pallidipennis. Here, we assessed how Cry1Ac/1Ab protein expression in Bt cotton affected Geocoris pallidipennis nymphs and adults. In addition to being found in the midgut of Geocoris pallidipennis nymphs and adults that feed on Bt-fed H. armigera, Cry1Ac protein was also found in the midgut of the cotton bollworm, Helicoverpa armigera, after it consumed Bt cotton. However, Geocoris pallidipennis survival rate, growth, development, and fertility were not negatively impacted. Additionally, the genes involved in immunological function,

18.10

Genetically Modified (GM) Crops

689

detoxification, antioxidant activity, and nutritional utilisation did not alter in expression in response to Cry1Ac exposure. Finally, we demonstrated that neither Geocoris pallidipennis nymphs nor adults could bind Cry1Ac to the proteins found on brush border membrane vesicles (BBMVs). In conclusion, these findings show that the insect predator Geocoris pallidipennis is barely affected by transgenic Cry1Ac/1Ab cotton. Using a tritrophic system made up of the crop plant Brassica napus, the pest mollusc Deroceras reticulatum, and the predatory carabid beetle Pterostichus melanarius, the potential effects of chemical pesticide control methods have been compared with those of transgenic plants expressing a protease inhibitor conferring insect resistance. As the most extensively used pesticide in UK oilseed rape (OSR) agriculture, cypermethrin was chosen as the standard therapy. The transgenic comparator was OSR, which expressed oryzacystatin-1 (OC-1), a cysteine protease inhibitor. Deroceras reticulatum in feeding trials did not exhibit any appreciable long-term effects from exposure to either the cypermethrin or OC-1 treatment on assessed life history parameters (survival, weight gain, or food consumption). However, after being exposed to OSR that expressed OC-1, Deroceras reticulatum was able to react to the dietary inhibitor by generating two novel proteases. Similar to Pterostichus melanarius, Deroceras reticulatum previously fed on either pesticide-contaminated or GM plant material did not exhibit any discernible changes in mortality, weight increase, or food consumption. Additionally, similar to the slug, a new protease of about Mr. 27 kDa was activated in the carabid after it consumed slugs fed OC-1-expressing OSR (Mulligan et al. 2006). With references, Table 18.4 displays the effects of several proteins used in transgenic plants on predators. Romeis et al. (2006) analysed laboratory, greenhouse, and field research that looked at how Bacillus thuringiensis toxin-expressing transgenic crops affected parasitoids and arthropod predators. They come to the conclusion that there were no overt toxic effects and that adverse reactions only happened when herbivores, which were sub-lethally harmed by Bt, were exploited as hosts or as prey. According to several reviews, Bt cotton has no effect on the world cotton’s beneficial insect ecosystems (Naranjo 2005a, b; Whitehouse et al. 2005). A variety of consequences on communities of ground-dwelling predators may result from altered population dynamics of pests targeted by the Cry1Ac toxin in Bacillus thuringiensis (Bt) transgenic cotton (Bt cotton) and possible reduced insecticide use in these transgenic cultivars. In commercial Bt and non-Bt cotton fields, a weekly survey of ground-dwelling arthropods was conducted each of the 3 years (during the cropping season). In the survey, 65 taxa of ground-dwelling arthropods were identified, which are crucial for controlling cotton pests, including carabids, cicindelines, staphylinids, dermapterans, heteropterans, and araneids. The analysis found no distinctions between the various cotton kinds in the populations of grounddwelling arachnids. Megacephala carolina made up 97% of cicindelines, 96% of all dermapterans, and 80% of carabid species, along with Selenophorus palliatus, Apristus latens, Harpalus gravis, and Anisodactylus merula. In each of the 3 years, Megacephala carolina outnumbered all other species that were collected. Anisodactylus merula, Calosoma sayi, Harpalus pennsylvanicus, and Stenolophus

Transgenic plant Maize (Zea mays)

Maize

Maize

Maize

Maize

Maize

Maize

Cotton (G. hirsutum)

Cotton

Bt protein Cry1Ab

Cry3Bb1

Cry1Ab

Cry1Ab

Cry1Ab

Cry3Bb1

VIP3A + Cry1Ab

Cy1Ac

Cry1Ac

Several

Aphis gossypii (hem. Aphididae)

Lepidopterous pests

Coleomegilla maculata (col: Coccinellidae) Several (13 arthropod orders) Chrysopa pallens (Neu: Chrysopidae) Predators (several)

Stethorus punctillum

C. carnea

Predators Coleoptera, Heteroptera, Neuroptera Araneae, Carabidae, Staphylinidae C. carnea

Slightly reduce predator

NE

No negative effects of stacked traits over conventional corn

No effects when fed with aphid prey

No effects

Naranjo (2005a, b)

Guo et al. (2008)

Dively (2005)

Dutton et al. (2002) AlvarezAlfageme et al. (2008) Lundgren and Wiedenmann (2005)

Hilbeck et al. (1998)

Bt-fed prey increased predator mortality and development times Effects, reduced prey quality

Bhatti et al. (2005)

References Pilcher et al. (1997)

No consistent negative effect

Impact No effects on predators in both laboratory and field experiments

18

Rhopalosiphum maidis (Hom: Aphididae)

Tetranychus urticae (Acari)

O. nubilalis; Spodoptera littoralis (Lepidoptera: Noctuidae) S. littoralis

Diabrotica spp. (col: Chrysomelidae)

Pest Feeding on pollen

Table 18.4 Impacts of transgenic crops and transgene products of predators

690 Controlled Field Cage and Field Evolution

Cotton

Cotton

Cotton

Potato (S. tuberosum) Potato

Potato

Potato

Potato

Injected prey

Potato Strawberry (Fragaria sp.)

Cry1Ac

Cry1Ac

Cry1Ac

Bt (Cry3Aa)

Bt (Cry3)

Bt (Cry3A)

Bt (Cry3 A)

CpTI

CpTI CpTI

Bt (Cry3Aa)

Cotton

Cry1Ac/Cry2Ab

L. oleracea Otiorhynchus sulcatus (Curculionidae)

L. oleracea

L. oleracea

Lepidoptera: Hem.

M. persicae

L. decemlineata

L. decemlineata

Lepidopterans

Spodoptera exigua, Helicoverpa zea (Lepidoptera: Noctuidae) S. exigua

Lepidopterous pests

C. carnea; Orius tristicolor (Anthocoridae) Heteropterans and spiders Coleoptera, Araneae Hippodamia convergens (Coccinellidae) Several Heteroptera H. axyridis, Nebria brevicollis Podisus maculiventris P. maculiventris Carabids and others

Geocoris punctipes (Het: Lygaeidae) P. maculiventris

Predators (several)

NEFE Not affected

Reduced growth of predators

No effects

No effects on development time

No effects on field pitfall trap capture numbers NEFE

No effects

NEFE

NEFE

NEFE

Predator numbers similar or higher in Bt fields

Genetically Modified (GM) Crops (continued)

Bell et al. (2003) Graham et al. (2002)

Bell et al. (2004)

Armer et al. (2000) Ferry et al. (2007)

Reed et al. (2001) Duan et al. (2004) Dogan et al. (1996)

Torres and Ruberson (2008) Sisterson et al. (2007)

Hagerty et al. (2005), Head et al. (2005) Torres and Ruberson (2006)

18.10 691

Oilseed rape (B. napus)

Oilseed rape (B. napus) Oilseed rape (B. napus) Potato (S. tuberosum)

MTI-2

OC-1

Deroceras reticulatum (Mollusca) P. Xylostella (Lepidoptera: Plutellidae) Leptinotarsa decemlineata (Colorado potato beetle)

Plutella xylostella

S. littoralis

Pest L. decemlineata

Perillus bioculatus

Pterostichus melanarius H. axyridis

Pterostichus madidus

P. maculiventris

Predators P. maculiventris

NEFE

NEO reproductive fitness; female weight gain reduced at first but compensated for later No effects on mortality, weight gain, or food consumption No effects on survival or development

NEFE

Impact NEFE

Mulligan et al. (2006) Ferry et al. (2003) Bouchard et al. (2003)

References AlvarezAlfageme et al. (2007) AlvarezAlfageme et al. (2007) Ferry et al. (2005)

18

NEFE no effects in field experiments

OC-1

OC-1

Potato

Transgenic plant Potato

HvCPI-1 C68 → G

Bt protein HvCPI-1 C68 → G

Table 18.4 (continued)

692 Controlled Field Cage and Field Evolution

18.11

Bio-Intensive Pest Management (BIPM)

693

ochropezus were the leading species when only predatory carabid species were taken into account, and the numbers caught were comparable between cotton varieties. Dermapterans, staphylinids, araneids, and heteropterans were more or less abundant depending on the sample date and the season but did not differ depending on the type of cotton. The frequent occurrence of Megacephala carolina, Selenophorus palliatus, and Pardosa pauxilla capture in both cotton fields and seasons infers that these species may be crucial for observing future changes in nearby communities as a result of agricultural operations (Torres and Ruberson 2007). Treatment with transgenic Cry1Ac/Cry2Ab cotton had no effect on the fitness of the Japanese ladybeetle, Propylaea japonica (Zhao et al. 2013). Consuming Trichoplusia ni on transgenic Cry1Ac/Cry2Ab cotton did not significantly influence Coleomegilla maculata’s mortality, growth, development, or fertility (Li et al. 2011). When comparing non-Bt (Brumoides suturalis, Cheilomenes sexmaculata, and Coccinella transversalis) and Bt (Brumoides suturalis, Cheilomenes sexmaculata, Coccinella transversalis, and Coccinella octomaculata) crops, the predatory arthropods in India during the 2019 -2020 cotton season were not altered (Mallesh and Sravanthy 2021). In the Telangana state of India, Bt and non-Bt cotton farms showed similar results (Mallesh and Sravanthy 2021). The impact of GM crops or Bt protein on predatory insects has been the subject of numerous studies up to this point. Chrysoperla sinica larvae grew and developed normally in GM cotton that expressed the Bt (Cry1Ac) and CpTI (cowpea trypsin inhibitor) proteins, and adult fecundity was not significantly different from that of the control treatment (Liu et al. 2021). Both the conventional cotton variety J14 and the GM cotton variety A26–5, which express the Cry1Ac/1Ab protein, were purchased from Biocentury Transgene Co. Ltd. (China). The effect of GM crops on non-target insects is one of the crucial aspects of environmental risk assessment (Zhang et al. 2022).

18.11 Bio-Intensive Pest Management (BIPM) This section will look at how predators interact with the different IPM programme strategies. Thomson and Hoffmann (2007) discovered that while mulches enhanced populations of parasitoids and predators that live in the canopy and in the soil, they had no impact on pest populations. Tillage has an impact on carabid populations either directly through tillage-related deaths or indirectly through loss of prey resources and microclimate changes (Thorbek and Bilde 2004). Tillage, however, has also been discovered to have an impact on predatory arthropods that live in foliage (Marti and Olson 2007). In contrast to the intended biting flies, Frick and Tallamy (1996) discovered that electric traps using ultraviolet light as an attractant killed almost exclusively non-target insects, with roughly 13.5% of the catch being predatory. At a farmer’s field in Tamil Nadu, India, the efficacy of the augmentative release of the reduviid predator Rhynocoris kumarii (Heteroptera: Reduviidae), hot water extract of Ipomea cornea (Convolvulaceae), and Vitex negundo (Verbenaceae)

694

18

Controlled Field Cage and Field Evolution

against a number of groundnut pests was assessed. The findings showed that endosulfan consistently and significantly lowers all pest populations. The combined use of both plant extracts in equal amounts significantly decreased the population of Aphis craccivora (Hemiptera: Aphididae) (54.40%), followed by Ipomea cornea (48.86%) and Vitex negundo (37.85%). Similar findings were made for Spodoptera litura (Lepidoptera: Noctuidae) and Helicoverpa armigera (Lepidoptera: Noctuidae). The effects of both plants on grasshoppers like C. trachypterus and A. crenulata are quite minimal. Atractomorpha crenulata, Aphis craccivora, Helicoverpa armigera, and Chrotogonus trachypterus are the species that most effectively eliminate Spodoptera litura. However, Mylabris pustulata can be restrained by the reduviid predator. The plot with the Rhynocoris kumarii-released field, endosulfan, and the combination of both plant extracts sprayed in equal proportion had the highest groundnut output among all the treatments examined. The majority of treatments consistently lower insect populations and boost production of groundnuts; therefore, they can all be used and incorporated into the control of groundnut pests (Sahayaraj and Ravi 2007). In order to increase the yield of groundnuts, Sahayaraj and Ravi (2007) combined Rhynocoris marginatus with a few botanicals in the groundnut field. In a different investigation, natural manure-applied control plants (24.52) had a higher proportion of pod damage did than urea (28.71), endosulfan (30.4), and a combination of urea- and endosulfan-treated (23.5) plotted plants. Natural manuresupplied control plants had a higher concentration of natural enemies (14.81) than did plants treated with urea (7.31), endosulfan (2.75), or a combination of urea and endosulfan (2.75). The highest re-sight of released Rhynocoris fuscipes was observed in the plots treated with natural manure (2.74), whereas a significant gradual decline or complete disappearance of Rhynocoris fuscipes was observed in the plots treated with urea (1.37), endosulfan (0.05), and a mixture of both urea and endosulfan (0.01). The evaluation of grain yield revealed that the chemically treated planted plants had a strong and consistent rise in production when compared to the control. For instance, the grain yield for the depicted plant treated with natural manure, urea, endosulfan, and a mixture of the two was 859.41 kg/ha, 778.43 kg/ ha, 937.6 kg/ha, and 954.7 kg/ha, respectively; these values were not statistically significant (Claver et al. 2018). The goal of the experiment was to determine how well natural enemies worked against the green apple aphid, Aphis pomi, in apple orchards, and the cabbage aphid, Brevicoryne brassicae, in a field of cabbage. For an experiment, second and third instar larvae of three coccinellid species—Coccinella septempunctata, Adalia tetraspilota, and Hippodamia variegata—and one chrysopid species—Chrysoperla z. sillemi—were released at a rate of 30 trees per week in apple orchards in Pattan, Baramulla, and at a rate of 5 plants per On the “Red Delicious” types of apples and the “Golden Acre” variety of cabbage, respectively; both stages were observed for their effectiveness against apple aphid and cabbage aphid, respectively. With the highest recovery rates of 52.00% and 54.4%, respectively, the third instar stage of Coccinella septempunctata demonstrated the best performance in terms of reduction of green apple aphid (62.00%) and cabbage aphid (63.98%). Therefore, Coccinella

References

695

septempunctata could be used as a biocontrol agent in Kashmir’s ecosystems to combat the green apple aphid and the cabbage aphid (Khan et al. 2017). In field surveys, generalist predators such as green lacewings and coccinellids have frequently been linked to Diaphorina citri (Hemiptera: Liviidae), but they have rarely been put to the test. Five predators were investigated for their capacity to prey under laboratory, greenhouse, and semi-field conditions: the coccinellids Diomus pumilio and Rhyzobius lophanthae, the green lacewings Chrysoperla comanche and Chrysoperla rufilabris, and the brown lacewing Sympherobius barberi (Neuroptera: Hemerobiidae). Under laboratory conditions, all of them were able to consume Diaphorina citri and lower its numbers, although important distinctions between greenhouse and field cages developed, with particular predators (the green lacewings and Diomus pumilio) being the most successful. We describe the potential use for these predators as an additional biological control tool to address the present challenges with Diaphorina citri control in commercial groves, in neighbourhoods, and at the interface between urban and agricultural areas (Gomez-Marco et al. 2022). Cruz-Esteban et al. (2022) examined the idea of combining the use of chromatic traps with the application of spinosad and the strategic release of Orius insidiosus in Mexico to successfully control Frankliniella occidentalis in organic berries without sacrificing fruit quality.

18.12 Recommendations Studies should be conducted in accordance with recommendations that a predator for a crop is essential. It is important to recognise and advise against the topological and season-specific natural enemies. Farmers need to know how to distinguish between pests and natural enemies so that they can protect the latter if they ever appear in their fields. A specific predator needs to be found and suggested for the preventative biological management of domestic or foreign invasive pests.

References Ables JR, McCommas DW Jr (1982) Efficacy of Podisus maculiventris as a predator of variegated cutworm on greenhouse cotton. J Ga Entomol Soc 17(2):204–206 Alvarez-Alfageme F, Martinez M, Pascual-Ruiz S, Castanera P, Diaz I, Ortego F (2007) Effects of potato plants expressing a barley cystatin on the predatory bug Podisus maculiventris via herbivorous prey feeding on the plant. Transgenic Res 16:1–13 Alvarez-Alfageme F, Ferry N, Castanera P, Ortego F, Gatehouse AMR (2008) Prey mediated effects of Bt maize on fitness and digestive physiology of the red spider mite predator Stethorus punctillum Weise (Coleoptera: Coccinellidae). Transgenic Res 17:943–954 Ambrose DP, Claver MA (1999) Suppression of cotton leafworm Spodoptera litura, flower beetle Mylabris pustulata and red cotton bug Dysdercus cingulatus by Rhynocoris marginatus (Fabr.) (Het., Reduviidae) in cotton field cages. J Appl Entomol 123(4):225–229 Ambrose DP, Kumar SP, Subbu GR, Claver MA (2003) Biology and prey influence on the postembryonic development of Rhynocoris longifrons (Stal) (Hemiptera: Reduvidae), a potential biological control agent. J Biol Control 17:113–119

696

18

Controlled Field Cage and Field Evolution

Amman GD, Speers CF (1971) Introduction and evaluation of predators from India and Pakistan for control of the balsam woolly aphid (Homoptera: Adelgidae) in North Carolina. Can Entomol 103(4):528–533 Ammar K, Jasman Aied N, Oueed Nassir A, Al-gamali (2013) Field survey and evaluation of the predators Coccinella undecimpunctata orchard (Coccinellidae: Coleoptera) in some vegetables and fields in Al–Mussiab/Babylon governorate. Euphrates J Agric Sci 5(3):193–202 Antony J, Danial M, Kurian C, Pillai GB (1979) Attempts on introduction and colonization of the exotic reduviid predator Platymeris laevicollis distant for the biological suppression of the coconut rhinoceros beetle, Oryctes rhinoceros. Placrosym 11:445–454 Armer CA, Berry RE, Kogan M (2000) Longevity of phytophagous heteropteran predators feeding on transgenic Bt-potato plants. Entomol Exp Appl 95:329–333 Ballal CR, Yamada K (2016) Anthocorid predators. In: Ecofriendly pest management for food security. Academic Press, London, pp 183–216 Bell HA, Down RE, Fitches EC, Edwards JP, Gatehouse AMR (2003) Impact of genetically modified potato expressing plant-derived insect resistance genes on the predatory bug Podisus maculiventris (Heteroptera: Pentatomidae). Biocontrol Sci Tech 13:729–741 Bell HA, Down RE, Kirkbride-Smith AE, Edwards JP (2004) Effect of microsporidian infection in Lacanobia oleracea (Lep., Noctuidae) on prey selection and consumption by the spined soldier bug Podisus maculiventris (Het., Pentatomidae). J Appl Entomol 128(8):548–553 Bhatti MA, Duan J, Head G, Jiang CJ, Mckee MJ, Nickson TE, Pilcher CL, Pilcher CD (2005) Field evaluation of the impact of corn rootworm (Coleoptera: Chrysomelidae)-protected Bt corn on ground-dwelling invertebrates. Environ Entomol 34:1325–1335 Boopathi T, Singh SB, Dutta SK, Dayal V, Singh AR, Chowdhury S, Lalhruaipuii K (2020) Harmonia sedecimnotata (F.): predatory potential, biology, life table, molecular characterization, and field evaluation against Aphis gossypii glover. Sci Rep 10(1):1–10 Bouchard E, Cloutier C, Michaud D (2003) Oryzacystatin I expressed in transgenic potato induces digestive compensation in an insect natural predator via its herbivorous prey feeding on the plant. Mol Ecol 12:2439–2446 Castillo J, Roda A, Qureshi J, Pérez-Hedo M, Urbaneja A, Stansly P (2022) Sesame as an alternative host plant to establish and retain predatory mirids in open-field tomatoes. Plant (Basel) 11:2779. https://doi.org/10.3390/plants11202779 Chambers RJ (1988) Syrphidae. In: Minks AK, Harrewijn P (eds) World crop pests: aphids, their biology, natural enemies and control, vol B. Elsevier, Amsterdam, pp 259–270 Chow A, Chau A, Heinz KM (2010) Compatibility of Amblyseius (Typhlodromips) swirskii (Athias-Henriot) (Acari: Phytoseiidae) and Orius insidiosus (Hemiptera: Anthocoridae) for biological control of Frankliniella occidentalis (Thysanoptera: Thripidae) on roses. Biol Control 53(2):188–196 Clausen CP (1978) Biological control of citrus insects. In: Reuther W, Calavan EC, Carman GE (eds) The citrus industry, vol IV. University of California Press, Berkeley, CA, pp 276–320 Claver MA, Kalyanasundaram M, David PMM, Ambrose DP (2003) Abundance of boll worm, flower beetle, predators and field colonization by Rhynocoris kumarii (Het., Reduviidae) following mulching and shelter provisioning in cotton. J Appl Entomol 127(7):383–388 Claver MA, Das SA, Ambrose DP (2018) Impact of urea and endosulfan on pod damage, gain yield, pest occurrence, abundance of natural enemies and colonization of released predator in the red gram agroecosytem. J Entomol Zool Stud 6(2):1325–1329 Cock MJW, van Lenteren JC, Brodeur J, Barratt BI, Bigler F et al (2010) Do new access and benefit sharing procedures under the convention on biological diversity threaten the future of biological control? BioControl 55:199–218 Cruz-Esteban S, Brito-Bonifacio I, Estrada-Valencia D, Garay-Serrano E (2022) Mortality of Orius insidiosus by contact with spinosad in the laboratory as well as in the field and a perspective of these as controllers of Frankliniella occidentalis. J Pestic Sci 47:93–99 Daane KM, Yokota GY (1997) Release methods affect egg survival and distribution of augmentated green lacewings (Chrysopidae: Neuroptera). Environ Entomol 26:455–464

References

697

Daane KM, Yokota GY, Zheng Y, Hagen KS (1996) Inundative release of common green lacewings (Neuroptera: Chrysopidae) to suppress Erythroneura variabilis and E. elegantula (Homoptera: Cicadellidae) in vineyards. Environ Entomol 25(5):1224–1234 Daane KM, Hagen KS, Mills NJ (1998) Predaceous insects for insect and mite control. In: Ridgway RL, Hoffmann MP, Inscoe MN, Glenister CS (eds) Mass-reared natural enemies: application, regulation, and needs, vol III. Thomas say publications in entomology. Entomological Society of America, Lanham, MD, pp 61–115 De Clercq P, Merlevede F, Mestdagh I, Vandendurpel K, Mohaghegh J, Degheele D (1998) Predation on the tomato looper Chrysodeixis chalcites (Esper)(Lep., Noctuidae) by Podisus maculiventris (Say) and Podisus nigrispinus (Dallas)(Het., Pentatomidae). J Appl Entomol 122(1–5):93–98 Dively GP (2005) Impact of transgenic VIP3A × cry 1Ab lepidopteran-resistant field corn on the nontarget arthropod community. Environ Entomol 34:1267–1291 Dogan EB, Berry RE, Reed GL, Rossignol PA (1996) Biological parameters of convergent lady beetle (Coleoptera: Coccinellidae) feeding on aphids (Homoptera: Aphididae) on transgenic potato. J Econ Entomol 89:1105–1108 Duan JJ, Head G, Jensen A, Reed G (2004) Effects of transgenic Bacillus thuringiensis potato and conventional insecticides for Colorado potato beetle (Coleoptera: Chrysomelidae) management on the abundance of ground-dwelling arthropods in Oregon potato ecosystems. Environ Entomol 33:275–281 Dumont F, Maisonhaute JÉ, Labrie G, Provost C (2021) Biological control of silverleaf whitefly Bemisia tabaci (Hemiptera: Aleyrodidae) using predatory bugs, Dicyphus hesperus (Hemiptera: Miridae) and Orius insidiosus (Hemiptera: Anthocoridae). Phytoprotection 101(1):38–45 Dutton A, Klein H, Romeis J, Bigler F (2002) Uptake of Bt-toxin by herbivores feeding on transgenic maize and consequences for the predator Chrysoperla carnea. Ecol Entomol 27: 441–447 Ehler LE, Long RF, Kinsey MG, Kelley SK (1997) Potential for augmentarive biological control of black bean aphid in California sugarbeet. Entomophaga 42:241–256 Elimem M, Chermiti B (2012) Use of the predators Orius laevigatus and Aeolothrips spp. to control Frankliniella occidentalis populations in greenhouse peppers in the region of Monastir, Tunisia. IOBC/WPRS Bull 80:141–146 El-Sebaey IAI, Abd El-Wahab HA (2010) Effect of different stages of reduviid predator corans africana El-Sebaey (Hemiptera: Heteroptera) on the population of Bemisia tabaci genn in tomato field at Qalubia and Bani-swif governorates. Egyptian J Agric Res 89(2):413–429 Ferry N, Raemaekers RJM, Majerus MEN, Jouanin L, Port G, Gatehouse JA, Gatehouse AMR (2003) Impact of oilseed rape expressing the insecticidal cysteine protease inhibitor oryzacystatin on the beneficial predator Harmonia axyridis (multicolored Asian ladybeetle). Mol Ecol 12:493–504 Ferry N, Jouanin L, Ceci LR, Mulligan EA, Emami K, Gatehouse JA, Gatehouse AMR (2005) Impact of oilseed rape expressing the insecticidal serine protease inhibitor, mustard trypsin inhibitor-2 on the beneficial predator Pterostichus madidus. Mol Ecol 14:337–349 Ferry N, Mulligan EA, Majerus MEN, Gatehouse AMR (2007) Bitrophic and Tritrophic effects of Bt cry 3A transgenic potato on beneficial, non-target, beetles. Transgenic Res 16:795–812 Frick TB, Tallamy DW (1996) Density and diversity of nontarget insects killed by suburban electric insect traps. Entomol News 107(2):77–82 Gomez-Marco F, Gebiola M, Simmons GS, Stouthamer R (2022) Native, naturalized and commercial predators evaluated for use against Diaphorina citri. Crop Prot 155:105907 Graham J, Gordon SC, Smith K, McNicol RJ, McNicol JW (2002) The effect of the cowpea trypspin inhibitor in strawberry on damage by vine weevil under field conditions. J Hortic Sci Biotechnol 77:L33–L40 Grundy PR, Maelzer D (2000) Assessment of Pristhesancus plagipennis (Walker) (Hemiptera: Reduviidae) as an augmented biological control in cotton and soybean crops. Aust J Entomol 9:305–309

698

18

Controlled Field Cage and Field Evolution

Grundy PR, Maelzer DA (2002) Augmentation of the assassin bug Pristhesancus plagipennis (Walker)(Hemiptera: Reduviidae) as a biological control agent for Helicoverpa spp. in cotton. Aust J Entomol 41(2):192–196 Guo JY, Wan FH, Dong L, Lovei GL, Han ZJ (2008) Tri-trophic interactions between Bt cotton, the herbivore Aphis gossypii glover (Homoptera: Aphididae), and the predator Chrysopa pallens (Rambur) (Neuroptera: Chrysopidae). Environ Entomol 37:263–270 Hagerty AM, Kilpatrick AL, Turnipseed SG, Sullivan MJ, Bridges WC (2005) Predaceous arthropods and lepidopteran pests on conventional, Bollgard, and Bollgard II cotton under untreated and disrupted conditions. Environ Entomol 34:105–114 Head G, Moar M, Eubanks M, Freeman B, Ruberson J, Hagerty A, Turnipseed S (2005) A multiyear, large-scale comparison of arthropod populations on commercially managed Bt and non-Bt cotton fields. Environ Entomol 34:1257–1266 Herrick NJ, Reitz SR (2004) Temporal occurrence of Podisus maculiventris (Hemiptera: Heteroptera: Pentatomidae) in North Florida. Fla Entomol 87(4):587–590 Hilbeck A, Baumgartner M, Fried PM, Bigler F (1998) Effects of transgenic Bacillus thuringiensis corn-fed prey on mortality and development time of immature Chrysoperla carnea (Neuroptera: Chrysopidae). Environ. Entomol 27:480–487 Hodek I, Evans EW (2012) Food relationships. In: Hodek I, van Emden HF, Honek A (eds) Ecology and behaviour of the ladybird beetles (Coccinellidae). John Wiley & Sons, Chichester, pp 141–274 Hosseini M, Ashouri A, Enkegaard A, Weisser WW, Goldansaz SH, Mahalati MN, Sarraf Moayeri HR (2010) Plant quality effects on intraguild predation between Orius laevigatus and Aphidoletes aphidimyza. Entomologia Experimentalis et Applicata 135(2):208–216 Hough-Goldstein J, McPherson D (1996) Comparison of Perillus bioculatus and Podisus maculiventris (Hemiptera: Pentatomidae) as potential control agents of the Colorado potato beetle (Coleoptera: Chrysomelidae). J Econ Entomol 89(5):1116–1123 Irvin NA, Pierce C, Hoddle MS (2021) Evaluating the potential of flowering plants for enhancing predatory hoverflies (Syrphidae) for biological control of Diaphorina citri (Liviidae) in California. Biol Control 157:104574 Kairo M, Paraiso O, Gautam RD, Peterkin DD (2013) Cryptolaemus montrouzieri (Mulsant) (Coccinellidae: Scymninae): a review of biology, ecology, and use in biological control with particular reference to potential impact on non-target organisms. CABI Rev 8:005 Khan I, Morse JG (1999) Field evaluation of Chrysoperla spp. as predators of citrus thrips. Sarhad J Agric (Pakistan) 15(6):607–610 Khan AA, Reyaz S, Kunndoo AA (2017) Evaluation of efficacy of predators against green apple aphid (aphis pomi) in apple orchards and cabbage aphid (Brevicoryne brassicae) in cabbage field of Kashmir. J Entomol Zool Stud 5:112–116 Koch RL, Galvan TL (2008) Bad side of a good beetle: the North American experience with Harmonia axyridis. BioControl 53:23–35 Kumar AG (2011) Mass multiplication, large scale release and biocontrol potential evaluation. Ph. D. dissertation. Manonmaniam Sundaranar University, India Li Y, Romeis J, Wang P, Peng Y, Shelton AM (2011) A comprehensive assessment of the effects of Bt cotton on Coleomegilla maculata demonstrates no detrimental effects by Cry1Ac and Cry2Ab. PLoS One 6(7):e22185 Liu ZF, Liang YY, Sun XT, Yang J, Zhang PJ, Gao Y, Fan RJ (2021) Analysis of differentially expressed genes of Chrysoperla sinica related to flight capacity by transcriptome. J Insect Sci 21 (1):18 Lundgren JG, Wiedenmann RN (2005) Tritrophic interactions among Bt (cry Mb1) corn, aphid prey, and the predator Coleomegilla maculata (Coleoptera: Coccinellidae). Environ Entomol 34:1621–1625. https://doi.org/10.1603/0046-225X-34.6.1621 Lyla KR, Beevi SP, Chandish B (2006) Field evaluation of anthocorid predator, Cardiastethus exiguus Poppius against Opisina arenosella Walker (Lepidoptera: Oecophoridae) in Kerala. J Biol Control 20(2):229–231

References

699

Lynch CA, Smith OM, Chapman EG, Crossley MS, Crowder DW, Fu Z, Snyder WE (2021) Alternative prey and farming system mediate predation of C olorado potato beetles by generalists. Pest Manag Sci 78:3769. https://doi.org/10.1002/ps.6553 Mallesh M and Sravanthy C (2021) Diversity of predatory arthropods in Bt and non-Bt cotton fields of Nalgonda district, Telangana state, India Manyuli T, Kyamanywa S, Luther GC (2008) Capability of Forficula auricularia linnaeus (Dermaptera: Forficulidae) to prey on Aphis craccivora koch (Homoptera: Aphididae) in eastern and Central Africa. Agron Afr 20(1):49–58 Marti OG, Olson DM (2007) Effect of tillage on cotton aphids (Homoptera: Aphididae), pathogenic fungi, and predators in south Central Georgia cotton fields. J Entomol Sci 42(3):354–367 McGregor RR, Gillespie DR, Quiring DM, Foisy MR (1999) Potential use of Dicyphus hesperus knight (Heteroptera: Miridae) for biological control of pests of greenhouse tomatoes. Biol Control 16(1):104–110 McPherson JE (1982) The Pentatomoidea (Hemiptera) of northeastern North America with emphasis on the fauna of Illinois. SIU Press Messelink GJ, Janssen A (2014) Increased control of thrips and aphids in greenhouses with two species of generalist predatory bugs involved in intraguild predation. Biol Control 79:1–7 Messelink GJ, Bloemhard CM, Sabelis MW, Janssen A (2013) Biological control of aphids in the presence of thrips and their enemies. BioControl 58:45–55 Messelink GJ, Bloemhard CMJ, Hoogerbrugge H, Van Schelt J, Ingegno BL, Tavella L (2015) Evaluation of mirid predatory bugs and release strategy for aphid control in sweet pepper. J Appl Entomol 139(5):333–341 Montemayor CO, Cave RD (2012) Evaluation of the predation capacity of Podisus maculiventris (Hemiptera: Pentatomidae) on Microtheca ochroloma (Coleoptera: Chrysomelidae) in field cages. J Econ Entomol 105(5):1719–1725 Mulligan EA, Ferry N, Jouanin L, Walters KF, Port GR, Gatehouse AM (2006) Comparing the impact of conventional pesticide and use of a transgenic pest-resistant crop on the beneficial carabid beetle Pterostichus melanarius. Pest Manag Sci 62(10):999–1012 Murata M, Imamura T, Miyanoshita A (2007) Suppression of the stored-product insect Tribolium confusum by Xylocoris flavipes and Amphibolus venator. J Appl Entomol 131(8):559–563 Nagarajan K (2010) Mass multiplication, large scale release and biocontrol potential evaluation of a Reduviid Predator Rhynocoris fuscipes (Fabricius)(Insecta: Heteroptera: Reduviidae) against Chosen Insect Pests, Ph.D. thesis. Manonmaniam Sundaranar University, Tirunelveli, Tamil Nadu, India Naranjo SE (2005a) Long-term assessment of the effects of transgenic Bt cotton on the function of the natural enemy community. Environ Entomol 34:1211–1223 Naranjo SE (2005b) Long-term assessment of the effects of transgenic Bt cotton on the abundance of non-target arthropod natural enemies. Environ Entomol 34:1193–1210 Pellizzari G, Porcelli F (2014) Alien scale insects (Hemiptera Coccoidea) in European and Mediterranean countries: the fate of new and old introductions. Phytoparasitica 42:713–721 Pérez-Aguilar DA, Martínez AM, Viñuela E, Figueroa JI, Gómez B, Morales SI, Pineda S (2019) Impact of the zoophytophagous predator Engytatus varians (Hemiptera: Miridae) on Bactericera cockerelli (Hemiptera: Triozidae) control. Biol Control 132:29–35 Pijnakker J, Overgaag D, Guilbaud M, Vangansbeke D, Duarte M, Wäckers F (2019) Biological control of the Japanese flower thrips Thrips setosus Moulton (Thysanoptera: Thripidae) in greenhouse ornamentals. IOBC/WPRS Bull 147:107–112 Pilcher CD, Obrycki JJ, Rice ME, Lewis LC (1997) Preimaginal development, survival, and field abundance of insect predators on transgenic Bacillus thuringiensis corn. Environ Entomol 26: 446–454 Ramos FLD, De Oño JL, Cosico SMQ (2019) Field evaluation and behavioral response of black earwig, Chelisoches morio (Fabricius) (Dermaptera: Chelisochidae) to the coconut leaf beetle, Brontispa longissima (Gestro) (Coleoptera: Chrysomelidae). AAJMRA 3(2):1

700

18

Controlled Field Cage and Field Evolution

Ravichandran B (2004) Biocontrol potential evaluation of a reduviid predator Rhynocoris kumarii ambrose and Livingstone insecta heteroptera reduviidae against three chosen cotton insect pests. Ph.D. Thesis. MS University, Tamil Nadu, p 101 Reed GL, Jensen AS, Riebe J, Head G, Duan JJ (2001) Transgenic Bt potato and conventional insecticides for Colorado potato beetle management: comparative efficacy and non-target impacts. Entomol Exp Appl 100:89–100 Reitz SR (2009) Biology and ecology of the western flower thrips (Thysanoptera: Thripidae): the making of a pest. Fla Entomol 92:7–13 Romeis J, Meissl M, Bigler F (2006) Transgenic crops expressing Bacillus thuringiensis toxins and biological control. Nat Biotechnol 24:63–71 Rondoni G, Borges I, Collatz J, Conti E, Costamagna AC, Dumont F, Cock MJ (2021) Exotic ladybirds for biological control of herbivorous insects–a review. Entomol Exp Appl 169(1): 6–27 Roy H and Migeon A (2010) Ladybeetles (Coccinellidae). Biodiversity and ecosystem risk assessment 4: 293–313 Sahayaraj K (1999) Field evaluation of Rhynocoris marginatus (Fab.) against two groundnut defoliators. Int Arachis News Lett 19:41–42 Sahayaraj K, Balasubramanian R (2016) Field evaluation. In: Artificial rearing of Reduviid predators for pest management. Springer, Singapore, pp 147–165 Sahayaraj K, Martin P (2003) Assessment of Rhynocoris marginatus (Fab.) (Hemiptera: Reduviidae) as aug[1]mented control in groundnut pests. J Central Euro Agr 4(2):103–110 Sahayaraj K, Ravi C (2007) Evaluation of reduviid predators and plant products against chosen groundnut pests. Arch Phytopathol Plant Protect 40(4):281–290 Sampson C, Kirk WD (2016) Predatory mites double the economic injury level of Frankliniella occidentalis in strawberry. BioControl 61:661–669 Sarkar SC, Wang E, Zhang Z, Wu S, Lei Z (2019) Laboratory and glasshouse evaluation of the green lacewing, Chrysopa pallens (Neuroptera: Chrysopidae) against the western flower thrips, Frankliniella occidentalis (Thysanoptera: Thripidae). Appl Entomol Zool 54(1):115–121 Sigsgaard L, Esbjerg P, Philipsen H (2006) Experimental releases of Anthocoris nemoralis F. and Anthocoris nemorum (L.)(Heteroptera: Anthocoridae) against the pear psyllid Cacopsylla pyri L.(Homoptera: Psyllidae) in pear. Biol Control 39(1):87–95 Simmons AM, Abd-Rabou S (2011) Inundative field releases and evaluation of three predators for Bemisia tabaci (Hemiptera: Aleyrodidae) management in three vegetable crops. Insect Sci 18 (2):195–202 Sisterson M, Biggs RW, Manhardt NM, Carriere Y, Dennehy TJ, Tabashnik BE (2007) Effects of transgenic Bt cotton on insecticide use and abundance of two generalist predators. Entomol Exp Appl 124:305–311 Sörensson A, Nedstam B (1993) Effect of Amblyseius cucumeris and Orius insidiosus on Frankliniella occidentalis in ornamentals. IOBC/wprs Bull 16:129–132 Srikanth J, Singaravelu B, Kurup NK, Mukunthan N, Santhalakshmi G, Nirmala R (2015) Predators as natural and applied biocontrol agents of sugarcane woolly aphid Ceratovacuna lanigera in India: an appraisal. J Sugarcane Res 5(2):53–72 Suasaard W (2010) Natural enemies of important insect pests of field crops and utilization as biological control agents in Thailand. In: Proceedings of international seminar on enhancement of functional biodiversity relevant to sustainable food production in ASPAC, pp 9–11 Tauber MJ, Tauber CA (1993) Adaptations to temporal variation in habitats: categorizing, predicting, and influencing their evolution in agroecosystems, 1n K. In: Kim C, McPheron BA (eds) Evolurion of insect pests. Wiley, New York, NY, pp 103–126 Thomson LJ, Hoffmann AA (2007) Effects of ground cover (straw and compost) on the abundance of natural enemies and soil macro invertebrates in vineyards. Agric For Entomol 9:173–179 Thorbek P, Bilde T (2004) Reduced management of generalist arthropods after crops management. J Appl Ecol 41:526–538

References

701

Tipping PW, Holko CA, Abdul-Baki AA, Aldrich JR (1999) Evaluating Edovum puttleri Grissell and Podisus maculiventris (say) for augmentative biological control of Colorado potato beetle in tomatoes. Biol Control 16(1):35–42 Torres JB, Ruberson JR (2006) Interactions of Bt-cotton and the omnivorous big-eyed bug Geocoris punctipes (say), a key predator in cotton fields. Biol Control 39:47–57 Torres JB, Ruberson JR (2007) Abundance and diversity of ground-dwelling arthropods of pest management importance in commercial Bt and non-Bt cotton fields. Faculty Publications: Department of Entomology, p 762. https://digitalcommons.unl.edu/entomologyfacpub/76 Torres JB, Ruberson JR (2008) Interactions of Bacillus thuringiensis cry 1Ac toxin in genetically engineered cotton with predatory heteropterans. Transgenic Res 17:345–354 Van De Veire M, Degheele D (1995) Comparative laboratory experiment with Orius insidiosus and Orius albidipennis (Het.: Anthocoridae), two candidates for biological control in glasshouses. Entomophaga 40:341–344 van Lenteren JC (2000) A greenhouse without pesticides: fact or fantasy? Crop Prot 19:375–384 Wallace MS and Hain FP (1999) Field surveys and evaluation of native predators of the hemlock woolly adelgid (Homoptera: Adelgidae) in the southeastern United States (Master's Thesis, North Carolina State University), 29, 638 Wang CL, Lee PC, Wu YJ (2001a) Field augmentation of Orius strigicollis (Heteroptera: Anthocoridae) for the control of thrips in Taiwan. ASPAC Food & Fertilizer Technology Center, Taipei, pp 1–9 Wang CL, Lee PC, Wu YJ (2001b) Field augmentation of Orius strigicollis (Heteroptera: Anthocoridae) for the control of thrips in Taiwan, vol 500. Food and Fertilizer Technology Center, Taipei, pp 1–9; ROC, Exten Bull Whitehouse MEA, Wilson LJ, Fitt GP (2005) A comparison of arthropod communities in transgenic Bt and conventional cotton in Australia. Environ Entomol 34:1224–1241 Wong SK, Frank SD (2012) Influence of banker plants and spiders on biological control by Orius insidiosus (Heteroptera: Anthocoridae). Biol Control 63(2):181–187 Zhang M, Ma Y, Luo J, Ji J, Gao X, Wu C, Cui J (2022) Transgenic insect-resistant Bt cotton expressing cry 1Ac/1Ab does not harm the insect predator Geocoris pallidipennis. Ecotoxicol Environ Saf 230:113129 Zhao C, Griffin JN, Wu X, Sun S (2013) Predatory beetles facilitate plant growth by driving earthworms to lower soil layers. J Anim Ecol 82(4):749–758

Commercially Available Predators

19

Contents 19.1 19.2

Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Commercial Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19.2.1 University Extension Service . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19.2.2 Commercial Producers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19.3 Quality Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19.3.1 Species Identification Is Verified . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19.3.2 Stock Deterioration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19.3.3 Shipping and Handling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19.4 Future Focus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Annexure 19.1: Contact Address for Selected Commercial Producers . . . . . . . . . . . . . . . . . . . . . . . . Annexure 19.2: Company Website . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Annexure 19.3: Government, University, Commercial and Non-profit Websites on Biological Control (O’Neil et al. 2003) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

19.1

703 704 705 705 707 712 715 715 716 724 730 731 732

Overview

The ability of insectaria to create an economically viable market for a highly dependable and reasonably priced supply of natural enemies is essential for the commercialisation of biological control. To begin achieving these goals, effective, standardised mass-rearing techniques are needed, including the following: (1) the use of affordable, nutrient-dense diets; (2) mechanised and space-efficient production systems; (3) trustworthy storage techniques; and (4) periodic evaluation of natural enemy quality. The mass breeding of insects, especially cannibalistic predators, demands a lot of room and work; at the moment, automated, spaceefficient technologies are used. Because all three instars of the Chrysoperla species are predatory, raising larvae currently represents the most expensive element of mass manufacturing. Most insectaria feed on mass-produced insect prey, which is more

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_19

703

704

19

Commercially Available Predators

Table 19.1 Estimated agricultural natural enemy pest control market from 2021 to 2029 Attribute Estimated agri natural enemy pest control market size in 2022 Projected agri natural enemy pest control (2029) market size Value CAGR (2022–2029) Top player share of agri natural enemy pest control market in 2021

Details US$ 16.6 billion US$ 25.1 billion 6.0% 5–10%

expensive than artificial diets (often lepidopteran eggs: Sitotroga, Anagasta, or Corcyra). The market for biocontrol agents in North America and Europe, which was estimated to be worth around USD 2359 million in 2018, is predicted to produce more than USD 7591 million by 2026, with a compound annual growth rate (CAGR) of about 15.7% between 2019 and 2026, as per a report (New York report 2019). According to recently published data from Future Market Insights (FMI), the market for agricultural natural enemy pest control had been expected to reach USD 16.6 billion in 2022 and USD 25.1 billion by 2029, growing at a CAGR of 5.0% from 2014 to 2021. Agri natural enemy pest control accounts for roughly 5–10% of the pest management market, which is its parent market (Table 19.1). Geographical divisions of the worldwide biocontrol agent market include North America, Europe, Asia-Pacific, and the rest of the world. During the projected period, North America is anticipated to have a large market share in the global market. The region is expanding as a result of rising customer demand for organic food. Due to the huge expansion of the organic product market in developing nations like China and India, Asia-Pacific is also anticipated to experience significant growth. Andermatt Biocontrol AG, BASF SE, Biobest Group NV, Marrone Bio Innovations, Inc., Vestaron Corp., Viridaxis SA, Precision Laboratories, LLC, Syngenta AG, and others are some of the major market participants for biocontrol agents worldwide.

19.2

Commercial Products

According to Lenteren (2003), over the past 30 years, the commercial use of biological control has grown extremely quickly. Currently, more than 125 types of natural enemies are produced by around 85 businesses globally. Although several species are also accessible in North America, the largest diversity of commercially produced natural enemy species is available in Europe. This is mostly due to a considerably larger greenhouse industry in Europe. Latin America, Asia, and (South) Africa are emerging markets. This chapter discusses the natural enemies that are most frequently sold. The target pest(s), optimum release rates, and unit of sale are all provided. Commercial mass production and marketing of two Chrysoperla species— Chrysoperla carnea and Chrysoperla rufilabris—is taking place in North America and Europe. The staphylinid beetles Orius majusculus and Orius laevigatus, which

19.2

Commercial Products

705

are both effective at controlling thrips, the predatory mite Hypoaspis miles, the staphylinid beetle Atheta coriaria, and the nematode Anthocoris nemoralis (a generalist predatory bug) were selected because they may potentially prey on Drosophila suzukii at its various life stages (larval, pupal, and adulthood). Syngenta Bioline (Little Clacton, UK) provided all of the control products (Cuthbertson et al. 2014). Initially purchased from Biobest (Biobest NV, USA Inc., Detroit, MI 482012311), Adalia bipunctata larvae were raised through adulthood on frozen Ephestia kuehniella eggs (Koppert Biological Systems, Romulus, MI 48174) (Khan et al. 2016).

19.2.1 University Extension Service Refer to “Commercial Sources of Predators, Parasitoids, and Pathogens (SP290-Z)”, The Agricultural Extension Service of the University of Tennessee, https://trace. tennessee.edu/utkagexcomhort/33 (Table 19.2) for more details.

19.2.2 Commercial Producers Commercial insectaria from North America (the United States and Canada), Europe (Germany, the United Kingdom, France, Spain, Italy, Greece, Sweden, Turkey, and Russia), Asia-Pacific (China, India, Japan, and South Korea), Latin America (Brazil and Mexico), the Middle East, and Africa (South Africa) are listed in Annexures 19.1, 19.2, and 19.3. In order to manage more than 16 pest species, Cryptolaemus montrouzieri (Coccinellidae) has been introduced into at least 64 nations and territories, according to Kairo et al. (2013). This coccinellid has been utilised in both traditional and augmentative biological control programmes.

Table 19.2 Zoological names and the common names of various predators and their target pests Zoological name Aphidoletes aphidimyza Feltiella acarisuga Cryptolaemus montrouzieri Cryptolaemus montrouzieri Delphastus catalinae (=pusillus) Macrolophus caliginosus Orius spp. Xylocoris flavipes

Common name Midge Midge Coccinellidae

Target pest Aphids Two-spotted spider mites Mealybugs

Coccinellidae

Mealybugs

Beetle

Silverleaf whiteflies

Mirid bug Minute pirate bugs Pirate bug

Whiteflies Aphids Eggs and larvae of beetle and moth pests in stored grains

706

19

Commercially Available Predators

19.2.2.1 India Chrysopids, Chrysopid eggs, Cryptolaemus montrouzieri (adults and larvae), Chilocorus nigrita (adults and larvae), Cheilomenes sexmaculata (adults, larvae), Curinus coeruleus (adults), Scymnus coccivora (adults), Coccinella septumpunctata (adults), Cardiastethus exiguus (adults/ny The Agricultural Promotion & Investment Corporation of Odisha Limited produced Cryptolaemus montrouzieri (for mealy bugs specifically on fruits) and Crysoperla carnea (controls larval pests in pulses, vegetables, and fruits). The following national biocontrol laboratories are accessible in various Indian states (Table 19.3) In these labs, insect predators such as Chrysoperla zastrowi sillemi, Mallada boniensis, and Cyrtorhynus feltiae (Neoaplectana carpocapsea) are raised on Corcyra cephalonica larvae. Reduviids, such as Isyndus heros, Endochus inornatus, Rhynocoris marginatus, Rhynocoris fuscipes, and Rhynocoris longifrons, are also widely spread in India and are efficient predators of numerous insect pests of pulses, cotton, soybean, maize, tomato, groundnut, peas, okra, and others. 19.2.2.2 Korea Since 1998, six businesses in Korea have been founded to generate or import natural enemies. In 1998, Korea IPM Co. Ltd in Geochang County and Nabis Co. Ltd in Mungyeong were both founded. In 2002, Nonsan’s Sesil Co. Ltd achieved the successful mass production of natural enemies. In Anseong, the Korea Beneficial Insect Lab Co. Ltd (KBIL), which mostly manufactures hygienic insects, was founded in 2002. The Netherlands-based Koppert Biological Systems launched their insect business in Korea in 2006. 19.2.2.3 Greenhouses In order to control whiteflies, predatory bugs (Hemiptera: Miridae) are widely used in Europe and Canada, but these all-purpose predators cannot be imported into the United States. The commercially accessible, highly effective aphid predator Aphidoletes aphidimyza (Diptera: Cecidomyiidae) has a wide host range. Aphidoletes aphidimyza is frequently used in vegetable crops to biologically control aphids, but it is rarely employed on ornamental plants because a substantial aphid population must first establish on the plants for biological control to be effective. Others that are commercially available or naturally migrate into attractive crops include lacewings, ladybeetles, tiny pirate bugs, and hover flies. For the same reasons as Aphidoletes aphidimyza, however, few gardeners utilise these in official aphid control programmes. Despite being widely available commercially and being regarded as good predators of a number of significant pests (including mealybugs), lacewings are rarely used in biological control programmes in greenhouses and nurseries. Their success in the greenhouse/nursery has been established in only a few research investigations (Parrella and Lewis 2017) (Table 19.4). At the international level, many companies have been coming forward to produce different natural enemies that target specific pests. Table 19.5 highlights various anthocorids, which have been utilised commercially worldwide.

19.3

Quality Control

707

Table 19.3 State Biocontrol Laboratories (SBCLs) in states and Union Territory States (UTS) established under the viii and x plans under grants-in-aid by the Indian government States Andhra Pradesh Arunachal Pradesh Andaman and Nicobar Islands Assam Bihar Chhattisgarh Goa Gujarat Haryana Himachal Pradesh Jammu and Kashmir Jharkhand Karnataka Kerala Lakshadweep Maharashtra Madhya Pradesh Meghalaya Mizoram Manipur Nagaland Orissa Pondicherry Punjab Rajasthan Sikkim Tamil Nadu Tripura Uttar Pradesh Uttarakhand West Bengal

19.3

Location Nidadavole, West Godavari Naharlagun, Papum Pare, Itanagar A&N Administration, Haddo, Port Blair Dalgaon, Darrang district, R.K. Mission Road, Ulubari, Guwahati-7 Mithapur, Patna Raipur Farmers’ Training Centre, Ela Farm, Old Goa Gandhinagar and Navsari Agricultural University, Navsari396450 Sirsa and Chandigarh Holta, Palanpur, Kangra district, and Mandi district Lal Mandi Campus, Srinagar Ranchi Kotnur “D”, Gulbarga-585102 Mannuthy, Thrissur-680655 and Thiruvananthapuram Andrott Islands Aurangabad district and Nandurbar district Barkhedi Kalan, Bhadbhada, Bhopal P.O. Nonglyer, East Khasi Hill, Upper Shillong-793009 Neihbawih, Siphir Mantripukhri, Imphal Medziphema, Kohima Baramunda, Post Delta Colony, Bhubaneshwar-751003 KVK Kurumbapett-9 Mansa Durgapura, Tonk Road, Jaipur Tadong, Gangtok Vinayapuram, Melur Taluk, Madurai Dutta Tilla, Badharghat, P.O. Arundhuti Nagar, Tripura West799003 Moradabad Haldwani and Dhakrani, Dehradun 230A, Netaji Subhash Chandra Road, Kolkata-700040

Quality Control

Both the practice of biological control and consumers’ view of biological control as a reliable pest management strategy depend on the standardised supply of high-quality natural enemies. However, because there are no stringent quality control regulations

708

19

Commercially Available Predators

Table 19.4 List of various predatory insects and their target pests as well as their hosts Predator Targeted pest Coleoptera: Coccinellidae Cryptolaemus Mealybugs montrouzieri

Delphastus catalinae (=Delphastus pusillus)

Whiteflies

Plants

Company

Citrus, ornamentals, and vegetables, including greenhouses and interiorscapes and pomegranate and banana vineyards

Buglogical Control Systems, Evergreen Growers Supply, Beneficial Insectary, Associates Insectary, ARBICO, Bio Control, S.A., BioBee, Biobest, Biobest Mexico S.A. de C.V., Bioline AgroSciences, and Everwood Farm Foothill Agricultural Research, Global Horticultural, Green Methods, HydroGardens, IPM Laboratories, Koppert, Natural Insect Control, Orcon, Plant Products, Rincon-Vitova, Sound Horticulture, and Tip Top Biocontrol are only a few of the organisations mentioned. Adana Biotar, Anatis Bioprotection, Crop Defenders for Turkey Anatis Bioprotection, Applied Bio-nomics, ARBICO, Beneficial Insectary, BioBee, Buglogical Control Systems, Dynamic Ecosystems Crop Supply, Evergreen Growers Supply, Hydro-Gardens, IPM Laboratories, Koppert, Natural Enemies, Natural Insect Control, and Plant Products are just a few examples of the companies we work with WestGrow Biological Solutions, Rincon-Vitova, Sound Horticulture, Crop Defenders, Orcon

Sweet potato, woolly aphid, azalea, hibiscus, cloudy winged, citrus ornamentals, vegetables, fruit, and citrus, including greenhouses and interiorscapes

(continued)

19.3

Quality Control

709

Table 19.4 (continued) Predator Rhyzobius lophanthae (=Lindorus lophanthae)

Targeted pest Scales, mealybugs

Coleoptera: Staphylinidae Dalotia Shore flies and thrips coriaria (=Atheta coriaria)

Diptera: Cecidomyiidae Aphidoletes Aphids aphidimyza

Plants Ornamentals, vegetables, citrus, and fruit

Company Natural Insect Control, Anatis Bioprotection, ARBICO, Buglogical Control Systems, Foothill Agricultural Research, IPM Laboratories, RinconVitova, Evergreen Growers Supply

Ornamentals, including greenhouses and interiorscapes

Crop Defenders, ARBICO, Evergreen Growers Supply, Tip Top Biocontrol, Sound Horticulture, WestGrow Biological Solutions, Natural Enemies, Bioline AgroSciences, Bio Control, S.A., Green Methods, Applied Bio-nomics, Dynamic Ecosystems Crop Supply, Beneficial Insectary, Buglogical Control Systems, Anatis Bioprotection, Everwood, and Natural Insect Control

Citrus, ornamentals, fruits, and vegetables, including greenhouses and interiorscapes

Beneficial Insectary, ARBICO, Anatis Bioprotection, Bio Control, S.A., BioBee, Biobest, Biobest Mexico S.A. de C.V., Bioline AgroSciences, Buglogical Control Systems, Crop Defenders, Dynamic Ecosystems, Crop Supply, Evergreen Growers Supply, Everwood Farm, Green Methods, HydroGardens, IPM Laboratories, Koppert, Natural Enemies, Natural Insect Control, Orcon, Plant Products, (continued)

710

19

Commercially Available Predators

Table 19.4 (continued) Predator

Feltiella acarisuga

Targeted pest

Spider mites

Hemiptera: Anthocoridae Orius Thrips, aphids, and insidiosus whiteflies

Plants

Ornamentals and vegetables, including greenhouses and interiorscapes

Ornamentals, vegetables, and citrus, including greenhouses and interiorscapes

Company Rincon-Vitova, Sound Horticulture, Tip Top Biocontrol, and WestGrow Biological Solutions are only a few of the companies mentioned Agricultural Insects, Bio Control, SA, Biotop, Systems for Buglogical Control Crop Protectors, Farm Everwood International Horticultural, Hydroponic Gardens, ARBICO and Koppert, Natural Opponents, Natural Pest Management, Natural Products, RinconVitova, Reliable Horticulture, S.A., Biobest Mexico de C. V. Anatis Biodefense, ARBICO, Agricultural Insects, Bio Control, SA BioBee, Biotop, S. A., Biobest Mexico de C.V., AgroSciences Bioline, Systems for Buglogical Control, Crop Protectors, Crop Supply in Dynamic Ecosystems, Providers of Evergreen Plants, Farm Everwood, Growing Life Biologicals (IPM Laboratories), Koppert, Natural Opponents, Natural Pest Management, Natural Products, RinconVitova Top Biocontrol Tip, and Reliable Horticulture (continued)

19.3

Quality Control

711

Table 19.4 (continued) Predator Orius armtus

Targeted pest Thrips (larvae and adults) aphids, spider mites, butterfly eggs Aphids, moth eggs, small caterpillars

Plants Sweet pepper, strawberries, Gerberas, eggplants Vegetables, sweet corn, field crops

Company Manichil IPM Services

General predators

All crops

ARBICO

Whiteflies

Greenhouse and tobacco

Anatis Bioprotection, ARBICO, BioBee, Biobest, Crop Defenders, GrowLiv Biologicals, Natural Insect Control, Sound Horticulture, Bioline AgroSciences

Colorado potato beetles and caterpillars

Ornamentals, vegetables, and citrus

ARBICO, Bioline AgroSciences, Buglogical Control Systems, Evergreen Growers Supply, Rincon-Vitova, Sound Horticulture

Neuroptera: Chrysopidae Chrysoperla Aphids and other small carnea soft-bodied insects

Ornamentals, citrus, fruit, and vegetables

Chrysoperla rufilabris

Ornamentals, citrus, fruit, and vegetables

Anatis Bioprotection, Bio Control, S.A., BioBee, Biobest, Buglogical Control Systems, Crop Defenders, Everwood Farm, GrowLiv Biologicals, Koppert, Natural Insect Control, Orcon, Plant Products, Beneficial Insectary BioBee, Buglogical Control Systems, Everwood Farm, Global Horticultural, IPM Laboratories, Natural Enemies, Natural Insect Control, Plant Products, RinconVitova, Sound Horticulture, Evergreen Growers Supply

Nabis kinbergii Reduviidae Zelus renardii Miridae Dicyphus hesperus

Pentatomidae Podisus maculiventris

Aphids and other small, soft-bodied insects

IPM Technologies

(continued)

712

19

Commercially Available Predators

Table 19.4 (continued) Predator Chrysoperla spp.

Targeted pest Aphids and other small soft-bodied insects

Plants Ornamentals, citrus, fruit, and vegetables

Company Biobest, Crop Defenders, ARBICO, Bioline AgroSciences, Kunafin, Natural Insect Control Bugs for Bugs

Aphids, moth eggs and small larvae, scales, and whiteflies Neuroptera: Hemerobiidae Micromus Aphids, whiteflies, and variegatus mealybugs

Field vegetables, tree crops, grapes

Sympherobius barberi

Grapes, citrus, tree crops, greenhouse crops

Anatis Bioprotection, Applied Bio-nomics, Crop Defenders, Dynamic Ecosystems Crop Supply, Evergreen Growers Supply, Natural Insect Control, WestGrow Biological Solutions, Everwood Farm Foothill Agricultural Research

Fruit trees

Rincon-Vitova

Mallada signata

Mealybugs, psyllids, thrips, mites, whiteflies, aphids, small caterpillars, leafhoppers, and insect eggs Thysanoptera: Thripidae Scolothrips Spider mites sexmaculatus

Vegetables and ornamentals

Source: https://edis.ifas.ufl.edu/pdf/IN/IN84900.pdf

in the United States, the quality of naturally occurring enemies that are sold commercially can vary. Examples include the fact that farmers’ orders for Chrysoperla carnea had not been regularly filled with the right species in a recent review of shipments from insectaries, and cannibalism dramatically decreased the survival of lacewings in transportation. Greater attention to maintaining correctly identified, pure colonies and enhanced processes during mass production and packing might help solve such issues. For Chrysoperla, Tauber et al. (2000) suggest the following:

19.3.1 Species Identification Is Verified At the beginning of the culture, the species of the Chrysoperla stock needs to be confirmed. To check for contamination, species identity should be reviewed occasionally throughout rearing (see the section above).

19.3

Quality Control

713

Table 19.5 Anthocorids utilised commercially worldwide Company name ARBICO

Country USA

Natural predators Orius insidiosus

Chrysoperla rufilabris, Chrysoperla sp. Hippodamia convergens

Aphidoletes aphidimyza Delphastus catalinae Dalotia coriaria (=Atheta coriaria) Stethorus punctillum

Associates Insectary BioBee USA Inc.

USA

BioBee USA Inc.

USA

BioBee USA Inc. BioBee USA Inc.

USA

BioBee USA Inc. BioBee USA Inc.

USA

USA

USA

Cryptolaemus montrouzieri Cryptolaemus montrouzieri Cryptolaemus Dalotia coriaria

Cryptolaemus montrouzieri Chrysoperla carnea Chrysoperla rufilabris Macrolophus pygmaeus Nephus bipunctatus Orius laevigatus

Targeted pests Thrips, whiteflies, spider mites, aphids, psyllids, small caterpillars, insect eggs, and others Aphids, mealybugs, spider mites, leafhopper nymphs, moth eggs, scale thrips, whiteflies, and more Aphids, chinch bugs, asparagus beetle larvae, thrips, alfalfa weevil larvae, bean thrips, grape root worms, Colorado potato beetle larvae, whiteflies, mites, and many other softbodied insects and eggs More than 60 different aphids, including green peach aphids (Myzus persicae) and hemlock woolly adelgid (Adelges tsugae) Aphids, scales, thrips, and, especially, greenhouse and outdoor sweet potato whiteflies, and other soft-bodied insect pests Fungus gnats (sciarid flies), thrips, and shore flies Two-spotted spider mites, European red mites, spruce spider mites, and southern red mites; various mites Mealybugs, aphids, mites, thrips, whiteflies, and other soft-bodied insect pests Scales, psyllids, and aphids and, of course, mealybugs Asian citrus psyllids (ACPs) Fungus gnats, thrips pupae, shore flies, moth fly larvae, root mealybugs, springtails, and other small arthropods Mealybugs Aphids, mites, and mealybugs Aphids, mites, and mealybugs Eggs and larvae of Tuta absoluta, whitefly eggs, thrips, aphids, leaf miners, and spider mites Mealybug life stages and eggs Aphids, spider mites, thrips, and whiteflies (continued)

714

19

Commercially Available Predators

Table 19.5 (continued) Company name Beneficial Insectary, Inc. Beneficial Insectary, Inc. Beneficial Insectary, Inc. Beneficial Insectary, Inc.

Country USA

Natural predators Adalia bipunctata

Targeted pests Aphid (all stages), spider mites, and mite eggs

USA

Chrysopa carnea

Aphids, mealybugs, spider mites, thrips, whiteflies, and other small caterpillars

USA

Cryptolaemus montrouzieri

Mealybugs, aphids, and scale bugs

USA

Delphastus catalinae Macrolophus pygmaeus

Tobacco whiteflies

Micromus angulatus Nesidiocoris tenuis Orius laevigatus Bioline AgroSciences

USA

Crop Defenders

FAR, Inc.

Natural Solution

Australia

Aphidoletes aphidimyza Chrysoperla sp. Feltiella acarisuga Dicyphus hesperus Atheta coriaria Orius insidiosus Delphastus catalinae Chrysoperla Cryptolaemus montrouzieri Cryptolaemus montrouzieri Brown lacewing Harmonia conformis Harmonia octomaculata Mallada signata (green lacewings)

Tobacco whiteflies, whiteflies, Tuta absoluta, spider mites, moth eggs, aphids, and red spider mites Mealybugs and aphids Whiteflies and red spider mites Thrips stages, aphids, whiteflies, and moth eggs Aphids Insects, mites, and aphids Spider mites Thrips, moth eggs and mites, and whiteflies Shore flies and thrips pupae Thrips Whiteflies Aphids Mealybugs Mealybugs Coccid insects, mealybugs, whiteflies, aphids, mites, psyllids, and insect eggs Aphids and whiteflies, thrips (including Western flower thrips), scales, caterpillars, and pest mites Aphids and whiteflies, thrips (including Western flower thrips), scales, caterpillars, and pest mites Aphids, whiteflies, mealybugs, and scales Thrips, caterpillars, and moth eggs (continued)

19.3

Quality Control

715

Table 19.5 (continued) Company name

Country

Natural predators Orius tantillus

Targeted pests Thrips (adult and larval stages, including Western flower thrips), spider mites, whiteflies, aphids, and caterpillars

19.3.2 Stock Deterioration The performance and survival of cultures should be periodically assessed during mass production. Recent research has suggested that (1) the timing of stock collection in the field and (2) the periodic intervention of diapause or cold storage may prevent Chrysoperla stock from degrading during continuous raising, even though this is possible. As an illustration, adults caught earlier in the season had better offspring than did those caught later in the season. Additionally, the introduction of diapause improved the ability of offspring descended from late-season cohorts to reproduce. These ecophysiological findings could be effectively used to standardise the calibre of commercial stock with some targeted investigation.

19.3.3 Shipping and Handling To avoid cannibalism and hatching while in transit, eggs that are ready to be sold should be delivered in insulated containers with cold packs as soon as possible after oviposition. Packaging with suitable packing material and food, such as Sitotroga or Anagasta eggs, can lower cannibalism and mortality when pre-fed larvae are ordered. Here, more in-depth investigation is required. It is crucial for the insectary sector to provide standards that support the dependability of commercially generated natural enemies when looking at the quality control issue as a whole. In this aspect, it appears that the European insectary industry and scientists work together more effectively and cooperatively than do their American counterparts. Both the use of biological control and users’ view of biological control as a dependable pest management strategy depend on the standardisation of high-quality natural enemy production (e.g. Tauber and Tauber 1975; Leppla and Fisher 1989; Bigler 1992). However, because there are no stringent quality control regulations in the United States, the quality of naturally occurring enemies that are sold commercially can vary. For instance, a recent analysis of shipments from insectaries has revealed that growers’ orders for C. carnea were not regularly completed with the right species, and cannibalism dramatically decreased the survival of lacewings while in transit (O’Neil et al. 1998). According to Leppla and Fisher (1989) and Bigler (1992), such issues can be resolved by improved mass production and packing techniques and increased caution in maintaining correctly labelled, pure colonies. Furthermore, they advise doing the following:

716

19

Commercially Available Predators

1. Verification of the species’ identity 2. Stock deterioration 3. Delivery and handling The following insects were used as an additional biological control measure globally based on the quality of the predators (van Lenteren 2012). The majority of predators have smaller (hundreds to a few thousands of individuals sold per week) market values than do those with larger (hundred thousand to millions of individuals sold per week) and more (ten thousand to a hundred thousand of individuals sold per week) moderate market values. Predators are used by farmers to control pests across Europe, North and South Africa, North and Latin America, Asia, Australia/New Zealand, and Turkey, which are the most well-known regions. Europe has the unique opportunity to practise sustainable agriculture by making use of the majority of predators that are sold commercially. Predators were used to manage pest insects such as aphids, coccids, coleopterans, diaspidids, dipterans, heteropterans, lepidopterans, margarodids, mites, pseudococcids, scales, thrips, and whiteflies. Between 1985 and 1992, thysanopteran predators such as Aeolothrips intermedius, Aleurodothrips fasciapennis, Franklinothrips megalops, Franklinothrips vespiformis, Karnyothrips melaleucus, and Scolothrips sexmaculatus were introduced to Asia, Europe, and North America to manage pest insects like thrips, diaspidids, and mites (Table 19.6). Nesidiocoris tenuis and Macrolophus pygmaeus are currently mass-produced and commercialised for augmentation, whereas the European Dicyphus species provide biological control services, mostly through conservation strategies (Castañé et al. 2004; Ingegno et al. 2017).

19.4

Future Focus

• A database should be constructed and maintained. • All predators should be subjected to quality control. • College students should be taught the value of commercial production of natural enemies, and technological transfer should be made available.

19.4

Future Focus

717

Table 19.6 Commercialised predators utilised at virus regions like South Africa is south of the Sahara, whereas North Africa is north of the Sahara. North America is made up of Canada and the United States, Australia, and New Zealand Predator Coleoptera Rodolia cardinalis Nephus includens Nephus reunioni Rhyzobius chrysomeloides Rhyzobius forestieri Scymnus rubromaculatus Hippodamia convergens Coccinella septempunctata Aleochara bilineata Clitostethus arcuatus Chilocorus bipustulatus Diomus sp. Exochomus laeviusculus Exochomus quadripustulatus Hippodamia variegata Chilocorus circumdatus Chilocorus baileyi Chilocorus nigritus Carcinops pumilio Cybocephalus nipponicus Dalotia (Atheta) coriaria Harmonia axyridis

Region where used

Target pest(s)

Initial year of use

Market value

Europe

Margarodids

1990

Small

Pseudococcids

2000 1990 1980

Coccids

1980 Aphids

1990 1993 1980

Root flies

1995

Whiteflies

1997

Diaspidids

1992–2005

Scales Aphids, scales

1990 1988

Small

2000 Australia

Aphids

2000

Europe, Australia

Diaspidids

1902 1992

Europe, Africa south

Diaspidids

1985

North America

Dipterans

1990

North America

Scales

2000

Europe, North America, Asia, Australia Europe, except France

Dipterans, thrips Aphids

2000 1995–2005

North America, Asia

Aphids

1990

Large

(continued)

718

19

Commercially Available Predators

Table 19.6 (continued) Predator Harmonia axyridis Rhyzobius lophanthae Adalia bipunctata Delphastus catalinae Delphastus pusillus Stethorus punctillum Cryptolaemus montrouzieri

Hymenoptera Aphelinus mali Aphelinus varipes Aphidius urticae Lysiphlebus testaceipes Anagrus atomus Anagyrus dactylopii Anagyrus fusciventris Anaphes iole Cales noacki Coccophagus cowperi Aphytis diaspidis Coccidencyrtus ochraceipes Comperiella bifasciata Dicyphus errans Dicyphus hesperus Encarsia citrina Encarsia guadeloupae Encarsia hispida Encarsia protransvena

Region where used

Target pest(s)

Initial year of use

Market value

Europe, North America

Coccids

1980

Small

Aphids Whiteflies

1998 1985 1993

Medium

Europe, North America, Asia Europe, Africa north and south, North and Latin America, Asia, Australia/New Zealand, Turkey

Mites

1984

Small

Coccids, pseudococcids

1917

Large

Europe

Aphids

1980, 1990, 2000

Small

Cicadellids Pseudococcids

1990 1995

Heteropterans Whiteflies Coccids, pseudococcids Diaspidids

1990 1970 1985

Small

1990 1995 1985

Dipterans Whiteflies

2000 2000–2005

Diaspidids Whiteflies

1984 1990–2000

Large Small

1990–2000 1990–2005 (continued)

19.4

Future Focus

719

Table 19.6 (continued) Predator Encarsia tricolor Eretmocerus eremicus Encyrtus infelix Encyrtus lecaniorum Metaphycus stanleyi Metaphycus swirskii Synacra paupera Leptomastix epona Leptomastix histrio Gyranusoidea litura Anthocoris nemorum Tetracnemoidea brevicornis Tetracnemoidea peregrina Tetrastichus coeruleus (asparagi) Meteorus gyrator Aphelinus abdominalis Aphelinus asychis Aphidius gifuensis Neochrysocharis formosa Aphidius colemani Aphidius ervi Aphidius transcaspinus Brontocoris tabidus Cephalonomia stephanoderis

Region where used

Target pest(s)

Coccids

Initial year of use 1985 1995–2002 1990 1985

Market value Large Small

1990 1995 Sciarids Pseudococcids

2000 1992 1995 1990

Medium

Psyllids, thrips

1992

Small

Pseudococcids, Margarodids Pseudococcids, Margarodids Coleopterans

1990

Small

2000

Lepidopterans Aphids

2005 1992

Medium

Aphids

2005 2005

Small Medium

Dipterans

1990

Medium

Europe, Africa north and south, North America, Asia, Australia/New Zealand Europe, Africa north, North and Latin America, Asia Africa south

Aphids

1991

Large

Aphids

1996

Large

Aphids

2005

Medium

Latin America

Lepidopterans

1990

Small

Latin America

Coleopterans

1990

Large

Europe

Europe, Africa north, North America, Asia Asia

1990

(continued)

720

19

Commercially Available Predators

Table 19.6 (continued) Predator Coccidoxenoides perminutus Dacnusa sibirica Diglyphus isaea

Dicyphus hesperus Urolepis rufipes Goniozus legneri Eretmocerus corni Telenomus remus Lixophaga diatraeae Metaphycus spp. Peru Diglyphus begini Lydella minense Thripobius semiluteus Metaphycus helvolus Metaphycus lounsburyi Eretmocerus eremicus Eretmocerus mundus Eretmocerus warrae Leptomastix dactylopii Leptomastidea abnormis Metaphycus flavus Muscidifurax zaraptor Nasonia vitripennis

Region where used Europe, Africa north and south Europe, Africa north, North and Latin America, Asia Europe, Africa north and south, North and Latin America, Asia North America

Target pest(s) Diaspidids, pseudococcids Dipterans

Initial year of use 1995

Market value Small

1981

Large

Dipterans

1984

Whiteflies

1995

Medium

Dipterans Lepidopterans Whiteflies

1990 1990 2000

Small

Lepidopterans Lepidopterans

1990 1980

Large

Coccids

1990

Small

Dipterans Coleopterans Thrips

2000 1990 1995

Medium large Small

Coccids

1943

Coccids

1902

Africa north and south, North and Latin America, Asia Europe, Africa north and south, North and Latin America, Asia Australia

Whiteflies

1995

Whiteflies

2001

Whiteflies

2000

Europe, Africa north, North America Europe, North America

Pseudococcids

1984

Medium

Pseudococcids

1984

Small

Coccids

1995

Small

Dipterans

1982

Medium

Dipterans

1970

Small

Pseudococcids

1995

Latin America

Europe, Australia

Large

(continued)

19.4

Future Focus

721

Table 19.6 (continued) Predator Anagyrus pseudococci Aphidius matricariae Muscidifurax raptorellus Ooencyrtus kuvanae Ooencyrtus pityocampae Ophelosia crawfordi Opius pallipes Praon volucre Pseudaphycus angelicus Pseudaphycus flavidulus Pseudaphycus maculipennis Psyttalia concolor Scutellista caerulea (cyanea) Spalangia cameroni Spalangia endius Pachycrepoideus vindemiae Paratheresia claripalpis Pheidole megacephala Prorops nasuta Prospaltella spp. Orgilus obscurator Pediobius foveolatus Peristenus digoneutis Spalangia gemini

Target pest(s)

Initial year of use

Market value

Aphids

1980

Medium

Africa north, North America

Dipterans

1970

Medium

Europe

Lepidopterans

1923

Small

Lepidopterans

1997

Small

Coccids, pseudococcids, margarodids Dipterans Aphids Pseudococcids

1980

Pseudococcids

1990

Pseudococcids

1980

Dipteran

1968–2000

Coccids

1990

Dipterans

1970

Dipterans

1970

Dipterans

1980

Lepidopterans

1980

Coleopterans

1990

Coleopterans Diaspidids Lepidopterans

1990 1990 1990

Medium Large

Coleopterans

1980

Small

Heteropterans

1980

Dipterans

1980

Region where used

Africa north, North and Latin America North and Latin America, Australia

North America

1980 1990 1990

Large

(continued)

722

19

Commercially Available Predators

Table 19.6 (continued) Predator Spalangia nigroaenea Aphidoletes aphidimyza, Diptera Coenosia attenuata Episyrphus balteatus Lysiphlebus fabarum Feltiella acarisuga Ophyra aenescens Neuroptera Chrysoperla carnea Chrysoperla externa Chrysoperla spp. Peru Chrysoperla rufilabris Coniopteryx tineiformis Conwentzia psociformis Sympherobius fallax Mallada signata

Region where used

Target pest(s) Dipterans

Initial year of use 1980

Market value

Europe, Africa north and south, North America, Asia

Aphids

1989

Large

Europe

Dipterans, whiteflies Aphids

1996

Small

1990

Medium

Aphids

1990

Small

Europe, North and Latin America Europe, North America

Mites

1990

Medium

Dipterans

1995

Small

Europe, Africa north, North and Latin America, Asia Latin America

Aphids

1970

Medium

Lepidopterans

1980

Large

Latin America

Aphids

1990

Europe, North America

Aphids

1970

Europe

Aphids, mites, scales Aphids, mites, scales Pseudococcids

1990–2005

Aphids, thrips, lepidopterans, mealybugs, whiteflies Aphids, thrips, lepidopterans, whiteflies, etc. Aphids

2000

Large

2000

Small

Australia

Micromus tasmaniae

Australia

Micromus angulatus Heteroptera Macrolophus pygmaeus Geocoris punctipes Nesidiocoris tenuis

Asia

Europe, Africa north and south North America Europe, Africa north, Asia

Small

1990–2005 1994

2005

Whiteflies

1994

Large

Lepidopterans, whiteflies Whiteflies, lepidopterans

2000

Small

2003

Large (continued)

19.4

Future Focus

723

Table 19.6 (continued) Predator Macrolophus caliginosus Nabis pseudoferus ibericus Orius albidipennis Orius minutus Orius tristicolor Orius insidiosus Orius majusculus Picromerus bidens Orius armatus Orius strigicollis Orius insidiosus Orius laevigatus Podisus maculiventris Podisus nigrispinus Thysanoptera Franklinothrips megalops Aeolothrips intermedius Aleurodothrips fasciapennis Karnyothrips melaleucus Franklinothrips vespiformis Scolothrips sexmaculatus

Region where used Europe Europe

Australia Asia North and Latin America Europe, Africa north, Asia Europe, North America

Target pest(s) Whiteflies, lepidopterans Lepidopterans

Initial year of use 2005

Market value Medium

2009

Small

Thrips

1993

Lepidopterans

1993 1995–2000 1991–2000 1993 1990

Thrips

1990 2000 1985 1993 1996

Latin America

Coleopterans, lepidopterans Lepidopterans

Europe

Thrips

1992

Thrips

2000

Diaspidids

1990

Diaspidids

1985

Europe, Asia

Thrips

1990

Europe, North America

Mites, thrips

1990

Source: van Lenteren (2012) and Carvalho et al. (2021)

Large Medium Small

Medium Large Small

1990

Small

724

19

Commercially Available Predators

Annexure 19.1: Contact Address for Selected Commercial Producers 1. ARBICO Arizona Biological Control, Inc., Rick Frey, PO Box 8910, Tucson, AZ 857380910, USA Phone: 800.827.2847/FX 520.825.208 Website: http://lwww.arbico-organics.com 2. Beneficial Insectary 9664 Tanqueray Ct., Redding, CA 96003 Phone: (530) 226-6300/(800) 477-3715 Fax: (530) 226-6310/(888) 472-0708 Email: [email protected] Website: http://www.insectary.com 3. Natural Insect Control Susan Cavey 3737 Netherby Road, Stevensville, Ontario LOS 1SO, Canada Phone: 905.382.2904 Fax: 905.382.4418 Website: http://www.naturalinsectcontrol.com 4. Associates Insectary Brett Chandler PO Box 969, Santa Paula, CA 93061-0969, USA Phone: 805.933.1301 Fax: 805.933.1304 Website: http://www.associatesinsectary.com 5. Biobest Biological Systems 2020 Fox Run Rd., Leamington, ON N8H 3V7, Canada Phone: (519) 322-2178 Fax: (519) 322-1271 Email: [email protected] Website: http://www.biobest.ca 6. National Bureau of Agricultural Insect Resources P. Bag No: 2491, H.A. Farm Post, Bellary Road, Bengaluru, 560 024, Karnataka, India Phone: +91(080)-2351 1982/98 Fax: +91(080)-2341 1961 Email: [email protected] Website: https://www.nbair.res.in/index.php/contact-us 7. A-1 Unique Insect Control 5504 Sperry Dr. Citrus Heights, CA 95621 Phone: 916-961-7945 Fax: 916-967-7082 Email: [email protected] Website: http://www.a-1unique.com

Annexure 19.1: Contact Address for Selected Commercial Producers

8. Biobee USA Inc. Rami Ben Dor 5126 S. Royal Atlanta Rd., Tucker, GA 30084, USA Phone: 770.274.2412 Website: http://www.biobee.com/ 9. Natural Enemies Shane Young 15648 SE 114th Ave. Suite #201, Clackamas, OR 97015, USA Phone: 503.342.6698 Website: http://naturalenemiesbiocontrol.com/ 10. American Insectaries, Inc. 243 S. Escondido Blvd., #318, Escondido, CA 92025 Phone: (760) 747-2920 Fax: (760) 498-0353 Email: [email protected] Website: http://www.betterbugs.com 11. Crop Defenders US Address 30600 Telegraph Rd., Suite 2345, Bingham Farms, MI 48025 Phone: 866.300.2929 Fax: 248.659.1911 Website: www.cropdefenders.com 12. Nature’s Control 3960 Jacksonville Hwy., P.O. Box 35, Medford, OR 97501 Phone: (541) 245-6033 Fax: (800) 698-6250 Email: [email protected] Website: http://www.naturescontrol.com 13. Applied Bio-Control P.O. Box 118 Waterford, CA 95386 209-874-1862 Fax: 209-974-1808 Email: [email protected] 4, 13, 15, 21, 22, 40 14. Dynamic Ecosystems Crop Supply Bejay Mills 829 Beaver Pt. Rd., Saltspring, BC V8K 1X9, Canada Phone: 250.217.0000 Website: http://www.dynamicecosystems.ca 15. Orcon Organic Control Inc. Steve Hazzard, 350 West Sepulveda Blvd., Carson, CA 90745, USA Website: https://organiccontrol.com/ 16. Anatis Bioprotection Silvia Todorova 278 rang Saint-Andre, St-Jacques-le-Mineur, Quebec J0J1Z0, Canada

725

726

17.

18.

19.

20.

21.

22.

23.

24.

19

Commercially Available Predators

Phone: 514.577.0817 Fax: 514.485.0323 Website: http://anatisbioprotection.com/en/ Evergreen Growers Supply, LLC. John Maurer 17592 S. Palmer Road, Oregon City, OR 97045, USA Phone: 503.908.1946 Fax: 503.908.1946 Website: http://www.evergreengrowers.com Peaceful Valley Farm Supply P.O. 2209, Grass Valley, CA 95945 Phone: (530) 272-4769/(888) 784-1722 Email: [email protected] Website: http://www.groworganic.com American Lake Doctors and Queen of the River Fish Co., Inc. 3810 N. 115th St. Longmont, CO 80501 Phone: 303-651-2514, 800-422-2514 Fax: 303-651-2224 Everwood Farm David Quillin 7937 Lakeside Dr. NE, Brooks, OR 97305, USA Website: http://www.everwoodfarm.com/ Plant Products Tom MacDonald 50 Hazelton St., Leamington, Ontario N8H 3W1, Canada Website: http://www.plantproducts.com Applied Bio-Nomics Ltd. Brian Spencer 11074 W. Saanich Rd., Sidney, BC V8L 5P5, Canada Phone: 250.656.2123 Fax: 250.656.3844 Website: http://www.appliedbio-nomics.com Evergreen Growers Supply 15875 SE 114th Ave, Building 1 Suite G, Clackamas, OR 97015 Phone: 503-908-1946 Email: [email protected] Website: https://www.evergreengrowers.com/aboutus Planet Natural 1612 Gold Ave., Bozeman, MT 59715 Phone: (800) 289-6656/(406) 587-5891 Fax: (406) 587-0223 Email: [email protected] Website: http://www.planetnatural.com

Annexure 19.1: Contact Address for Selected Commercial Producers

25. ARBICO Environmentals P.O. Box 8910 Tucson, AZ 85738 Phone: 800-827-2847, 602-825-9785 (consultant) Beneficial Insectary 9664 Tanqueray Ct. Redding, CA 96003 Phone: 530-226-6300, 800-477-3715 Fax: 530-226-6310 Website: http://www.insectary.com 26. FAR, Inc. Joe Barcinas 550 Foothill Pkwy., Corona, CA 92882, USA Phone: 951.371.0120 Fax: 951.279.5150 Website: https://www.far-inc.com/ 27. Rincon-Vitova Insectaries, Inc. P.O. Box 1555, Ventura, CA 93022 1555 Phone: (800) 248-2847 Fax: (805) 643-6267 Email: [email protected] Website: http://www.rinconvitova.com 28. Andermatt Biocontrol Suisse Ag Stahlermatten 6, 6146 Grosdietwii, Switzerland Website: http://www.biocontrol.ch/ 29. GrowLiv Biologicals Ghulam Mustafa 3665 3rd Concession rd N, Amherstburg, Ontario N9V2Y9, Canada Phone: 151.999.23547 Website: www.growliv.com/ 30. Spalding Laboratories, Inc. Tom Spalding P. O. Box 10,000, Reno, NV 89510, USA Phone: 408.929.9577 Website: http://www.spalding-labs.com 31. Bio Control, S.A. Freddy Piedra Apartado postal 164-7050, Cartago, Costa Rica Phone: 506.2592.4292 Mobile: 506.8305.0505 Fax: 506.2591.9607 Website: http://biocontrol.cr/ 32. Global Horticultural, Inc Jennifer Blom 4222 Saan Road, Beamsville, Ontario L0R 1B1, Canada Phone: 905.563.3211 Fax: 905.563.3191 Website: http://www.globalhort.com/

727

728

19

Commercially Available Predators

33. Sound Horticulture Alison Kutz 1050 Larrabee Ave. Suite 104 #365, Bellingham, WA 98229, USA Phone: 360.739.9095 Fax: 360.715.9264 Website: http://www.soundhorticulture.com 34. Biocontrol Network 5116 Williamsburg Rd., Brentwood, TN 37027 Phone: 615-370-4301, 1-800-441-2847 Fax: 615-370-0662 Email: [email protected], [email protected] Website: http://www.biocontrolnetwork.com 35. GrowLiv Biological 3665 3rd Concession rd North, Amherstburg, ON, Canada, N9V2Y9 Phone: 519-992-3547 Email: [email protected] Website: https://www.growliv.com 36. Tip Top Biocontrol David Duddy PO Box 7614, Westlake Village, CA, USA Website: https://tiptopbiocontrol.com/ Phone: 805.445.9001 37. Beneficial Insectary Canada 60 Taggart St. Guelph, Ontario, Canada N1H6H8 519-763-8653 Fax: 519-763-9103 13, 21, 40 Better Yield Insects 44 Bristol Rd. Narragansett, RI 02882 Phone: 401-792-3416, 800-662-6562 Fax: 401-792-8058 38. Green Methods.Com 9664 Tanqueray Ct., Redding, CA 96003, USA Phone: 800.477.3715 Website: http://www.greenmethods.com 39. The Agricultural Promotion & Investment Corporation of Odisha Limited Baramunda, Bhubaneswar, Odisha, PIN-751003, India Email: [email protected] 40. Buglogical Control Systems P.O. Box 32046, Tucson, AZ 85751-2046 Phone/Fax: (520) 298-4400 Email: [email protected] Website: http://www.buglogical.com 41. Hydro-Gardens P.O. Box 25845, Colorado Springs, CO 80936-5845 Phone: (888) 693-0578 Email: [email protected] Website: http://www.hydro-gardens.com

Annexure 19.1: Contact Address for Selected Commercial Producers

42. UAV-IQ Precision Agriculture Neuman, Andreas 10250 Constellation Blvd, Ste. 100, Los Angeles, CA 90067, USA Phone: 831.272.4786 Website: https://www.uaviq.com 43. Bioline Agrosciences Ascencion Urquidez P. O. Box 2430, Oxnard, CA 93034 Website: https://www.biolineagrosciences.com/ 44. Hydro Gardens-HGI Worldwide Mike Morton 8765 Vollmer Rd., Colorado Springs, CO 80908, USA Phone: 719.495.2266 Website: http://www.hydro-gardens.com 45. Wainwright-Evans, Suzanne Buglady Consulting 4660 Trestle Lane, Slatington, PA 18080, USA Website: http://www.bugladyconsulting.com 46. BioWorks 100 Rawson Road, Suite 205, Victor, NY 14564, USA Phone: 585.433.9537 Website: http://www.bioworksinc.com/ 47. Kunafin “The Insectary” Frank Junfin PO Box 190, 13955 U.S. Hwy 277, Quemado, TX 78877, USA Phone: 830.757.1181 Fax: 830.757.1468 Website: http://www.kunafin.com 48. Bioyolojik Tarim Dan ve Hiz Koll Sti (Biotar) Cemal Pasa Mahallesi Bahar Caddesi Eras Apart. No27 K1 D6 01120, Seyhan-Adana Turkiye Turkey 49. IPM Laboratories Carol Glenister 980 Main Street, Locke, NY 13092, USA Phone: 315.497.2063 Fax: 315.497.3129 Website: http://www.ipmlabs.com 50. Natural Solutions Australia Pty Ltd 11 Jurgens Place, Bowen Queensland (QLD) 4805, Australia 51. Project Directorate Biological control, ICAR, 2HGM + QPR Bellary Road, H.A. Farm Post, Hebbal, Bengaluru, Karnataka 560024 Phone: +91(080)-2351 1982/98 Fax: +91(080)-2341 1961 Email: [email protected]/[email protected] Website: https://aicrp.icar.gov.in

729

730

19

Commercially Available Predators

52. Kimberly Stoner Department of Entomology Connecticut Agricultural Experiment Station 123 Huntington Street, P.O. Box 1106, New Haven, CT 06504-1106 Phone: (203) 974-8458 Fax: (203) 974-8502 Email: [email protected] 53. Central Institute for Cotton Research (CICR) Post Bag No.2, Shankar Nagar Post Office, Nagpur, 440 010, Maharashtra, India Phone: (07103) 275536 Office: (07103) 275537, 275538, 275539, 275549, 275617 Fax: (07103) 275529 Email: [email protected] Website: www.cicr.org.in

Annexure 19.2: Company Website 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27.

A-1 Unique Insect Control, http://a-1unique.com/ American Insectaries, Inc., http://www.americaninsectaries.com/ Anatis Bioprotection, http://anatisbioprotection.com/en/ Andermatt Biocontrol Suisse Ag, http://www.biocontrol.ch/ Applied Bio-nomics, http://www.appliedbio-nomics.com ARBICO Organics, http://www.arbico-organics.com Arbico Organics, https://www.arbico-organics.com/ Associates Insectaries, http://www.associatesinsectary.com/ Associates Insectary, http://www.associatesinsectary.com BASF Agricultural Specialties, https://agriculture.basf.com/us/en.html Beneficial Insectary, http://www.insectary.com BFG Supply, https://www.bfgsupply.com/ Bio Control, S. A., http://biocontrol.cr/ BioBee USA, http://www.biobee.com/ Biobest Canada, http://www.biobestgroup.com Biobest, http://www.biobestgroup.com/en/ Biobest Mexico, http://www.biobestgroup.com Biobest USA, http://www.biobestgroup.com Biofac, http://www.biofac.com/ Bioline AgroSciences, https://www.biolineagrosciences.com/ Biotactics, http://www.benemite.com BioWorks, http://www.bioworksinc.com Buglogical Control Systems, http://www.buglogical.com Buglogical Tucson Arizona, https://www.buglogical.com/ Crop Defenders, http://www.cropdefenders.com/ Dynamic Ecosystems Crop Supply, http://www.dynamicecosystems.ca Evergreen Growers Supply, http://www.evergreengrowers.com

Annexure 19.3: Government, University, Commercial and Non-profit. . .

28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58.

731

Everwood Farm, http://www.everwoodfarm.com/ Foothill Agriculture Research (FAR Inc.), https://www.far-inc.com/ Garden Alive, https://www.gardensalive.com/ Global Horticultural, http://www.globalhort.com/ Green Methods, http://greenmethods.com Greenmethods Redding, https://greenmethods.com/ GrowLiv, http://www.growliv.com/ Hydro-Gardens, http://www.hydro-gardens.com IPM Laboratories, http://www.ipmlabs.com Koppert Biological Systems, http://www.koppert.com Koppert Howell Michigan, https://www.koppertus.com/ Kunafin “The Insectary” http://www.kunafin.com Kunafin Quemado, http://www.kunafin.com/ Natural Enemies, https://naturalenemies.com/ Natural Insect Control, http://www.naturalinsectcontrol.com/ Natural Pest Control, http://www.natpestco.com/ Nature’s Control, http://www.naturescontrol.com/ Orcon (Organic Control), http://organiccontrol.com/ Organic Control, https://www.organiccontrol.com/ Peaceful Valley, https://www.groworganic.com/ Planet Natural, https://www.planetnatural.com/ Plant Products, https://www.plantproducts.com/ Rincon-Vitova, http://www.rinconvitova.com Rincon-Vitova Insectaries, http://www.rinconvitova.com/ Sierra Biological Inc., http://www.sierrabiological.com/ Sound Horticulture, http://soundhorticulture.com/ Spalding Laboratories, https://www.spalding-labs.com Tip Top Bio-Control, http://www.tiptopbiocontrol.com/ WestGrow Biological Solutions, https://www.thebuglady.ca/ BITOTAR Biolojik Tarim Dan ve muh Hiz Koll Sti, http://www.biotar.com Natural Solutions Australia., http://www.naturalsolutions.com.au

Annexure 19.3: Government, University, Commercial and Non-profit Websites on Biological Control (O’Neil et al. 2003) United States 1. National Biological Control Institute, www.aphis.usda.gov/ppq/nbci/ 2. USDA Insect Biological Control Laboratory, www.barc.usda.gov/psi/ibl/ iblhome.htm 3. Cornell University Biological Control Guide, www.nysaes.cornell.edu/ent/ biocontrol/

732

19

Commercially Available Predators

4. Purdue Entomology Biological Control, www.entm.purdue.edu/Entomology/ research/bclab/BCMAC.HTML 5. Biological Control at Michigan State University, www.cips.msu.edu/biocontrol/ 6. UC Berkeley Center for Biological Control, www.CNR.Berkeley.EDU/biocon/ 7. University of California Statewide IPM Project, www.ipm.ucdavis.edu/ 8. Midwest Institute for Biological Control, www.inhs.uiuc.edu/cee/biocontrol/ home.html 9. Biological Virtual Information Center, http://ipmwww.ncsu.edu/biocontrol/ 10. Biological Control News, www.entomology.wisc.edu/mbcn/mbcn.html 11. Insect Parasitic Nematodes, www.oardc.ohio-state.edu/nematodes/ 12. The Association of Natural Bio-control Producers, www.anbp.org/ 13. International 14. The Biotechnology and Biological Control Agency, www.e-bbca.net/main.htm 15. International Organisation for Biological Control 16. Integrated Control of Noxious Animals and Plants (IOBC), www.iobc-wprs.org/ 17. CAB International, www.cabi.org/ 18. Centre de Recherche de l’Est sur les cereals et oleagineux, res2.agr.ca/ecorc/ isbi/biocont/libhomf.htm 19. Institut National de la Recherche Agronomique (INRA), inra.fr/Internet/ Hebergement/OPIE-Insectes/luttebio.htm T 20. Centro de Control Biologico de Plagas y Enfermedades, www.usfq.edu.ec/1 AGROEMPR/HOME.HTML 21. CPL Worldwide Directory of Agrobiologicals, www.cplscientific.co.uk/press/ wda-features.html 22. Institute of Arable Crops Research (Great Britain), www.iacr.bbsrc.ac.uk/iacr/ tiacrhome.html 23. Embrapa (Brazil), www.embrapa.br/ 24. FAO—Community Integrated Pest Management, www.communityipm.org/ 25. The Consortium for International Crop Protection (CICP), www.ipmnet.org/ 26. IPM Europe: IPM Working for Development Newsletter, www.nri.org/ IPMForum/ipmwd.htm 27. CIAT—Integrated Pest and Disease Management (IPDM), www.ciat.cgiar.org/ ipm/index.htm 28. European and Mediterranean Plant Protection Organization (EPPO), www. eppo.org/index.html 29. The Center for Agroecology and Sustainable Food Systems, www.agroecology. org/index.html

References Bigler F (1992) Quality control of mass reared arthropods. Proceedings, 5th Workshop of the IOBC Global Working Group, 25-28 March 1991. Swiss Federal Research Station for Agronomy, Z, Wageningen

References

733

Carvalho LM, Diniz AJF, Sul PE, São João Del Rei MG (2021) Antocorídeos como agentes de controle biológico: avanços e desafios. Controle alternativo de pragas e doenças: opção ou necessidade? EPIMAG, Belo Horizonte, p 47 Castañé C, Alomar O, Goula M, Gabarra R (2004) Colonization of tomato greenhouses by the predatory mirid bugs Macrolophus caliginosus and Dicyphus tamaninii. Biol Control 30:591– 597 Cuthbertson AG, Blackburn LF, Audsley N (2014) Efficacy of commercially available invertebrate predators against Drosophila suzukii. Insects 5(4):952–960 Globe Newswire (2019) Biocontrol Agents Market in North America and Europe Will Reach USD 7,591 Million By 2026: Zion Market Research Ingegno BL, Candian V, Psomadelis I, Bodino N, Tavella L (2017) The potential of host plants for biological control of Tuta absoluta by the predator Dicyphus errans. Bull Entomol Res 107: 340–348 Kairo MTK, Paraiso O, Gautam RD, Peterkin DD (2013) Cryptolaemus montrouzieri (Mulsant) (Coccinellidae: Scymninae): a review of biology, ecology, and use in biological control with particular reference to potential impact on non-target organisms. CAB Rev Perspect Agric Vet Sci Nutr Nat Resour 8:1–20 Khan AA, Qureshi JA, Afzal M, Stansly PA (2016) Two-spotted ladybeetle Adalia bipunctata L. (Coleoptera: Coccinellidae): a commercially available predator to control Asian citrus psyllid Diaphorina citri (Hemiptera: Liviidae). PLoS One 11(9):e0162843. https://doi.org/10.1371/ journal.pone.0162843 Lenteren JV (2003) Commercial availability of biological control agents. In: Quality control and production of biological control agents: theory and testing procedures. CABI Publishing, Wallingford, pp 167–179 van Lenteren JC (2012) The state of commercial augmentative biological control: plenty of natural enemies, but a frustrating lack of uptake. BioControl 57:1–20. https://doi.org/10.1007/s10526011-9395-1 Leppla NC, Fisher WR (1989) Total quality control in insect mass production for insect pest management 1. J Appl Entomol 108(1–5):452–461 O’Neil RJ, Giles KL, Obrycki JJ, Mahr DL, Legaspi JC, Katovich K (1998) Evaluation of the quality of four commercially available natural enemies. Biol Control 11(1):1–8 O’Neil RJ, Yaninek JS, Landis DA, Orr DB (2003) Biological control and integrated pest management. Cabi Publishing, Wallingford, pp 19–30 Parrella MP, Lewis E (2017) Biological control in greenhouse and nursery production: present status and future directions. Am Entomol 63(4):237–250 Tauber MJ, Tauber CA (1975) Criteria for selecting Chrysopa carnea biotypes for biological control: adult dietary requirements. Can Entomol 107:589–595 Tauber MJ, Tauber CA, Daane KM, Hagen KS (2000) Commercialization of predators: recent lessons from green lacewings (Neuroptera: Chrysopidae: Chrosoperla). Am Entomol 46(1): 26–38

Biosafety Assessment of Synthetic Pesticides

20

Contents 20.1 20.2 20.3

Other Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Systemic Insecticides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hemiptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.3.1 Pentatomidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.3.2 Lygaeidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.3.3 Anthocoridae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.3.4 Reduviidae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.4 Lacewings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.5 Dermapterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.6 Coleopterans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.7 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.8 Syrphids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.9 Field Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.10 Insect Growth Regulators (IGRs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.10.1 Botanical IGRs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.11 Neuropeptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.12 Indirect Effects of Pesticides on Natural Enemies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.12.1 Development, Adult Longevity, and Reproduction . . . . . . . . . . . . . . . . . . . . . . . . 20.12.2 Functional Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.12.3 Biological Traits: Bioefficacy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.13 Fungicides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.14 Beneficial Impacts: Hormesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20.15 Future Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

20.1

735 737 738 738 740 741 746 751 752 754 756 757 758 759 764 764 765 767 768 770 771 773 775 775

Other Molecules

Pesticides are poisonous substances with particular modes of action (MoAs) that are intended to destroy living things. The number of pesticides on the market, which come in thousands of chemical variants, is a reflection of the wide variety of target

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_20

735

736

20

Biosafety Assessment of Synthetic Pesticides

organisms and their distinctive biochemical and physiological features. The most popular way for eradicating insect pests is chemical management, and its application has grown. Compatibility with other biological control agents is necessary since insecticides are used all over the world to eradicate insect pests and are crucial to agricultural production. Pesticide exposure can occur through direct contact with the insecticide during application, interaction with residues on surfaces, and consumption of contaminated food. Pesticides are the most powerful weapons for managing pests because they are extremely effective, quick to work, handy to use, and typically affordable. However, the improper, inadequate, and indiscriminate use of insecticides has led to major problems such the development of resistance to pests, the revival of target species, the appearance of insect pests, the extermination of beneficial organisms, health concerns, and environmental harm. One must therefore carefully select the best components for one’s plant protection strategy. In an integrated control strategy, some insecticides with minimal toxicity to insects’ natural enemies had to be used. It is widely accepted that when pesticides are used in agriculture to control pests, weeds, and fungal diseases, they have an impact on the non-target species of plants and animals that are potentially susceptible to toxic substances. Studies on these unanticipated consequences on arthropod groups have mostly focused on the direct effects that pesticides have on terrestrial and aquatic species. The main devastating effect of synthetic pesticides is the direct death of beneficial insects. Various recent studies have looked at the non-lethal effects of pesticides on insects in reaction to findings of the harmful effects of low levels of these chemicals on pollinators. These inquiries are properly covered in several publications and articles on the topic. However, rigorously controlled mesocosms and microcosms that replicate the wild communities of plants and animals are where the majority of what we know about indirect effects comes from (Sánchez-Bayo 2021). These studies sometimes document indirect consequences. It is highly desired for pest species’ natural enemies to show signs of pesticide resistance. This makes it possible to combine more efficient biological management with synthetic and organic compounds used to control particular pests, especially secondary pest outbreaks. It is interesting to note that no heteropteran predator has been shown to exhibit field-evolved resistance. However, lady beetles and lacewings have been found to exhibit pesticide resistance in their predatory insect populations. Imidacloprid, a systemic nitromethylene analogue insecticide, is the most frequently used pesticide in apple orchards all over the world to eliminate aphids. When compared to traditional surfactants, organosilicone additives have lower-equilibrium surface tension and surface energy, which is crucial for effective leaf wetting and spray-drop retention on leaf surfaces. Insecticides and herbicides are routinely added to organosilicone surfactants to improve their physical and chemical qualities. Coccinellidae, Staphylinidae, and other common predator groups discovered via visual sampling during corn-growing seasons were unaffected by the imidacloprid treatment, but Carabidae only had moderate effects during 1 of the 5 years under study. In contrast, imidacloprid significantly reduced Heteroptera in Spain; however,

20.2

Systemic Insecticides

737

the reduction varied depending on the year (1997–2001) (Albajes et al. 2003). The formulas that follow are closely related to this chapter. . Percentage egg survival = number of adults emerged/number of eggs treated times 100 . Egg survival is equal to 100% of egg mortality. . The ratio of adults emerging to treated larvae, expressed as a percentage, is 100. . 100% larval mortality equals 100% larval survival . Adults emerged/pupae treated × 100 = percentage of pupae surviving . 100% of eggs are lost to mortality, whereas 100% of pupae survive. . Adult mortality rate = adult deaths/adult releases multiplied by 100%. Mortality is equal to 100 divided by the number of dead larvae. . Corrected percentage mortality = ((XY) (100Y ))/100. . where X is the percentage of deaths from treatment, and Y is the percentage of deaths from controls. . The number of mites in predator treatment divided by the number of mites in mite-alone treatment is 100. Percentage of inset suppression = 1. . Insect reduction percentage (IRP) = number of mites under control minus number of mites receiving treatment in which 100 mites are under control.

20.2

Systemic Insecticides

Due of the lack of any direct exposure, systemic insecticides have been touted as being largely non-toxic to natural enemies when administered to the soil or growth medium as drenches or granules (Mizell and Sconyers 1992). This might is not the case, though, as systemic poisons have the potential to have indirect effects on natural enemies like Podisus maculiventris through an assortment of mechanisms, including the death of prey, the pollutant of floral parts with the active substance, the consumption of the ingredient while consuming plant fluids, and the contamination of prey ingesting either lethal or sub-lethal concentrations of the active ingredient (Tillman and Mullinix 2004). Systemic insecticides may not have many direct effects on aboveground predators when administered to the soil or growing medium, but if prey numbers are decreasing quickly, then they could have an indirect influence on natural predators. As a result, there are fewer or perhaps no longer any available prey that natural enemies can use as sustenance (Kiritani 1979), which makes it more challenging for them to find any survivors. The small pirate bug Orius spp., which may devour plants at some stage in their life cycle, could be negatively affected unintentionally by the systemic insecticide’s active ingredient dispersing into floral parts (petals and sepals) (Kiman and Yeargan 1985). Any indirect effects on parasitic insects that consume floral nectar and nectar may depend on the application type (soil vs. foliar) and potential timing (spatially and temporally) of the application. Research on the green lacewing Chrysoperla carnea and the pink lady beetle Coleomegilla maculate shows that systemic insecticides applied to flowers may have an indirect impact on natural enemies by altering when they

738

20

Biosafety Assessment of Synthetic Pesticides

forage (Rogers et al. 2007). Compared to the other treatments, the acephate treatment significantly reduced predator concentrations and was generally detrimental to the arthropod complex (says Bordini et al. 2021).

20.3

Hemiptera

20.3.1 Pentatomidae Numerous traditional pesticides have gradually been taken off the industry over the previous few decades and replaced with new synthetic insecticides. Spinosad, a macrocyclic lactone that relates to the spinosyns group, has an impact on insects by contact and ingestion. Saccharopolyspora spinosa yielded this chemical, which kills insects by over-excitating their nervous systems (Actinomycetales: Pseudonocardiaceae). In terms of its neurotoxic effect, spinosad is similar to the neonicotinoid pesticides class, but it differs at the site of nicotinic acetylcholine receptor binding, affecting the GABA (gamma-aminobutyric acid) neurotransmitter agonist. Spinosad is said to be selective and low-risk when used in agriculture in small-holder farms in rotation with neonicotinoid pesticides; however, it has been found to have adverse effects on predatory and parasitic insects, bees, and other beneficial insects in agriculture. Insecticide tolerance is common among predatory insects, which highlights how crucial it is for integrated pest management (IPM) programmes to be successful. In contrast, exposure to insecticides can have unfavourable impacts on insect development, longevity, fecundity, and behaviour without actually killing the insects. Soybean leaves previously exposed to diamide chlorantraniliprole, deltamethrin, and methamidophos. According to the findings, tested insecticides can have an indirect impact on caterpillar predators by causing them to consume treated Anticarsia gemmatalis food (Lepidoptera: Erebidae). Over the course of 5 days, adults of Podisus nigrispinus were presented with caterpillars of Anticarsia gemmatalis that had been fed on for 12 h using sprayed soybean leaves. In IPM initiatives in the agro-ecosystem of soybeans, Podisus nigrispinus was found to be incompatible with spinosad and methamidophos. Podisus nigrispinus had a low toxicity to deltamethrin. However, more information might be required before recommending it for IPM. Because it is least poisonous to this predator, chlorantraniliprole is considered to be the most promising (de Castro et al. 2015). Permethrin toxicity was seen in Podisus nigrispinus in a different investigation, with a median lethal concentration (LC50) of 0.46 g L-1 and a survival rate of 47% after 72 h of treatment. After 6 h of pesticide treatment, the midgut’s histological alterations included cytoplasmic vacuolisation, an unevenly lined epithelium, and apocrine secretions in the lumen. Cytotoxic effects were seen in the three different regions of the midgut, including the release of granules and vacuoles into the lumen, the presence of autophagosomes, and expansion of the basal plasma membrane infolds. After being exposed for 12 h, midgut cells started to die. Permethrin has toxic effects, reduces longevity, and changes the histology and cytology of the

20.3

Hemiptera

739

Table 20.1 Impact of various insecticides on Podisus nigrispinus LC50/biosafety with references Insecticides Permethrin Deltamethrin Spinosad Permethrin Tebufenozide Thiamethoxam Imidacloprid

LC50/biosafety 0.46 μg L-1 (72 h) Safer 3.15 μg L-1 0.25 mg mL-1 5.71 mg mL-1 0.04 mg mL-1 3.75 mg mL-1

References Martínez et al. (2018) de Castro et al. (2015) dos Santos-Junior et al. (2019) Silva et al. (2020) Silva et al. (2020) Silva et al. (2020) Martínez et al. (2019)

midgut in the predator Podisus nigrispinus, suggesting that the cell stress brought on by this pesticide can compromise physiological processes like digestion and endanger the predator’s the capacity to act as a biological pest control agent. Permethrin’s low selectivity against non-target organisms, such as the predatory bug Podisus nigrispinus, suggests that its application in biological control should be more thoroughly scrutinised (Martínez et al. 2018) (Table 20.1). The pesticide spinosad was hazardous to Podisus nigrispinus; 32% of individuals survived after 48 h of exposure to LC50. The main histological changes in the salivary system were epithelial disorganisation, cytoplasmic vacuolisation, and apocrine secretion into the gland lumen. Cytotoxic consequences included the presence of autophagosomes, the release of granules and vacuoles into the lumen, an elevation in basal plasma membrane infoldings, and apoptosis. The Podisus nigrispinus salivary complex’s histology and cytology are altered by spinosad, which also induces toxicity and reduces the lifespan. The findings imply that the insecticide’s effects on cellular stress compromise Podisus nigrispinus’s ability to function as a biological pest control agent by affecting extra-oral digesting (dos Santos-Junior et al. 2019). In the first 6 h following pesticide exposure, this sub-lethal dose of imidacloprid caused histological changes in the midgut epithelium and cytotoxic consequences like irregular border epithelium, cytoplasmic vacuolisation, and apocrine secretions. After 12 h, the digestive cells in the midgut programmed cell death (apoptosis). These findings imply that imidacloprid may alter the physiology of Podisus nigrispinus, a biological control agent, reducing its capacity to successfully eat other insects (Martínez et al. 2019). The adverse impact on third instar nymphs of the predator Podisus nigrispinus and the effects of Bacillus thuringiensis (Bt), permethrin, tebufenozide (TEB), and thiamethoxam (TMX) were observed by Silva et al. in a laboratory study in 2020. By measuring their lethal amounts, pesticide toxicity for this insect was ascertained. A video tracking system and a respirometer were used to study the behaviour of the Podisus nigrispinus after exposure to insecticides. The consumption of prey and nymphs was measured 24 h after fasting. In free-choice testing, whether Podisus nigrispinus nymphs preferred prey treated with pesticides or those not treated with the same was assessed. Podisus nigrispinus nymphs are poisonous to the insecticides Bt (LC50 = 1.10 (0.83–1.46) mg mL-1), permethrin (LC50 = 0.25 (0.17–0.34) mg mL-1), tebufenozide (LC50 = 5.71 (4.17–7.57) mg mL-1), and

740

20

Biosafety Assessment of Synthetic Pesticides

thiamethoxam (LC50 = 0.04 (0.02–0.06) mg mL-1). The respiration rate of Podisus nigrispinus was lowered by Bt and the insecticides tebufenozide, permethrin, and thiamethoxam. The nymphs of this bug are affected by the insecticides thiamethoxam, tebufenozide, and permethrin. Podisus nigrispinus prefers less Bt-, permethrin-, and thiamethoxam-treated prey. In the control, tebufenozide, Bt, permethrin, and thiamethoxam treatments, the survival rates for the nymphs of this predator were 93.3%, 66.7%, 56.6%, 0%, and 0%, respectively. In addition, the use of neurotoxic pesticides to treat prey decreases the predatory power of this natural opponent (Silva et al. 2020). Pests and their natural enemies are not the only ones that insecticides affect. However, there is also some insect ineactin when we use insecticides. Indoxacarb’s lethality (LC50 = 2.62 g L-1 and maximum lethal concentration (LC90) = 6.11 g L-1) was demonstrated in concentration–mortality bioassays on adults of the Podisus distinctus species. When predatory insects were not exposed to indoxacarb, the survival rate was 100%; this rate dropped to 40.7% when they were exposed to 2.62 g L-1 and to 0.1% when they were treated with 6.11 g L-1. When given preys were tainted with Bt, permethrin, and thiamethoxam, Podisus nigrispinus survival and prey consumption are reduced. The midgut lumen of Podisus nigrispinus is histologically affected by tebufenozide, one of the four insecticides under investigation, and exhibits aberrant epithelial architecture, cytoplasm vacuolisation, and cell fragment release (Silva et al. 2021). After exposure to a pesticide, indoxacarb decreased Podisus distinctus respiration from 18.45 to 14.41 L CO2 h-1 at 2.62 g L-1 for up to 3 h, preventing them from eating and causing them to act hyperactively. Because indoxacarb harms the natural enemy, it should be more thoroughly evaluated before being used with Podisus distinctus for pest management in forestry (Batista et al. 2022). Indoxacarb’s effects on Podisus distinctus’s mortality, survival, respiration, preference, prey consumption, and locomotor activity were investigated by Batista et al. in 2022. When predatory insects were not exposed to indoxacarb, the survival rate was 100%; this rate dropped to 40.7% when they were exposed to 2.62 g L-1 and to 0.1% when they were treated with 6.11 g L-1. After exposure to an insecticide for up to 3 h, indoxacarb decreased Podisus distinctus’s respiration from 18.45 to 14.41 L CO2 h-1 at 2.62 g L-1, preventing it from eating and causing it to exhibit hyperexcitation (Batista et al. 2022).

20.3.2 Lygaeidae Broad-spectrum pesticides are expected to highly affect generalist predators. Imidacloprid, tebufenozide, azinphos-methyl, and spinosad are compatible with integrated pest control and can be safely used while preserving the community of beneficial insects, such as big-eyed bugs, as per experimental data reported by Elzen (2001). According to studies by Boyd and Boethel (1998) and Elzen (2001), malathion, chlorfenapyr, and fipronil residues are the most harmful to big-eyed bugs. Satoh et al. (2012) evaluated the toxicity of 11 agricultural pesticides on Geocoris varius nymphs in their second instar through topical treatment and oral

20.3

Hemiptera

741

administration. They discovered that six agricultural pesticides, including chlorfenapyr, were either fully risk-free or barely dangerous when applied topically. It has been demonstrated that some chemicals, including dichlorvos, methomyl, acetamiprid, pyridaben, and two insect growth regulators (IGRs), chlorfluazuron and lufenuron, are dangerous or just moderately harmful. The oral toxicity of dichlorvos, methomyl, acetamiprid, and pyridaben had no impact on nymphal survival. When compared to the untreated control, applying sub-lethal doses of triflumizole, pyridalyl, sulphur, and chlorfenapyr had no discernible impact on nymphal development (Kóbor 2020).

20.3.3 Anthocoridae The Orius species eat mites and a wide variety of tiny insects, including aphids, scales, thrips, and insect eggs. Since 1991, Orius laevigatus has been commercially sold in Europe and the United States, but Orius minutus, a Korean native, has not. Experimental conditions were used to study the effects of formamidines, avermectins, and norpyrethrates on Anthocoris nemoralis. The median lethal dose (LD50) values of anthocorid bugs were typically lower than those of pear psyllas, indicating that predators are much more susceptible to prey. Different ways to boost to the beneficial insect are indicated by the determination of residual activities of the test insecticides. Based on the results of this study, pear psylla can be controlled by spraying abamectin alone while anthocorid bug populations are active (Berrada et al. 1996). Orius insidiosus is the subject of information on the direct effects of insecticides and miticides (Pietrantonio and Benedict 1999; Shipp et al. 2000; Elzen 2001; Studebaker and Kring 2003a, b; Ashley et al. 2006; Bostanian and Akalach 2006), but there is little data on the direct effects of fungicides, pesticide combinations, and surfactants on Orius insidiosus. Furthermore, Petri dish experiments (bioassays) with exposure periods ranging from 24 (Stark et al. 1995; Pietrantonio and Benedict 1999; Studebaker and Kring 2003b), 48 (Shipp et al. 2000; James 2004; Gradish et al. 2011), and 72 h (Delbeke et al. 1997; Elzen 2001) can influence the survival or death of adults. Using a residual insecticide bioassay, two lab-raised predators—the sly bower bug, Orius insidiosus, and the large-eyed bug, Geocoris punctipes—were subjected to ten insecticides, including three more recent ones with unique mechanisms of action. These insects play a significant role as carnivores of various cotton economic pests. Azinphos-methyl, imidacloprid, spinosad, tebufenozide, pronil, endosulfan, chlorfenapyr, cybuthrin, profenofos, and malathion were among the insecticides examined. The responses of the two tested species to the pesticides varied significantly. Male Orius insidiosus was significantly less harmful to tebufenozide and cythrin than to malathion. Additionally, compared to malathion, tebufenozide was significantly less harmful to female Orius insidiosus. Chlorfenapyr, endosulfan, and pronil were substantially more harmful to male Geocoris punctipes than were imidacloprid, tebufenozide, and spinosad. Female Geocoris punctipes were significantly less toxic to spinosad, tebufenozide, and azinphos-methyl compared to pronil

742

20

Biosafety Assessment of Synthetic Pesticides

and endosulfan. When compared to other treatments, including the control, the spinosad therapy significantly increased Orius insidiosus fecundity. When compared to other treatments, including the control, consumption of eggs of the bollworm Helicoverpa zea by Orius insidiosus was much lower in the pronil, profenofos, and cybuthrin treatments. Malathion, profenofos, endosulfan, pronil, azinphos-methyl, and imidacloprid treatments significantly reduced the consumption of Helicoverpa zea eggs by Geocoris punctipes in comparison to the control. The number of eggs consumed by Geocoris punctipes did not significantly differ between the tebufenozide treatment and the control group. Spinosad’s decreased toxicity to Geocoris punctipes is in line with earlier reports. Based on these findings, the pesticides malathion, endosulfan, profenofos, pronil, and cybuthrin are not compatible with IPM of cotton pests, whereas imidacloprid, tebufenozide, azinphos-methyl, and spinosad should provide pest control while protecting beneficial species (Elzen 2001). In an experiment, the effectiveness of imidacloprid 4.6 SC (suspension concentrate [= flowable concentrate]) and cyantraniliprole 10 SE (soluble emulsions) and 20 SC against Orius insidiosus and Orius pumilio, the two most important thripscontrolling natural enemies in pepper, was assessed. Four to 6 weeks after being systemically administered, cyantraniliprole 20 SC exhibited no discernible effects on populations of tiny pirate beetles in the flowers. Prior to pepper flowering, applications comprised of two rates, each administered twice. Although thrips numbers in the flowers were not much affected, cyantraniliprole can be utilised in a similar way in a conservation biological control programme for thrips. However, when used as a transplant drench, imidacloprid significantly reduced minute pirate bug numbers in the flowers 4–6 weeks later and is incompatible with the conservation biological management scheme (Funderburk et al. 2013). The lethal and sub-lethal effects of insecticides on Orius insidiosus were examined in this study using a life table with eight treatments and 40 duplicates. Orius insidiosus eggs were submerged in aqueous solutions of the chemical compounds after being normally deposited in plant stems. Daily evaluations were made on egg viability, embryonic period length, nymphal survival, and nymphal period length. Mature insects were partnered, and the quality of their procreation was evaluated. Every day, both the quantity of eggs laid and the adult survival rates were evaluated. Abamectin, cartap hydrochloride, spirotetramat + imidacloprid, and flubendiamide were categorised as dangerous pesticides. Pymetrozine was classified as slightly dangerous, whereas pyriproxyfen (PYR) and rynaxypyr were classified as safe. While other pesticides had no effect on the population parameters, rm, GT, DT (tablets for direct application), and pyriproxyfen did (Moscardini et al. 2013). The two important predators of the tomato pinworm Tuta absoluta are two predatory pirate bugs, namely, Amphiareus constrictus and Blaptostethus pallescens (Hemiptera: Anthocoridae). A series of research toxicity tests were carried out to evaluate the risk of azadirachtin and chlorantraniliprole on the eggs, nymphs, and adults of these two predatory pirate bugs (Lepidoptera: Gelechiidae). A labelsuggested field rate of these two insecticides, which is intended for the control of Tuta absoluta, was applied to all three stages. The safe, acute toxicity for these stages

20.3

Hemiptera

743

is indicated by the fact that neither azadirachtin nor chlorantraniliprole altered the mortality of adult predators or the capacity of their eggs to hatch. However, the ability of predatory nymphs to mature was significantly impacted by azadirachtin and chlorantraniliprole. This decline may jeopardise the biological control of pests like Tuta absoluta and have a direct impact on the size of the predator population in the next cycle. Our findings generally recommend using these low-risk insecticides with caution in integrated pest management programmes that combine chemical and biological methods (Gontijo et al. 2015a, b). Orius insidiosus was tested with three aphicides: flonicamid, sulfoxaflor, and flupyradifurone. Both species were unaffected by flonicamid independent of life stage or exposure method. Orius insidiosus eggs hatch effectively when placed in sunflower stems sprayed with these two substances; however, several nymphs died when subjected to the treated stems, perhaps as a result of phytophagous behaviour that led to some insecticide absorption. Despite these effects, we come to the conclusion that both sulfoxaflor and flupyradifurone are probably quite safe compared to other broad-spectrum insecticides, making them potentially compatible with biological control and overall management of Melanaphis sacchari in sorghum grain (Barbosa et al. 2017). A commercially available natural adversary of western flower thrips, including Frankliniella occidentalis, is the sneaky flower bug, Orius insidiosus (Pergande). Chlorfenapyr, dinotefuran, spinosad, azadirachtin, azoxystrobin, buprofezin, cyantraniliprole, fenhexamid, kinoprene, rosemary, rosemary oil, peppermint oil, and cottonseed oil (trade name: Eco-Mite) were among the 16 pesticides tested that significantly reduced the survival of adult Orius insidiosus. There were no direct effects on the survival of adult Orius insidiosus across the four surfactant treatments (alcohol ethoxylate; polyether-polymethylsiloxane copolymer; alkylphenol ethoxylate, propylene glycol, alkyl amine ethoxylate, and sulphuric acid; and alkylphenol ethoxylate, alkylphenol polymerised resin, and linoleic) (Herrick and Cloyd 2017). The four pollutants (abamectin, spinosad, pyridalyl, and chlorfenapyr) commonly used to control western flower thrips populations had a direct impact on Orius insidiosus adults after 96 h (0–20% adult survival), indicating that they should not be used in conjunction with Orius insidiosus in western flower thrips management programmes. Orius insidiosus is harmed by abamectin (Shipp et al. 2000; Bostanian and Akalach 2006). The findings of our study with abamectin, which indicated 60% and 20% adult survival after 72 h and 96 h of exposure, respectively, coincide with Studebaker and Kring’s (2003b) finding that abamectin is severely detrimental to Orius insidiosus adults in Petri dish studies. Spinosad was not hazardous to adult Orius insidiosus when it was exposed to residues on cotton (Gossypium hirsutum L.) leaves, according to Studebaker and Kring (2003a), but it was damaging when it was exposed to adults in treated Petri dishes (Studebaker and Kring 2003b). Spinosad was not hazardous to Orius insidiosus, according to Pietrantonio and Benedict (1999); however, Jones et al. (2005) reported that adult Orius insidiosus exposed to 1-day and 8-day spinosad residues had 35% and 17% mortality rates, respectively. When exposed to variously aged residues on tomato (Solanum lycopersicum) (Biondi et al. 2012), spinosad and abamectin were both found to be severely damaging to Orius laevigatus in Petri dish trials (van de Veire and Tirry 2003).

744

20

Biosafety Assessment of Synthetic Pesticides

Spinosad exposure caused 47% (males) and 54% (females) of Orius insidiosus adults to die over a 72-h period (Elzen 2001). In our investigation, adult Orius insidiosus survived for 40% and 20% of the time after being exposed to spinosad for exposures of 72 h and 96 h, respectively. Elzen (2001) discovered that exposure to chlorfenapyr in Petri dish studies resulted in 57% (males) and 31% (females) mortality rates of adult Orius insidiosus. Pietrantonio and Benedict (1999) showed that exposure to chlorfenapyr yielded 25–50% mortality of Orius insidiosus. In our investigation, adult Orius insidiosus survived 40% and 20% after 72 h and 96 h of exposure to chlorfenapyr, respectively. Therefore, combining Orius insidiosus with abamectin, spinosad, and chlorfenapyr may not be feasible without affecting the control of western flower thrips population. The neonicotinoid (imidacloprid, dinotefuran, acetamiprid, and thiamethoxam) and pyrethroid (tau-fluvalinate) pesticides were extremely harmful to adult Orius insidiosus survival (0–60% survival), implying that these pesticides should be excluded from management initiatives for western flower thrips that include Orius insidiosus. Orius insidiosus was found to be completely killed by the pyrethroids k-cyhalothrin and fenpropathrin, according to Studebaker and Kring (2003a, b) and Ashley et al. (2006). Therefore, it is not recommended to combine pyrethroids with Orius insidious to control populations of western flower thrips. The adult survival of Orius insidious was considerably impacted by imidacloprid (neonicotinoid). Insidious tests were conducted using Petri dishes with about 100% mortality (Studebaker and Kring 2003b). Imidacloprid exposure was shown by Delbeke et al. (1997) to be particularly damaging to Orius laevigatus, whereas Elzen (2001) reported mortality rates of 47% (males) and 62% (females) for adult Orius insidiosus exposed to imidacloprid. We discovered that adult Orius insidiosus survived imidacloprid exposure for 60% (24 and 48 h) and 20% (72 and 96 h). The results of our investigation and those of previous studies indicate that integrating imidacloprid with Orius insidiosus is not practical. Nowadays, Frankliniella occidentalis and other small pest arthropods are biologically managed by the predatory insect Orius sauteri, one of the most substantial positive arthropods in Northeast Asia (Pergande). Acetamiprid and imidacloprid, two neonicotinoid chemical insecticides, are mostly utilised in China to combat Frankliniella occidentalis. Despite being used in the field or greenhouse at sub-lethal doses to save helpful arthropods, F. occidentalis may nevertheless have an impact on the predator Orius sauteri. This study’s goal is to evaluate the long-term effects of a 24-h exposure to these two insecticides on Orius sauteri’s life cycle at application rates that are similar to the organism’s 24-h LC10, LC20, and LC30 values in the lab. The effects of acetamiprid and imidacloprid on female Orius sauteri’s fecundity were significantly greater than those of imidacloprid at all dosages examined, according to the results. Additionally, Orius sauteri’s lifetime and oviposition period were both decreased by the two pesticides. Additionally, it was discovered that the sub-lethal impacts on the first generation increased nymphal mortality, decreased adult longevity, and decreased fertility. However, none of the acetamiprid and imidacloprid administrations at the concentrations of LC10, LC20, and LC30 had any appreciable impact on the length of time that various nymphal stages took to mature or the sex ratio of the F1 generation. This study is the first to evaluate how

20.3

Hemiptera

745

well Orius sauteri and neonicotinoid insecticides get along and demonstrates that the use of acetamiprid and imidacloprid is likely detrimental to the population trends of this organism (Lin et al. 2020). Mahmoudi-Dehpahni et al. (2021) evaluated the long-term lethal and sub-lethal effects of thiamethoxam (TMX) on the predatory insect, Orius albidipennis, which preys on Aphis gossypii (Hemiptera: Aphididae). The maximum field recommended concentrations (MFRCs) of 12 and 14 of TMX were administered to first instar nymphs for 24 h. The therapy using soil treatment (bottom-up effect: plant–aphid– predator) had the smallest impact on plant survival, according to the findings. In fact, leaf-dip (residual contact method) and aphid-dip (through the oral exposure route) treatments significantly reduced survival. Adult longevity and egg production were negatively impacted by all evaluated concentrations of TMX in the leaf-dip and aphid-dip treatments, whereas neither feature was significantly influenced by the soil application treatment. In spite of all exposure routes, the insecticide had no effect on egg hatchability. In comparison to the control, the egg incubation period was decreased by oral treatment and residual contact treatments using one-fourth MFRC of TMX but not by other treatments. Finally, systemic (soil) injection of TMX proved safe for this predator based on the International Organization for Biological Control (IOBC)’s toxicity classification standards. Even at one-fourth MFRC, it was still somewhat hazardous and damaging (depending on the concentration) to the predator through oral exposures and lingering contact. According to our findings, TMX applied to soil is compatible with Orius albidipennis and can enhance predator conservation efforts in the integrated management of Aphis gossypii (Mahmoudi-Dehpahni et al. 2021). Buprofezin, emamectin benzoate, flonicamid, flupyradifurone, spirotetramat, spiromesifen, and sulfoxaflor were the seven insecticides that had earlier been licenced in Korea and were shown to be less hazardous to the Orius species based on earlier studies (Tang et al. 2007; Bielza et al. 2009; Biondi et al. 2012; Broughton et al. 2014; Barbosa et al. 2017). Except for pirimicarb, which was non-toxic, older chemical pesticides (methamidophos, dimethoate) were poisonous to Orius armatus. The appropriateness of insecticides made with newer chemistry varied. Both adults and nymphs were killed with abamectin. Spirotetramat, imidacloprid, and chlorantraniliprole were non-toxic. Orius armatus was mildly harmful to spinetoram and spinosad. Additionally, when compared to the untreated control, spinosad decreased fertility by 20%. Although pyrethroid proved non-toxic, female beans subjected to it produced 30% fewer eggs and 20% fewer nymphs than the control treatment (Broughton et al. 2014). In a glass vial assay, the acute toxicity against adult females of Orius minutus (Hemiptera: Anthocoridae) and Orius laevigatus of five insecticides and two acaricides (spiromesifen and emamectin benzoate) with low toxicity to natural enemies that are employed to limit those pest groups was assessed. The three objectives were: (1) to find out the lethal effects of these seven pesticides against Orius minutus and Orius laevigatus, (2) to record the sublethal effects of spirotetramat and spiromesifen against Orius minutus, and (3) to examine the efficacy of the combined use of a miticide and Orius minutus against Tetranychus urticae. The results show that spirotetramat and spiromesifen, two of the seven

746

20

Biosafety Assessment of Synthetic Pesticides

pesticides, were the least acutely hazardous to the evaluated predators. We discovered that neither substance adversely affected any of the life history parameters of Orius minutus, with the exception of fecundity, in an assay on the sub-lethal toxicity of spirotetramat and spiromesifen solely against Orius minutus. The concurrent usage of spiromesifen and Orius minutus was proven to be effective against Tetranychus urticae (Acari: Tetranychidae) in a potted plant experiment, and the Tetranychus urticae was decreased. The Tetranychus urticae population was substantially greater than the single treatment or untreated control. As a result, we draw the conclusion that spiromesifen is a selective acaricide with a less detrimental impact on Orius minutus biological parameters than other poisons we evaluated (Rahman et al. 2022). The impact of pyrethroids on Orius laevigatus showed substantial heterogeneity in pyrethroid sensitivity in wild populations of the heteropteran predator Orius laevigatus, with a median fatal dose (LC50) of 1.6–77.0 mg L-1. We were successful in identifying an Orius laevigatus strain with a high pyrethroid resistance (LC50 = 1059.9 mg L-1). The resilience of the entire population present in the crop facing pyrethroid application was also increased by the fact that such resistance was demonstrated in every instar, especially in the last nymphal stages. Orius laevigatus adults and nymphs subjected to the maximum field rate of numerous pyrethroids and natural pyrethrins, which are widely used to control a number of pests in organic and integrated pest management crops, may be able to survive due to the level of resistance attained. This strain’s resistance to pyrethrins and pyrethroids would therefore increase the biocontrol protocols’ robustness, which is a crucial condition for the widespread adoption of biological control (Balanza et al. 2021). The mortality of female adults of Orius laevigatus was evaluated against Proteus, matrine, and pyridalyl at 24, 48, 72, and 96 h after exposure to 14-day residues of the treatments on strawberry leaves. The residue of Proteus was less toxic; the remaining products caused the lowest mortality at different times after exposure. Against female adults, the most toxic insecticide was Proteus (LC50 = 44.3 μL L-1), followed by pyridalyl (LC50 = 83.8 μL L-1) and matrine (LC50 = 102.7 μL L-1). Sub-lethal treatments (LC25) significantly prolonged the developmental duration of total immature stages from 17.6 days in control to 21.6 days and 20.0 days in Proteus and pyridalyl treatments, respectively. Moreover, the fecundity of Orius laevigatus treated with Proteus, pyridalyl, and matrine decreased to 58.8%, 75.6%, and 96.7%, respectively, in comparison to the control. Compared with the control population (0.118 days-1), the intrinsic rate of increase (r) of the F1 generation decreased by 0.053 days-1, 0.095 days-1, and 0.110 days-1 in Proteus, pyridalyl, and matrine treatments, respectively. The consumption rate of control bugs reached 14.0 thrips during 24 h. The adults fed on Proteus treatment had the lowest consumption rate in this period (9.4 preys). Overall, matrine proved to be harmless with reproductive capacity and r similar to what was recorded in control bugs (Kordestani et al. 2022).

20.3.4 Reduviidae Reduviid predators play an important role in bio-intensive integrated pest management (Ambrose 1999). George and Ambrose (1998) recorded the impact of various

20.3

Hemiptera

747

pesticides against these reduviid predators. According to Sahayaraj (1991), Rogor EC 30%, decis EC 2.8%, and fenvalerate EC 20% (Sumicidin) are the safer insecticides for Acanthaspis pedestris life phases (Fig. 20.1). The impact of insecticides on predatory assassin bugs is little understood. For instance, quinalphos and other organophosphate (OP) insecticides adversely impacted the stadial body weight, fertility, and longevity of Rhynocoris marginatus (George and Ambrose 1998). Another species of assassin insect, Acanthaspis pedestris, saw its predatory effectiveness lowered by cypermethrin (Claver et al. 2003). Only 1.17% of Rhynocoris marginatus nymphal (fourth instar) deaths were caused by contact poisoning when nimbecidine was used (Sahayaraj and Selvaraj 2003). Staubli et al.’s (1984) research established the adverse effects of numerous organophosphate insecticides on coccinellid predators. In general, synthetic pyrethroids and organophosphates were more toxic to spiders (Wakeil et al. 2013). According to Patel (2018), lady beetles and spiders had minimally negative field effects when exposed to flubendiamide + thiacloprid and spirotetramat + imidacloprid, respectively (Patel and Sarkar 2019). Chlorantraniliprole was found to work well with biocontrol agents, making it a less hazardous insecticide (Larson et al. 2012). Chlorantraniliprole’s marginally detrimental effects on Chrysoperla carnea larva were discovered in a laboratory setting (Sabry et al. 2014). With the neonicotenoids imidacloprid and acetamiprid, the effect on the adult and larva of Cryptolaemus montrouzieri was somewhat detrimental (Halappa et al. 2013). Chrysopa lacciperda larvae were adversely affected by fipronil in a laboratory setting (Singh et al. 2010). The survival, lifespan, fecundity, egg hatching, and progeny survival of Cryptolaemus montrouzieri larvae and adults after direct spirotetramat treatment in the laboratory were unaffected (Planes et al. 2013). During the field application of bifenthrin, cartap hydrochloride, and emamectin benzoate, coccinellids and spiders posed a risk (Karthick et al. 2014). Orius insidiosus, the main predator of thrips in pepper, has been shown to be reduced

Decis

Fenvolerate

Rogor

0.08 0.07

LC50 (%)

0.06 0.05 0.04 0.03

0.02 0.01 0 First Instar

Second Instar

Third Instar Life stages

Fourth Instar

Fifth Instar

Fig. 20.1 The effects of three common insecticides on the LC50 values (%) of the several life stages of Acanthaspis pedestris

748

20

Biosafety Assessment of Synthetic Pesticides

Table 20.2 The LC50 values of Rhynocoris fuscipes, Rhynocoris kumarii, and Rhynocoris marginatus are affected by monocrotophos, dimethoate, methyl parathion, quinalphos, and endosulfan (%) (after George and Ambrose 1998) Reduviids Rhynocoris fuscipes Rhynocoris kumarii Rhynocoris marginatus

Monocrotophos 0.022

Dimethoate 0.043

Methyl parathion 0.010

Quinalphos 0.045

Endosulfan 0.114

0.047

0.053

0.017

0.061

0.128

0.048

0.085

0.011

0.082

0.189

in population by the broad-spectrum insecticide tolfenpyrad (Srivastava et al. 2014). The coccinellid population was reduced in the rice ecosystem by acephate (Sharanappa et al. 2019). However, in an onion crop, chlorfenapyr was found to be safe against coccinellids (Yadav et al. 2020). When used against the rice plant hopper predator Cyrtorhinus lividipennis, ethiprole had an average level of toxicity (Nagalingam et al. 2009). All of these earlier discoveries generally concur with the findings of the current research publication. In his dissertation from 1996, George examined how three reduviid species— Rhynocoris kumarii, Rhynocoris marginatus, and Rhynocoris fuscipes—were affected by pesticides (Table 20.2). Here are some comprehensive results. Rhynocoris kumarii and Rhynocoris marginatus were 4.5 and 4.0 times more resistant to monocrotophos, respectively, than was Rhynocoris fuscipes after 24 h, causing 23.33% mortality at doses of 0.0162%, 0.0144%, and 0.0036% on the nymphal instars of each species, respectively. Monocrotophos induced 20.00% mortality among individuals during a 24-h period in Rhynocoris kumarii, Rhynocoris marginatus, and Rhynocoris fuscipes at concentrations of 0.054%, 0.054%, and 0.0288%, respectively. From this, it was shown that both Rhynocoris kumarii and Rhynocoris marginatus exhibited identical monocrotophos resistance and that they did so 1.875 times better than Rhynocoris fuscipes. A similar pattern was observed in extended exposure. For instance, at 48 h, 72 h, and 96 h, monocrotophos concentrations of 0.072% induced mortality rates of 46.67%, 46.67% and 86.67%, 56.67%, 80.00%, and 93.33%, and 70.00%, 70.00%, and 100.00% in the adults of Rhynocoris kumarii, Rhynocoris marginatus, and Rhynocoris fuscipes, respectively. The mortality depended on the duration. Dimethoate exposure for 24 h resulted in 46.67% mortality for Rhynocoris kumarii, Rhynocoris marginatus, and Rhynocoris fuscipes III nymphal instars. This suggests that these species are 1.25 and 2.50 times more resistant to dimethoate, respectively, than is Rhynocoris fuscipes. Similar trends were seen after exposure times of 48 h, 72 h, and 96 h (0.03% concentration caused mortality of 56.67%, 33.33%, and 75.00%; 66.67%, 43.33%, and 85.71%; and 80.00%, 53.33%, and 100.00% in Rhynocoris kumarii, Rhynocoris marginatus, and Rhynocoris fuscipes,

20.3

Hemiptera

749

respectively). Rhynocoris kumarii and Rhynocoris marginatus were 1.33 and 2.00 times more resistant, respectively, than was Rhynocoris fuscipes at 24 h, as seen by the 13.33% mortality rate in concentrations of 0.06%, 0.09%, and 0.045%, respectively, among the adults of each species. A similar pattern was seen for exposure times of 48 h, 72 h, and 96 h (0.09% concentration caused mortality of 43.33%, 30.00%, and 76.67%; 5.67%, 43.33%, and 93.33%; and 67.74%, 56.67%, and 100.00% in Rhynocoris kumarii, Rhynocoris marginatus, and Rhynocoris fuscipes, respectively). The findings unmistakably show that Rhynocoris marginatus was the most resistant to dimethoate, whereas Rhynocoris fuscipes was the least resistant. The percentage of mortality caused by methyl parathion reveals that in the 111 nymphal instars of Rhynocoris kumarii, Rhynocoris marginatus, and Rhynocoris fuscipes, 26.67% mortality was observed at concentrations of 0.00445, 0.00755, and 0.003%, respectively, indicating that Rhynocoris kumarii and Rhynocoris marginatus were 1.33 and 2.50 times more resistant, respectively, than was Rhynocoris fuscipes. Similar trends were seen in all other exposure durations and in all people. Rhynocoris kumarii and Rhynocoris marginatus were 1.25 and 2.00 times more resistant, respectively, than was Rhynocoris fuscipes at 24 h, according to the 33.33% mortality rates found at concentrations of 0.025%, 0.040%, and 0.020%, respectively, in the nymphal instars of each species. The other durations showed a similar pattern as well. Rhynocoris kumarii and Rhynocoris marginatus were 1.33 and 2.0 times more resistant to quinalphos, respectively, than was Rhynocoris fuscipes, according to the 13.33% mortality rates among adults observed at doses of 0.05%, 0.075%, and 0.0375% for each species, respectively. Table 20.2 displays the endosulfan fatality percentage of Rhynocoris kumarii, Rhynocoris marginatus, and Rhynocoris fuscipes’ 111 nymphal instars, which all experienced 16.67% mortality in 24 h, with concentrations of 0.035%, 0.0875%, and 0.021%, respectively. This suggests that Rhynocoris kumarii and Rhynocoris marginatus are 1.667 and 4.167 times more resistant, respectively, than is Rhynocoris fuscipes. A similar pattern was seen in adults as well (after 72 h of exposure, 53.3% mortality was seen in Rhynocoris kumarii, Rhynocoris marginatus, and Rhynocoris fuscipes). In this instance, Rhynocoris kumarii and Rhynocoris marginatus were more resilient than was Rhynocoris fuscipes, and Rhynocoris marginatus was the most resilient of the three species (George and Ambrose 1998). The effects of Acanthaspis pedestris, a non-target potential reduviid biological control agent, on predatory behaviour, mating behaviour, and functional response were examined. The intensity of the aberrant behaviour increased as the cypermethrin concentration increased. The operational reaction events, such as the attack ratio, handling time, and rate of discovery, were adversely affected by the pesticide. Additionally, cypermethrin prolonged the mating procedures and reduced Acanthaspis pedestris predatory efficacy. Cypermethrin transformed the type II decelerating curve of the functional response into a type IV dome-shaped curve (Claver et al. 2003). According to Edward George and Ambrose (2004), prohaemocytes, plasmatocytes, granular haemoglobin-containing cells, cystocytes, and oenocytoid cells are all present in the hemogram of Rhynocoris kumarii. They provide immunity to animals. It was investigated how five pesticides, including

750

20

Biosafety Assessment of Synthetic Pesticides

monocrotophos, dimethoate, methyl parathion, quinalphos, and endosulfan, affected the total and differential haemoglobin counts (THC and DHC, respectively). Except for endosulfan, all insecticides initially decreased prohaemocytes and plasmatocytes, increased granular haemoglobin, changed the proportion of cystocytes and oenocytoid cells, and increased the number of haemoglobin-containing cells (THC). However, endosulfan initially raised plasmatocytes and prohaemocytes while decreasing granular haemocytes and THC. Methyl parathion and monocrotophos had the most effect on both DHC and THC, whereas endosulfan had the least effect. Thus, among these five insecticides, endosulfan is regarded as the safest to use with Rhynocoris kumarii, followed by dimethoate and quinalphos. Similar observation was also recorded in Rhynocoris marginatus. On the non-target reduviid predator Rhynocoris marginatus, the relative toxicity of two regularly used insecticides, cypermethrin (Cymbush 10 EC [emulsifiable concentration]) and fenvalerate (Sumicidin 20 EC), was assessed. Fenvalerate was the more toxic of the two insecticides tested than was cypermethrin. Both cypermethrin and fenvalerate reduced the quantity of plasmatocytes while increasing prohaemocytes and granular haemoglobin. They also affected the density of oenocytoid and cystocyte cells (George et al. 2021). Although both pesticides raised the total number of haemoglobin-containing cells, fenvalerate had the most effect. Using dipping techniques, the effects of cypermethrin, deltamethrin, and trichlorfon were evaluated on Sycanus dichotomus of the third instar. These insecticides are frequently used to stop leaf defoliators on oil palm trees. The developmental stages of the third and fifth instars that had been exposed to insecticides were significantly longer under different treatments, according to the results of the sub-lethal dose treatment on the Sycanus dichotomus third instar. Additionally, the percentage of Sycanus dichotomus larvae that survived after being treated with insecticides was significantly lower than that of control insects. The female adult insects that survived the treatments, regardless of the type exposed during the third instar of the Sycanus dichotomus, live longer than did the male insects. The longevity and fertility of Sycanus dichotomus adults from the third instar subjected to insecticides, however, do not differ noticeably from control insects. The findings of this study indicate that it is important to use insecticides carefully in order to control pest populations without endangering the environment of oil palms’ natural enemies (Ismail and Abd Ghani 2018). A total of 22 pesticides were put to the test against predatory Rhynocoris marginatus nymphs in their fourth instar in a laboratory environment. The interventions were classified as harmless (E 30%), somewhat detrimental (30%E 79%), moderately harmful (80%E 99%), and harmful (E > 99%) in agreement with the IOBC standards for corrected mortality (E). In the findings, no insecticide was labelled as dangerous. Chlorpyriphos, acephate 95 SG, and quinalphos are three moderately toxic organophosphate insecticides that cause a high level of at-par mortality (85.74–81.85%). With E ranging from 76.48% to 51.30%, seven insecticides, including bifenthrin, tolfenpyrad, fipronil, acephate 75 SP (soluble powder), lambda-cyhalothrin, cartap hydrochloride, and ethion + cypermethrin, had a marginally negative impact. Mortality rates for the 12 remaining insecticides

20.4

Lacewings

751

with assassin predators ranged from 27.22% to 5.37%. Acetamiprid, azadirachtin, spirotetramat, imidacloprid, chlorfenapyr, chlorantraniliprole, flubendiamide, flubendiamide + thiacloprid, emamectin benzoate, spirotetramat + imidacloprid, and ethiprole + imi were all found to have lower levels of safety than thiacloprid (Patel 2020).

20.4

Lacewings

Insecticides are commonly employed in farming systems where lacewings are found. They are crucial biological control agents due to their remarkable capacity for the establishment of pesticide resistance. The insecticides pirimicarb, chlorpyrifos, fenvalerate, and demeton-S-methyl were all applied topically in acetone to the lacewing Austromicromus tasmaniae, the beetle Coccinella undecimpunctata, and two of their natural enemies, Acyrthosiphon kondoi and Acyrthosiphon pisum. The LD50 values were established. Pirimicarb showed a toxic effect on aphids 1000–10,000 times greater than that on predators, whereas fenvalerate and demetonS-methyl produced effects of similar magnitude (5–1000 and 16–400, respectively). Aphids and predators were both damaging to chlorpyrifos. When pirimicarb, chlorpyrifos, and fenvalerate were treated as aqueous solutions, the lacewing confirmed the order of toxicity of these three chemicals when applied topically. LD50 rates of pirimicarb, fenvalerate, and demeton-S-methyl were topically applied to adult lacewings before the insect groups were exposed to a post-treatment temperature of 10 or 25 °C. The toxicity of fenvalerate significantly increased at a lower temperature (at around 3 °C) (Syrett and Penman 1980). Chrysoperla carnea is a powerful, voracious predator (Neuroptera: Chrysopidae). It is frequently used to control insect pests such whiteflies, aphids, jassids, thrips, mites, mealybugs, and lepidoptera eggs. It has demonstrated outstanding resistance to many pesticide classes, including conventional pyrethroids, insect growth regulators (IGRs), and new modes of action (Pathan et al. 2008, 2010; Mansoor et al. 2017; Mansoor and Shad 2019a, b, 2020a, b, c). Flupyradifurone, sulfoxaflor, and flonicamid were three aphicides tested against the lacewing, Chrysoperla carnea (Neuroptera: Chrysopidae). Regardless of life stage or exposure technique, flonicamid had no effect on either species. Lacewing adults were more susceptible to sulfoxaflor and flupyradifurone than were larvae and exhibited greater mortality rates when fed contaminated honey solution compared to when they came into contact with residues on an inert surface (Barbosa et al. 2017). The green lacewing Chrysoperla carnea (Neuroptera: Chrysopidae) is a powerful biocontrol agent that feeds on a wide range of insect pests. It possesses a significant level of resistance to numerous pesticide classes. The degree, stability, and pace of reversion of resistance to specific new chemical insecticides were investigated in this study using a population of Chrysoperla carnea collected from the field. Chrysoperla carnea strain from the field displayed resistance to imidacloprid, indoxacarb, and chlorfenapyr by 313.44, 216.50, and 276.83 times, respectively, in comparison to the susceptible population. Imidacloprid, indoxacarb, and

752

20

Biosafety Assessment of Synthetic Pesticides

chlorfenapyr each had initial susceptibility values for Chrysoperla carnea of 2627.31 ppm (part per million), 502.10 ppm, and 581.89 ppm, respectively. Insecticide resistance declined at a rate of 0.039 ppm, 0.048 ppm, and 0.040 ppm, respectively. Rotation of these pesticides involves quality control to ensure Chrysoperla carnea’s survival in a hostile environment where repeated applications are required to prevent economic damage from pests (Mansoor and Shad 2019a, b). Syrphid fly, green lacewing, and ladybird beetle populations had all been the highest in the control plots while being the lowest in the cypermethrin-treated plots, according to recent field studies at an egg facility (Ullah et al. 2022). In a different investigation, the effects of sub-lethal dosages of the neonicotinoid insecticides acetamiprid and dinotefuran on the neuropteran species Chrysopa pallens were assessed and determined. Dinotefuran and acetamiprid significantly lengthened the larval and pupal stages relative to control at LD10 (8.18 ng active ingredients (a.i.) per insect and 9.36 ng a.i. per insect, respectively) and LD30 (16.84 ng a.i. per insect and 15.01 ng a.i. per insect, respectively), according to the results (Su et al. 2022). Methoxy-SEL did not develop cross-resistance to acetamiprid (1.17-fold), lambda-cyhalothrin (1.40-fold), or profenofos, in contrast to the field strain (95% FLs [field levels] overlapped) (2.09-fold). It is helpful to conduct genetic crosses between populations of naturally occurring enemies that are susceptible to insecticides and those that are resistant to them in order to better understand the mechanism of inheritance, effective dominance, degree of dominance, and number of genes supporting resistance development.

20.5

Dermapterans

Earwigs are frequently seen as an effective all-natural controller of pests for crop insects because the bulk of them are multipurpose predators. In spite of the fact that earwigs are prized in neotropical maize fields, little is known about them. The earwig Doru luteipes of the Forficulidae (order: Dermaptera) family is one of the main predators of the fall armyworm Spodoptera frugiperda. The single study that has been conducted on the effects of insecticides on Doru luteipes so far focuses on the acute fatal sensitivity of old organophosphates, carbamates, and the pyrethroid deltamethrin to this predator (Bacci et al. 2001). Prior to this, chlorfenapyr showed an intermediate safety profile for Doru luteipes that seemed to vary according to the species and crop in question (Black et al. 1994; Argentine et al. 2002). Studies on Doru luteipes and its prey species examined the toxicity and selectivity of the insecticides chlorantraniliprole, chlorfenapyr, chlorpyrifos, cyhalothrin, deltamethrin, etofenprox, methomyl, and spinosad. Doru luteipes’s behavioural response to these drugs in terms of movement was also assessed. Concentration– response bioassays revealed that etofenprox (>1100 less toxic), spinosad (>3500 less toxic), and chlorantraniliprole (>550,000 less damaging) have much lower potencies compared to chlorpyrifos, the most toxic pesticide tested against this type of earwig. The most selectivity was demonstrated by these three compounds when comparing the earwig to its prey, the fall armyworm. Time–response

20.5

Dermapterans

753

bioassays, using the pesticide label rates recommended against the fall armyworm, also showed the good selectivity of chlorantraniliprole and etofenprox in addition to deltamethrin and methomyl. Chlorpyrifos once more showed the least amount of selectivity. The earwig was exposed to surfaces that were exposed to pesticides, demonstrating that spinosad reduced the adults’ locomotory activity and most likely increased their exposure to pesticides, whereas they avoided surfaces that had been treated with chlorfenapyr, etofenprox, and chlorpyrifos. The pesticides chlorantraniliprole and etofenprox seem to be the most effective against Spodoptera frugiperda while safeguarding Doru luteipes populations (Campos et al. 2011). Forficula auricularia responds to the MoA-classed insecticides that were put it to the test. The mortality rate for the control group was initially modest but increased over the course of the study: 0% at 0 Days After Treatment (DAT), 2.5% at 12 DAT, 5% at 20 and 26 DAT, and 7.5% at 34 DAT. In all, 9 of the 17 represented classes of insecticides—carbamates, juvenile hormone mimics, feeding blockers, mite growth inhibitors, microbial disruptors, organotin acaricides, benzoylurea, diacylhydrazines, and tetronic acid derivatives—as well as biological insecticides, paraffin oils, emulsifiers, protectants, and herbicides—proved harmless to adults. Organophosphates, pyrethroids, pyrethrins, neonicotinoids, avermectins, mitochondrial electron transport indicator (METI) acaricides, and oxadiazines were among the eight classes that were just marginally harmful to adults (spinosyns). There were a few mild-to-severe side effects; however, they took time to manifest and did not happen immediately (minimum of 12 days). It was amazing that the paralysing effects of some products (60–90% on the first day of observation) resulted in up to 50% death after 34 days (Peusens and Gobin 2008). The European earwig, Forficula auricularia, is the all-purpose predator in orchards. It exhibited considerably higher mortality and erratic behaviour after exposure to indoxacarb, spinosad, and chlorpyrifos-methyl than did earwigs from the control group. Earwigs from orchards that had applied chlorpyrifos-methyl over a long period of time (more than 10 years) perished much less frequently when exposed to this insecticide than did earwigs from orchards that had not been sprayed (Jana et al. 2021). Nymphs and adults of Euborellia annulipes were exposed to the pesticides by contact with dried residues on either treated plants carrying prey or inert surfaces as well as through consumption of infected prey. Pymetrozine, chlorantraniliprole, and spinetoram revealed no effect on the predator regardless of the tested earwig population, life stage with developmental period and survival, or exposure method (ingestion and residual). In order to prevent nymphs from completing their development, dried cyantraniliprole residue caused just 27% of adults to survive for 20 days after exposure. Pyriproxyfen was benign to nymphs and adult earwigs through acute toxicity (70–100% survival 72 h after exposure) but impeded normal nymphal development, leading to chronic toxicity. Chlorfenapyr, indoxacarb, lambdacyhalothrin, chlorpyrifos, dimethoate, and malathion were harmful regardless of the predator’s life stage or manner of exposure. Thiamethoxam, lambda-cyhalothrin, and indoxacarb had less of an adverse effect on plants exposed to them when predators were allowed to hide in the soil (Potin et al. 2022). According to these

754

20

Biosafety Assessment of Synthetic Pesticides

results, an insecticide’s selectivity is influenced by the insecticide, the stage of the predator’s life cycle, and the predator’s behaviour. As a result, it is possible to evaluate different predator life stages via a range of exposure methods without prohibiting the insect from engaging in its natural behaviour in order to acquire more precise estimates of pesticide sensitivity.

20.6

Coleopterans

Since the coccinellids’ predatory activity normally starts at a medium-to-high level of pest density, natural treatment is not always swift but is frequently successful. Aphids are the primary prey of the predator Coccinella undecimpunctata (Coleoptera: Coccinellidae). There have been investigations into the sensitivity of Coccinella undecimpunctata to several pesticides; however, every one of them has had detrimental impacts on this species (Lowery and Isman 1995; Omar et al. 2002). Recent studies have shown that pirimicarb and pymetrozine, when sprayed on the insects, generally had no adverse effects on the biological features of the immature or adult stages of Coccinella undecimpunctata. As a result, these chemicals may be used in conjunction with Coccinella undecimpunctata for integrated control of sucking pests. These biological characteristics include the rate of egg hatching, reproduction, fertility, and maturation time (Cabral et al. 2011). According to Mollah et al. (2013), esfenvalerate 5 EC at 1 mL L-1 water treatment, deltamethrin 2.5 EC at 1 mL L-1 water treatment, cypermethrin 10 EC at 1 mL L-1 water treatment, and fenitrothion and cartap hydrochloride 50 SP at 2 g L-1 water treatment produced the highest number of dead ladybird beetles (4.45), followed by the other treatments (2.97). At a concentration of 2 mL L-1 water, the effects of diazinon and Nogos on Menochilus sexmaculatus larvae and adults resulted in mortality rates of 86.7%, 83.3%, 86.7%, and 93.3%, respectively (Islam and Sardar 1997). In general, Carabidae are not acutely toxic to herbicides against Diabrotica undecimpunctata howardi (Brust and House 1990), Harpalus rufipes (Zhang et al. 1997), and fungicides against them (Sotherton and Moreby 1984), though they may have an indirect impact on survival through the removal of food sources and habitat modification. However, mounting data suggest that pesticides can severely decrease populations of non-target creatures, such as predatory beetles like Hippodamia convergens (Coleoptera: Coccinellidae) and coccinellids (Barbosa et al. 2018). Olszak et al. (2004) examined the effects of seven pesticides (six insecticides and one acaricide) on the adults of A. bipunctata and Coccinella septempunctata at various life stages (adulthood, larval, egg hatching). The longevity and fecundity of females of both studied species were shown to be unaffected by food (aphids) tainted with pesticides such pirimicarb, novaluron, pyriproxyfen, and fenpyroximate. In the laboratory, acute negative effects of pyriproxyfen, imidacloprid, deltamethrin + heptenophos, lambdacyhalothrin, and Bacillus thuringiensis subsp. tenebrionis at field rates on mature seven-spot ladybird beetles, Coccinella septempunctata, were observed (Bozsik 2006). The acute surface contact

20.6

Coleopterans

755

impacts (dry spray on leaves of Philadelphus coronarius) were employed to test the toxicity of the preparations, with the exception of Bacillus thuringiensis, which was treated with mixed pollen. Each of the 4–6 dosages evaluated was administered to 22 individuals (pyriproxifen 12.5, 25, 50, 100, 200, 400 mg AI litre-1; imidacloprid 62.4, 125, 250, 500 mg AI litre-1; deltamethrin + heptenophos 26.4, 53.1, 106.3, 212.5 mg AI litre-1; lambda-cyhalothrin 1.1, 3.4, 10, 30 mg AI litre-1; B. thuringiensis 1.5, 3.0, 12.0, 48, 192, 768 mg AI litre-1), According to various evaluation categories, pyriproxyfen, imidacloprid, and B. thuringiensis subsp. tenebrionis appear to be safe for Coccinella septempunctata adults. However, the other two treatments were mildly toxic to them, requiring additional semi-field or field tests to determine their true impact under field circumstances (Bozsik 2006). In a second investigation, Pasqualini and Civolani (2003) looked at the efficacy of six pesticides on the adults of the aphidophagous coccinellids Adalia bipunctata, Coccinella septempunctata, and Oenopia conglobata in apple, pear, and peach orchards. The insecticides studied included the organophosphates (OPs) chlorpyrifos, chlorpyrifos-methyl, azinphos-methyl, and malathion as well as the carbamatederived methomyl and nereistoxin equivalents. Azinphos-methyl consistently caused between 76% and 90% of coccinellid death in four studies. Chlorpyrifos EC caused mortality in five tests that ranged from 40.2% to 63%. Over three investigations, chlorpyrifos Wdg mortality ranged from 50.8% to 70%. In 1999, chlorpyrifos-methyl caused a 31% mortality rate in apples and an 86.1% mortality rate in pears. In a single trial on apples, methomyl and cartap were compared, and the death rates were 66.7% and 10%, respectively. In a different investigation, malathion was examined and found to be 43.5% deadly. According to Chandra et al. (2014), imidacloprid and thiamethoxam were shown to be safe for Coccinella beetles because they reduced beetle mortality in treated chemicals. According to Maula et al. (2010), plots treated with dimethion had the lowest mortality of Coccinella septempunctata. According to Bana et al. (2014), imidacloprid had a minimum population of 2.67 and 2.10/10 plants. After 7 days of spray application, imidacloprid showed a 57.34% reduction in the population of Coccinella septempunctata, whereas dimethoate showed a 52.77% drop in the population of syrphid fly larvae. Coccinella septempunctata larvae and adults were given several dosages of nitenpyram—10%, 25%, 50%, 100%, and 150% of the maximum recommended limit field rate (MRFR) of 3 g a.i. ha-1, 7.5 g a.i. ha-1, 15 g a.i. ha-1, 30 g a.i. ha-1, and 45 g a.i. ha-1, respectively—and a blank control based on a preliminary acute 72-h toxicity study. The median lethal rate (LR50) (application rate that results in the death of 50% of the population) for nitenpyram for Coccinella septempunctata decreased from 73.43 to 63.0 g a.i. ha-1 in the long-term test, whereas the hazard quotient (HQ) values stayed under the threshold value of 2. The lowest LR50 value for nitenpyram (lowest LR50 = 63.0 g a.i. ha-1) is more than 45 g a.i. ha-1, and it did not significantly alter the survival rate, fecundity, pupation, or adult emergence at 150% of the label rate. Similarly, neither the total developmental time nor egg hatchability was significantly impacted at 100% of the label rate (NOER [no

756

20

Biosafety Assessment of Synthetic Pesticides

observed effect application rates] = 30 g a.i. ha-1). When provided below or at a dose of 30 g a.i. ha-1, nitenpyram may be deemed safe for Coccinella septempunctata based on the assessment of the total effect (E). The minimum suggested field application rate for nitenpyram (30 g a.i. ha-1) for controlling aphids in China was exceeded by both the lowest LR50 and NOER values. All tests showed that nitenpyram can be used as prescribed and is safe around natural enemies like Coccinella septempunctata (Jiang et al. 2018). A prominent predator in the neotropics, Eriopis connexa is linked to pests on horticulture crops in Argentina that are economically significant. When eggs and two larval instars of E. connexa were exposed to insecticides under laboratory conditions, Noelia et al. (2022) evaluated the lethal and sub-lethal effects of two IGR insecticides on Eriopis connexa. The results indicate that pyriproxyfen and cypermethrin had significant effects on egg hatching, with respective effects of 28.8% and 70.4%. Pyriproxyfen lengthened the developmental duration of both the larval and pupal stages and decreased the survival of larvae that emerged approximately 52% from the third day after hatching. Teflubenzuron, on the other hand, sped up the pupae stage’s development without affecting hatching or survival of Eriopis connexa. Cypermethrin increased the developmental duration of the second larval instar while decreasing the survival of the second and fourth larval instars between 36.4% and 74.6%, respectively. Pyriproxyfen decreased female fecundity and fertility while also lengthening the fourth larval instar’s developmental time. Teflubenzuron reduced the survival of the second and fourth larval instars by 46.9% and 28.6%, respectively, while lengthening the time required for larval stage of evolution. Teflubenzuron also decreased the fertility and fecundity of females. Eriopis connexa (Coleoptera: Coccinellidae) is an essential biological control agent in wheat crops. In this study, we investigated the lethal effects of 15 pesticides used in the wheat crop on adults of the predator Eriopis connexa, following the methodology advised by the International Organization for Biological Control (IOBC). The potential for deadly effects has not been studied. On the glass plates that were used to create the exposed areas, the insecticides were sprayed. The chitin synthesis inhibitors diflubenzuron (DFB) SC, diflubenzuron WP, lufenuron, triflumuron, and the pyrethroid beta-cyfluthrin were found to have no effect on adult predators, and their use should be highlighted. On the other hand, it is preferable to avoid employing methomyl, lambda-cyhalothrin, and thiamethoxam + lambda-cyhalothrin A and B while adults of the predators are present in the crop (Pasini et al. 2022).

20.7

Neuroptera

The common green lacewing, Chrysoperla carnea (Neuroptera: Chrysopidae), is one of the most frequent arthropod predators. It eats a range of soft-bodied insects, including aphids, lepidopteran eggs and neonates, scales, whiteflies, and mites. It has long been regarded as a strong option for pest control strategies due to its wide prey range, geographic dispersion, pesticide resistance or tolerance, voracious larval feeding capability, and commercial availability. Chrysoperla carnea inundated discharges were successful in reducing pest complex numbers in a variety of

20.8

Syrphids

757

crops. It was investigated how pesticides affected Chrysoperla carnea fertility and acute mortality. The insecticides’ overall mortality rate differed considerably from that of the control group. Fecundity was significantly decreased by propargite and pymetrozine but not by imidacloprid. Additionally, a comparison of the overall impacts of the pesticides showed that, whereas propargite and pymetrozine were somewhat detrimental according to the IOBC classification, imidacloprid was categorised as a harmless substance (Rezaei et al. 2007). Chrysoperla carnea reacts differently to different insecticides. Spinosad, according to Medina et al. (2001), had little-to-no effect on the fecundity and longevity of adult Chrysoperla carnea and no effect on eggs or pupae. Chrysoperla carnea’s third instar larvae were toxic to azadirachtin and diflubenzuron, but pyriproxyfen and tebufenozide were safe when used at the approved field rates (Medina et al. 2003; Güven and Göven 2003). Using an organic farming method, spinosad was used to manage a range of pest species in greenhouses, whereas Chrysoperla carnea and Coccinella undecimpunctata were introduced to manage aphid numbers on pepper and cabbage (Mandour 2009). The green lacewing Chrysoperla nipponensis (Neuroptera: Chrysopidae) is one of the most major native natural predators that is widely distributed in China, Japan, and Korea. Aphids, thrips, leafhoppers, psyllids, caterpillars, and mites are just a few of the nuisance species that this species’ larvae feed on in both farmlands and woodlands. Because of its excellent foraging and predatory skills, it is believed that this animal could function as a biological control agent in a range of field crops. In a study by Barbosa et al. (2017), two generalist predators, Chrysoperla carnea (Neuroptera: Chrysopidae) and Orius insidiosus, were utilised to assess the lethality of three aphicides. After 24 h of exposure, imidacloprid alone was the least effective therapy (LC50 = 18.59 mg L-1), followed by imidacloprid combined with Silwet 618 (LC50 = 8.13 mg L-1), STIK2 (LC50 = 6.74 mg L-1), ECO (LC50 = 6.40 mg L-1), and Silwet L-77 (LC50 = 6.09 mg L-1). These results show that organosilicones considerably enhanced imidacloprid’s synergistic effects ( p = 0.001). After 24 h of exposure, Silwet 618 had the lowest synergistic impact, whereas Silwet L-77 had the highest. The concentration of imidacloprid utilised going forward was 10.95 mg L-1 (95% confidence interval (95% CI): 8.16–13.63 mg L-1), which was the drug’s LC30 value (Liu et al. 2021) (Table 20.3).

20.8

Syrphids

While many syrphids have little-to-no great economic importance, numerous species in the ten genera are important as predators when aphids are in their larval stage: Allograpta, Baccha, Mesograpta, Melanostoma, Paragus, Pipiza, Scaeva, Syrphus, Metasyrphus, and Sphaerophoria. There is a dearth of information on how various insecticides affect Syrphidae. The inability to reduce harm beneficial insects when instant pesticide control of a pest is required, the inability to recognise the critical role that Syrphidae play in the control of aphids and other pests of ornamental and agricultural plants, the lack of funding for research into this issue, and the absolute refusal of insecticide producers to perform experiments are likely a few reasons.

758

20

Biosafety Assessment of Synthetic Pesticides

Table 20.3 Impact of various pesticides on Chrysoperla nipponensis with references Neuropterans Chrysoperla nipponensis Chrysoperla nipponensis Chrysoperla nipponensis Chrysoperla nipponensis Chrysoperla nipponensis

Pesticides Imidacloprid

Hours of exposure 24

LC50 18.59 mg L1

Silwet L-77

24

6.09 mg L-1

Imidacloprid + Silwet 618 STIK2

24

8.13 mg L-1

24

6.74 mg L-1

ECO

24

6.40 mg L-1

References Liu et al. (2021) Liu et al. (2021) Liu et al. (2021) Liu et al. (2021) Liu et al. (2021)

Different insecticides have the capacity to kill syrphids by contact or consumption of tainted aphid prey (Azab et al. 1971; Kirknel 1975). When 40 different pesticides were examined by Hassan et al. in 1983, they found that diflubenzuron caused comparatively low mortality (less than 50%). It has been found that pupae are less sensitive than are eggs and larvae (Laska 1973). A herbicide identical to pirimicarb has also been found to affect a number of syrphid species (Proctor and Baranyovits 1969). Researchers in India have studied the effects of several botanicals and pesticides on Eupeodes confrater in order to create safer insecticides for the sugarcane crop (Likhil and Mallapur 2009).

20.9

Field Conditions

The populations of helpful insects in agricultural fields will increase as a result of greater comprehension of the impact of chemical pesticides on repellency and how these chemicals alter the physiology of beneficial insects (Ndakidemi et al. 2016). Studies conducted in the field on the cotton (cv. Bt. 207) variety in 2009 at Sindh, show that neem oil, imidacloprid, and profenofos were shown to have the greatest effects on the control plot’s natural enemy activity, as suggested by Sahito et al. (2011). Beneficial arthropods can directly come in contact with spray droplets, absorb residues when they touch polluted plant surfaces, or eat prey, nectar, or honeydew that has been exposed to insecticides in the fields (i.e. uptake of insecticide-contaminated food sources). In field conditions, Costa et al. (2018) investigated the effects of cyhalothrin on the lady beetle Eriopis connexa. Insecticides cyantraniliprole, flupyradifurone, pyrifluquinazon, and sulfoxaflor were tested in the field for their selectivity on the cotton arthropod community (27 taxa measured), which includes the important generalist predator taxa Collops spp., Orius tristicolor, Geocoris spp., Misumenops celer, Drapetis nr. divergens, Chrys Predator numbers were infrequently impacted by insecticides, and the entire arthropod predator community was preserved when compared to positive controls

20.10

Insect Growth Regulators (IGRs)

759

that had been treated with acephate and an untreated check. Instead of a direct harmful effect, rare, large reductions in predator abundances were probably caused by decreasing prey availability after insecticide spraying (Bordini et al. 2021). The studied pesticides are selective and suitable for long-term pest management in the Arizona cotton system, according to Bordini et al.’s (2021) study. Usually, by enhancing changes in predator-to-prey ratios novel possibilities for insect pest control arise, which protect natural adversaries and assist in biological control.

20.10 Insect Growth Regulators (IGRs) Insect growth regulators come in three different varieties: ecdysone antagonists, chitin synthesis inhibitors, and juvenile hormone mimics (Table 20.4). The larvae, nymphs, grubs, or naiads of some insect pests are directly impacted by insect growth regulators. The polymer known as chitin, which is found in the exocuticle, trachea, ovarioles, and peritrophic membrane of the insect midgut, is made up of N-acetyl-Dglucosamine units. Changes in chitin biosynthesis have physiological effects that have an impact on both insect growth and reproduction. These chemicals are regarded to be generally safe for natural enemies since they have an indirect effect on adults’ chitin production, particularly during the growth stages. A benzoylurea called lufenuron is more effective when consumed than when applied topically. It has a long shelf life on leaves and is regarded as a product with little negative environmental impact. Numerous research studies have examined, in both lab and field settings, the indirect effects of insect growth regulators on natural enemies, including parasitoids and carnivores. It has been considered that insect growth regulators are acceptable with minimal indirect impacts on natural enemies. However, the indirect effects of insect growth regulators on natural enemies differ substantially and mostly vary depending on the type of natural enemy, the insect growth regulators, the life stage being taken into account, and the timing of application (spatially and temporally). Buprofezin has a negligible impact on the developing stages of Cryptochaetum iceryae (Diptera: Cryptochaetidae). Buprofezin, fenoxycarb, or pyriproxyfen were applied to none of the larvae of Rodolia cardinalis Mulsant (Coleoptera: Coccinellidae), which later matured into adults. Chilocorus bipustulatus egg hatching was entirely suppressed by buprofezin and chlorfluazuron (Coleoptera: Coccinellidae). Both egg hatching and larval development of Elatophilus hebraicus (Hemiptera: Anthocoridae) were unaffected by buprofezin but were completely suppressed by fenoxycarb or pyriproxyfen given before or after oviposition on pine needles (Mendel et al. 1994). Buprofezin (Applaud®) and triflumuron (Alsystin®), two chitin synthesis inhibitors, were tested in a lab assay against coccinellids Rodolia cardinalis, native Rodolia spp., and Chilocorus nigrita and egg viability of the coccinellids Chilocorus nigrita and Cryptolaemus montrouzieri. Hat carried out these studies in 1994. Exposure to residue-bearing foliage had no impact on the number of eggs laid by Chilocorus nigrita, but, when adults were exposed to weathered residues that were 3, 7, or 19 weeks old from a single

760

20

Biosafety Assessment of Synthetic Pesticides

Table 20.4 Examples for ecdysone antagonists, chitin synthesis inhibitors, and juvenile hormone mimics IGR Cyromazine, diflubenzuron

Pyriproxyfen, fenoxycarb Buprofezin (Applaud®) and triflumuron (Alsystin®), methoxyfenozide, pyriproxyfen, tebufenozide, lufenuron, chlorfluazuron, teflubenzuron, flufenoxuron, hexaflumuron and lufenuron Kinoprene

Activity Moulting is disrupted by an insect growth regulator that alters cuticle sclerotisation by making insects’ cuticles stiffer A juvenile hormone analogue Chitin synthesis inhibitor

Juvenile hormone mimic

application of pyriproxyfen or triflumuron, a complete failure of eggs to hatch was ensured. Three-week-old buprofezin residues exhibited the same effect, but 7- and 19-week-old residues no longer greatly decreased egg viability. Both Cryptolaemus montrouzieri and Cryptolaemus nigrita adults began to lay viable eggs 20 days after being deprived of all residue-bearing leaves. Remains of triflumuron and pyriproxyfen from a week ago both significantly reduced Chilocorus montrouzieri’s capacity to create progeny. Ten-week-old triflumuron residues were still detrimental to this species, whereas similarly aged pyriproxyfen residues were not (Hattingh and Tate 1995). A different study found that exposing Podisus maculiventris’s fifth instars to pyriproxyfen had no adverse impact on reproduction (De Clercq et al. 1995). Buprofezin (100 mg active component per L) had no direct impact on Chrysoperla rufilabris’s egg viability or subsequent development. However, the higher rates (500 and 1000 mg active component per L) when used on first instars did slow down their general development till adulthood; II and III instars and pupae were unaffected (Liu and Chen 2000). Similar to this, an Orius sp. exposed to pyriproxyfen in a laboratory context did not exhibit any unintended effects on developmental cycles, female life, or reproduction (Nagai 1990); however, these findings may not be conclusive given that control mortality was close to 70%. Additionally, pyriproxyfen has been shown to significantly slow down Chrysoperla rufilabris’s immature growth (Chen and Liu 2002), although it had no noticeable effects on Delphastus catalinae’s female fecundity after adults consumed treated Bemisia tabaci (sweet potato whitefly) eggs (Liu and Stansly 2004). Medina et al. (2002) used [14C]-labelled isotopes to conduct an experiment to understand the mechanism of action of IGR. When Chrysoperla carnea (Neuroptera: Chrysopidae) adults were topically treated with diflubenzuron (DFB) at dosages based on the maximum field recommended concentration before and after the start of oviposition, egg hatching was completely inhibited due to the death of the embryo. Tebufenozide (TEB) and pyriproxyfen (PYR) had no effect on fecundity or egg fertility, in contrast. Using [14C]-labelled isotopes of each insect growth regulator, the patterns of penetration through the cuticle, distribution inside the insect body, and excretion were examined to help

20.10

Insect Growth Regulators (IGRs)

761

explain these variations in toxicity (IGR). DFB and TEB penetration levels were around 16% and 26%, respectively, after 7 days, compared to 88% for PYR after 24 h. In contrast to DFB and TEB, PYR had an extremely high excretion rate. With the exception of DFB, the ovaries, and the eggs that were deposited over the course of a week, little radioactivity that had been absorbed was found in the female body. The concentration of DFB and PYR peaked in the eggs laid on the second and fourth post-treatment days, respectively. According to the most recent evidence, it is important to penetrate the insect cuticle. However, other mechanisms are probably at play in the current IGRs’ preference for this helpful bug (Medina et al. 2002). IGRs are not compatible with integrated pest management (IPM) for citrus in southern Africa, where coccinellid biocontrol agents play a significant role, according to the authors’ conclusion. The effects of two insect growth regulators (IGRs) and conventional insecticides on natural enemies were compared in field tests from 1997 to 1999. Insecticide regimes based on the initial application of the IGR buprofezin or pyriproxyfen decreased the densities of 8 of the 20 predator taxa examined in at least 1 year when compared to an untreated control, including common species like Geocoris punctipes, Nabis alternatus, Chrysoperla carnea, and the empidid fly Drapetis nr. divergens. Predator and pest population change patterns in relation to IGR application dates indicate that variables other than direct harmful impacts, such as decreased prey availability, may also be at play; eggs caused lethal and sub-lethal effects in adult lady beetles of the Delphastus catalinae species. Feeding on whitefly eggs treated with pyriproxyfen did not significantly shorten the lifespan of male or female beetles, whereas feeding on eggs treated with buprofezin did. The pre-oviposition period was also unaffected by pyriproxyfen but was extended by buprofezin by 3–6 days. Buprofezin decreased Delphastus catalinae egg production and oviposition times; however, a 28-day pyriproxyfen treatment actually enhanced these factors. Both IGRs, particularly the higher doses of pyriproxyfen and both doses of buprofezin, decreased the fertility of Delphastus catalinae eggs. The transfer of Delphastus catalinae to water-treated whitefly eggs in the case of pyriproxyfen, however, largely reversed the process. There was no overall difference in viable egg production after a 28-day exposure to the low amount of pyriproxyfen since Delphastus catalinae fecundity was actually increased (Liu and Stansly 2004). In addition, eating eggs from the insect growth regulator buprofezin-treated sweet potato whitefly (B. tabaci) reduced female fertility and fecundity and sterilised the males of the predatory coccinellid Delphastus catalinae (Liu and Stansly 2004), indicating that this insect growth regulator is incompatible with these insects. Buprofezin, however, had no effect on the Geocoris species, Harmonia axyridis, Stethorus punctum picipes (Coleoptera: Coccinellidae), or the predatory bug, Orius tristicolor, from nymphs to adults (Hemiptera: Geocoridae) (James 2004). Prior studies have shown that Chrysoperla carnea larvae are susceptible to topical lufenuron treatments during their first instars (Hussain et al. 2012; Golmohammadi and Hejazi 2014). However, other impacts of lufenuron, such as the impact of feeding on lufenuron-treated prey on predators, predator preferences for lufenurontreated and untreated prey, and the effects of lufenuron consumption on adults, have

762

20

Biosafety Assessment of Synthetic Pesticides

not been researched in this species. To determine the true impact of this insecticide on Chrysoperla carnea, evaluation of these predator–prey interactions is crucial (Suárez-López et al. 2020). Six insect growth regulators—buprofezin, methoxyfenozide, pyriproxyfen, diflubenzuron, tebufenozide, and lufenuron—were sprayed onto first instar Ceraeochrysa cincta larvae, and Rugno et al. (2016) evaluated their lethal and sub-lethal effects. The first instar larvae of Ceraeochrysa cincta were severely harmed by lufenuron and diflubenzuron, which resulted in 100% mortality before they reached the second instar. Buprofezin significantly improved the fraction of females in the Ceraeochrysa cincta population that survived while also causing about 25% mortality of the larvae and significantly reducing the fertility and longevity of the insects. The duration and survival of the juvenile stages were unaffected by methoxyfenozide and tebufenozide, whereas methoxyfenozide drastically decreased the insects’ fertility and lifespan. Pyriproxyfen decreased the larval stage’s chance of survival by 19.5%, but it had no effect on the development, survival, or ability of the surviving individuals to reproduce. Diflubenzuron and lufenuron were determined to be toxic to Ceraeochrysa cincta based on the reduction coefficient, but buprofezin and methoxyfenozide were marginally dangerous, and tebufenozide and pyriproxyfen were innocuous (says Rugno et al. 2016). Buprofezin and methoxyfenozide significantly decreased the R o, r and λ of Ceraeochrysa cincta, according to the estimation of life table parameters, whereas pyriproxyfen and tebufenozide had no negative effects on population parameters, suggesting that these insecticides might be used in pest management programmes to protect and boost the population of the predator in agro-ecosystems. The usage of pesticides that prevent the synthesis of chitin has risen, and this includes benzoylureas (like lufenuron), which have a detrimental effect on the development of phytophagous insects’ juvenile stages through touch and ingestion. In this investigation, interactions between the drug lufenuron and the predator Chrysoperla carnea were examined. Lufenuron treatment given to eggs 24 or 48 h after oviposition had no impact on the percentage of eggs that hatched successfully or on the survival of the neonate larvae that resulted. When lufenuron was applied topically to Chrysoperla carnea second instar larvae (L2), it caused substantial mortality rates; the LC50 value was 0.0153 mL L-1 (0.0086–0.0236 mL L-1). Additionally, the third instar larvae of Chrysoperla carnea that ate prey treated with lufenuron 24 h earlier (at a dose of 1 mL L-1) had their growth severely shortened, and, when they reached the pupal stage, there was a high proportion of mortality. In choice bioassays, a significant proportion of Chrysoperla carnea larvae preferred lufenuron-treated prey to untreated prey. The viability of the resultant eggs was significantly decreased when lufenuron was consumed by Chrysoperla carnea adults, but it had no influence on female fertility or adult longevity (Suárez-López et al. 2020). These findings suggest that the developmental stage of Chrysoperla carnea should be taken into account when determining whether to apply lufenuron within IPM techniques. On the biology and behaviour of Chrysoperla carnea, the effects of four IGRs were investigated. IGRs, together with their comparatively low and high

20.10

Insect Growth Regulators (IGRs)

763

dosages, were sprayed on eggs, larvae (24-h-old), and pupae. When Chrysoperla carnea pupae were treated, the effects on egg, larval, and pupal survival as well as adult Chrysoperla carnea fecundity and fertility were evaluated. The IGRs were categorised using the proposed IOBC (International Organization for Biological and Integrated Control) toxicity scale. Diflubenzuron, pyriproxyfen, and lufenuron were classified as just mildly detrimental (class 2) to the eggs, but buprofezin was classified as moderately dangerous (class 3). To Chrysoperla carnea larvae, lufenuron and diflubenzuron were rated as slightly dangerous (class 2), whereas pyriproxyfen and buprofezin were rated as innocuous (class 1). The amount of time that the larvae spent in the treated and untreated zones was equal, indicating that buprofezin had no effect on the larvae’s locomotor behaviour. However, all other behaviours were adversely affected. Pyriproxyfen and buprofezin, which were shown to be somewhat detrimental (class 2) and moderately harmful (class 3), in relation to the pupae, significantly decreased the fecundity and fertility of the Chrysoperla carnea adults (Table 20.5). When pupae were treated, lufenuron and diflubenzuron did not have a substantial impact. In comparison, did the combined effects of the IGRs lufenuron and diflubenzuron have a substantial impact on population parameters? This may help with the integration of IGRs and Chrysoperla carnea for the purpose of preserving both species in agro-ecosystems. The two most important neotropical predators of agricultural pests are Chrysoperla externa and Eriopis connexa. Rimoldi et al. (2017) exposed their eggs to the insecticides pyriproxyfen and acetamiprid. They found that Eriopis connexa’s cumulative survival was decreased at all concentrations of both insecticides, whereas Chrysoperla externa’s survival was considerably diminished Table 20.5 Influence of insect growth regulators on various predators with references Predators Ceraeochrysa cincta

IGR Buprofezin, methoxyfenozide, pyriproxyfen, diflubenzuron, tebufenozide, lufenuron,

Safer IGR Pyriproxyfen, tebufenozide

Delphastus catalinae

Buprofezin, triflumuron

Triflumuron

Chrysoperla externa

Acetamiprid

Acetamiprid

Eriopis connexa

Acetamiprid

Acetamiprid

Harmonia axyridis Stethorus punctum picipes Geocoris spp.

Acetamiprid

Buprofezin

Acetamiprid

Buprofezin

Acetamiprid

Buprofezin

References Rugno et al. (2016) Liu and Stansly (2004) Rimoldi et al. (2017) Rimoldi et al. (2017) James (2004) James (2004) James (2004)

764

20

Biosafety Assessment of Synthetic Pesticides

by acetamiprid and pyriproxyfen at levels of 50 mg L-1 and 37.6 mg L-1, respectively. In both species, the reductions primarily impacted the eggs and first larval instar. The mean Eriopis connexa survival time in pesticide treatments was significantly shorter than that in controls, and survival curves generally deviated from those in controls. Various pesticide doses lengthened the first larval instar’s developmental periods of Eriopis connexa and Chrysoperla externa, respectively. Furthermore, pyriproxyfen prolonged the fourth instar of Eriopis connexa while shortening the first larval instar’s period. Acetamiprid was more toxic to Eriopis connexa than to Chrysoperla externa at the two highest concentrations (Rimoldi et al. 2017). They came to the conclusion that the relative toxicity of pyriproxyfen to the two species had been reversed.

20.10.1 Botanical IGRs Both primary and secondary metabolites are present in plants. Secondary metabolites aid in protecting plants from herbivore harm, but to sustain the fundamental physiological functions that are also important sources of food for herbivores, primary metabolites are required. Only 4 of the 43 plant families—Meliaceae, Asteraceae, Labiatae, and Leguminosae—having phyto-antifeedants, a class of bioactive compounds—have shown signs of stress. The isoprene units are used to classify terpenes. The four subgroups of terpenes—monoterpenes, sesquiterpenes, diterpenes, and triterpenes—all contain a variety of chemicals that behave as antifeedants. Plant-derived coumarins, alkaloids, steroids, and flavonoids may potentially exhibit antifeedant properties. Azadirachtin A, which comes from the Azadirachta indica plant, is one of the most widely used antifeedants and is used to treat 400 different insect species from the families of Blattodea, Coleoptera, Diptera, Dermaptera, Ensifera, Hymenoptera, Lepidoptera, Isoptera, Phasmida, Thysanoptera, and Siphonaptera (Shin-Foon 1989). Azadirachta indica, Melia toosendan, Melia azedarach, Tripterygium wilfordii, Tripterygium hypoglaucum, and Tephrosia vogelii were used in his laboratory and outdoor tests.

20.11 Neuropeptides One or more tiny molecular (classic) neurotransmitters invariably coexist in nerve cells with neuropeptides, which are auxiliary messenger molecules. Neuropeptides function as both transmitters and trophic factors, and they are especially important when the nervous system is under stress due to an injury, discomfort, or tension. Neuropeptides include, for instance, oxytocin, vasopressin, thyroid-stimulating hormone (TSH), luteinising hormone (LH), growth hormone (GH), insulin, and glucagon. Acetylcholine, dopamine, serotonin, and histamine are all examples of neurotransmitters. The creation and use of neuropeptide synthetic analogues or mimics is one approach that may be taken in the hunt for more environmentally friendly and target-specific insecticidal medications. As regulating peptides that play

20.12

Indirect Effects of Pesticides on Natural Enemies

765

a role in metabolism, homeostasis, growth and development, behaviour and reproduction, and muscle movement in insects, neuropeptides are classified as such. Because of their great specificity, neuropeptides and their corresponding receptors can be included into an insecticidal victim system to specifically reduce the fitness of target pest insects while reducing adverse environmental consequences. Insect kinins are multifunctional neuropeptides that have a conserved C-terminal pentapeptide motif: Tyr-X-X-X-X (His, Asn, Ser, or Phe)-Trp-Gly-NH2. Insect kinins, which are found in the majority of insects, carry out a number of functions, including releasing digestive enzymes, inhibiting the growth of larval weight, and responding to desiccation and starvation stress. Kinins are found in insects that are attached to certain tissues and are quickly broken down by peptidases in the insect haemolymph. A single aminoisobutyric acid (AIB) or hydroxycinnamic acid (HCA) can be added in the second or third places of the kinin active core to fix this issue. By preventing tissue-bound peptidase from accessing the main hydrolysis site, the kinin’s biostability is improved. On the survival of the common green lacewing Chrysoperla carnea (Neuroptera), three topically applied kinin analogues (1728 ((Aib)FF(Aib)WGa), 2139 (FF(Aib)WGa), and 2460 (HCA-R(Aib)WGa) were assessed (Shi et al. 2022a). The authors discovered that the 2460 analogue (HCA-R(Aib)WGa) efficiently decreased aphids, resulting in a 46% mortality rate within 5 days of treatment. On the other hand, 2460 had no appreciable impact on Chrysoperla carnea, a beneficial insect, in terms of survival, food intake, or weight gain. Three natural enemies were investigated in a lab environment using the caffeic acid phenethyl amide–pyrokinin (CAPA–PK) analogues 1895 (2Abf-Suc-FGPRLamide) and 2315 (ASG-[3L]-VAFPRVamide) (Chrysoperla carnea, Nasonia vitripennis, and Adalia bipunctata). There were no discernible changes in food consumption, weight gain, or survival between the pollinator and the three representative natural adversaries (Shi et al. 2022b). These results may facilitate the development of CAPA analogues as safer and more effective broad-spectrum pesticide substitutes.

20.12 Indirect Effects of Pesticides on Natural Enemies Studies on the indirect effects of pesticides on natural enemies have not been as thorough as those on the direct impacts, and those studies mostly focused on assessing fecundity and longevity (Cloyd 2012). . . . .

Adult longevity, fecundity, reproduction Time for incubation Time for development (stadial period) Mobility, prey-searching effectiveness, feeding habit, predation rate (number of preys ingested in a 24-h period), sex ratio (number of a certain gender to the total number of adults that have emerged), and predation rate are all factors. . Emergence rates (the proportion of adults born after hatching to the total number of adults raised)

766

20

Biosafety Assessment of Synthetic Pesticides

. Consumption of prey (during a particular stage or whole life stage) . Changes in the population . Prey acceptance, repellency, orientation behaviour (for oviposition by female parasitoids) Indirect effects may impede the ability of natural enemies to establish populations, suppress their capacity to utilise prey, affect parasitism (for parasitoids) or consumption (for predators) rates, reduce female reproduction, reduce the availability of prey, hinder the ability of natural enemies to recognise prey, affect the sex ratio (females:males), and restrict mobility. Additionally, more than one physiological and/or behavioural feature may be indirectly altered following pesticide exposure. Understanding the indirect effects of different pesticide doses on fecundity, fertility, reproduction, adult and larval longevity, and prey consumption is essential for successfully integrating natural enemies with pesticides and avoiding any indirect effects on population dynamics. The critical physiological and behavioural elements stated above allow natural enemies to manage arthropod pest populations. Age, kind (parasitoid vs. predator), life phases (immature vs. adult), and sex (male vs. female) of natural enemies are a few factors that may influence the indirect effects of pesticides (Cloyd and Bethke 2011). In addition, depending on exposure, the kind of pesticide (nerve toxin vs. non-nerve toxin) and the method of administration (foliar vs. systemic) may have a significant impact on the severity of any indirect effects on natural enemies (immediate vs. chronic). For example, even while foliar pesticide applications do not directly influence natural enemies, they may nevertheless have indirect effects because they typically indicate immediate exposure. Remains from foliar treatments may also have an indirect effect by preventing plants from producing volatile cues that some natural enemies employ to identify victims (prey patches) over considerable distances within plant groups. This might have an impact on foraging behaviours and search efficiency (Morgan and Hare 2003; Teodoro et al. 2009). The following are a few of the indirect effects (Riaza et al. 2022): . Limited capacity for prey capture: Doses of cypermethrin made it harder for predators to find and catch victims. In a study, it was discovered that parasitoids treated with carbamates have less ability to guide themselves to plants with aphid infestations. . There are fewer food sources for predators, parasitoids, and pollinators: Insecticides can lower the population of insects that feed other beneficial insects. The elimination of hosts or prey would lead to a dearth of food sources for beneficial insects and natural enemies due to insecticidal effects. They may be forced to travel in search of fresh prey or hosts due to this scarcity. For instance, bumblebee and honeybee movement towards food sources was inhibited by imidacloprid dosages. . Negative effects on the growth of parasitoids and predators: The sub-lethal effects of the important aphid parasitoid Aphidius ervi may result from insecticidal residues.

20.12

Indirect Effects of Pesticides on Natural Enemies

767

20.12.1 Development, Adult Longevity, and Reproduction Notably, imidacloprid alone increased the duration of the third larval instar. Furthermore, the organosilicones Silwet 618, ECO, and Silwet L-77 significantly reduced this impact. Comparing pupae exposed to imidacloprid alone to control pupae, the pupal developmental duration was considerably prolonged. A longer length of pupal development was also observed when imidacloprid was administered to C. nipponensis coupled with Silwet 618, ECO, or Silwet L-77 during the second larval instar stage in contrast to imidacloprid alone. Similar results were observed for pupal weight following treatment with imidacloprid, both on its own and in combination with Silwet 618, ECO, or Silwet L-77. Imidacloprid treatment alone and treatment combining imidacloprid with Silwet 618, STIK2, ECO, or Silwet L-77, in particular, differed significantly from each other. Imidacloprid had no effect on the emergence rate of C. nipponensis in comparison to the control group. Silwet L-77 in conjunction with imidacloprid lowered the emergence rate of C. nipponensis, but Silwet 618, STIK2, or ECO had no impact on it (Liu et al. 2021). During the second larval instar stage, when C. nipponensis female adults were subjected to imidacloprid at LC30, their longevity was significantly decreased. Female longevity also varied significantly depending on whether imidacloprid was taken alone or in combination with Silwet L-77. Additionally, compared to other treatments, the adult pre-oviposition period (APOP) of C. nipponensis became much longer when exposed to imidacloprid mixed with Silwet L-77 at the second larval instar stage. The duration of oviposition was greatly reduced when imidacloprid was used alone or in combination with the organosilicones Silwet 618, Silwet L-77, STIK2, or ECO. Additionally, when C. nipponensis was treated with the combination of imidacloprid and Silwet L-77 in contrast to imidacloprid therapy, the oviposition time was dramatically decreased. Additionally, the total quantity of eggs laid by each female drastically decreased after treatment with the medication combination of imidacloprid and Silwet L-77 (Liu et al. 2021). The generalist, voracious predator Cycloneda sanguinea is found in both cultivated and uncultivated environments. This is the only insecticide that proved extremely harmful to Cycloneda sanguinea via touch (mortality rate of 90%) yet is non-toxic when consumed through consumption of aphids tainted with chlorantraniliprole. It is interesting to note that Cycloneda sanguinea was highly poisonous (100% mortality) after ingesting methomyl- and chlorfenapyrcontaminated prey. Through exposure to polluted surfaces and after ingesting contaminated prey, spinosad was not hazardous to Cycloneda sanguinea. However, whereas chlorfenapyr only had an impact on vertical flight, direct interaction of the insects with methomyl and spinosad dramatically reduced Cycloneda sanguinea’s flying activity (vertical flight and free-fall flight) (da Silva et al. 2022). Nesidiocoris tenuis is one of the species that is most often used for biological control in the Palaearctic (Hemiptera: Miridae). It plays a complex role in greenhouse pest control because of its potent effectiveness against a wide range of pests, including aphids, whiteflies, and lepidopterans like the South American tomato pinworm, Tuta absoluta (Lepidoptera: Gelechiidae). Predator survival varied with

768

20

Biosafety Assessment of Synthetic Pesticides

pesticide types and concentrations; for spinosad and lambda-cyhalothrin, the LC30/ label rate ratios ranged from 8.45% to 65.40%. When exposed to all pesticides at estimated low-lethal concentrations, female Nesidiocoris tenuis was less fertile. Lambda-cyhalothrin and spinosad only had the same effect at LC30, but chlorpyrifos significantly reduced predator propensity for a host plant even at LC1. Lambdacyhalothrin (at all doses) and chlorpyrifos (at LC10 and LC30) also affected how long Nesidiocoris tenuis females took to decide (Passos et al. 2022). Su et al. found that acetamiprid and dinotefuran at LD30 significantly lengthened the Chrysopa pallens adult pre-oviposition period (APOP) and total pre-oviposition period (TPOP) (Su et al. 2022). In contrast, there was no discernible difference between the two pesticides’ longevity, fecundity, reproductive days, preadult survival rate (%), intrinsic rate of rise (r), net reproductive rate (R0), and finite rate of growth at LD10 and LD30. These results provide scientific foundations for the responsible use of these two pesticides as well as for the use and storage of Chrysopa pallens.

20.12.2 Functional Response Serangium japonicum is a significant Bemisia tabaci predator in China. He et al. (2012) examined the effects of imidacloprid’s toxicity on Serangium japonicum’s functional response to Bemisia tabaci eggs. Adult Serangium japonicum plants subjected for 24 h to dried residues of imidacloprid at the recommended field rate (4 g active ingredient per 100 L, or 40 ppm (part per million)) on cotton against Bemisia tabaci revealed substantial mortality rates. The lowest rate, 5 ppm, was determined to be sub-lethal because it did not significantly increase mortality compared to the control group. The risk for Serangium japonicum in treated fields (HQ > 2) was determined to be 11.54 ppm for the lethal rate of 50 and 3.47 for the hazard quotient (HQ), respectively. Serangium japonicum’s functional reaction to Bemisia tabaci eggs was impacted by exposure to dried imidacloprid residues at the sub-lethal rate (5 ppm) on cotton leaves. This resulted in longer handling times and lower egg consumption peaks. The amount of Bemisia tabaci eggs devoured on treated leaves was substantially less than that on untreated leaves, indicating that imidacloprid residues also disrupted predator voracity. After being transferred to untreated cotton leaves, all effects vanished within a few hours. Serangium japonicum demonstrated no toxicity to imidacloprid systemically given at the recommended field rate (for cotton), and the predator’s functional response was unaffected. The sub-lethal effects of imidacloprid on Serangium japonicum demonstrated in our study are probably detrimental to the development and reproductive ability of Serangium japonicum and may ultimately slow the increase of the predator population. These findings suggest the necessity of investigating imidacloprid’s possible impacts on Serangium japonicum in order to create successful integrated pest management strategies for Bemisia tabaci in China (He et al. 2012). On nymphs in the fifth instar, diazinon, fenitrothion, and chlorpyrifos were examined for their sub-lethal effects on the functional response of the predatory

20.12

Indirect Effects of Pesticides on Natural Enemies

769

insect Andrallus spinidens (Pentatomidae), a potential biological control agent. The experiment used the last instar larvae of Chilo suppressalis (Lepidoptera: Pyralidae) as prey in densities of 2, 4, 8, 16, 32, and 64. The outcomes of logistic regressions showed type-II functional responses in all pesticide treatments and the control. The mean number of preys ingested by Andrallus spinidens was significantly decreased by tested pesticides, according to a comparison of functional response curves. In comparison to the other treatments, the functional response curve of Andrallus spinidens to chlorpyrifos treatment was noticeably lower. When compared to the control, the spraying of insecticides in this study resulted in a drop in the attack rate and an increase in the handling time of exposed bugs. Treatments with chlorpyrifos and fenitrothion showed the longest handling times (3.97 h) and the lowest attack rates (0.023 preys/predator), respectively (Gholamzadeh Chitgar et al. 2014). Two mirid predatory nymphs, Macrolophus pygmaeus and Nesidiocoris tenuis, had their functional response curves tested to see how the insecticides thiacloprid and chlorantraniliprole affected them (Martinou and Stavrinides 2015). They claimed that when feeding on the eggs of the moth Ephestia kuehniella in the absence of pesticides, both predators showed a type-II functional response. The model estimated handling time was substantially lower for Nesidiocoris tenuis than for Macrolophus pygmaeus, suggesting that the former is a more effective predator. Macrolophus pygmaeus residual exposure to sub-lethal doses of either insecticide was connected to a change in the asymptote but not the functional response curve’s type. Given that it resulted in both a considerable decrease in attack frequency and an increase in handling time, thiacloprid appears to be the herbicide that is least compatible with Macrolophus pygmaeus. In contrast, exposure to chlorantraniliprole considerably lengthened the predator’s handling time but not its assault frequency. Nesidiocoris tenuis residual exposure to either insecticide at sub-lethal doses had no discernible impact on the functional response model’s type or its parameters (Martinou and Stavrinides 2015). The effects of the commonly used pesticides abamectin, imidacloprid, and chlorpyrifos on the functional response type and characteristics of Macrolophus pygmaeus, fed on Tuta absoluta eggs, were examined. Using dried residues, the 24-h sub-lethal concentration (LC30) of pesticides was assessed. According to the findings, the most harmful insecticide for Macrolophus pygmaeus adults was chlorpyrifos. The results of a functional response trial that used various densities of prey eggs—including 1, 2, 3, 4, 5, 8, 10, 15, 20, and 30—showed that predatory bugs exhibited type-II functional responses both before and after being treated with pesticides at the LC30 concentration. In contrast to abamectin, two insecticides (imidacloprid and chlorpyrifos) extended the predator’s handling time while decreasing the assault rate (Sharifian et al. 2017). The functional response of second instar Chrysoperla nipponensis larvae to Corcyra cephalonica eggs was discovered. The results showed that at all tested densities, exposure to imidacloprid alone or in combination with organosilicones Silwet 618, STIK2, ECO, or Silwet L-77 considerably decreased the amount of prey items consumed by Chrysoperla nipponensis larvae. As demonstrated by the logistic regression for all prey consumed by C. nipponensis’s second instar larvae, the

770

20

Biosafety Assessment of Synthetic Pesticides

functional response of C. nipponensis treated with imidacloprid alone or in combination with organosilicones was a type-II functional response. The attack rate and prey handling time were also measured using the random predator equation 2 to evaluate the predatory capabilities of the Chrysoperla nipponensis second instar larvae on Corcyra cephalonica eggs. The attack rate was significantly reduced by imidacloprid, especially when combined with the organosilicones Silwet 618, STIK2, ECO, or Silwet L-77. Additionally, Chrysoperla nipponensis given the imidacloprid–Silwet L-77 combination treatment had the lowest attack rate. However, there was no obvious difference in the prey handling time between the imidacloprid-treated and control groups. To tackle Chrysoperla nipponensis prey, imidacloprid must be combined with an organosilicone (such as Silwet 618, STIK2, ECO, or Silwet L-77) (Liu et al. 2021). The results indicated that integrated pest control programmes should consider the harmful effects of these pesticides on various predators (IPM).

20.12.3 Biological Traits: Bioefficacy Eocanthecona furcellata’s trehalose concentration was enhanced, and juvenile hormone (JH) titre was decreased by λ-cyhalothrin. Eocanthecona furcellata’s predation capacity and theoretical maximum predation were both reduced by cyhalothrin by 31.08% and 48.90%, respectively. In addition, JH supplementation may considerably increase trehalase and detoxifying enzyme activities following λ-cyhalothrin therapy as well as revive Eocanthecona furcellata’s predatory capacity. Our findings improve insecticide selection, boost the ecological functions of predators, and provide insight into the toxicological mechanisms by which predatory stink bug species respond to insecticides (Qiong et al. 2022). It is crucial in insecticide toxicity research, but mostly unexplored, to determine whether and how exposure to insecticides may influence the biological control effectiveness of predatory arthropods. Although other studied pesticides have been shown to promote the nymphal developmental processes of Rhynocoris kumarii and Rhynocoris marginatus, quinalphos or endosulfan are safer for Rhynocoris fuscipes (George and Ambrose 1998) (Fig. 20.2). A total of 540 individuals underwent treatments with deltamethrin, acephate, and thiamethoxam, with 180 of those insects coming from the adult stage and the other 180 from the five larval stages. The obtained data demonstrated that the median lifespan of those with various stages of Rhynocoris albopilosus varied depending on the treatments. The length of the longest middle development has been noted at the level of the individuals of the V stage for the individuals treated with acephate and thiamethoxam, with respective values of 22.91 days and 65.05 days. A general “knock-down effect,” excretions, and a cleansing of the rostrum and antennas have also been noted following therapy. In contrast to acephate, which had a survival rate that exceeded 70%, the deltamethrin and thiamethoxam on Rhynocoris albopilosus do not exceed 50%. The acephate’s role would be to provide guidance in an

20.13

Fungicides

Stadial Periods (days)

Rhynocoris fuscipes

771 Rhynocoris kumarii

Rhynocoris marginatus

100 80 60 40 20 0

Pesticides

Fig. 20.2 Rhynocoris fuscipes, Rhynocoris kumarii, and Rhynocoris marginatus stadial periods (days) are affected by the chemicals monocrotophos, dimethoate, methyl parathion, quinalphos, and endosulfan. (After George and Ambrose 1998)

integrated strategy against Rhynocoris albopilosus, but it must do so in accordance with its developmental cycle (Ghislaine et al. 2021). According to the IOBC methodology on average mortality, chlorpyriphos (85.74%), acephate (85.56%), and quinalphos (81.85%) were placed in toxicity class 3, whereas the insecticides from toxicity class 2 were listed in the following order, going from the lowest to the highest Thiacloprid, Azadirachtin, Spirotetramat/ Actamiprid, Imidacloprid, Chlorfenapyr/Chlorantraniliprole 18.5 SC/Flubendiamide 20 WG, Flubendiamide 19.92 + Thiacloprid 19.92 SC, Spirotetramat 11.01 + Imidacloprid 11.01 SC/Emamectin (Patel 2020).

20.13 Fungicides Copper-based fungicides have raised environmental concerns about the possibility of soil accumulation and effects on soil biota such earthworms, phytonematodes, microbes, etc. Some research studies have indicated adverse impacts on beneficial mite species, whereas other studies have found no negative consequences. Fungicides are frequently employed in horticultural and agricultural production systems, so it is reasonable to consider how they can affect natural enemies indirectly. Despite the fact that fungicides typically pose less of a threat to natural enemies than do insecticides and miticides (Wright and Verkerk 1995), it is crucial to assess any potential side effects and, consequently, compatibility with natural enemies. It is probable that the type of fungicide will determine compatibility with natural enemies, given that “older” fungicides may be more indirectly harmful to natural enemies than “newer,” which may be related to the way of action or any metabolites. The kind and species of natural enemies, the application’s timing (both

772

20

Biosafety Assessment of Synthetic Pesticides

geographically and temporally), and the life stage exposed can all affect how this pesticide behaves, despite being identical to other pesticides in some ways. The potential impacts of copper on the biology and ecology of insects have received little attention in study. For instance, the copper treatment lengthened Olla v-nigrum’s larval development by 8.5% (8.2 days) but had no noticeable effects on the other two species, Curinus coeruleus (16.5 days) and Harmonia axyridis (9.4 days). Because of their increased levels of activity throughout treatments, Olla v-nigrum larvae might have been exposed to copper sulphate at higher concentrations. Additionally, female Olla v-nigrum treated with copper as larvae laid fewer eggs and had a 6.6% longer pre-reproductive phase than did their control counterparts. There were no discernible differences between control and treatment females for either species in terms of the mean percentage of eggs that actually hatched. The percentage of Olla v-nigrum eggs that hatch was lower than those of the other two species as a result of some inevitable egg cannibalism by the early hatching larvae of this species. According to Edwards and Hodgson (1973), copper oxychloride did not quickly poison the mite predator Stethorus nigripes (Coccinellidae). Mani and Thorntakarya (1988) found that copper oxychloride and the Bordeaux mixture are both safe for the coccinellid Scymnus coccivora. According to Mani et al.’s (1997) observations, exposure to copper oxychloride had no negative effects on the lifespan or reproduction of the mealybug predator Cryptolaemus montrouzieri. Niranjan et al. (1998) claim that after being subjected to sub-lethal dosages of copper sulphate, the beetle Hydrophilous olivaceous underwent significant changes in the protein concentration of the haemolymph, adipose bodies, and gonads (Hydrophilidae). Bayley et al. (1995) measured both acute and chronic toxicity, including increased larval mortality and depressed adult locomotor functions while rearing Pterostichus cupreus (Carabidae) larvae on copper-contaminated food and soil in one of the more thorough investigations into how copper affects beetle biology. When Leptinotarsa decemlineata larvae were fed mineral salts on potato leaves, Izhevskii (1976) discovered that copper and sulphate salts hindered the beetle’s enzymatic activity and delayed the growth of the larvae. According to Edwards and Hodgson (1973), copper oxychloride did not quickly poison the mite predator Stethorus nigripes (Coccinellidae). Scymnus coccivora. According to Mani et al. (1997), exposure to copper oxychloride had no negative impact on the lifespan or reproduction of the mealybug predator Cryptolaemus montrouzieri. Niranjan et al. (1998) claim that after being subjected to sub-lethal dosages of copper sulphate, the beetle Hydrophilous olivaceous underwent significant changes in the protein concentration of the haemolymph, adipose bodies, and gonads (Hydrophilidae). Bayley et al. (1995) measured both acute and chronic toxicity, including increased larval mortality and depressed adult locomotor functions, while rearing Pterostichus cupreus (Carabidae) larvae on copper-contaminated food and soil in one of the more thorough investigations into how copper affects beetle biology. However, Izhevskii (1976) found that feeding mineral salts to Leptinotarsa decemlineata (Chrysomelidae) larvae on potato leaves caused the beetles’ enzymatic activity to be blocked and delayed larval development. Eight pesticides significantly reduced

20.14

Beneficial Impacts: Hormesis

773

the survival of the insects, whereas four fungicides—aluminium tris, azoxystrobin, fenhexamid, and kresoxim-methyl—had no indirect effects on the predatory flower bug Orius insidiosus (Hemiptera: Anthocoridae) (Herrick and Cloyd 2017).

20.14 Beneficial Impacts: Hormesis “Hormesis” is the term used for biological stimulation brought on by modest dosages of toxins or other stresses. Hormonal stimulation of pest insect life cycle features can have a negative influence on agriculture, but it can also be used to boost the effectiveness of biological control treatments by stimulating beneficial insects. Another possible application of sub-lethal (hormetic) stress that has not yet been thoroughly researched is its potential to maximise the generation of natural enemies (artificial or otherwise) for mass release. The following publications, however, were accessible in the literature. Secondary poisoning frequently results in the natural enemies’ demise in laboratory and field research. In studies conducted in the lab, Helicoverpa armigera (Noctuidae) cotton bollworm larvae in their second instar that had been treated with azadirachtin were less likely to survive after being consumed by Mallada signatus (Neuroptera: Chrysopidae) lacewing larvae (Qi et al. 2001). Three carabid predators, Pterostichus madidus, Pterostichus melanarius, and Nebria brevicollis (Carabidae), had high mortality levels as a result of exposure to dimethoate residues in prey aphids treated at field exposure rates, according to Scarpellini and Andrade (2011). Additionally, the ladybug Cycloneda sanguinea (Coccinellidae) showed this decrease in mortality (Mauchline et al. 2004). In the laboratory, huge percentages of Micromus tasmaniae (Neuroptera: Hemerobiidae) lacewing larvae perished after feeding on Nasonovia ribisnigri (Hemiptera: Aphididae) lettuce aphids, with imidacloprid treatments producing 96% mortality and pirimicarb treatments 30–40% mortality (Walker et al. 2007). Similarly, systemic pesticides may have negative effects on predatory insects like the lacewing Chrysoperla carnea (Neuroptera: Chrysopidae), which also consumes extra-floral nectar-carrying residues, and Orius insidiosus (Hemiptera: Anthocoridae), which frequently feeds on plant sap (Gontijo et al. 2014). Orius insidiosus, a hemipteran predator exposed to thiamethoxam residues through treated vegetative tissue and insect prey, was tested for toxicity by Camargo et al. (2017). According to the results, the concentrations that were needed to kill more than half of the examined insects were greater than the concentrations that the insects are likely to encounter in the field. Orius insidiosus consumption of Aphis glycines Matsumura was impacted at 10 ng mL-1 and 5 ng mL-1 of thiamethoxam at 24 h of evaluation. Orius insidiosus experienced considerable mortality 24 h after being exposed to aphids treated with thiamethoxam at these doses (Camargo et al. 2017). The consumption of polluted food suggests that carabids eating in treated fields and field borders could die as a result of an indirect exposure pathway. Due to emergence failure, parasitoid wasp populations frequently decline because their larvae feed on hosts that have been exposed to insecticides or insect growth

774

20

Biosafety Assessment of Synthetic Pesticides

regulators. Pyriproxyfen is an illustration of the latter since it only affects insect eggs and larvae rather than adult insects. Thus, the parasitoid wasp Trissolcus japonicus (Hymenoptera: Scelionidae) was found ovipositing when given brown stink bug eggs produced by females treated with pyriproxyfen (Pentatomidae); however, adult wasps failed to emerge from the host eggs (Penca and Hodges 2017). Target pests and other non-target insects are often eliminated when pesticides are used on a crop, although individual insects may not always die right away. Depending on the treatment dose that each insect receives, the time it takes to die can be anywhere between a few minutes and a few days. Natural predators of the affected species might become poisoned once more, alter their hunting behaviour, or even die in the interim. For instance, in imidacloprid-treated cabbage plots, the spined soldier bug, Podisus maculiventris (Hemiptera: Pentatomidae), was blocked and its body weight was lowered as the bugs consumed diamond moths, Plutella xylostella (Lepidoptera: Plutellidae); curiously, the insecticide did not significantly reduce the moth numbers despite being used (Resende-Silva et al. 2019). Insects’ biological processes and functions can be stimulated by specific low doses of insecticide along the dose–response continuum, which are sub-lethal concentrations below those that have inhibitory effects. Hormesis is the name for the bi-phasic dose–response paradigm, which is characterised by biological suppression at high doses of stress and activation at low concentrations of stress. Hormetic reactions are generally believed to increase survival and resilience to environmental stress, regulating physiology and the allocation of resources to preserve organismal stability under challenging circumstances, regardless of the organism or stressor. Although other effects such as increased longevity and immunity, heightened behaviour, and preconditioning for stress have been observed, insecticide-induced hormesis is typically seen in insects as increased reproduction (Rix and Cutler 2020). Podisus maculiventris (Pentatomidae) fecundity was enhanced at 0.5 and 1.0 mg L-1 imidacloprid (2% of the field rate) when treated as young adults, according to Rix and Cutler’s experiments (Rix and Cutler 2020), with no modifications in time to oviposition, fertility, or survival. Nymphs treated with imidacloprid at concentrations of 0.15 and 1.5 mg L-1 boosted fecundity at the price of fertility and survival, whereas nymphs exposed to 0.015 mg L-1 (less than 1% of the field rate) had stimulated reproduction as well. In a different experiment, we discovered that trans-generational reproductive stimulation is possible without significantly lowering fertility or survival. Our findings imply that producers of biocontrol agents may be able to deliberately administer low doses of stress to natural enemies during culture without endangering fitness in succeeding generations. Insect growth regulators can also reduce the survival of parasitoid wasps by altering the biochemistry and histology of the midgut epithelium, as is the case with lufenuron on the stink bug predator Podisus nigrispinus (Hemiptera: Pentatomidae), which serves as the main biological control agent of the cotton leafworms Alabama argillacea (Lepidoptera: Erebidae) (Lira et al. 2020). The persistence of systemic insecticides in plant and fungal tissues acts as a death trap for non-target mycophagous insects like the 20-spotted lady beetle, Psyllobora vigintimaculata (Coleoptera: Coccinellidae), which eats powdery mildew conidia

References

775

and hyphae (Erysiphales). These dangerous fungi grow on treated plants and act as storage sites for the pesticides used, such as imidacloprid, harming the ladybugs that act as the fungi’s natural pest controllers (Choudhury et al. 2020).

20.15 Future Recommendations . It is possible to conserve predators in the field by minimising chemical and physical habitat disturbances. . Insecticides should be used in combinations. . Insecticides should be combined with either botanicals or microbicides. . Biosafety evaluation is crucial before recommending any insecticides for field application. . It is necessary to study the sub-lethal effects of as many insecticides against as many predatory insects. . Through research is essential. Future research should heavily focus on how pesticide and fungicide interactions affect predatory insects.

References Albajes R, López C, Pons X (2003) Predatory fauna in cornfields and response to imidacloprid seed treatment. J Econ Entomol 96(6):1805–1813 Ambrose DP (1999) Assassin bugs. Science Publication, Enfield, NH, p 337 Argentine JA, Jansson RK, Halliday WR et al (2002) Potency, spectrum and residual activity of four new insecticides under glasshouse conditions. FL Entomol 85:552–556 Ashley JLD, Ames Herbert EE, Lewis CC et al (2006) Toxicity of three acaricides to Tetranychus urticae (Tetranychidae: Acari) and Orius insidiosus (Anthocoridae: Hemiptera). J Econ Entomol 99:54–59 Azab AK, Megahed MM, El Mirsawi DH (1971) On the biology of Bemisia tabaci (Genn.) (Hemiptera-Homoptera: Aleyrodidae). Bull Soc Entomol Egypte 55:305–315 Bacci L, Picanço MC, Gusmão MR et al (2001) Seletividade de Inseticidas a Brevicoryne brassicae (L.) (Hemiptera: Aphididae) e ao predador Doru luteipes (Scudder) (Dermaptera: Forficulidae). Neotrop Entomol 30:707–713 Balanza V, Mendoza JE, Cifuentes D et al (2021) Selection for resistance to pyrethroids in the predator Orius laevigatus. Pest Manag Sci 77(5):2539–2546 Bana JK, Jat BL, Chaudhary SK (2014) Effect of insecticides on Coccinella septempunctata L. J Entomol 76(4):352–354 Barbosa PRR, Michaud JP, Bain CL et al (2017) Toxicity of three aphicides to the generalist predators Chrysoperla carnea (Neuroptera: Chrysopidae) and Orius insidiosus (Hemiptera: Anthocoridae). Ecotoxicology 26:589–599 Barbosa PRR, Oliveira MD, Barros EM et al (2018) Differential impacts of six insecticides on a mealybug and its coccinellid predator. Ecotoxicol Environ Saf 147:963–971 Batista CH, Plata-Rueda A, Zanuncio JC et al (2022) Indoxacarb effects on non-target predator, Podisus distinctus (Hemiptera: Pentatomidae). Environ Sci Pollut Res 29(20):29967–29975 Bayley M, Baatrup E, Heimbach U et al (1995) Elevated copper levels during larval development cause altered locomotor behaviour in the adult carabid beetle Pterostichus cupreus L (Coleoptera: Carabidae). Ecotoxicol Environ Saf 32:166–170

776

20

Biosafety Assessment of Synthetic Pesticides

Berrada S, Nguyen TX, Fournier D (1996) Comparative toxicities of some insecticides to Cacopsylla pyri L (Hom., Psyllidae) and one of its important biological control agents, Anthocoris nemoralis F (Het., Anthocoridae). J Appl Entomol 120(1–5):181–185 Bielza P, Fernandez E, Gravalos C et al (2009) Testing for non-target effects of spiromesifen on Eretmocerus mundus and Orius laevigatus under greenhouse conditions. BioControl 54(2): 229–236 Biondi A, Desneux N, Siscaro G et al (2012) Using organic-certified rather than synthetic pesticides may not be safer for biological control agents: selectivity and side effects of 14 pesticides on the predator Orius laevigatus. Chemosphere 87(7):803–812 Black BC, Hollingworth RM, Ahammadsahib KI et al (1994) Insecticidal action and mitochondrial uncoupling activity of AC-303,630 and related halogenated pyrroles. Pestic Biochem Physiol 50:115–128 Bordini I, Ellsworth PC, Naranjo SE et al (2021) Novel insecticides and generalist predators support conservation biological control in cotton. Biol Control 154:104–502 Bostanian NJ, Akalach M (2006) The effect of indoxacarb and five other insecticides on Phytoseiulus persimilis (Acari: Phytoseiidae), Amblyseius fallacis (Acari: Phytoseiidae) and nymphs of Orius insidiosus (Hemiptera: Anthocoridae). Pest Manag Sci 62:334–339 Boyd ML, Boethel DJ (1998) Residual toxicity of selected insecticides to heteropteran predaceous species (Heteroptera: Lygaeidae, Nabidae, Pentatomidae) on soybean. Environ Entomol 27 (1):154–160 Bozsik A (2006) Susceptibility of adult Coccinella septempunctata (Coleoptera: Coccinellidae) to insecticides with different modes of action. Pest Manag Sci 62(7):651–654 Broughton S, Harrison J, Rahman T (2014) Effect of new and old pesticides on Orius armatus (Gross) - an Australian predator of western flower thrips, Frankliniella occidentalis (Pergande). Pest Manag Sci 70(3):389–397 Brust GE, House GJ (1990) Effects of soil moisture, no-tillage and predators on southern corn rootworm (Diabrotica undecimpunctata howardi) survival in corn agroecosystems. Agric Ecosyst Environ 31(3):199–215 Cabral S, Soares AO, Garcia P (2011) Voracity of Coccinella undecimpunctata: effects of insecticides when foraging in a prey/plant system. J Pest Sci 84:373–379 Camargo C, Hunt TE, Giesler LJ et al (2017) Thiamethoxam toxicity and effects on consumption behavior in Orius insidiosus (Hemiptera: Anthocoridae) on soybean. Environ Entomol 46 (3):693–699 Campos MR, Picanço MC, Martins JC et al (2011) Insecticide selectivity and behavioral response of the earwig Doru luteipes. Crop Prot 30(12):1535–1540 de Castro AA, Poderoso JCM, Ribeiro RC et al (2015) Demographic parameters of the insecticideexposed predator Podisus nigrispinus: implications for IPM. BioControl 60(2):231–239 Chandra A, Malik YP, Kumar A (2014) Efficacy and economics of new insecticides for management of aphid (Lipaphis erysimi) in Indian mustard. Curr Adv Agric Sci 6(1):88–90 Chen TY, Liu TX (2002) Susceptibility of immature stages of Chrysoperla rufilabris (Neurop., Chrysopidae) to Pyriproxyfen, a Juvenile Hormone. J Appl Entomol 126:125 Choudhury RA, Sutherland AM, Hengel MJ et al (2020) Imidacloprid movement into fungal conidia is lethal to mycophagous beetles. Insects 11:496 Claver MA, Ravichandran B, Khan MM et al (2003) Impact of cypermethrin on the functional response, predatory and mating behaviour of a non-target potential biological control agent Acanthaspis pedestris (Stål) (Het., Reduviidae). J Appl Entomol 127(1):18–22 Cloyd RA (2012) Indirect effects of pesticides on natural enemies. In: Soundararajan RP (ed) Pesticides—advances in chemical and botanical pesticides. InTech, Rijeka, pp 127–150 Cloyd RA, Bethke JA (2011) Impact of neonicotinoid insecticides on natural enemies in greenhouse and interiorscape environments. Pest Manag Sci 67(1):3–9 Costa PMG, Torres JB, Rondelli VM et al (2018) Field-evolved resistance to λ-cyhalothrin in the lady beetle Eriopis connexa. Bull Entomol Res 108:380–387

References

777

De Clercq P, De Cock A, Tirry L et al (1995) Toxicity of diflubenzuron and pyriproxyfen to the predatory bug Podisus maculiventris. Entomol Experiment Appl 74:17 Delbeke F, Vercruysse P, Tirry L et al (1997) Toxicity of diflubenzuron, pyriproxyfen, imidacloprid and diafenthiuron to the predatory bug Orius laevigatus (Het.: Anthocoridae). Entomophaga 42: 349–358 Edward George PJ, Ambrose DP (2004) Impact of insecticides on the haemogram of Rhynocoris kumarii Ambrose and Livingstone (Hem., Reduviidae). J Appl Entomol 128(9–10):600–604 Edwards BAB, Hodgson PJ (1973) The toxicity of commonly used orchard chemicals to Stethorus nigripes (Coleoptera: Coccinellidae). J Aust Entomol Soc 12:222–224 Elzen GW (2001) Lethal and sublethal effects of insecticide residues on Orius insidiosus (Hemiptera: Anthocoridae) and Geocoris punctipes (Hemiptera: Lygaeidae). J Econ Entomol 94(1):55–59 Funderburk J, Srivastava M, Funderburk C et al (2013) Evaluation of imidacloprid and cyantraniliprole for suitability in conservation biological control program for Orius insidiosus (Hemiptera: Anthocoridae) in field pepper. FL Entomol 96(1):229–231 George PE, Ambrose DP (1998) Effect of insecticides on the post-embryonic development in Rhynocoris marginatus (Fabricius) (Heteroptera: Reduviidae). J Biol Control 12(2):113–118 George PE, Ravichandran B, Jesudurai A et al (2021) Relative toxicity and haematological impact of insecticides cypermethrin and Fenvalerate to a non-target reduviid predator Rhynocoris marginatus (Fabricius) (Heteroptera: Reduviidae). J Entomol Zool Stud 9(1):2040–2043 Ghislaine BLG, Mamadou D et al (2021) Effets de trois insecticides sur la longévité, la durée de développement et le comportement de Rhynocoris albopilosus Signoret (Hétéroptères: Reduviidae). J Appl Biosci 159:16420–16428 Gholamzadeh Chitgar M, Hajizadeh J, Ghadamyari M et al (2014) Sublethal effects of diazinon, fenitrothion and chlorpyrifos on the functional response of predatory bug, Andrallus spinidens Fabricius (Hem.: Pentatomidae) in the laboratory conditions. J King Saud Univ Sci 26(2): 113–118 Golmohammadi G, Hejazi MJ (2014) Toxicity and side effects of three insecticides on adult Chrysoperla carnea (Neu.: Chrysopidae) under laboratory conditions. J Entomol Soci Iran 33 (4):23–28 Gontijo PC, Moscardini VF, Michaud JP et al (2014) Non-target effects of chlorantraniliprole and thiamethoxam on Chrysoperla carnea when employed as sunflower seed treatments. J Pest Sci 87:711–719 Gontijo LM, Celestino D, Queiroz OS et al (2015a) Impacts of azadirachtin and chlorantraniliprole on the developmental stages of pirate bug predators (Hemiptera: Anthocoridae) of the tomato pinworm Tuta absoluta (Lepidoptera: Gelechiidae). Fla Entomol 98(1):59–64 Gontijo PC, Moscardini VF, Michaud JP et al (2015b) Non-target effects of two sunflower seed treatments on Orius insidiosus (Hemiptera: Anthocoridae). Pest Manag Sci 71:515–522 Gradish AE, Scott-Dupree CD, Shipp L et al (2011) Effect of reduced risk pesticides on greenhouse vegetable arthropod biological control agents. Pest Manag Sci 67:82–86 Güven B, Göven MA (2003) Side effects of pesticides used in cotton and vineyard areas of Aegean Region on the green lacewing, Chrysoperla carnea (Steph.) (Neuroptera: Chrysopidae), in the laboratory. Pest Benef Organ IOBC-WPRS Bull 26:21–24 Halappa B, Awaknavar JS, Archana D (2013) Toxicity of newer insecticides to lady bird beetle, Cryptolaemus montruzeiri Mulsant (Coccinellidae: Coleoptera) under laboratory condition. J Entomol Res 37(2):145–148 Hattingh V, Tate B (1995) Effects of field-weathered residues of insect growth regulators on some Coccinellidae (Coleoptera) of economic importance as biocontrol agents. Bull Entomol Res 85(4):489–493 He Y, Zhao J, Zheng Y et al (2012) Lethal effect of imidacloprid on the coccinellid predator Serangium japonicum and sublethal effects on predator voracity and on functional response to the whitefly Bemisia tabaci. Ecotoxicology 21(5):1291–1300

778

20

Biosafety Assessment of Synthetic Pesticides

Herrick NJ, Cloyd RA (2017) Direct and indirect effects of pesticides on the insidious flower bug (Hemiptera: Anthocoridae) under laboratory conditions. J Econ Entomol 110(3):931–940 Hussain D, Ali A, Tariq R, Mushtaq-ul-Hassan M, Saleem M (2012) Comparative toxicity of some new chemistry insecticides on Chrysoperla carnea (Stephens) under laboratory conditions. J Agric Res 50(4):509–515 Islam ATMF, Sardar MA (1997) Toxic effects of insecticides on Bean Aphid, Aphis craccivora (Koch), and its predator Menochilus sexmaculatus (F.) (Coleoptera: Coccinellidae). Bangladesh J Entomol 7:13–19 Ismail NF, Abd Ghani DMNI (2018) Kesan dos separa mati racun serangga ke atas kadar kemandirian hidup dan kesuburan Sycanus dichotomus (Hemiptera: Reduviidae). Serangga 23(3):17–28 Izhevskii SS (1976) The physiological effect of mineral salts on the Colorado beetle, Leptinotarsa decemlineata Say. Ekologiya 4:90–92 James DG (2004) Effect of buprofezin on survival of immature stages of Harmonia axyridis, Stethorus punctum picipes (Coleoptera: Coccinellidae), Orius tristicolor (Hemiptera: Anthocoridae), and Geocoris spp. (Hemiptera: Geocoridae). J Econ Entomol 97:900–904 Jana N, Michal S, Jana O et al (2021) Forficula auricularia (Dermaptera) in orchards: monitoring seasonal activity, the effect of pesticides, and the perception of European fruit growers on its role as a predator or pest. Pest Manag Sci 77(4):1694–1704 Jiang J, Ma D, Zhang Z et al (2018) Favorable compatibility of nitenpyram with the aphid predator, Coccinella septempunctata L. (Coleoptera: Coccinellidae). Environ Sci Pollut Res 25(27): 27393–27401 Jones T, Scott-Dupree C, Harris R et al (2005) The efficacy of spinosad against the western flower thrips, Frankliniella occidentalis, and its impact on associated biological control agents on greenhouse cucumbers in southern Ontario. Pest Manag Sci 61:179–185 Karthick KS, Kandibane M, Kumar K (2014) Safety of newer insecticides to natural enemies in the coastal rice ecosystem of Karaikal, U.T. of Puducherry. J Biopest 7(2):195–198 Kiman ZB, Yeargan KV (1985) Development and reproduction of the predator Orius insidiosus (Hemiptera: Anthocoridae) reared on diets of selected plant material and arthropod prey. Ann Entomol Soc Am 78(4):464–467 Kiritani K (1979) Pest management in rice. Annu Rev Entomol 24(1):279–312 Kirknel E (1975) The effect of various insecticides in laboratory experiments with two aphid predators, the seven-spotted ladybird (Coccinella septempunctata (L.)) and larvae of hover flies (Metasyrphus corollae Fabr.). Tidsskrift Planteavl 79:393–404 Kóbor P (2020) A review on biology and agricultural significance of big-eyed bugs (Heteroptera: Lygaeoidea: Geocoridae). Hungarian Agric Res 29(2):4 Kordestani M, Mahdian K, Baniameri V et al (2022) Proteus, matrine, and pyridalyl toxicity and their sublethal effects on Orius laevigatus (Hemiptera: Anthocoridae). J Econ Entomol 115(2): 573–581 Larson JL, Redmond CT, Potter DA (2012) Comparative impact of an anthranilic diamide and other insecticidal chemistries on beneficial invertebrates and ecosystem services in turfgrass. Pest Manag Sci 68(5):740–748 Laska P (1973) Toxicity of pirimicarb and other pesticides to coccinetlids and syrphids. In: Proceedings of the 7th British Insecticide and Fungicide Conference, 2, pp 681–685 Likhil EK, Mallapur CP (2009) Studies on syrphid species complex and their natural enemies in sugarcane ecosystem in northern Karnataka. Entomon 34(1):53–56 Lin R, He D, Men X et al (2020) Sublethal and transgenerational effects of acetamiprid and imidacloprid on the predatory bug Orius sauteri (Poppius) (Hemiptera: Anthocoridae). Chemosphere 255:126778 Lira ACS, Wanderley-Teixeira V, Teixeira ÁC et al (2020) Physiological and behavioral interactions of a predator with its prey under indirect exposure to the insect growth regulator lufenuron. Crop Prot 137:105289

References

779

Liu T-X, Chen T-Y (2000) Effects of the chitin synthesis inhibitor buprofezin on survival and development of immatures of Chrysoperla rufilabris (Neuroptera: Chrysopidae). J Econ Entomol 93:234 Liu TX, Stansly PA (2004) Lethal and sublethal effects of two insect growth regulators on adult Delphastus catalinae (Coleoptera: Coccinellidae), a predator of whiteflies (Homoptera: Aleyrodidae). Biol Control 30(2):298–305 Liu Z, Zhang P, Yang J et al (2021) Effects of imidacloprid applied alone or in combination with organosilicone surfactants on biological traits and predatory feeding of Chrysoperla nipponensis (Neuroptera: Chrysopidae). J Econ Entomol 114(5):2069–2075 Lowery DT, Isman MB (1995) Toxicity of neem to natural enemies of aphids. Phytoparasitica 23: 297–306 Mahmoudi-Dehpahni B, Alizadeh M, Pourian HR (2021) Exposure route affects the toxicity class of thiamethoxam for the predatory bug, Orius albidipennis (Hemiptera: Anthocoridae) by changing its fitness. J Econ Entomol 114(2):684–693 Mandour NS (2009) Influence of spinosad on immature and adult stages of Chrysoperla carnea (Stephens) (Neuroptera: Chrysopidae). BioControl 54:93–102 Mani M, Thorntakarya TS (1988) Studies on the safety of different pesticides to the grape mealybug natural enemies, Anagyrus dactylopii (How.) and Scymnus coccivora Ayyar. Indian J Plant Protect 16:205–210 Mani M, Lakshmi VJ, Krishnamoorthy A (1997) Side effects of some pesticides on the adult longevity, progeny production and prey consumption of Cryptolaemus montrouzieri Mulsant (Coccinellidae, Coleoptera). Indian J Plant Protect 25:48–51 Mansoor MM, Shad SA (2019a) Resistance of green lacewing, Chrysoperla carnea (Stephens), to buprofezin: cross resistance patterns, preliminary mechanism and realized heritability. Biol Control 129:123–127 Mansoor MM, Shad SA (2019b) Resistance, its stability and reversion rate of resistance to imidacloprid, indoxacarb and chlorfenapyr in a field population of green lacewing Chrysoperla carnea (Stephens) (Neuroptera: Chrysopidae). Arch Phytopathol Plant Protect 52(9–10):884– 894 Mansoor MM, Shad SA (2020a) Biochemical mechanism, inheritance and cross-resistance to cyromazine in a non-target Chrysoperla carnea: a potential predator of whiteflies and aphids. Chemosphere 260:127620. PMID: 32758770 Mansoor MM, Shad SA (2020b) Genetics, cross-resistance and realized heritability of resistance to acetamiprid in generalist predator, Chrysoperla carnea (Steph.) (Neuroptera: Chrysopidae). Egypt J Biol Pest Contr 30(1):1–8 Mansoor MM, Shad SA (2020c) Inheritance of polygenic but stable pyriproxyfen resistance in a bio-control agent Chrysoperla carnea (Neuroptera: Chrysopidae): cross-resistance and realized heritability. Pest Manag Sci 76:4009. https://doi.org/10.1002/ps.5952 Mansoor MM, Raza ABM, Abbas N et al (2017) Resistance of green lacewing, Chrysoperla carnea Stephens to nitenpyram: cross-resistance patterns, mechanism, stability, and realized heritability. Pestic Biochem Physiol 135:59–63. PMID: 28043332 Martinez DG, Pienkowski RL (1982) Laboratory studies on insect predators of potato leafhopper eggs, nymphs, and adults. Environ Entomol 11(2):361–362 Martínez LC, Plata-Rueda A, da Silva Neves G et al (2018) Permethrin induces histological and cytological changes in the midgut of the predatory bug, Podisus nigrispinus. Chemosphere 212: 629–637 Martínez LC, Plata-Rueda A, Gonçalves WG et al (2019) Toxicity and cytotoxicity of the insecticide imidacloprid in the midgut of the predatory bug, Podisus nigrispinus. Ecotoxicol Environ Saf 167:69–75 Martinou AF, Stavrinides MC (2015) Effects of sublethal concentrations of insecticides on the functional response of two mirid generalist predators. PLoS One 10(12):e0144413 Mauchline AL, Osborne JL, Powell W (2004) Feeding responses of carabid beetles to dimethoatecontaminated prey. Agric For Entomol 6:99–104

780

20

Biosafety Assessment of Synthetic Pesticides

Maula AKMM, Shah MMR, Siddique NA et al (2010) Effectiveness of three insecticides against mustard aphid and predator under field condition. Bangladesh J Agric Res 35:179–187 Medina P, Budia F, Smagghe G et al (2001) Activity of spinosad, diflubenzuron and azadirachtin on eggs and pupae of Chrysoperla carnea (Stephens) under laboratory conditions. Biocontrol Sci Tech 11:597–610 Medina P, Smagghe G, Budia F et al (2002) Significance of penetration, excretion, and transovarial uptake to toxicity of three insect growth regulators in predatory lacewing adults. Arch Insect Biochem Physiol 51(2):91–101 Medina P, Smagghe G, Budia F et al (2003) Toxicity and absorption of azadirachtin, diflubenzuron, pyriproxyfen, and tebufenozide after direct spray in predatory larvae of Chrysoperla carnea. Environ Entomol 32:196–203 Mendel Z, Blumberg D, Ishaaya I (1994) Effects of some insect growth regulators on natural enemies of scale insects (Hom.: Coccoidea). Entomophaga 39(2):199–209 Mizell RF, Sconyers MC (1992) Toxicity of imidacloprid to selected arthropod predators in the laboratory. Florida Entomol 75(2):277–280 Mollah MI, Rahman M, Alam Z (2013) Effect of insecticides on lady bird beetle (Coleoptera: Coccinellidae) in country Bean Field. Middle-East J Sci Res 17:1607–1610 Morgan DJW, Hare JD (2003) Volatile cues used by the parasitoid, Aphytis melinus, for host location: California red scale revisited. Entomol Expe Appl 88(3):235–245 Moscardini VF, da Costa Gontijo P, Carvalho GA et al (2013) Toxicity and sublethal effects of seven insecticides to eggs of the flower bug Orius insidiosus (Say)(Hemiptera: Anthocoridae). Chemosphere 92(5):490–496 Nagai K (1990) Effects of a juvenile hormone mimic material, 4-phenoxyphenol (RS)-2(2-pyridyloxy) propyl ether, on Thrips palmi Karny (Thysanoptera: Thripidae) and its predator Orius sp. (Hemiptera: Anthocoridae). Appl Entomol Zool 25:199–204 Nagalingam K, Bhojan V, Boomathi N (2009) Non-target effect of ethiprole 10 SC to predators of rice planthoppers. Madras Agric J 96(1–6):208–212 Ndakidemi B, Mtei K, Ndakidemi PA (2016) Impacts of synthetic and botanical pesticides on beneficial insects. Agric Sci 7(6):364 Niranjan C, Chanda K, Saiffullah K et al (1998) Effect of copper sulfate on protein and free amino acid concentrations of aquatic beetle Hydrophilous olivaceous (Hydrophilidae: Coleoptera). Environ Ecol 16:573–578 Noelia FM, Scorsetti AC, Minardi G et al (2022) Toxicity assessment of two IGR insecticides on eggs and larvae of the ladybird Eriopis connexa (Coleoptera: Coccinellidae). Pest Manag Sci 79: 1316 Olszak RW, Ceryngier P, Warabieda W (2004) Influence of some pesticides on fecundity and longevity of Coccinella septempunctata and Adalia bipunctata (Col., Coccinellidae) under laboratory conditions. Pest Benef Organ IOBC-WPRS Bull 27:105 Omar BA, El-Kholy MI, Tohamay TH (2002) Field evaluation of certain insecticides on Pegomya mixta and related predators inhabiting sugar beet fields. Egypt J Agric Res 80:1055–1063 Pasini RA, Rakes M, De Armas FS et al (2022) Lethal effects of insecticides on adults of Eriopis connexa (Germar, 1824) (Coleoptera: Coccinellidae). Entomol Commun 4:ec04007 Pasqualini E, Civolani S (2003) Studies on side effects of some insecticides on aphid-feeding Coccinellidae in Emilia-Romagna fruit crops. Pest Benef Organ IOBC-WPRS Bull 26:51–55 Passos LC, Ricupero M, Gugliuzzo A et al (2022) Does the dose make the poison? Neurotoxic insecticides impair predator orientation and reproduction even at low concentrations. Pest Manag Sci 78(4):1698–1706 Patel LC (2018) Bioefficacy of a new novel combined insecticide, flubeniamide 24 + thiacloprid 24 SC against the caterpillars and sucking pest complex of blackgram. Indian J Plant Protect 46(1):15–26 Patel LC (2020) Laboratory contact effect of some insecticides on predatory assassin bug, Rhynocoris marginatus Fabricius (Reduviidae: Hemiptera). Int J Chem Stud 8:767–770

References

781

Patel LC, Sarkar A (2019) Efficacy of Spirotetramat 11.01 + Imidacloprid 11.01 SC against jassid, red mite and general predators in Tomato. J Crop Weed 15(3):192–200 Pathan AK, Sayyed AH, Aslam M et al (2008) Evidence of field-evolved resistance to organophosphates and pyrethroids in Chrysoperla carnea (Neuroptera: Chrysopidae). J Econ Entomol 101(5):1676–1684. PMID: 18950051 Pathan AK, Sayyed AH, Aslam M et al (2010) Resistance to pyrethroids and organophosphates increased fitness and predation potential of Chrysoperla carnae (Neuroptera: Chrysopidae). J Econ Entomol 103(3):823–834. PMID: 20568629 Penca C, Hodges AC (2017) Pyriproxyfen treatment terminates Halyomorpha halys reproductive diapause, with an indirect mortality effect on its egg parasitoid Trissolcus japonicus. J Pest Sci 90:1269–1276 Peusens G, Gobin B (2008) Side effects of pesticides on the European earwig Forficula auricularia L.(Dermaptera: Forficulidae). IOBC-WPRS Bull 39:40–43 Pietrantonio PV, Benedict J (1999) Effect of new cotton insecticide chemistries, tebufenozide, spinosad, and chlorfenapyr on Orius insidiosus and two Cotesia species. Southwest Entomol 24: 21–29 Planes L, Catalan J, Tena A (2013) Lethal and sublethal effects of spirotetramat on the mealybug destroyer, Cryptolaemus montrouzieri. J Pest Sci 86:321–327 Potin DM, Machado AVA, Barbosa PRR et al (2022) Multiple factors mediate insecticide toxicity to a key predator for cotton insect pest management. Ecotoxicology 31(3):490–502 Proctor JH, Baranyovits FL (1969) Pirimicarb: a new specific aphicide for use in integrated control programmes. In: Proceedings of the 5th Insecticide and Fungicide Conference, 2, pp 546–549 Qi B, Gordon G, Gimme W (2001) Effects of neem-fed prey on the predacious insects Harmonia conformis (Boisduval) (Coleoptera: Coccinellidae) and Mallada signatus (Schneider) (Neuroptera: Chrysopidae). Biol Control 22:185–190 Qiong Y, Linfa Q, Shu X et al (2022) Detrimental impact of λ-cyhalothrin on the biocontrol efficacy of Eocanthecona furcellata by affecting global transcriptome and predatory behavior. J Agric Food Chem 70(4):1037–1046 Rahman MA, Sarker S, Jang L et al (2022) Residual selectivity of some pesticides on the predatory bug, Orius minutus (Hemiptera: Anthochoridae). Biocontrol Sci Tech 32(4):467–483 Resende-Silva GA, Turchen LM, Guedes RNC et al (2019) Imidacloprid soil drenches affect weight and functional response of spined soldier bug (Hemiptera: Pentatomidae). J Econ Entomol 112: 558–564 Rezaei M, Talebi K, Naveh VH et al (2007) Impacts of the pesticides imidacloprid, propargite, and pymetrozine on Chrysoperla carnea (Stephens)(Neuroptera: Chrysopidae): IOBC and life table assays. BioControl 52(3):385–398 Riaza GM, Rahmana MM, Hassanb J (2022) Insecticidal exposures modulates the beneficial insect’s biodiversity in agroecosystem. Trop Agroecosyst 3(1):1–6 Rimoldi F, Fogel MN, Ronco AE et al (2017) Comparative susceptibility of two Neotropical predators, Eriopis connexa and Chrysoperla externa, to acetamiprid and pyriproxyfen: short and long-term effects after egg exposure. Environ Pollut 231:1042–1050 Rix RR, Cutler GC (2020) Low doses of a neonicotinoid stimulate reproduction in a beneficial predatory insect. J Econ Entomol 113(5):2179–2186 Rogers MA, Krischik VA, Martin LA (2007) Effects of soil application of imidacloprid on survival of adult green lacewing, Chrysoperla carnea (Neuroptera: Chrysopidae) used for biological control in greenhouses. Biol Control 42:172–177. pmid:17208560 Rugno GR, Zanardi OZ, Bajonero Cuervo J et al (2016) Impact of insect growth regulators on the predator Ceraeochrysa cincta (Schneider) (Neuroptera: Chrysopidae). Ecotoxicology 25(5): 940–949 Sabry AH, Hassan KA, Rahman AA (2014) Relative toxicity of some modern insecticides against the pink bollworm, Pectinophora gossypiella (Saunders) and their residues effects on some natural enemies. Int J Sci Environ Technol 3(2):481–491

782

20

Biosafety Assessment of Synthetic Pesticides

Sahayaraj K (1991) Bioecology, ecophysiology and ethology of chosen predatory Hemipterans and their potential in biological control (Insecta, Heteroptera, Reduviidae). PhD thesis. Madurai Kamaraj University, Madurai, India Sahayaraj K, Selvaraj P (2003) Side effects of selected biopesticides on reduviid predator Rhynocoris marginatus Fab. Entomol Croatia 7(1–2):43–50 Sahito HA, Abro GH, Syed TS et al (2011) Screening of pesticides against cotton mealybug Phenacoccus solenopsis Tinsley and its natural enemies on cotton crop. Int Res J Biochem Bioinformatics 1(9):232–236 Sánchez-Bayo F (2021) Indirect effect of pesticides on insects and other arthropods. Toxics 9(8): 177 dos Santos-Junior VC, Martínez LC, Plata-Rueda A et al (2019) Exposure to spinosad induces histopathological and cytotoxic effects on the salivary complex of the non-target predator Podisus nigrispinus. Chemosphere 225:688–695 Satoh Y, Kobori Y, Oida H, Nomura M, Tanaka H, Tezuka T (2012) Effect of agrichemicals on the polyphagous predatory bug, Geocoris varius (Uhler) (Heteroptera: Geocoridae). Japanese J App Entomol Zool 56(2):43–48 Scarpellini JR, Andrade DJD (2011) The effect of insecticides on the lady beetle Cycloneda sanguinea L. (Coleoptera, Coccinellidae) and on the aphid Aphis gossypii Glover (Hemiptera, Aphididae) on cotton plants = Efeito de inseticidas sobre a joaninha Cycloneda sanguinea L. (Coleoptera, Coccinellidae) e sobre o pulgao Aphis gossypii Glover (Hemiptera, Aphididae) em algodoeiro. Arq Inst Biol 78:393–399 Sharanappa AK, Sahu R et al (2019) Effect of certain insecticide on natural enemies of rice stem borer, Scirpophaga incertulas (walker) on rice, Oryza sativa L. J Entomol Zool Stud 7(1):1100– 1104 Sharifian I, Sabahi Q, Bandani AR (2017) Effect of some conventional insecticides on functional response parameters of Macrolophus pygmaeus (Hem.: Miridae) on Tuta absoluta (Lep.: Gelechiidae). Bihar Biol 11:10–14 Shi Y, Nachman RJ, Gui SH et al (2022a) Efficacy and biosafety assessment of neuropeptide CAPA analogues against the peach-potato aphid (Myzus persicae). Insect Sci 29(2):521–530 Shi Y, Pandit A, Nachman RJ et al (2022b) Transcriptome analysis of neuropeptides in the beneficial insect lacewing (Chrysoperla carnea) identifies kinins as a selective pesticide target: a biostable kinin analogue with activity against the peach potato aphid Myzus persicae. J Pest Sci 96:253–264 Shin-Foon C (1989) Studies on plants as a source of insect growth regulators for crop protection. J Appl Entomol 107(1–5):185–192 Shipp JL, Wang K, Ferguson G (2000) Residual toxicity of avermectin b1 and pyridaben to either commercially produced beneficial arthropod species used for control of greenhouse pests. Biol Control 17:125–131 Silva WM, Martínez LC, Plata-Rueda A et al (2020) Respiration, predatory behavior and prey consumption by Podisus nigrispinus (Heteroptera: Pentatomidae) nymphs exposed to some insecticides. Chemosphere 261:127720 Silva WM, Martínez LC, Plata-Rueda A et al (2021) Exposure to insecticides causes effects on survival, prey consumption, and histological changes in the midgut of the predatory bug, Podisus nigrispinus (Hemiptera: Pentatomidae). Environ Sci Pollut Res 28(41):57449–57458 da Silva WR, Pereira RC, Mendonça LVP et al (2022) Lethal and sublethal effects of insecticides used in the management of Plutella xylostella (Lepidoptera: Plutellidae) on the predator Cycloneda sanguinea L.(Coleoptera: Coccinellidae). Pest Manag Sci 78:4397. https://doi.org/ 10.1002/ps.7060 Singh JP, Jaiswal AK, Monobrullah M et al (2010) Effect of selected pesticides on larval mortality of the neuropteran predator, Chrysopa lacciperda Kimmins of the lac insect, Kerria lacca (Kerr). J Asia Pac Entomol 13:69–72 Sotherton NW, Moreby SJ (1984) Contact toxicity of some foliar fungicide sprays to three species of polyphagous predators found in cereal fields. Ann Appl Biol 104:16–17

References

783

Srivastava M, Funderburk J, Olson S et al (2014) Impacts on natural enemies and competitor thrips of insecticides against the western flower thrips (Thysanoptera: Thripidae) in fruiting vegetables. FL Entomol 97(2):337–348 Stark JD, Jepson PC, Mayer DF (1995) Limitations to use of topical toxicity data for predictions of pesticide side effects in the field. J Econ Entomol 88:1081–1088 Staubli A, Hachler M, Antonin P et al (1984) Tests de nocivite de divers pesticides envers les enemuies natural des principaux ravageurs de poiriers en suissae romande. Revue Suissae Viticulture, Arboriculture, Horticulture 16(5):279–286 Studebaker GE, Kring TJ (2003a) Effects of various residues in cotton on gender and development stage of the insidious flower bug (Hemiptera: Anthocoridae). J Entomol Sci 38:409–419 Studebaker GE, Kring TJ (2003b) Effects of insecticides on Orius insidiosus (Hemiptera: Anthocoridae), measured by field, greenhouse, and petri dish bioassays. Fla Entomol 86:178– 185 Su Y, Ren X, Ma X et al (2022) Evaluation of the toxicity and sublethal effects of acetamiprid and dinotefuran on the predator Chrysopa pallens (Rambur) (Neuroptera: Chrysopidae). Toxics 10(6):309 Suárez-López YA, Hatem AE, Aldebis HK et al (2020) Lethal and sublethal effects of lufenuron on the predator Chrysoperla carnea (Stephens) (Neuroptera: Chrysopidae). Crop Prot 135:105217 Syrett P, Penman DR (1980) Studies of insecticide toxicity to lucerne aphids and their predators. N Z J Agric Res 23(4):575–580 Tang F, Li SC, Kong XB et al (2007) Studies on selective toxicity of acetamiprid, imidacloprid and buprofezin to Trialeurodes vaporariorum and two natural enemies. Chin J Pest Sci 9:88–91 Teodoro AV, Pallini A, Sub OC (2009) Sub-Lethal effects of Fenbutatin oxide on prey location by the predatory mite Iphiseiodes zuluagai (Acari: Phytoseiidae). Exp Appl Acarol 47:293–299 Tillman PG, Mullinix BG Jr (2004) Comparison of susceptibility of pest Euschistus servus and predator Podisus maculiventris (Heteroptera: Pentatomidae) to selected insecticides. J Econ Entomol 97(3):800–806 Ullah M, Ullah F, Khan MA et al (2022) Efficacy of various natural plant extracts and the synthetic insecticide cypermethrin 25EC against Leucinodes orbonalis and their impact on natural enemies in brinjal crop. Int J Trop Insect Sci 42(1):173–182 van de Veire M, Tirry L (2003) Side effects of pesticides on four species of beneficials used in IPM in glasshouse vegetable crops: “worst case” laboratory tests. IOBC/WPRS Bull 26:41–50 Wakeil NE, Nawal G, Ahmed S et al (2013) Side effects of insecticides on natural enemies and possibility of their integration in plant protection strategies. In: Insecticides – development of safer and more effective technologies. InTech, Rijeka, pp 3–56. https://doi.org/10.5772/54199 Walker MK, Stufkens MAW, Wallace AR (2007) Indirect non-target effects of insecticides on Tasmanian brown lacewing (Micromus tasmaniae) from feeding on lettuce aphid (Nasonovia ribisnigri). Biol Control 43:31–40 Wright DJ, Verkerk RHJ (1995) Integration of chemical and biological control systems for arthropods: evaluation in a multitrophic context. Pestic Sci 44:207–218 Yadav D, Khinchi SK, Sharma P et al (2020) Bio-efficacy of insecticides against Thrips tabaci on Onion and adverse effect on natural enemy. Ann Plant Protect Sci 28(1):12–14 Zhang J, Drummond FA, Liebman M et al (1997) Phenology and dispersal of Harpalus rufipes DeGeer (Coleoptera: Carabidae) in agroecosystems in Maine. J Agric Entomol 14(2):171–186

Biocompatibility of Biopesticides with Predatory Insects

21

Contents 21.1 21.2

Role of Biopesticides in Insect Pest Management . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Microbial Pesticides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.1 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.2 Heteroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.3 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.4 Other Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.5 Microbial Metabolites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2.6 Genetically Modified (GM)/Recombinant Microorganisms . . . . . . . . . . . . . . . . . . . 21.3 Botanical Biopesticides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.1 Coleoptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.2 Neuroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.3 Heteroptera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.4 Different Predators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.5 Combined Effects of Different Botanicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3.6 Botanical Products, Including Oil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.4 Field Trials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.4.1 Botanicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.4.2 Botanical Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.4.3 Post-harvest Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.5 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

21.1

785 786 786 789 792 793 793 798 802 806 807 809 810 811 811 812 812 813 814 815 815

Role of Biopesticides in Insect Pest Management

Major living things are vulnerable to competition from other creatures, parasitism, and predation. In order to protect agricultural crops against insect pests, fungal, bacterial, and viral diseases, weeds, nematodes, and mollusc pests, the study of these interactions has shown numerous potential options for the use of living creatures as biopesticides. Worldwide, there is growing concern over the unfavourable effects of synthetic chemical pesticides, including their influence on the environment, toxicity

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7_21

785

786

21 Biocompatibility of Biopesticides with Predatory Insects

to non-target animals including people, and the emergence of insect populations that are resistant to them. This has made it necessary to produce newer, environmentally benign pesticides, such as biopesticides, which have a broad spectrum of pesticidal activities but less environmental persistence and mammalian and avian toxicity. Today, a variety of biopesticide treatments are commercially available for the control of insect pests, fungal and bacterial diseases, and weeds. These products include as active agents, bacteria, fungi, nematodes, protozoa, viruses, and beneficial insects. For bio-intensive integrated pest management (BIPM), researchers must put forth a lot of effort to discover new tools that might be implemented in BIPM programmes. Despite this, there are many ways to manage plant nematodes, but the most common and widely utilised approach still involves chemical pesticides. One emerging trend in BIPM programmes, namely, the use of biorational agents, still has to be developed. The use of plant extracts, soil amendments, and metabolites of microbes and plants, either alone or in combination with other BIPM components like predators, have all been suggested as methods of controlling pests, including entomopathogenic fungi, bacteria, and nematodes. By releasing predators along with either microbicides or botanicals that could be utilised in BIPM programmes, this experiment hoped to shed light on how biocompatibility has been maintained while being safe for the environment, domestic and wild animals, and humans.

21.2

Microbial Pesticides

Bacillus thuringiensis (Bt), Lysinibacillus sphaericus (formerly Bacillus sphaericus), Brevibacillus laterosporus, a wide variety of fungi (Beauveria bassiana, Metarhizium anisopliae, Verticillium lecanii, Lecanicillium spp., Hirsutella spp., Paecilomyces spp.), viruses (baculovirus), protozoa (e.g. Nosema), and some beneficial nematodes (Heterorhabditis and Steinernema (Rhabditida), Photorhabdus, and Xenorhabdus) are considered the best biocontrol agents for pest management (Table 21.1). These organisms, their by-products, and their metabolites have been utilized for pest management. In this chapter, the biocompatibility of biopesticides with predatory insects is highlighted.

21.2.1 Coleoptera According to Mahmoud (2010), Spodoptera littoralis larvae infected with Bt were attacked and eaten by Coccinella undecimpunctata (fourth larval instar and adults of the coccinellid predator, Coleoptera) (Lepidoptera: Noctuidae). In order to ascertain the effects of M-One™ (Bacillus thuringiensis var. san diego) on the larval stages of Coleomegilla maculata lengi, Giroux et al. undertook a laboratory experiment in 1994. On pollen treated with suspensions of M-One™ at 20 mL L-1 (5.6108 CPBIU L-1) and at 200 mL L-1, coccinellid larval development (from egg hatching to adulthood) was completed in 29.3 days and 38.5 days, respectively, as opposed to 21.9 days for the control (water). There was no larval mortality due to M-One™.

21.2

Microbial Pesticides

787

Table 21.1 Biopesticidal organisms, their commercial names, and target pests Active substances Bacteria Bacillus thuringiensis aizawai Bacillus thuringiensis kurstaki

Bacillus thuringiensis israelensis Bacillus thuringiensis tenebrionis Bacillus thuringiensis sphaericus Bacillus firmus Burkholderia spp. Saccharopolyspora Chromobacterium subtsugae Fungi Beauveria bassiana

Beauveria brongniartii Hirsutella thompsonii Isaria fumosorosea Metarhizium anisopliae Metarhizium brunneum Paecilomyces lilacinus Paecilomyces fumosoroseus Verticillium lecanii Lecanicillium lecanii Myrothecium verrucaria

Commercial names

Main targets

Able WG, Agree WP, Florbac, XenTari Biobit, Cordalene, Costar WG, Crymax WDG, Deliver, DiPel, Foray, Javelin WG, Lepinox Plus, Lipel, Rapax Teknar, VectoBac, Vectobar

Armyworms, diamondback moths Lepidoptera

Novodor, Trident

Colorado potato beetles

VectoLex, VectoMax

Mosquitoes

BioNemagon Majestene, Venerate

Nematodes Chewing and sucking insects and mites; nematodes Conservation of insects

Spinosa Tracer™ 120, Spintor 480 SC Grandevo

Bio-Power, Biorin/Kargar, BotaniGard, Daman, Naturalis, Nagastra, Beauvitech WP, Bb-Protec, Racer, Mycotrol Bas-Eco – – Biomet/Ankush, Bio-Magic, Devastra, Kalichakra, Novacrid, Met52/BIO1020 granular Attracap BioNematon, MeloCon, Mytech WP, Paecilo Bioact WG, NOFLY WP BioCatch, Mealikil, Bioline/VertiStar Lecatech WP, Varunastra DiTera

Mosquitoes and black flies

Chewing and sucking insects and mites Wide range of insects and mites

Helicoverpa armigera, berry borer, root grubs Spider mites Whiteflies Pacer beetles and caterpillar pests; grasshoppers and termites Agriotes spp. Plant pathogenic nematodes Paecilomite insects, mites, nematodes, thrips Mealybugs and sucking insects Aphids, leaf miners, mealybugs, scale insects, thrips, whiteflies Nematodes (continued)

788

21 Biocompatibility of Biopesticides with Predatory Insects

Table 21.1 (continued) Active substances Virus Helicoverpa zea nucleopolyhedrovirus Helicoverpa armigera nucleopolyhedrovirus (HearNPV) Helicoverpa zea nuclear polyhedrosis virus Spodoptera litura nucleopolyhedrovirus (SLNPV) Spodoptera littoralis nucleopolyhedrovirus (SpliNPV) Spodoptera exiguan nucleopolyhedrovirus (SeNPV) Adoxophyes Orana granulovirus (AoGV) Cryptophlebia leucotreta granulovirus Plutella xylostella granulovirus Lymantria dispar multiple nucleopolyhedrovirus (LdMNPV) Cydia pomonella granulovirus (CpGV) Neodiprion abietis nucleopolyhedrovirus (NeabNPV) Cydia pomonella L., granulovirus, CpGV

Commercial names

Main targets

Heligen

Gemstar

Helicoverpa spp. and Helicoverpa virescens African cotton bollworm (Helicoverpa armigera), corn earworm (H. zea), H. virescens, Helicoverpa punctigera Heliothis and Helicoverpa spp.

Biovirus-S, Somstar-SL

Spodoptera litura

Littovir

African cotton leaf worm (Spodoptera littoralis)

Spexit, Spod-X

Spodoptera exigua

Capex

Summer fruit tortrix moth (Adoxophyes orana) False codling moth (Thaumatotibia leucotreta)

Biovirus-H, Helicovex, Helitec, Somstar Ha

Cryptex

Plutellavex

Plutella xylostella

Gypchek

Lymantria dispar

CYD-X, Madex, Carpovirusine

Cydia pomonella

Neodiprion abietis NPV

Neodiprion abietis

Cyd-X®

Cacopsylla pyricola

Third instar C. maculata did not distinguish between the eggs of Leptinotarsa decemlineata treated with M-One™ at 20 mL L-1 and those treated with water. However, 48 h after the test started, at a concentration of 200 mL M-One™ L-1, the number of eggs attacked was 34.7% fewer than the eggs treated with water alone. These findings suggest that the application of M-One™ at the manufacturer’s suggested field rate of 20 mL L-1 poses little danger to populations of larval C. maculata (Giroux et al. 1994). Two species of Coccidioides, specifically the pink-spotted lady beetle and the 12-spotted lady beetle, are affected by specific Beauveria bassiana spores. Smith

21.2

Microbial Pesticides

789

and Krischik (2000) identified Coleomegilla maculata, mealybug ladybird or mealybug destroyer Cryptolaemus montrouzieri (Coccinellidae) (Ipheseius degenerans (Acari: Phytoseiidae), insidious flower bug Orius insidiosus (Anthocoridae), and Phytoseiulus per (Ludwig and Oetting 2001)). Similar findings were made with B. thuringiensis kurstaki’s effects on Geocoris punctipes and Nabis capsiformis (Boyd and Boethel 1998). However, studies of Bajwa and Aliniazee (2001) (B. thuringiensis kurstaki effects on Eris marginata and Cheiracanthium inclusum), Smith and Krischik (2000) (B. bassiana GHA effects on Harmonia axyridis), and Ludwig and Oetting (2001) (Metarhizium anisopliae and Verticillium lecanii effects on Ipheseius degenerans) were in contrast to the previous views. Bharani et al. (2015) conducted a field study to test the efficacy of Beauveria bassiana (2.5 kg ha1 ) and Verticillium lecanii (2.5 kg ha-1) combined with two novel insecticides (imidacloprid 30.5 Systemic and Contact (SC) (100 mL ha-1) and thiamethoxam 25 Wettable Granules (WG) (75 g active ingredients (a.i.) ha-1)) in field tomatoes. The insecticides imidacloprid 30.5 SC and thiamethoxam 25 WG (2.64 and 2.79 coccinellid plants, respectively) showed lower toxicity compared to control treatment (3.53 coccinellids plant-1) for coccinellids, and Beauveria bassiana and Verticillium lecanii (3.26 and 3.12 coccinellids plant-5, respectively) were found to be less toxic (Bharani et al. 2015).

21.2.2 Heteroptera 21.2.2.1 Reduviids Sycanus leucomesus, a hemipteran predator, was subjected to a laboratory assessment by Sajap et al. (1999) to determine the effects of nuclear polyhedrosis virus (NPV). They discovered that the NPV treatment decreased nymphal survival rate, adult longevity, and fertility. Under controlled laboratory settings, the biosafety of a significant soil-dwelling entomopathogenic fungus, Metarhizium anisopliae, on a widespread soil-dwelling reduviid predator, Acanthaspis pedestris, occurring in the same ecosystem, was investigated (Sahayaraj et al. 2008). In the experiment, 1.8 × 107, 2.1 × 106, 3.7 × 105, and 2.9 × 104 Metarhizium anisopliae spores mL-1 and cypermethrin (0.05% and 0.10%) were used to treat eggs. Acanthaspis pedestris eggs’ capacity to hatch was significantly decreased by cypermethrin. Both the incubation period and the growth of Acanthaspis pedestris were unaffected by Metarhizium anisopliae. Compared to the control category, the survival rate and the particle-carrying capacity were lower (Sahayaraj et al. 2008). In a laboratory setting, additional entomopathogenic microorganisms (EPMs) were tested against the third, fourth, and fifth nymphal stages of the reduviid predator Rhynocoris kumarii (Hemiptera: Reduviidae). These EPMs included Spodoptera litura nucleopolyhedrovirus (SpliNPV), Metarhizium anisopliae, and Pseudomonas fluorescens (Sahayaraj et al. 2008). The amount of time spent handling the prey, the number of predators per day, the developmental stage, and the survival rate were all noted. When Rhynocoris kumarii larvae were fed on Spodoptera litura larvae infected with entomopathogens, their predation rate was equal to or greater (2.9

790

21 Biocompatibility of Biopesticides with Predatory Insects

preys) than that of the untreated control (2.0 preys). After consuming Spodoptera litura larvae infected with Pseudomonas fluorescens, the juvenile predator developed substantially more slowly (2–4 days) than did those that consumed larvae infected with other microbial control agents. However, neither the adult sex ratio nor the predator nymphal survival rate was impacted by feeding exclusively on Pseudomonas fluorescens. The results of this study imply that integrating reduviids with the tested entomopathogens is a suitable and potentially effective technique for the control of Spodoptera litura populations, even though three entomopathogens had some impact on the biological characteristics of Rhynocoris kumarii (Sahayaraj et al. 2008). The effects of two neem-based insecticides, Nimbecidine and Vijay Neem, as well as the viral pathogen Spodoptera-Nuclear polyhedrosis virus (NPV), on the incubation period, egg hatching, nymphal development, and survival and predatory potential of the reduviid predator Rhynocoris marginatus were examined in another laboratory experiment. The incubation duration had no discernible effect, whereas neem-based biopesticides dramatically slowed down egg hatching, extended nymphal development, and lowered nymphal survival rates. Further research showed that topical toxicity affected this reduviid more adversely than did contact toxicity. S-NPV biopesticides were shown to have the greatest effect on Rhynocoris marginatus. Field-level analysis could provide more information about how these biopesticides affect this reduviid (Sahayaraj et al. 2008). For the first time, an entomopathogenic fungus, Aspergillus flavus, was discovered in 2012 in the natural agro-ecosystems of Tirunelveli district, Tamil Nadu, as a pathogen of the reduviid predator, Rhynocoris marginatus (Hemiptera: Reduviidae) (Sahayaraj et al. 2012). In recent studies, the effects of entomopathogenic microorganisms on Rhynocoris kumarii (Hemiptera: Reduviidae) have been examined in the laboratory. These microbes included Spodoptera litura nucleopolyhedrovirus (SpliNPV), Metarhizium anisopliae, and Pseudomonas fluorescens. The third, fourth, and fifth nymphal stages of Rhynocoris kumarii were given reduviid, which was then given Spodoptera litura nucleopolyhedrovirus (SpliNPV), Metarhizium anisopliae, and Pseudomonas fluorescens-infected Spodoptera litura larvae. The prey handling time and predation rate (number/day/predator) increased. The predation rate of Rhynocoris kumarii life stages was equal to or greater than that of the untreated control when they were given entomopathogen-infected S. litura larvae. The developmental period of the juvenile predator was much longer (2–4 days) after feeding on Pseudomonas fluorescens-infected S. litura larvae than it was for juvenile predators fed on larvae infected with other microbial control agents. However, neither the adult sex ratio nor the predator nymphal survival rate was impacted by feeding exclusively on Pseudomonas fluorescens. The results of this study imply that integrating reduviids with the tested entomopathogens is a compatible and potentially effective strategy for the management of Spodoptera litura populations, despite the fact that three entomopathogens had some impact on the biological parameters of Rhynocoris kumarii (Sahayaraj et al. 2018). This integrated approach needs to be evaluated in the field, despite its potential, in order to support the laboratory results.

21.2

Microbial Pesticides

791

In a different investigation, the effects of an entomopathogenic fungus (EPF) on Rhynocoris marginatus (Hemiptera: Reduviidae) were examined in a laboratory setting (Ullah et al. 2019). The handling time, predation rate, consumption rate, and survival rate of EPF-infected prey S. litura on the adult predator were noted. There was no discernible impact of the EPF species on the predation rate. Furthermore, there was no discernible difference in the survival rate of predators that consumed either healthy larvae or those that had the EPF. These EPFs interacted with a reduviid predator, and the results suggested that both species of EPF, particularly I. fumosorosea, might be utilised in conjunction with the predator to increase the biological control of S. litura in commercial crops (Ullah et al. 2019). The studies’ findings showed that the deterrent activity of microbes decreased the preys’ feeding activity, which, in turn, decreased the efficiency of the consumption of ingested and digested food, decreased weight gain, decreased the production of faecal pellets, and ultimately slowed down the natural enemies’ relative growth rate and precipitated their rapid demise. Additionally, EPF infection may change the chemical cues or nutritional make-up of the prey, which could impact the effectiveness of biological control and the reduviid predator’s performance.

21.2.2.2 Pentatomidae When combined, Anticarsia gemmatalis infected with NPV and Podisus maculiventris, a predatory stink bug, had a favourable effect on each other, according to Abbas and Boucias (1984). De Nardo et al. (2001) carried out the following two tests: In the first experiment, Anticarsia gemmatalis raised on an artificial diet was exclusively given to Podisus nigrispinus. It was then treated with one formulation of Ag-Nuclear polyhedrosis virus (AgNPV) (for infected prey) or with water (for healthy prey) as the control. The artificial diet was given an extra treatment in the second experiment, which involved an inactive (autoclaved) AgNPV formulation. In the first and second experiments, the predators consumed these exclusive prey for their whole lifespans for three or four successive generations. To gather information for the creation of fertility life tables, daily oviposition patterns, survival distribution curves, daily oviposition, and mortality observations were made. In addition, the predators were fed prey that had consumed an infected or inactivated AgNPV diet as opposed to healthy prey. Because the AgNPV has been shown to be extremely specific to Lepidoptera and does not replicate or have any negative effects on P. nigrispinus, it is possible that the negative effects reported are not caused by any inactive components contained in the commercial formulation. Under laboratory conditions, the fertility life tables have been demonstrated to be an effective tool for assessing the possible effects of biopesticides on non-target populations (De Nardo et al. 2001). In their study of the impact of Bt-infected Plutella xylostella larvae as food on the nymphs and adults of the pentatomid predator Podisus nigrispinus (Hemiptera), Carvalho et al. (2012) discovered that Bt had no biological differences between the Bt-infected and control groups. Similar to this, Magalhães et al. (2015) looked into how Bt var. kurstaki and the commercial product “Agree” (a mix of Bt var. kurstaki and Bt var. aizawai) affected the biological characteristics of Podisus nigrispinus

792

21 Biocompatibility of Biopesticides with Predatory Insects

when it consumed infected Plutella xylostella larvae. Throughout the nymphal development, they discovered that the infected larvae were devoured more than the untreated ones, with no impact on nymphal survival or population growth characteristics. Carvalho et al. (2018) noted that rearing P. nigrispinus nymphs on Tenebrio molitor (Coleoptera: Tenebrionidae) larvae treated with Bt had an impact on the predatory female’s oviposition times and overall egg production. Third instar nymphs of Podisus nigrispinus (Heteroptera: Pentatomidae), a non-target natural enemy, were exposed to Bt in a laboratory setting. Bt was poisonous to Podisus nigrispinus nymphs, and it significantly decreased their respiratory rate, survival rate (56.6%), and predatory potential (median lethal concentration (LC50) = 1.10 (0.83–1.46) mg mL-1) (Silva et al. 2020). Jacobsen et al. (2019) investigated the relationship between Orius majusculus and the entomopathogenic fungi Metarhizium brunneum and Neozygites floridana (bioagents of Tetranychus urticae infesting strawberry). Orius majusculus was found to spend less time looking at leaf discs containing Metarhizium brunneum spores than it did on discs without any fungal spores and more time looking at discs with Neozygites floridana spores. According to Ladurner et al. (2012) and Gao et al. (2012), Orius laevigatus and Beauveria bassiana (strain ATCC 74040) are compatible. Orius sauteri and Beauveria bassiana (Bals-Criv) are likewise compatible (strain Bb-RSB). Orius albidipennis was found to be able to recognise and stay away from areas that had been treated with the entomofungal bioagent Metarhizium anisopliae in a study by Pourian et al. (2011). In greenhouse-grown sweet pepper, a combination of Orius laevigatus and Macrolophus pygmaeus releases seemed to be the best method for thrips and aphid control (Messelink and Janssen 2014). The use of Orius majusculus in combination with predatory midges, predatory thrips, and parasitoids significantly improved the suppression of aphids and thrips infesting sweet pepper, supporting the idea that intraguild predation, which may be detrimental to biocontrol, may be offset by the beneficial effects of generalist predators for control of a variety of pests (Messelink et al. 2011, 2013). Combining Orius laevigatus and the predatory mite Amblyseius swirskii releases in Tunisia helped control the Frankliniella occidentalis infestation of greenhouse pepper (Elimem and Chermiti 2012). In France and Spain, the combined use of Neoseiulus spp. and Orius sp. effectively reduced F. occidentalis in protected strawberry crops (Sampson et al. 2011). When released simultaneously on four ornamental plants (Saintpaulia, Impatiens, Gerbera, and Brachyscome multifida) infested with F. occidentalis in greenhouses, the predators Amblyseius (=Neoseiulus) cucumeris and O. insidiosus were quite efficient (Sorensson and Nedstam 1993).

21.2.3 Neuroptera With 82 genera, the neotropical green lacewing fauna exhibits tremendous variety. Brazil is a frequent location for Chrysoperla externa (Neuroptera: Chrysopidae). First, second, and third instars of Chrysoperla externa were tested by three

21.2

Microbial Pesticides

793

entomopathogenic fungi at 24, 48, 72, 96, and 120 h: Metarhizium rileyi (strain UFMS 03), Beauveria bassiana (strain ESALQ PL63), and Metarhizium anisopliae (strain ESALQ E9). The results show that after 120 h following administration, Chrysoperla externa larvae showed considerably greater susceptibility to the (P) (HAA [hours after application]). First instar survival at ages between 24 and 120 HAA was unaffected by application techniques (A), concentrations (B), or interactions. The selectivity of these isolates in green lacewing larvae is indicated by the low mortality rate of Chrysoperla externa larvae treated with Metarhizium anisopliae (ESALQ E9) and Metarhizium rileyi (UFMS 03). At all Chrysoperla externa larval stages, fungal concentrations or application techniques had little effect (Mingotti Dias et al. 2020). The authors came to the conclusion that even while the B. bassiana strain caused more death in Chrysoperla externa larvae than in Metarhizium anisopliae and Metarhizium rileyi, these three entomopathogenic fungi might still be utilised in conjunction with Chrysoperla externa for long-term pest control.

21.2.4 Other Predators Field research has shown that before illness directly kills grasshoppers and locusts, the fungal entomopathogen Metarhizium anisopliae var. acridum may produce a significant increase in the host’s sensitivity to predation. Prior to that, laboratory tests were conducted to explore some probable processes underlying this phenomenon by observing how the behaviour of the desert locust Schistocerca gregaria Forsk. changed after infection by Metarhizium anisopliae var. acridum. In the first experiment, which tracked general locust activity in little cages for the course of the disease’s incubation period, infected locusts were seen to move more often and for longer periods of time until they died (average bodily movement from 3 days). Infection-related mating behaviour and a decreased feeding survival time of 11 hours were partially demonstrated. In a follow-up experiment, locusts were individually subjected to a predator attack simulation, and the commencement and potency of any evasive responses were assessed. Reduced escape ability was reported in infected locusts (both the propensity to escape and the strength of the response). This was only noticeable at the last stages of infection, just before death, in contrast to the rather early alterations in the overall activity reported in the first experiment. Early on in the infection phase, both an increase in mobility and overall apparency and a decrease in escape capabilities suggest methods by which the susceptibility of locusts and grasshoppers to predation might be exacerbated after infection with Metarhizium anisopliae var. acridum (Arthurs and Thomas 2001).

21.2.5 Microbial Metabolites 21.2.5.1 Spinosad Spinosad is a naturally occurring substance produced by the fermentation of Saccharopolyspora spinosa that is made up of two identical molecules called

794

21 Biocompatibility of Biopesticides with Predatory Insects

spinosyn A and spinosyn D in about a 5:1 ratio. Spinosad is the most effective against species of the two orders Lepidoptera and Diptera. It is largely a stomach poison with some contact activity. It is a neurotoxin with a new method of action that appears to involve both the gamma-aminobutyric acid (GABA) receptors and the nicotinic acetylcholine receptor (Salgado 1998). Although several microbial metabolites have been used for pest control, spinosad is the most well-known and often used metabolite. As an illustration, consider the tetracycline–macrolides complex spinosad, which was created by the actinomycete Saccharopolyspora spinosa and was recovered from a Jamaican soil sample (Sparks et al. 1998). According to conventional toxicity studies, spinosad is essentially non-toxic to mammals and birds. Chrysoperla carnea (Chrysopidae) and Forficulidae doru taeniatum are both affected by spinosad (Cisneros et al. 2002; Medina et al. 2001, 2003). Additional research feeding coccinellid and chrysopid larvae treated with spinosad aphids found no predator mortality (Schoonover and Larson 1994). Following the cessation of eating after exposure, paralysis and death occur about 24 h later. However, other research indicates that the spinosad insecticide affects Geocoris punctipes (Lygaeidae) (Boyd and Boethel 1998; Tillman and Mulrooney 2000; Elzen et al. 1998), Aleochara bilineata (Staphylinidae) (Cisneros et al. 2002), and Orius (Boyd and Boethel 1998). Insect natural enemies such as the Orius species, Chrysops species, and coccinellids have also been observed to be virtually unaffected by spinosad (Bret et al. 1997). For numerous predators, like the psylla predator Deraeocoris brevis, spinosad was safe (according to Arthurs et al. 2007). The authors further claimed that Cyd-X®, which contained the granulovirus CpGV and Cydia pomonella, is less safe than is the spinosad (Entrust®) medication. The US Environmental Protection Agency considers spinosad (marketed under the brand “Naturalyte”) to be another biological insecticide that poses a low risk to the environment and is less toxic (Saunders and Bret 1997). Given its biological effects on both target and non-target arthropod species, it is regarded as a biological pesticide (Croft 1990). Spinosad exhibits a significant increase in the level of insect control when employed as baits, phagostimulants, or feeding stimulators. According to Schoonover and Larson’s (1994) findings, there was no observed predator mortality when coccinellid adults and chrysopid larvae were fed aphids that had previously been treated with 200 ppm (parts per million) spinosad (Bret et al. 1997). Furthermore, the significance of relatively low-concentration toxicity studies may be questioned, given that the product label maximum application rates surpass 4000 ppm. In the field test, earwigs were kept on experimental maize plants, but, under normal circumstances, they might have abandoned spinosad-treated plants. However, we found no indication of touch repellency, and earwigs and staphylinid parasitoids readily ate the spinosad-containing granules. Contrarily, the mineralbased granular formulation of chlorpyrifos produced only moderate mortality and was likely repulsive or unpleasant to earwigs. The granular formulation was not consumed by Chrysoperla carnea (Neuroptera: Chrysopidae) larvae, and they experienced minimal overall mortality. The earwig Doru taeniatum (Dermaptera: Forficulidae) suffered mortality after 14 days of exposure of 48% in the 1.2 ppm

21.2

Microbial Pesticides

795

spinosad treatment and of 98% in the 1200 ppm spinosad treatment, compared to 20% in controls. Seventy-two hours after consuming spinosad-contaminated Spodoptera frugiperda (Lepidoptera: Noctuidae) larvae, earwigs experienced an 86% mortality/intoxication rate. On earwigs housed in gauze bags, a field experiment was conducted to examine the effects of commercial granular chlorpyrifos and spinosad applied in maize flour granules (200 ppm a.i. and 2000 ppm a.i.; 4.8–48 g a. i. ha-1, respectively) or as an aqueous spray (160 ppm a.i.; 48 g a.i. ha-1, respectively). In contrast to 33% on plants treated with granular chlorpyrifos, 83% on plants sprayed with 160 ppm spinosad, and 91–95% on plants treated with 200–2000 ppm spinosad granules, the mortality of earwigs on control plants was less than 15% at 2 days after application. Additional mortality over the 24-h postsampling period varied between 5% in control treatments, 9% in the chlorpyrifos treatment, and 55–65% in the spinosad spray and granular treatments. We come to the conclusion that spinosad does not share the same environmental safety profile as the majority of well-known biological insecticides (Cisneros et al. 2002). Spinosad is not necessarily safe and is harmless only because it is made from a naturally occurring soil bacterium. 21.2.5.1.1 Cry Proteins Hilbeck et al. (1998) supplied Chrysoperla carnea larvae an artificial meal laced with the toxin Cry1Ab and discovered that there was a much greater total death rate in the larvae (57%) than in the untreated control (30%). Additionally, when given Cry1Ab later in their development than the control group (17%), considerably more larvae died (29%) than the control group (17%). Between treated and untreated larvae, no variations in developmental period were found. When Chrysoperla carnea larvae were fed on Spodoptera littoralis larvae fed on feed mixed with Cry1Ab and Cry2A toxins at varying concentrations, Hilbeck et al. (1999) obtained nearly identical findings. Three commercial formulations of Bacillus thuringiensis var. kurstaki for rangeeni and the kusmi lac culture were evaluated in the field for their ability to significantly reduce the incidence of Eublemma amabilis and Pseudohypatopa pulverea attacking Kerria lacca. During the rangeeni rainy and summer crops, there were decreases in the occurrence of Eublemma amabilis ranging from 42.47% to 96.24% and 65.00–95.00%, respectively, and in that of Pseudohypatopa pulverea ranging from 52.50% to 97.50% and 66.67–100.00%, respectively. On the kusmi winter lac crop, the incidence of Eublemma amabilis and Pseudohypatopa pulverea decreased by 61.02–93.55% and 12.86–98.57%, respectively, whereas on the kusmi summer crop, it decreased by 56.16–95.52% and 66.17–98.50%, respectively. The productivity of the lac crop in kusmi improved from 78.95% to 168.42%, 11.40–36.84%, and 33.33–166.67% in rangeeni rainy lac crop, respectively. From treated crops, higher fecundity and superior-quality brood milk were achieved. Spinosad, indoxacarb, fipronil, ethofenprox, and other insecticides, as well as biopesticides, can be used to effectively inhibit predators attacking lac production (Singh et al. 2014).

796

21 Biocompatibility of Biopesticides with Predatory Insects

21.2.5.2 Abamectin The macrocyclic lactones (avermectins and milbemycins), which are created by a natural fermentation of the bacteria Streptomyces avermitilis, include avermectins, one of the suggested alternative biological tools. Avermectin B1a makes up more than 80% of the avermectin mixture, whereas avermectin B1b makes up less than 20%. On vegetables, fruits, and field crops, it is applied as a substitute for insecticides, acaricides, and nematicides. It has four major component pairs (A1a, A2a, B1a, and B2a) and four minor component pairs (A1b, A2b, B1b, and B2b). Ivermectin, abamectin, doramectin, eprinomectin, and selamectin are members of the family of 16-membered macrocyclic lactones known as avermectins. Another illustration is spinosad (Dow AgroSciences), a blend of the tetracycline–macrolides substances known as spinosyns A and D generated by the actinomycete Saccharopolyspora spinosa Mertz and Yao, which was isolated from a soil sample in Jamaica. 21.2.5.2.1 Hemiptera Using a laboratory and an extended laboratory approach, Van De Veire et al. (2002) evaluated the residual effect of abamectin and spinosad on the predatory insect Orius laevigatus (Hemiptera: Anthocoridae) over the spring and summer of 2000. According to the reported median lethal dose (LD50) values, abamectin had less of an effect on the second and fifth nymphal instars as well as on the adult stages than did spinosad. Additionally, the toxicity of abamectin and spinosad was markedly higher in the second instar stages and adults than in the fifth nymphal stages. A further finding was that the effects of abamectin were significantly more enduring in the spring than in the summer. Spray deposits were hazardous for 1 month in the spring and for 2 weeks in the summer when applied at the appropriate dose of 10 ppm a.i. for controlling leaf miners. After 5 residual days, spinosad summer spray deposits (recommended rate of 250 ppm a.i. for controlling leaf miners and caterpillars) were not harmful (Van De Veire et al. 2002). 21.2.5.2.2 Coleoptera The high mortality rates of adult Aleochara bilineata were observed in Gyllenhal (Coleoptera: Staphylinidae) after consuming 1000 or 2000 ppm spinosad active ingredients (a.i.), whereas 200 ppm was the least mortality level (Cisneros et al. 2002). 21.2.5.2.3 Larvae of the Neuroptera Chrysoperla carnea (Neuroptera: Chrysopidae) did not consume the granular formulation of spinosad-active ingredient-based product, and its overall mortality was not excessive (Cisneros et al. 2002). In a different study, the effectiveness of spinosad against the green lacewing Chrysoperla carnea’s second instar stages and adults was assessed along with those of chlorpyrifos, lambda-cyhalothrin, cypermethrin, and buprofezin (with the exception of buprofezin) (Sabry and El-Sayed 2011). The results showed that lambda-cyhalothrin (LC50 = 8.81 ppm), cypermethrin (LC50 = 26.9 ppm), spinosad (LC50 = 294.36 ppm), and buprofezin

21.2

Microbial Pesticides

797

(LC50 = 997.05 ppm) were all less hazardous to the second instar stages of the green lacewing Chrysoperla carnea than were chlorpyrifos (in contrast to other insecticides, lambda-cyhalothrin was extremely harmful to Chrysoperla carnea adults). The study demonstrates that the toxicity of pesticides was divided into four categories: dangerous pesticides (chlorpyrifos), moderately harmful pesticides (lambda-cyhalothrin and cypermethrin), mildly harmful pesticides (spinosad), and innocuous pesticides (buprofezin). Because of this, buprofezin and spinosad are better insecticides for integrated pest management (IPM) and can be applied when C. carnea population density is at its highest (Sabry and El-Sayed 2011). Although Bacillus thuringiensis (Bt) is typically considered to be safe for natural adversaries, transgenic Bt-expressing maize has been shown to have adverse effects on a significant predator, Chrysoperla carnea. DiPel (Bt sprays) was evaluated by Dutton et al. (2003) for its impact on Chrysoperla carnea larvae. DiPel (field concentrations) was applied to maize plants, and the aphid Rhopalosiphum padi, the spider mite Tetranychus urticae, and the Lepidoptera larva Spodoptera littoralis used it as food. Aphids were unaffected by DiPel; however, spider mites experienced adverse consequences (lower intrinsic rate of natural increase). Similar to Spodoptera littoralis larvae, DiPel substantially hampered the development of Spodoptera littoralis larvae. Additionally, when Chrysoperla carnea was fed Spodoptera littoralis larvae that were “Bt-contaminated,” DiPel dramatically increased mortality, extended maturation time, and exhibited a modest decrease in weight (Dutton et al. 2003). 21.2.5.2.4 Dermaptera The earwig Doru taeniatum (Dermaptera: Forficulidae) suffered 48% mortality (at 1.2 ppm) and climbed to 98% at 1200 ppm after 14 days of exposure to a product based on the active ingredient in spinosad, as opposed to 20% in controls (Cisneros et al. 2002). Additionally, Cisneros et al. (2002) observed that 72 h after consuming spinosad-contaminated Spodoptera frugiperda (Lepidoptera: Noctuidae) larvae, earwigs suffered 86% mortality/intoxication. The effects of Beauveria bassiana, Metarhizium anisopliae, and their products Boveril® and Metarril®, as well as a control (0.0 conidia L-1), were studied on the survival of the predator Euborellia annulipes (Dermaptera: Anisolabididae). In clutches that had Metarhizium anisopliae applied, high rates of Euborellia annulipes nymphal hatching have been observed. Additionally, Beauveria bassiana has had a deleterious impact on the rate at which Euborellia annulipes nymphs hatch. All concentrations of conidia Beauveria bassiana in clutches of Euborellia annulipes have been shown to have a marginally negative effect. In every concentration of Metarhizium anisopliae tested on female Euborellia annulipes, the survival rate was 100%. Beauveria bassiana was topically applied to females, and the survival rates ranged from 80% (12.50 × 109 conidia) to 100% (5.00 × 109 conidia). B. bassiana treatments resulted in a survival rate of 95.00 (12.50 × 109 conidia) to 100% (5.00 × 109 conidia) in male Euborellia annulipes insects, whereas Metarhizium anisopliae treatments resulted in a survival rate of 96.02 (12.50 × 109 conidia) to 100% (5.00 × 109 conidia). Beauveria bassiana and Metarhizium anisopliae, two

798

21 Biocompatibility of Biopesticides with Predatory Insects

entomopathogenic fungi, had little impact on the mortality of Euborellia annulipes nymphs and adults. However, Beauveria bassiana applications straight into clutches of Euborellia annulipes should be carried out with caution (De Oliveira et al. 2011). 21.2.5.2.5 Various Insects A generalist predator, Chrysoperla carnea, an egg parasitoid, Trichogramma chilonis, and two species of beneficial insects, Bombyx mori and honeybee Apis mellifera, were chosen as non-target insects against whom the impact of Helicoverpa armigera NPV (HearNPV) suspensions was assessed. The results showed that using HearNPV’s wettable powder and suspension formulations against non-target insects had no negative effects. However, the formulations could be used in associated IPM programmes as eco-friendly pest control agents (Mehrvar). Abamectin demonstrated a 39% reduction in egg hatching when tested against Cheilomenes sexmaculata at concentrations of 9, 4.5, 2, 25, 1, 12, and 0.56 mg L-1 (Azod et al. 2019).

21.2.6 Genetically Modified (GM)/Recombinant Microorganisms Worldwide, Bt’s cry protein has been developed into a crucial tool for integrated pest management (IPM), particularly when used against lepidopteran larvae. Recombinant baculoviruses have undergone genetic engineering to shorten the infected pests’ time to death, minimising agricultural harm. The vulnerability of natural enemies (predators and parasitoids) to GM crops is a universal concern; however, tritrophic interactions generally have a minimal impact on natural enemies’ life histories and behaviours. Changes in tritrophic interactions (plant herbivorous parasitoids/ predators) in various food chains caused by GM plants influence natural enemies by changing the volume and nutritional quality of prey as well as the emission of volatile compounds that attract them.

21.2.6.1 Hemiptera Because they can eat eggs and/or larvae, generalist predators like stink bugs, which are prevalent in many crops, can be harmed by Bt plants (Symondson et al. 2002). The predators Orius insidiosus (Hemiptera: Anthocoridae) and Geocoris punctipes (Hemiptera: Geocoridae) consumed larvae in cotton containing Cry1Ac and Cry2Ab proteins or corn containing Cry1F, compared to prey that fed on cotton or isogenic corn, with similar survival, development, fecundity, and fertility. In particular, when exposed to the proteins through their meal, feeding Orius spp. with various preys fed with Bt toxins does not harm Anthocoridae predators (Zhang et al. 2008) (Table 21.2). 21.2.6.2 Chrysoperla spp. Propylea japonica is a common pollen feeder and non-target predatory insect that lives in Bt cotton fields and comes into touch with Bt proteins on a daily basis. The life traits and fertility of the ladybird Propylea japonica (Coleoptera: Coccinellidae) and the predator Chrysoperla rufilabris (Neuroptera: Chrysopidae) fed on fake diets

21.2

Microbial Pesticides

799

Table 21.2 Impact of GM crops fed preys on the life traits of predatory insects Orius insidiosus, Orius sauteri, Orius majuscules, and Orius tantillus Predators Orius insidiosus Orius sauteri

Orius majusculus Orius tantillus a b

Bt/non-Bt Non-Bt Bt Simian 3 (non-Bt) GK-12 (Bt) NuCOTN 33B (Bt) Bt maize Non-Bt maize Bt maize rice Non-Bt maize rice

Nymphal period (days) 16.8 16.9 9.8

Nymphal survival (%) 68.3 70.0 96.3

10 9.8

85.2 92.3

14.07 13.94 13.0 13.6

96.67a 3.33 34.1b 35.7b

References

Zhang et al. (2008)

Mortality longevity

with Bt toxin incorporation were unaffected (Zhao et al. 2016). Mealybugs, Phenacoccus solenopsis, whiteflies, Bemisia tabaci, leafhoppers, and Amrasca biguttula biguttulla, as well as their predator Chrysoperla zastrowi sillemi, were fed herbivores by Bt cotton that did not express single (Cry1Ac) and dual toxins (Cry1Ac and Cry2Ab). However, it reduces the ability of Chrysoperla zastrowi sillemi’s first, second, and third instar stages to feed (22.83 preys/predator) on Phenacoccus solenopsis. However, it has no appreciable impact on life characteristics such the larval period, adult length, or survival rate (Shera et al. 2018) (Table 21.3). However, the consumption of aphids by Chrysoperla carnea larval space predator individuals raised on Bt corn (222.5 individuals) was lower than that of those raised on non-Bt maize (253.5 individuals). When predator larvae fed on aphids reared on non-Bt corn plants (9.0 days) were contrasted with larvae fed on aphids reared on Bt corn plants, the larval duration was dramatically impacted (9.9 days). The percentages of pupation were 60% and 50%, respectively, and the pupal duration on Bt corn was much longer than that on non-Bt corn. On the other hand, compared to those on non-Bt maize, Chrysoperla carnea larvae finished their larval growth and pupal stage in less time (19.4 days and 19.9 days, respectively). Alternatively, Chrysoperla carnea that was fed on aphids raised on Bt corn saw a lower mortality rate (40%) than did those that fed on aphids raised on non-Bt corn (50%). The results showed that Bt plants had no negative effects on the biological activity of Chrysoperla carnea (Moussa et al. 2018). On Solenopsis invicta, Geocoris punctipes, and Hippodamia convergens, common predators of Heliothis in Texas cotton, genetically modified recombinant baculoviruses (Autographa californica nuclear polyhedrosis viruses (AcNPVs), one Helicoverpa zea nuclear polyhedrosis virus (HzNPV), and two corresponding

800

21 Biocompatibility of Biopesticides with Predatory Insects

Table 21.3 Life traits of Chrysoperla zastrowi sillemi, Phenacoccus solenopsis reared on Bt (Cry1Ac, Cry1Ac, and Cry2Ab) and non-Bt cotton cultivars Preys Phenacoccus solenopsis

Bemisia tabaci

Amrasca biguttula

Parameters Larval period (days) Pupal period (days) Adult longevity (days) Larval survival (%) Larval period (days) Pupal period (days) Larval survival (%) Larval period (days) Pupal period (days) Larval survival (%)

Bollgard (Cry1Ac) 11.13

Bollgard II (Cry1Ac and Cry2Ab) 11.12

NonBt 11.08

7.19

7.14

7.24

29.19

29.14

29.38

80

83

80

10.85

10.84

10.80

7.14

7.10

7.05

86.7

83.3

83.3

14.48

14.45

14.43

76.7

73.3

76.7

82.6

86.4

82.6

wild-type NPVs) were assessed. The findings show that no appreciable changes in the life history traits of predators that consumed Helicoverpa virescens larvae infected with any of the seven viruses—rate of food consumption, travel speed, fecundity, and survival—were found. In all, 13.4% of adults and 0.5% of eggs in G. punctipes were positive for viruses. Adult H. convergens were reported to be Polymerase Chain Reaction (PCR)-positive in 12% of cases. Recombinant virus particles may become more persistent in the environment thanks to residence in all three predators that were examined, which may result in an undetermined level of harm from their unintentional mobility (Li and Kevin 1999).

21.2.6.3 Indian Perspective In response to larval infestations of Helicoverpa zea and Helicoverpa virescens, Smith et al. (2000) prepared wild-type viruses and recombinant viruses that have the ability to express a scorpion toxin when applied to cotton. When treatments were administered 3–4 days following the peak of larval emergence, there was no difference in effectiveness between recombinant and wild-type baculoviruses. Similar to this, plots treated with the chemical standard consistently showed fewer predator populations than did those treated with the recombinant and wild-type baculoviruses. However, also, it was shown that 13.5–105 m of 13 of the 26 predators dispersed 2–5 days after the original virus applications. The findings also show that there is little chance for recombinant viral DNA to accidentally

21.2

Microbial Pesticides

801

propagate through dispersed predators (Smith et al. 2000). However, recent research has revealed that, regrettably, the proliferation of fertilizer-intensive Bt seeds has worsened the predation by pests other than lepidopterans (Kranthi and Stone 2020). For the usage of Bt crops in the future, this must be considered. From a US standpoint, Al-Deeb et al. (2003) discovered that the populations of common predators (Hippodamia convergens, Orius insidiosus, and Scymnuss spp.) did not alter noticeably between those in a conventional corn field and those in a Bt corn field. There were just two unusual predatory taxa found on Bt cotton in the United States: Stiretrus anchorago and Geocoris floridanus (Torres and Ruberson 2005). The authors also noted that during the observation periods (2002–2004), predatory heteropterans—including the damsel bug species Nabis roseipennis, Nabis americoferus, Tropiconabis capsiformis, and Nabis alternatus—were rare; however, Geocoris punctipes and Orius insidiosus were more numerous. Studies of a similar nature were also conducted in India in the years 2001–2002 and 2009–2010. Generalist predators, including spiders, chrysopids, and coccinellids, were unaffected (Dhillon et al. 2011).

21.2.6.4 Spain’s Perspective In 1998, comparable research on Bt maize was carried out in Spain. According to the studies, Carabidae and Araneae were the two most common predator groups (85–90% of all predators) gathered in pitfall traps, whereas Anthocoridae, Coccinellidae, and Araneae made up around 90% of the total number of predators documented in visual samplings. They fluctuated in number from one year to the next and between different areas, but no obvious trends associated to Bt maize were noted (de la Poza et al. 2005). 21.2.6.5 Turkey’s Perspective Between 2001 and 2002, a different study was carried out in Turkey using Bt-transgenic (DK-626 Bt) and non-transgenic maize hybrids (DK-626 and Pioneer P-3394) to examine the population growth of generalist predatory insects, including Chrysoperla carnea (Neuroptera: Chrysopidae) and Scymnus levaillanti (Coleoptera: Coccinellidae) (Gullu et al. 2004). According to the findings, predator numbers changed over the course of the maize-growing season. Predator populations did not vary amongst the hybrids of maize, with one exception. The lone exception was the 2002 crop of Chrysoperla carnea, which was found to be substantially more common in Bt maize plots than in non-Bt hybrids. 21.2.6.6 Philippines’ Viewpoint In the Philippines, Bt eggplants (Solanum melongena) expressing Cry1Ac were planted during three cropping seasons in 2010, and the natural enemies were observed. According to a survey, coccinellids (Coccinellidae) were the most numerous predators (Navasero et al. 2016).

802

21 Biocompatibility of Biopesticides with Predatory Insects

21.2.6.7 Brazil and China’s Viewpoints In Brazil, Bt cotton had no effect on the number of larvae and adults of dominating predators (Thomazoni et al. 2013); however, in China, Yang et al. (2015) described changes in agricultural fauna over the previous 15 years as a result of the use of cotton expressing Bt Cry1A toxins. However, it was also reported that using Bt cotton on a large scale led to an increase in the populations of the three important groups of generalist predators in China (ladybirds, lacewings, and spiders) in both Bt cotton and three common neighbouring crops, namely, maize, peanut, and soybean (Lu and Kongming 2012). Additionally, because generalist predators often have significant dispersal abilities, increased predator abundance on Bt cotton eventually enhanced predator-meditated biocontrol services throughout the Chinese agroecosystem. GM cotton (Bt cotton) was first commercialised in Mexico in 1996. Recently, it has been shown that Bt cotton has been impacting the food chains of parasitoids, pollinating insects, and predatory beetles (Rocha-Munive et al. 2018). Because the majority of the generalist predators are phloem feeders, they can continue to provide biological control services in the agro-ecosystem dominated by transgenic Bt cotton crops. 21.2.6.8 Korean Perspective From AD 126, a non-transgenic control line, namely, a transgenic line (C95) of cabbage (Brassica oleracea var. capitata), was created to include Cry1Ac1. Kim and Lee (2018) conducted two experiments back to back in the same experimental field in Korea in 2014. They claimed that the presence of Cry1Ac1 in cabbage had no impact on Coleopteran predator adults, such as Tachyura laetifica and Harmonia axyridis, as well as Propylea japonica and Paederus fuscipes predator larva wild populations.

21.3

Botanical Biopesticides

Pesticidal plants have been seen as a crucial part of BIPM since they have active chemicals that are similar to synthetic pesticides and can harm natural enemies and pollinators who are not the intended targets. Some plant extracts have been known to have insecticidal qualities for a long time. A literature search turned up 283 plant species from 44 plant families that had been used in investigations of habitat alteration. Apiaceae, Apocynaceae, Asteraceae, Boraginaceae, Brassicaceae, Campanulaceae, Fabaceae, Lamiaceae, Myrtaceae, Papaveraceae, Polygonaceae, Primulaceae, Proteaceae, Rosaceae, Rubiaceae, and Scrophulariaceae are just a few of the plant families that feature species that have pesticidal/insecticidal properties. The most species utilised as both botanical pesticides and habitat modification are found in three families: Apiaceae, Asteraceae, and Lamiaceae. The bulk of secondary metabolites, including steroids, alkaloids, tannins, terpenes, phenols, flavonoids, and resins, are the typical bioactive chemicals in botanical pesticides and exhibit antifungal, antibacterial, antioxidant, or insecticidal activities.

21.3

Botanical Biopesticides

803

Table 21.4 Plant bioactive principles and their target pests Plant product used as a biopesticide Limonene and linalool Neem A Pyrethrum/pyrethrins Rotenone

Ryania Sabadilla

Target pests Fleas, aphids and mites, kill fire ants, several types of flies, paper wasps, and house crickets A variety of sucking and chewing insects Ants, aphids, roaches, fleas, flies, and ticks Leaf-feeding insects, such as aphids, certain beetles (asparagus beetles, bean leaf beetles, Colorado potato beetles, cucumber beetles, flea beetles, strawberry leaf beetles, and others) and caterpillars as well as fleas and lice on animals Caterpillars (European corn borers, corn earworms, and others) and thrips Squash bugs, harlequin bugs, thrips, caterpillars, leafhoppers, and stink bugs

Nevertheless, its usage has mostly been restricted to crops grown for subsistence in less developed nations. Research into plant-derived pesticides has increased in part because of adverse environmental impacts on non-target creatures and the development of resistance brought on by the misuse of pesticides in developed nations. Nowadays, a growing number of species from the families Meliaceae and Lamiaceae, which are rich in secondary metabolites, are among plants with a higher potential to be employed for the creation of active products against insects. Reduced predator mobility, lethal (direct effects), non-lethal (indirect effects), decreased ability to recognise and capture prey, decreased food availability, oviposition and feeding repellency, reproductive and developmental problems, and biochemical and molecular changes are some of the consequences. As a result, these natural pesticides offer a novel crop protection alternative that poses fewer hazards to human health and the environment. Therefore, it is important to support their usage, and it is hoped that coordinated efforts would help end these restrictions so that the tremendous potential of these pesticides can be used. Numerous ancient plant bioactive compounds that were examined were effective against a variety of pest insects (see the following Table 21.4). Few species have been studied commercially, even though more than 1500 plants have been designated as botanical insecticides. In addition to their ability to kill pests, botanical pesticides such as Azadirachta indica, Pongamia pinnata (pongamol and pongapin), Derris chinensis and Derris elliptica (Rotenone), Annona squamosa (annonin, squamocin), Ryania speciosa (ryanodine), and others demonstrated the following pest-killing properties (Table 21.5). Neem, Azadirachta indica, has distinguished itself as the best insecticide of plant origin among the available choices. Neem and its compounds are undoubtedly being used more and more in agriculture, but there are ongoing concerns about their safety for the environment. This section describes how to employ plant-based pesticides in BIPM and focuses on their biological effects on entomophages, their predators, and the environment. In comparison to synthetic pesticides, plant-derived pesticides

804

21 Biocompatibility of Biopesticides with Predatory Insects

Table 21.5 Prominent plants used as biopesticides, their active principles, and actions Plant Acorus calamus

Active principles Calamol

Azadirachta indica

Azadirachtin A-G, nimbin, deacetylsalannin, salannin, and their derivatives Sulphur compounds, e.g. diallyl trisulphide, diallyl disulphide, methyl allyl trisulphide allicin

Allium sativum L. (Liliaceae)

Annona squamosa (Annonaceae)

Debitterised annona oil annonin, squamocin

Capsicum annum (Solanaceae)

Proto alkaloids, e.g. capsaicin

Celastrus angulatus maxim. (Celastraceae)

Sesquiterpene pyridine alkaloids

Chrysanthemum cinerariifolium

Pyrethrin 1 and 2, cinerin 1 and 2, jasmolin 1 and 2

Commercial formulations –

Align, Azter, NeemAmin EC, Stardoor and B.P. 20/S, NeemAzal T/S AjoNey (Invernaderos Hidroponicos Neisi, Mexico); EcoA-Z®, L’EcoMix® or CapsiAlil® (Ecoflora Agro, Colombia) ANOSOM® (AgriLife, India)

Hot Pepper Wax (Rincon-Vitova Insectaries, USA); ChileNey (Invernaderos Hidroponicos Neisi, Mexico) CELAN-X SL (Marketing Arm International, Inc., USA) Not identified Citrus sinensis (L.) Osbeck (Rutaceae) limonene and linalool Demize EC (Paragon Professional Pest Control Products, USA); Prev-Am (Oro Agri SA (Pty) Ltd, South Africa) Spruzit® (Neudorff, Germany); PyGanic®Crop Protection EC 5.0 or AZERATM (MGK®, USA); 1.5% Aphkiller

Mode of action Insecticidal and ovicidal, antifeedant, repellent, ovipositiondeterrent, chemosterilant effect Insecticidal, ovicidal, oviposition-deterrent, antifeedant, repellent, insect growth regulators Insecticidal ovicidal, antifeedant, repellent

Insecticidal inhibiting mitochondrial complex III antifeedant, oviposition-deterrent, insect growth regulators In insects, capsaicin causes metabolic disruption, membrane damage, and nervous system failure. It also exhibits physical repellent action Neurotoxic mode of action (hyperactivity followed by hyperexcitation leading to rapid knockdown and immobilisation). Linalool was identified as an inhibitor of acetylcholinesterase

Insecticidal ovicidal, antifeedant disrupting the sodium (Na) and potassium (K) ion exchange process in insect nerve fibres and (continued)

21.3

Botanical Biopesticides

805

Table 21.5 (continued) Plant

Active principles

Commercial formulations AS (Beijing Kingbo Biotech Co., Ltd, China)

Datura stramonium

Atropine

Pongamia pinnata

Pongamol, pongapin karanjin, debitterised karanjin oil

DERISOM® (Agri Life, India); Rock Effect (Agro CS a.s., The Czech Republic)

Derris chinensis, Derris elliptica, Lonchocarpus spp.

Rotenone

5% Rotenone ME (Beijing Kingbo Biotech Co., Ltd, China); Rotenone Dust (Bonide Products, Inc., USA)

Ryania speciosa

Ryanodine

Dust

Schoenocaulon officinale (Melanthiaceae)

Cevadine, veratridine

Veratran D® (MGK®, USA)

Vitex negundo, Vitex trifolia

Vitexin and negundoside

Ocimum sanctum

Juvocimene 1 and 2, ocimin

Nicotiana tabacum

Nicotine and nornicotine

Tagetes erecta

Tagetone, mycene

Dust or Nico Neem (Nico Orgo Manures, India); 10% Nicotine AS (Beijing Kingbo Biotech Co., Ltd, China)

Mode of action interrupting the normal transmission of nerve impulses Insecticidal and ovicidal, antifeedant, repellent Feeding deterrent, growth regulator, oviposition (egg deposition)suppressant or oviposition-sterilant Insecticidal and antifeedant, inhibitor of cellular respiration (mitochondrial complex electron transport inhibitor), causing rapid cessation of feeding. Death occurs several hours to a few days after exposure Contact and stomach poison Affects nerve cell membrane action, causing loss of nerve cell membrane action, loss of nerve function, paralysis, and death Insecticidal, antifeedant, and repellent Insecticidal and ovicidal, ovipositiondeterrent, antifeedant, repellent, insect growth regulators Insecticidal and ovicidal, ovipositiondeterrent, disruption of normal nerve impulse activity Insecticidal, nematicidal, repellent, fungicidal

806

21 Biocompatibility of Biopesticides with Predatory Insects

generally spare the beneficial fauna due to their inherent low mammalian toxicity and capacity for biodegradation. This is true even when they are efficient against agriculturally significant insect pests. While foliar application may significantly minimise the danger, several plant-derived chemicals, either in crude form or in formulation, demonstrated little-to-moderate negative impacts on beneficial fauna (parasitoids, predators, and pollinators). Azadirachta indica, sometimes known as the neem tree, is a member of the Meliaceae family and was first found in India and Myanmar. It can be grown in both subtropical and tropical climates. The primary triterpenoid molecule found in neem oil is azadirachtin (A, B, C, D, E, F, G, H, I, etc.). Among the remaining triterpenoids were salannin and nimbin. It consists of more than 200 allelochemicals, which are present in different areas of the plant in varying amounts and exhibit a range of pesticidal activities. Because of their insecticidal, fungicidal, bactericidal, and nematicidal capabilities, neem components have garnered interest on a global scale. Similar to Azadirachta indica, Melia azedarach has limonoids in common. Meliantriol, melianone, melianol, meliacin (1-cinnamoyl melianone), meliacarpin, and meliartenin are a few of the limonoids that have been isolated from the fruits of M. azedarach. One of the most intriguing components of neem, which is made from seed kernels, is azadirachtin, which has been proven to alter insect feeding behaviour and insect development as well as to have strong efficacy against a variety of insect pests (Table 21.5). Additionally, azadirachtin has a generally acceptable toxicological profile. The most researched azadirachtins are azadirachtin A, azadirachtin B, nimbin, and salannin. Azadirachtins have several isomers ranging from A to I. Neem products have proven to be particularly successful at lowering phytophagous insects and enhancing plant health. Neem also interacts with other BIPM elements, such as predatory insects. An in-depth discussion of the benefits and drawbacks of neem’s bioactive ingredients for predatory insects is covered in this chapter.

21.3.1 Coleoptera According to Medina et al. (2001; Vinuela et al. 2000), azadirachtin has an impact on the biology of Coccinella septempunctata (Coccinellidae) and Chrysoperla carnea (Chrysopidae) (Banken and Stark 1997). Coccinella undecimpunctata L. and Eupeodes fumipennis (Thompson) (Coccinellidae) were not affected, though (Lowery and Isman 1995). A commercial neem insecticide was found to be non-toxic to adult coccinellid predators by Hoelmer et al. in 1990. When directly sprayed on Coccinella septempunctata (seven-spot ladybug) larvae in the lab, azadirachtin (Neemix) was essentially harmless (Banken and Stark 1997). No pesticide, according to Tunca et al. (2020, 2021), is completely risk-free and non-toxic to natural enemies. However, compared to synthetic chemical pesticides, botanical pesticides typically have a far higher margin of safety. The findings of this investigation concur with those reported by Tunca et al. (2021), who discovered that new chemistry insect. Feeding adult (spotted ladybird beetle) and Mallada signatus

21.3

Botanical Biopesticides

807

(Neuroptera) larvae and their prey (Helicoverpa armigera’s second instar stages) that had consumed neem oil solution allowed researchers to test the effects of predators being indirectly exposed to neem seed extracts (50 and 200 ppm azadirachtin) (Qi et al. 2001). Additionally, it was discovered that azadirachtin treatments at concentrations of 50 and 200 ppm were not hazardous to adult Harmonia conformis or M. signatus larvae. It also manifested some impacts on M. signatus metamorphosis at the same time. Additionally, M. signatus larvae were found to consume more second instar Helicoverpa armigera after being exposed to 200 ppm azadirachtin. Treatments with azadirachtin had an adverse effect on M. signatus, slowing pupation and reducing pupal survival. An important biological control agent called Harmonia axyridis feeds on Aphis glycines Matsumura, an invasive insect pest of North American soybeans (Coleoptera: Coccinellidae). Neem is occasionally used to control populations of harmful pests. Neem seed oil (Ahimsa Botanicals, Bloomington, MN; 1% v:v) emulsified with organic castile soap (Dr. Bronner’s Soaps, Menomonee Falls, WI; 0.1% v:v) and a deionised water-only control served as the treatments. Azadirachtin 45 g L-1 EC (Neemix4.5 EC; Certis USA, Columbia) on a crop of legumes, Neemix was treated at the maximum labelled field rate for aphids. The results show that there were substantial differences between treatments in the proportion of Harmonia axyridis’s first and third instars that survived to adulthood. Both azadirachtin and neem oil considerably decreased first instar pairwise comparisons of survivorship to adulthood (Kraiss and Cullen 2008), whereas azadirachtin dramatically decreased third instar pairwise comparisons of survivorship to adulthood when compared to neem oil. Other coccinellids showed elevated mortality rates for second instar Hippodamia variegate treated with azadirachtin and neem seed powder combined with water, compared to a water control, of 40% and 27%, respectively (Hamd et al. 2005). Similar to the current study, only 42% of first instar Coccinella undecimpunctata treated with 1% neem oil survived to adulthood (Lowery and Isman 1995). Recently, in India, it has been discovered that feeding of Coccinella septempunctata (28.1%) and Coccinella sexmaculata (34.7%) was decreased by 3% neem oil (Meena et al. 2020).

21.3.2 Neuroptera The generalist predator, the lacewing Chrysoperla carnea, was investigated in 2003 to see how it responded to the insect growth regulators azadirachtin, diflubenzuron, pyriproxyfen, and tebufenozide. Third instars were directly exposed to prepared materials of each drug in varying doses for topical treatment. Pyriproxyfen and tebufenozide were safe for Chrysoperla carnea at the maximum field recommended dose, whereas azadirachtin and diflubenzuron were not (the respective LD90 values were 24.5 and 6.9 ng active ingredients (a.i.) per insect). Females that produced viable eggs were given sub-lethal doses of azadirachtin and diflubenzuron, but azadirachtin had a minor negative impact on oviposition. Ovarian development and egg fertility were unaffected by pyriproxyfen and tebufenozide. The rate of

808

21 Biocompatibility of Biopesticides with Predatory Insects

penetration and excretion following topical application is explored in relation to toxicity data as a second strategy of this investigation. Pyriproxyfen exhibited a penetration rate of 80% 1 h after delivery, compared to only percentages of 10–20% for diflubenzuron and tebufenozide during the same period. Although the penetration of pyriproxyfen was rapid and high, the majority of the drug was also promptly excreted. Our findings imply that careful consideration should be given to the use of azadirachtin and diflubenzuron in conjunction with Chrysoperla carnea in integrated pest management (IPM) programmes. Tebufenozide and pyriproxyfen are safe for Chrysoperla carnea. In a different study, Medina et al. (2004) examined the effects of azadirachtin (Align), a commercial formulation, on reproduction in Chrysoperla carnea (Neuroptera: Chrysopidae) adults. Under laboratory conditions, they used three exposure techniques (residual, direct contact, and ingestion) at a concentration of 48 mg AI L-1. Regardless of the method of exposure, Align proved to be safe for newly emerging adults. However, in females that took azadirachtin after the start of oviposition, fertility was decreased in a manner that could be reversed. In all cases of therapy, fertility was comparable to that of controls. Additional research revealed that men were not responsible for the decrease in oviposition. Furthermore, electron microscopy research shows that Align interfered with vitellogenin synthesis and/or uptake by developing oocytes, resulting in considerably smaller growing follicles in treated females than in controls. This is because azadirachtin may affect several protein and hormone levels that are involved in reproduction. Under laboratory and field circumstances, the neem-based pesticide Azter (azadirachtin 0.15% EC) did not harm non-target organisms such Mallada desjardinsi (Neuroptera) and Oligota pygmaea (Coccinellidae), which are natural adversaries of red spider mite (RSMs) (according to Vasanthakumar et al. 2013). A lab test was carried out in Pakistan in 2014 to determine the impact of neem and datura leaf extracts on the biology of Chrysoperla carnea. The findings show incubation times (2.2 days, 2.5 days, and 3.6 days, respectively), total larval developmental times (17.03, 13.3, and 15.09 days), pupal times (8.82, 10.9, and 12.33 days), oviposition times (34.42, 30.6, and 26.4 days), pre-oviposition times (6.35, 5.5, and 3.6 days), and post-oviposition times (8.5, 6.9, and 4.7 days) on neem, datura and confidor, respectively. However, neem leaf extracts caused the least amount of mortality in the first, second, and third instar larvae of Chrysoperla carnea, followed by datura and confidor. The neem pesticide had the highest adult death rate, followed by the datura and confidor insecticides (Ali et al. 2015). Cyrtorhinus lividipennis was used as a test subject for the bioactivities of methyl eugenol (ME), which revealed contact toxicity (LC50 values of 527 mg L-1). However, in contrast to synthetic insecticides, the toxicity is lower (Xu et al. 2015). Neem oil is an efficient pest management method, but it also has unintended side effects on unintended pests, such as green lacewings and eggs of Diatraea saccharalis that were pre-treated with 0.5%, 1%, and 2% neem oil (Garcia and Scudeler 2019). Additionally, the outcomes demonstrated that neem oil alters the pattern of cyst distribution and organisation in the testes as well as the usual progression of cyst growth, delaying spermatogenesis in the testes of treated insects. There was no cellular modification in the treated groups, according to tests for

21.3

Botanical Biopesticides

809

cellular stress and DNA fragmentation. This biopesticide has an adverse effect on the spermatogenic process and may reduce the persistence of this species in the agroecosystem (Garcia and Scudeler 2019). On the larvae of C. carnea, cinnamon oil and its active component, cinnamaldehyde, were examined at concentrations of 500, 1000, 5000, and 10,000 ppm. The results show that employing the active component cinnamaldehyde rather than cinnamon oil increased C. carnea’s predation efficiency (Ghada and Naglaa 2020).

21.3.3 Heteroptera Blaptostethus pallescens, a biocontrol agent found in Brazilian tropical tomato fields that manages fruit borer populations, belongs to the family Anthocoridae. Celestino et al. (2014) evaluated the lethal response of B. pallescens to the bioinsecticide azadirachtin and to two synthetic insecticides (chlorpyrifos and deltamethrin). The assessment of its potential sub-lethal effects was prompted by the discovery that the median lethal time (LT50) of azadirachtin (0.006 mg a.i. mL-1) on the predator is 27 days, as opposed to deltamethrin (0.02 mg a.i. mL-1) and chlorpyrifos (1.44 mg a.i. mL-1). The male and female parents of the pirate bug were both exposed to azadirachtin, which reduced daily fecundity, adult progeny output, and sex ratio but did not result in behavioural avoidance. These outcomes slowed the predator’s population increase in succeeding generations. Azadirachtin warrants caution when applied in fields with this biological control agent, the authors suggested, despite being safer than the standard synthetic insecticides tested. Each predator of the tomato pinworm Tuta absoluta (Lepidoptera: Gelechiidae), Amphiareus constrictus, or Blaptostethus pallescens (Hemiptera: Anthocoridae) received 200 μL of either insecticide or control solution applied with an artist’s air brush (Sagyma SW440A, Brazil) attached to an air pump (Prismatec 131A Tipo 2 VC, Brazil) at a pressure of 6.9 × 104 Pa. Azadirachtin appeared to alter adult predator mortality or the capacity of their eggs to hatch, indicating a safe acute toxicity for these stages. Azadirachtin, however, adversely affected the predatory nymphs’ ability to mature. This decline may jeopardise the biological control of pests like T. absoluta and have a direct impact on the size of the predator population in the next generation. Our findings generally recommend using these low-risk insecticides with caution in integrated pest management programmes that combine chemical and biological methods (Gontijo et al. 2015).

21.3.3.1 Miridae Neem-Amin EC, Stardoor, and B.P. 20/S (neem formulations) were examined in the lab utilising direct toxicity testing on Macrolophus caliginosus’s first instar stages (Tedeschi et al. 2001). The findings showed that all the products were harmful to insects, with LD50 values much lower than the maximum recommended rate (1.217, 0.264, 1.083 mg a.i./l instead of 15, 31.5, and 80 mg a.i./l for Neem-Amin EC, Stardoor, and B.P. 20/S respectively). With Neem-Amin EC and B.P. 20/S, the fertility of the remaining females was also seen to decrease. Three citrus essential

810

21 Biocompatibility of Biopesticides with Predatory Insects

oils (EOs; lemon, mandarin, and sweet orange) were tested against Nesidiocoris tenuis (Hemiptera: Miridae) using two formulations (emulsion and poly (ethylene glycol) (PEG) nanoparticles). The outcome revealed that the insecticide based on mandarin EO lowered the survival and reproduction of predators (Campolo et al. 2020). The effects of Mentha pulegium (Papadimitriou et al. 2019), garlic (Allium sativum), and other citrus species (Campolo et al. 2020; Ricupero et al. 2022) were tested. Nesidiocoris tenuis (Hemiptera: Miridae) was used as a test subject for EOs, and the results indicate both acute toxicity and sub-lethal effects. Recently, anise (Pimpinella anisum: Apiaceae), fennel (Foeniculum vulgare: Apiaceae), lavender (Lavandula angustifolia: Lamiaceae), and garlic have been used to make pesticide nano-emulsions (A. sativum). The IPM has not endorsed garlic-based nanoemulsions due to their increased toxicity (LC30 = 1.34 mg mL-1) (Passos et al. 2022).

21.3.4 Different Predators Citronella oil and eucalyptus oil, two natural pesticides, were extremely poisonous to the pumpkin pest. They proved to be less damaging to the predatory ants Paratrechina sp. and Diaphania hyalinata. Garlic extract, eucalyptus oil, and citronella oil all prevented the Diaphania hyalinata larvae from eating. Rotenone, andiroba oil, garlic extract, and eucalyptus oil all prevented Diaphania hyalinata from laying eggs. Pumpkin plants were not phytotoxic to any of the pesticides. Citronella oil and eucalyptus oil significantly reduced target pest mortality and changed their behaviour. Additionally, they did not display phytotoxicity on pumpkin plants and were selective for the predator Paratrechina sp. Because they caused significant mortality, changed the behaviour of the target pests, and were selective for the predator Paratrechina sp., eucalyptus oil and citronella oil appear to be the most promising chemicals for use in melonworm management strategies (Silva and Adriano 2016). We investigated the resistance of tobacco and chilli extracts to red ants (Wasmannia auropunctata) and 28-spotted ladybird beetles (Epilachna spp.). This study has demonstrated that natural predators and other natural enemies can peacefully coexist with botanical pesticides in most cases (Iamba and Yoba 2019). Predatory insects like Podisus nigrispinus (Heteroptera: Pentatomidae), which feed on agricultural crops and forest plantations in America, can be controlled. Lepidopteran pests like Euprosterna elaeasa (Limacodidae), Spodoptera exigua Hübner, and Trichoplusia ni (Noctuidae), which are these insects’ natural prey in Brazilian agricultural crops, have been found to be poisonous to Podisus nigrispinus by essential oils. However, essential oils’ fatal and sub-lethal effects on this predatory beetle have also been demonstrated. In various crop systems, Podisus nigrispinus is a predator of pests that cause defoliation; nevertheless, in order to prevent damaging this natural ally, more research is needed to determine how essential oils acting as insecticides affect this pest’s natural opponent. Four commercial formulations—pyrethrin (McLaughlin Gormley King Co., USA), spinosad (Dow AgroSciences, USA), azadirachtin + pyrethrin (McLaughlin Gormley King

21.3

Botanical Biopesticides

811

Co., USA), and chlorantraniliprole (DuPont, USA)—were tested using contact and oral toxicity bioassays on Podisus maculiventris nymphs. According to a contact toxicity test, pyrethrin and azadirachtin + pyrethrin were more harmful to Podisus maculiventris than was spinosad. Spinosad was the most harmful pesticide to Podisus maculiventris in an oral toxicity test, followed by pyrethrin, azadirachtin + pyrethrin, and chlorantraniliprole. The results demonstrate how important a role bioassays play in a substance’s toxicity to predators (de Castro et al. 2018). In an experiment on Podisus nigrispinus, Brügger et al. (2019) sought to determine the lethal and sub-lethal effects of lemongrass essential oil and its terpenoid components (geranyl acetate and citral). The results indicate that the lethal doses of lemongrass oil increased from the first to the fifth instars, with LD50 values ranging from 1.08 to 139.30 g insect-1 and LD90 values ranging from 2.02 to 192.05 g insect-1. For third instar Podisus nigrispinus nymphs, the LD50 and LD90 values of the lemongrass were 21.58 g insect-1 and 28.35 g insect-1, respectively. In the control, mortality was always less than 1% (Brügger et al. 2019).

21.3.5 Combined Effects of Different Botanicals The Syrphidae, Coccinellidae, and Chrysopidae family of predators have kept the cabbage aphid, Brevicoryne brassicae L., one of the most significant insect pests harming cabbage crops in Poland, in check. Spinosad 96 g a.i. ha-1 (as a Spintor 480 SC: 0.2 L ha-1) and azadirachtin 24 mL a.i. ha-1 (as a NeemAzal T/S: 2.4 L ha1 ) were used in a field experiment in 2004 and 2006, respectively. The amount of Coccinellidae larvae, Coccinellidae beetles, and Syrphidae spp. on plants treated with the tested pesticides and on untreated control plants did not differ substantially over the course of the study. Similar to this, there was no statistically significant difference in the quantity of Chrysoperla sp. eggs laid on cabbage plants treated with spinosad or azadirachtin compared to untreated plants (Nawrocka 2008). In North America, pyrethrins and azadirachtin from two botanical insects were observed by Pezzini and Koch (2015) to be compatible. In North America, the soybean aphid Aphis glycines may be preyed upon by Chrysoperla rufilabris, Orius insidiosus, and Hippodamia convergens. Toxic to both adult Orius insidiosus and the larvae of Chrysoperla rufilabris was the pyrethrin–azadirachtin combination. Additionally, the combination of pyrethrins and azadirachtin prolonged Chrysoperla rufilabris’s growth. These findings suggest that flonicamid and the combination of pyrethrins and azadirachtin may be able to improve the compatibility between chemical and biological controls.

21.3.6 Botanical Products, Including Oil Thymol is a primary component of Lippia sidoides oil. Studying its effects was the key predator of the defoliating caterpillar Spodoptera frugiperda, Podisus nigrispinus (Lima et al. 2020). It has a 93-h lethal time (LT50). Additionally, nymphs

812

21 Biocompatibility of Biopesticides with Predatory Insects

exposed to thymol displayed altered behaviour, including variations in speed and distance travelled. More insects remained on the untreated side of the arenas than did those that had received thymol treatment. Products Prev-Am® oil

Natural enemy Nesidiocoris tenuis

Thymol

Podisus nigrispinus

Neem, pungam, and madhuca

Cheilomenes sexmaculata, Chrysoperla carnea Cyrtorhinus lividipennis Coccinella septempunctata to Toxoptera aurantii

Methyl eugenol (ME) Pyrethrin, oxymatrine, celangulin, and matrine

21.4

Type of consequences Reduction in predatory voracity Changes in predation and prolong mortality Modulation of body carbohydrates

References Soares et al. (2019) Lima et al. (2020)

Mortality

Razak and Sivasubramanian (2007) Xu et al. (2015)

Mortality, predatory behaviour

Wang et al. (2014)

Field Trials

Studies Conducted in Greenhouses Dutton et al. (2003) evaluated the impact of spraying Chrysoperla carnea larvae in a greenhouse with DiPel (a marketed Bt treatment against Spodoptera littoralis larvae infesting maize). The findings indicated adverse effects on the predator’s larvae, including a notable rise in mortality, a lengthened developing period, and a somewhat lower adult weight.

21.4.1 Botanicals IR42 rice seeds were soaked in 0% neem seed kernel extract (NSKE) solution, seedling roots were deepened in 5% NSKE solution for 12 h, and 150 L ha-1 of 4% NSKE solution emulsified with 1% Teepol was sprayed six times at intervals of 10 days starting 5 days after transplanting (DT). Insecticide-treated plots had significantly ( 5%) fewer predatory mirid and spider populations than neem-treated areas, with a difference of 48 DT (Kareem et al. 1988). According to Ahmad et al. (2003), the first instar stages of Coccinella septempunctata exhibited increased mortalities (at 100%) when aphid-infested fields were sprayed with neem oil. Jones et al. (2005) noted that natural predatory fauna were unaffected by bacteriaand neem-based pesticides. When a bean field was treated with fresh neem oil at 2.5 mL L-1 water, stored neem oil at 2.5 mL L-1 water killed lady beetles (Mollah et al. 2013). When pungam oil and neem oil were applied to mulberry plants, the number of predatory coccinellid beetles decreased by 29.72% and 35.20%, respectively. The efficacy of three natural pesticides (neem, datura, and bitter apple) against

21.4

Field Trials

813

onion thrips (Thrips tabaci) was examined in Pakistani experimental field plots in 2014 by Abdul Khaliq et al. The results indicate that acephate had a more severe negative impact on lady beetle, hover fly, and lacewing populations than did the other novel chemicals and botanicals. Additionally, there was a noticeable difference in lady beetle, hover fly, and lacewing populations between the treated and untreated plots overall. The post-host testing, however, revealed that these variations were caused by the various impacts of the chemical pesticides. Citrullus colocynthis, Azadirachta indica, tobacco Nicotina tabbacium, and Movento (Spirotetramat) botanical pesticides were less hazardous when applied in the field to coccinellid predators C. septempunctata, Brumoides suturalis, and M. sexmaculatus (Kunbhar et al. 2018). The populations of the predator’s coccinellids and C. carnea were significantly impacted by the use of two herbal remedies in 2019 (A. indica and M. azedarach). Intriguingly, the plot where A. indica was used had a greater population of coccinellids (1.07 individuals per plant), whereas M. azedarach had a lower population (0.93 individuals per plant). The coccinellid population was lower (0.53 individuals per plant), and there was no discernible difference between the positive and negative controls ( p > 0.05). In the plots where M. azedarach was applied, C. carnea was found to be more abundant (1.13 individuals per plant), followed by A. indica (0.87 individuals per plant) and the negative controls (0.87 individuals per plant). The population of C. carnea exposed to bifenthrin was extremely low (0.40 individuals per plant). Overall, predators were more numerous in plots where botanicals were used, whereas bifenthrin insecticide application in the positive control treatment resulted in the lowest number of predators (Arshad et al. 2019). In a different study, it was discovered that the treatment of one spray of botanicals in a mulberry plantation was more effective than were sprays of just botanicals, such as neem oil and pongamia oil. When applied directly to predators in tomato (Solanum lycopersicum) fields, karanja oil was lethal to both, including the predatory lady beetle Delphastus catalinae, suggesting that in integrated management programmes, natural enemies should be introduced after the application of these botanical derivatives (Cabra 2020).

21.4.2 Botanical Products Coccinella septempunctata (Coleoptera: Coccinellidae) responds to neem commercial insecticides like Neemix (Banken and Stark 1997). A 2-year trial was conducted in 2005–2006 at the Lithuanian Institute of Horticulture to evaluate the toxicity of biopesticides such as BioNature R2000 (Azadirachta indica 210 g L-1, Pinus resinosa 180 g L-1, and Ricinus communis), Bioshower (100% fatty acids), and insecticidal soap (20% fatty acids) against Chrysoperla L. (Raudonis et al. 2010). The tested products had no harmful effects on Chrysoperla larvae. While Bioshower and insecticidal soap demonstrated no toxicity to the larvae of C. septempunctata, BioNature R2000 was mildly to moderately harmful. While BioNature R2000 was only marginally poisonous to Amblyseius andersoni, Bioshower and insecticidal

814

21 Biocompatibility of Biopesticides with Predatory Insects

soap were both moderately to extremely toxic. Many nations permit the use of biopesticides BioNature R2000, Bioshower, and insecticidal soap in organic farming, and these products could be used to safely control aphids (Raudonis et al. 2010).

21.4.3 Post-harvest Storage Researchers are looking for alternatives to synthetic pesticides because of the rising worry regarding the amount of pesticidal residues in food. Their uncontrolled usage has caused a variety of environmental and human health issues as well as the emergence of insect strains that are resistant to them. Recently, attention has been focused on using higher plant products as innovative chemotherapeutics in plant protection in many parts of the world. Plant products have the potential to be used in pest management because of their non-phytotoxicity, systemicity, simple biodegradability, and stimulatory nature of the host metabolism. These organic pesticides, which were commonly used up until the 1940s, were replaced by synthetic pesticides of the present, which at the time seemed more convenient, affordable, and longlasting. Botanical pesticides are once more becoming more and more popular, and some plant products are being employed as green insecticides on a global scale. Even though there is a growing quantity of the scientific literature demonstrating the bioactivity of plant derivatives against various pests, only a small number of botanicals are now utilised in agriculture. Because of their generally safe status and widespread customer acceptance, pyrethroids and neem products are wellestablished in the market as botanical pesticides. More recently, several essential oils of higher plants have also been utilised as antimicrobials against storage pests. Some volatile oils have been suggested as plant-based antimicrobials to prevent microbial contamination and reduce food spoiling since they frequently contain the main aromatic and flavouring components of herbs and spices. Additionally, several plant products that are antibacterial also have significant antioxidant characteristics, which are advantageous for preventing free radical-mediated organoleptic deterioration of plant commodities and lengthening their shelf life. Botanical pesticides are most suited for use in the cultivation of organic foods in industrialised nations, but they can play a far larger role in the production and post-harvest protection of food goods in developing nations (says Dubey et al. 2008). Spinosad vs. Dermaptera In a field trial in 2002, Cisneros et al. (2002) compared the effects of applying commercial granular chlorpyrifos and spinosad on earwigs housed in gauze bags (200 ppm and 2000 ppm at 4.8–48 g a.i. ha-1 and 160 ppm at 48 g a.i. ha-1, respectively). Compared to 33% on plants treated with granular chlorpyrifos, 83% on plants sprayed with 160 ppm spinosad, and 91–95% on plants treated with 200–2000 ppm spinosad granules, the mortality of earwigs, Doru taeniatum, on control plants was less than 15% at 2 days post-application. Additional mortality over the 24-h post-sampling period varied from 5% in control treatments to 9% in the chlorpyrifos treatment to 55–65% in the spinosad spray and granular treatments (Cisneros et al. 2002). The authors come to the conclusion

References

815

that spinosad does not share the majority of well-known biological pesticides’ environmental safety profile. The authors concluded that spinosad does not share a majority of the environmental safety profiles of well-known biological pesticides. Consideration of botanical pesticides in the future.

21.5

Recommendations

The aforementioned discussion reveals that while biopesticides have been acknowledged as a crucial element of BIPM, it is crucial to assess their effectiveness against biological control agents in a laboratory setting as well as in potted plants, controlled field cages, screen houses, cages, and fields before introducing them. Additionally, the same can be researched for various crops and agro-ecological zones, with recommendations made at the whole national level.

References Abbas MST, Boucias DG (1984) Interaction between nuclear polyhedrosis virus-infected Anticarsia gemmatalis (Lepidoptera: Noctuidae) larvae and predator Podisus maculiventris (Say)(Hemiptera: Pentatomidae). Environ Entomol 13(2):599–602 Ahmad M, Ossiewatsch HR, Basedow T (2003) Effects of neem-treated aphids as food/hosts on their predators and parasitoids. J Appl Entomol 127:458–464. https://doi.org/10.1046/j. 0931-2048.2003.00779.x Al-Deeb MA, Wilde GE, Blair JM (2003) Effect of Bt corn for corn rootworm control on nontarget soil microarthropods and nematodes. Environ Entomol 32(4):859–865 Ali SS, Khaskheli MY et al (2015) Effects of bio-pesticides on biology of Chrysoperla carnea F. (Neuroptera: Chrysopidae). J Basic Appl Sci 11:559–566 Arthurs S, Thomas MB (2001) Effects of temperature and relative humidity on sporulation of Metarhizium anisopliae var. acridum in mycosed cadavers of Schistocerca gregaria. J Invertebr Pathol 78(2):59–65 Arthurs SP, Lacey LA, Miliczky ER (2007) Evaluation of the codling moth granulovirus and spinosad for codling moth control and impact on non-target species in pear orchards. Biol Control 41(1):99–109 Arshad M, Ullah MI, Çağatay NS, Abdullah A, Dikmen F, Kaya C, Khan RR (2019) Field evaluation of water plant extracts on sucking insect pests and their associated predators in transgenic Bt cotton. Egypt J Biol Pest Control 29(1):1–6 Azod F, Shahidi-Noghabi S, Mahdian K (2019) The impact of the pesticides abamectin and spirotetramat on biological parameters of the coccinellid predator Cheilomenes sexmaculata (Fabricius) via egg exposure. Pistachio Health J 2(1):62–70 Bajwa WI, Aliniazee MT (2001) Spider fauna in apple ecosystem of western Oregon and its field susceptibility to chemical and microbial insecticides. J Econ Entomol 94(1):68–75 Banken JAO, Stark JD (1997) Stage and age influence on the susceptibility of Coccinella septempunctata (Coleoptera: Coccinellidae) after direct exposure to Neemix, a neem insecticide. J Econ Entomol 90:1102–1105 Bharani GN, Kohilambal H, Sivasubramanian P (2015) Comparative efficacy of bio pesticides and insecticides against tomato thrips (Thrips tabaci Lind.) and their impact on coccinellid predators. The Bioscan 10(1):207–210 Boyd ML, Boethel DJ (1998) Susceptibility of predaceous hemipteran species to selected insecticides on soybean in Louisiana. J Econ Entomol 91(2):401–409

816

21 Biocompatibility of Biopesticides with Predatory Insects

Bret BL et al (1997) Biological properties of spinosad. Down Earth 52(1):6–13 Brügger BP, Martínez LC, Plata-Rueda A et al (2019) Bioactivity of the Cymbopogon citratus (Poaceae) essential oil and its terpenoid constituents on the predatory bug, Podisus nigrispinus (Heteroptera: Pentatomidae). Sci Rep 9(1):1–8 Cabra C (2020) Toxicity across food webs: effect of a botanical extract on tri-trophic interactions in tomato (Doctoral dissertation, University of British Columbia) Campolo O, Puglisi I, Barbagallo RN et al (2020) Side effects of two citrus essential oil formulations on a generalist insect predator, plant and soil enzymatic activities. Chemosphere 257:127–252 Carvalho VFP, Vacari AM, Pomari AF (2012) Interaction between the predator Podisus nigrispinus (Hemiptera: Pentatomidae) and the entomopathogenic bacteria Bacillus thuringiensis. Environ Entomol 41(6):1454–1461 Carvalho GASD, Martins DJ, Brito IMCD et al (2018) Can Bacillus thuringiensis affect the biological variables of natural enemies of Lepidoptera? Arquivos do Instituto Biológico 85: e0052018 de Castro AA, Legaspi JC, Tavares WDS et al (2018) Lethal and behavioral effects of synthetic and organic insecticides on Spodoptera exigua and its predator Podisus maculiventris. PLoS One 13 (11):e0206789 Celestino D, Braoios GI, Ramos RS et al (2014) Azadirachtin-mediated reproductive response of the predatory pirate bug Blaptostethus pallescens. BioControl 59(6):697–705 Cisneros J, Goulson D, Derwent LC (2002) Toxic effects of spinosad on predatory insects. Biol Control 23(2):156–163 Croft BA (1990) Arthropod biological control agents and pesticides. John Wiley and Sons Inc, New York, NY De Nardo EA, Maia AH, Watanabe MA (2001) Effect of a formulation of Anticarsia gemmatalis (Lepidoptera: Noctuidae) nuclear polyhedrosis virus on the predator Podisus nigrispinus (Heteroptera: Pentatomidae: Asopinae), using the fertility life table parameters. Environ Entomol 30(6):1164–1173 De Oliveira FQ, Batista JDL, Malaquias JB (2011) Susceptibility of the predator Euborellia annulipes (Dermaptera: Anisolabididae) to mycoinsecticides. Rev Colomb Entomol 37 (2):234–237 Dhillon MK, Gujar GT, Kalia V (2011) Impact of Bt cotton on insect biodiversity in cotton ecosystem in India. Pak Entomol 33(2):161–165 Dubey NK, Srivastava B, Kumar A (2008) Current status of plant products as botanical pesticides in storage pest management. J Biopest 1(2):182–186 Dutton A, Klein H, Romeis J (2003) Prey-mediated effects of Bacillus thuringiensis spray on the predator Chrysoperla carnea in maize. Biol Control 26(2):209–215 Elimem M, Chermiti B (2012) Effect of Orius laevigatus and Amblyseius swirskii releases on Frankliniella occidentalis populations in pepper crop greenhouses in the Bekalta region of Tunisia. IOBC/WPRS Bull 80:147 Elzen GW, Elzen PJ, King EG (1998) Laboratory toxicity of insecticide residues to Orius insidiosus, Geocoris punctipes, Hippodamia convergens, and Chrysoperla carnea. Southwest Entomol 23:335–342 Gao YL, Reitz SR, Wang J, Tamez-Guerra P, Wang ED, Xu XN, Lei ZR (2012) Potential use of the fungus Beauveria bassiana against the western flower thrips Frankliniella occidentalis without reducing the effectiveness of its natural predator Orius sauteri (Hemiptera: Anthocoridae). Biocontrol Sci Tech 22:803–812 Garcia ASG, Scudeler EL (2019) Can exposure to neem oil affect the spermatogenesis of predator Ceraeochrysa claveri? Protoplasma 256(3):693–701 Ghada AAE, Naglaa YM (2020) Efficacy of cinnamon oil and its active ingredient (cinnamaldehyde) on the cotton mealy bug Phenacoccus solenopsis Tinsley and the predator Chrysoperla carnea. Bull Natl Res Centre 44(1):1–4

References

817

Giroux S, Coderre D, Vincent C (1994) Effects of Bacillus thuringiensis var. San Diego on predation effectiveness, development and mortality of Coleomegilla maculata lengi (Col.: Coccinellidae) larvae. Entomophaga 39(1):61–69 Gontijo LM, Celestino D, Queiroz OS (2015) Impacts of azadirachtin and chlorantraniliprole on the developmental stages of pirate bug predators (Hemiptera: Anthocoridae) of the tomato pinworm Tuta absoluta (Lepidoptera: Gelechiidae). Fla Entomol 98:59–64 Gullu M, Tatli F, Kanat AD (2004) Population development of some predatory insects on Bt and non-Bt maize hybrids in Turkey. IOBC WPRS Bull 27(3):85–92 Hamd EEA, El Shafie HAF, Basedow T (2005) The different effects of two preparations of neem (Azadirachta indica) and of sumicidin on the aphid predator Hippodamia variegate (Goeze) (Coleoptera: Coccinellidae). J Plant Dis Prot 112:580–585 Hilbeck A, Moar WJ, Pusztai-Carey M (1998) Toxicity of Bacillus thuringiensis CrylAb toxin to the predator Chrysoperla carnea (Neuroptera: Chrysopidae). Environ Entomol 27(5):1255– 1263 Hilbeck A, Moar WJ, Pusztai-Carey M (1999) Prey-mediated effects of Cry1Ab toxin and protoxin and Cry2A protoxin on the predator Chrysoperla carnea. Entomol Experiment Appl 91(2):305– 316 Hoelmer K, Osborne L, Yokomi R (1990) Effects of neem extracts on beneficial insects in greenhouse culture. ARS US Department of Agriculture, Agricultural Research Service Aphidophagous Coccinellids. J Bio-pest 3:81–84 Iamba K, Yoba S (2019) Compatibility of predators and botanical extracts against Plutella xylostella on round cabbage. IOSR J Agric Vet Sci 12(11):64–68 Jacobsen SK, Klingen I, Eilenberg J, Markussen B, Sigsgaard L (2019) Entomopathogenic fungal conidia marginally affect the behavior of the predators Orius majusculus (Hemiptera: Anthocoridae) and Phytoseiulus persimilis (Acari: Phytoseiidae) foraging for healthy Tetranychus urticae (Acari: Tetranychidae). Exp Appl Acarol 79:299–307 Jones T, Scott-Dupree C, Harris R (2005) The efficacy of Spinosad against western flower thrips Frankliniella occidentalis, and its impact on associated biological control agents on greenhouse cucumbers in Southern Ontario. Pest Manag Sci 61:179–185 Kareem AA, Saxena RC, Malayba MT (1988) Effect of sequential neem (Azadirachta indica) treatment on green leafhopper (GLH), rice tungro virus (RTV) infection, predatory mirid and spiders in rice. Int Rice Res Newsl 13:37 Kim Y-J, Lee J-H (2018) Levels of Cry1Ac1 protein in herbivorous and predatory arthropods in fields of Bacillus thuringiensis cabbage. J Asia Pac Entomol 21(3):1048–1053 Kraiss H, Cullen EM (2008) Insect growth regulator effects of azadirachtin and neem oil on survivorship, development and fecundity of Aphis glycines (Homoptera: Aphididae) and its predator, Harmonia axyridis (Coleoptera: Coccinellidae). Pest Manag Sci 64(6):660–668 Kranthi KR, Stone GD (2020) Long-term impacts of Bt cotton in India. Nat Plants 6(3):188–196 Kunbhar S, Rajput LB, Gilal AA (2018) Impact of botanical pesticides against sucking insect pests and their insect predators in brinjal crop. J Entomol Zool Stud 6(2):83–87 Ladurner E, Benuzzi M, Franceschini S, Sterk G (2012) Efficacy of Beauveria bassiana strain ATCC 74040 against whiteflies on protected tomato and compatibility with Nesidiocoris tenuis and Orius laevigatus. IOBC/WPRS Bull 80:277–282 Li J, Kevin M (1999) Effects of recombinant baculoviruses on three nontarget heliothine predators. Biol Control 15(3):293–302 Lima APS, Santana EDR (2020) Insecticide activity of botanical compounds against Spodoptera frugiperda and selectivity to the predatory bug Podisus nigrispinus. Crop Prot 136:105230 Lowery DT, Isman MS (1995) Toxicity of neem to natural enemies of aphids. Phytoparasitica 23: 297–306 Lu Y, Kongming W (2012) Widespread adoption of Bt cotton and insecticide decrease promotes biocontrol services. Nature 487(7407):362–365

818

21 Biocompatibility of Biopesticides with Predatory Insects

Ludwig SW, Oetting RD (2001) Susceptibility of natural enemies to infection by Beauveria bassiana and impact of insecticides on Ipheseius degenerans (Acari: Phytoseiidae). J Agric Urban Entomol 18(3):168–178 Magalhães GO, Vacari AM, Laurentis VL et al (2015) Interactions of Bacillus thuringiensis bioinsecticides and the predatory stink bug Podisus nigrispinus to control Plutella xylostella. J Appl Entomol 139(1–2):123–133 Mahmoud T (2010) Effect of pirimor on some biological processes of Coccinella septempunctata L. with special care to the integument and fat bodies as deffensive barrier. Mesopotamia J Agric 38:19–27 Medina P, Budia F, Tirry L et al (2001) Compatibility of Spinosad, Tebufenozide, Azadirachtin with eggs and pupae of the predator Chrysoperla carnea (Stephens) under laboratory conditions. Biocontrol Sci Tech 11:597–610 Medina P, Budia F, Del Estal P (2003) Effects of three modern insecticides, pyriproxyfen, spinosad and tebufenozide, on survival and reproduction of Chrysoperla carnea adults. Ann Appl Biol 142(1):55–61 Medina P, Budia F, Del Estal P et al (2004) Influence of azadirachtin, a botanical insecticide, on Chrysoperla carnea (Stephens) reproduction: toxicity and ultrastructural approach. J Econ Entomol 97(1):43–50 Meena A, Sharma RK, Chander S (2020) Efficacy of biopesticides against major arthropod pests and their natural enemies on okra. Indian J Entomol 82(1):189–194 Messelink GJ, Janssen A (2014) Increased control of thrips and aphids in greenhouses with two species of generalist predatory bugs involved in intraguild predation. Biol Control 79:1–7 Messelink GJ, Bloemhard CMJ, Vellekoop R (2011) Biological control of aphids by the predatory midge Aphidoletes aphidimyza in the presence of intraguild predatory bugs and thrips. Acta Hortic 915:171–177 Messelink GJ, Bloemhard CMJ, Sabelis MW, Janssen A (2013) Biological control of aphids in the presence of thrips and their enemies. BioControl 58:45–55 Mingotti Dias P, de Souza LE, Amorim Pessoa LG (2020) Selectivity of entomopathogenic fungi to Chrysoperla externa (Neuroptera: Chrysopidae). Insects 11(10):716 Mollah MI, Rahman M, Alam Z (2013) Effect of insecticides on lady bird beetle (Coleoptera: Coccinellidae) in country bean field. Middle-East J Sci Res 17:1607–1610 Moussa S, Baiomy F, Abouzaid K (2018) Potential impact of host pest fed on Bt-modified corn on the development of Chrysoperla carnea (Stephens) (Neuroptera: Chrysopidae). Egypt J Biol Pest Contr 28(1):1–6 Navasero MV, Candano RN, Hautea DM, Hautea et al (2016) Assessing potential impact of Bt eggplants on non-target arthropods in the Philippines. PLoS One 11(10):e0165190 Nawrocka B (2008) The influence of spinosad and azadirachtin on beneficial fauna naturally occurring on cabbage crops. Veg Crops Res Bull 69:115–124 Papadimitriou DM, Petrakis EA, Arvaniti KA et al (2019) Comparative bioactivity of essential oils from two Mentha pulegium (Lamiaceae) chemotypes against Aphis gossypii, Aphis spiraecola, Tetranychus urticae and the generalist predator Nesidiocoris tenuis. Phytoparasitica 47:683– 692 Passos LC, Ricupero M, Gugliuzzo A, Soares MA, Desneux N, Campolo O, Zappalá L (2022) Sublethal effects of plant essential oils toward the zoophytophagous mirid Nesidiocoris tenuis. J Pest Sci 95:1609–1619 Pezzini DT, Koch RL (2015) Compatibility of flonicamid and a formulated mixture of pyrethrins and azadirachtin with predators for soybean aphid (Hemiptera: Aphididae) management. Biocontrol Sci Tech 25(9):1024–1035 Pourian HR, Talaei-Hassanloui R, Kosari AA, Ashouri A (2011) Effects of Metarhizium anisopliae on searching, feeding and predation by Orius albidipennis (Hem., Anthocoridae) on Thrips tabaci (Thy., Thripidae) larvae. Biocontrol Sci Tech 21:15–21 de la Poza M, Pons X et al (2005) Impact of farm-scale Bt maize on abundance of predatory arthropods in Spain. Crop Prot 24(7):677–684

References

819

Qi B, Gordon G, Gimme W (2001) Effects of neem-fed prey on the predacious insects Harmonia conformis (Boisduval) (Coleoptera: Coccinellidae) and Mallada signatus (Schneider) (Neuroptera: Chrysopidae). Biol Control 22(2):185–190 Raudonis L, Duchovskiene L, Valiuskaite A et al (2010) Toxicity of biopesticides to green apple aphid, predatory insects and mite in an apple-tree orchard. Zemdirbyste 97:49–54 Razak TA, Sivasubramanian P (2007) Effect of three botanical oils on carbohydrate content in Cheilomenes sexmaculata Fabricius and Chrysoperla carnea Stephens. Asian J Biochem 2 (2):124–129 Ricupero M, Biondi A, Cincotta F et al (2022) Bioactivity and physico-chemistry of garlic essential oil nanoemulsion in tomato. Entomol Generalis 42:921 Rocha-Munive MG, Soberón M, Castañeda S et al (2018) Evaluation of the impact of genetically modified cotton after 20 years of cultivation in Mexico. Front Bioeng Biotechnol 6:82 Sabry KH, El-Sayed AA (2011) Biosafety of a biopesticide and some pesticides used on cotton crop against green lacewing, Chrysoperla carnea (Stehens) (Neuroptera: Chrysopidae). J Biopest 4 (2):214 Sahayaraj K, Lalitha C, Balasubramaniam R et al (2008) Biosafety of Metarhizium anisopliae (Metschnikoff) Sorokin on a reduviid predator Acanthaspis pedestris (Hemiptera: Reduviidae). Hexapoda 15(1):46–48 Sahayaraj K, Borgio JAF, Kumar SM (2012) First record of Aspergillus flavus as a fungal pathogen of reduviid predator, Rhynocoris marginatus (Fab.) (Hemiptera: Reduviidae). Entomo Brasilis 5 (1):80–81 Sahayaraj K, Subash N, Allingham RW, Kumar V, Avery PB, Mehra LK et al (2018) Lethal and sublethal effects of three microbial biocontrol agents on Spodoptera litura and its natural predator Rhynocoris kumarii. Insects 9(3):101 Sajap AS, Kotuli JR, Kadir HA (1999) Impact of prey infected by nuclear polyhedrosis virus on a predator, Sycanus leucomesus Walk. (Hemiptera: Reduviidae). J Appl Entomol 123:93–97 Salgado VL (1998) Studies on the mode of action of Spinosad: insect symptoms and physiological correlates. Pestic Biochem Physiol 60:91–102 Sampson C, Boullenger A, Puerto Garcia F, Hernandez Parra R (2011) Implementing integrated pest management programmes in protected strawberry crops across Europe. IOBC/WPRS Bull 70:171–180 Saunders DG, Bret BL (1997) Fate of spinosad in the environment. Down Earth 52(1):14–20 Schoonover JR, Larson LL (1994) Laboratory activity of spinosad on non-target beneficial arthropods. Arthropod Manag Tests 20(1):357 Shera PS, Karmakar P, Sharma S (2018) Impact of Bt cotton expressing single (Cry1Ac) and dual toxins (Cry1Ac and Cry2Ab) on the fitness of the predator Chrysoperla zastrowi sillemi (EsbenPetersen): prey-mediated tri-trophic analysis. Egypt J Biol Pest Contr 28(1):98 Silva RS, Adriano C (2016) Toxicity of botanical insecticides on Diaphania hyalinata, their selectivity for the predatory ant Paratrechina sp., and their potential phytotoxicity on pumpkin. Int J Pest Manag 62(2):95–104 Silva WM, Martínez LC et al (2020) Respiration, predatory behavior and prey consumption by Podisus nigrispinus (Heteroptera: Pentatomidae) nymphs exposed to some insecticides. Chemosphere 261:127720 Singh JP, Jaiswal AK, Monobrullah M (2014) Impact of some selected insecticides and bio-pesticides on incidence of predators, parasitoid and productivity of lac insect, Kerria lacca (Kerr). Indian J Agric Sci 84(1):64–72 Smith SF, Krischik VA (2000) Effects of biorational pesticides on four coccinellid species (Coleoptera: Coccinellidae) having potential as biological control agents in interiorscapes. J Econ Entomol 93:732–736 Smith CR, Heinz KM, Sansone CG et al (2000) Impact of recombinant baculovirus applications on target heliothines and nontarget predators in cotton. Biol Control 19(3):201–214 Soares MA, Campos MR, Passos LC et al (2019) Botanical insecticide and natural enemies: a potential combination for pest management against Tuta absoluta. J Pest Sci 92:1433–1443

820

21 Biocompatibility of Biopesticides with Predatory Insects

Sorensson A, Nedstam B (1993) Effect of Amblysieus cucumeris and Orius insidiosus on Frankliniella occidentalis in ornamentals. Bulletin IOBC/WPRS 16:129–132 Sparks C, Thompson D, Kirst A et al (1998) Biological activity of spinosyns, new fermentation derived insect control agents, on tobacco budworm (Lepidoptera: Noctuidae) larvae. J Econ Entomol 91:1277–1283 Symondson WOC, Sunderland KD, Greenstone MH (2002) Can generalist predators be effective biocontrol agents. Annu Rev Entomol 47:561–594 Tedeschi R, Alma A et al (2001) Side-effects of three neem (Azadirachta indica A. Juss) products on the predator Macrolophus caliginosus Wagner (Het., Miridae). J Appl Entomol 125(7):397– 402 Thomazoni D, Soria MF, Degrande PE et al (2013) Arthropods biodiversity index in bollgard cotton (Cry1Ac) in Brazil. Interciencia 38:849–856 Tillman PG, Mulrooney JE (2000) Effect of selected insecticides on the natural enemies Coleomegilla maculata and Hippodamia convergens (Coleoptera: Coccinellidae), Geocoris punctipes (Hemiptera: Lygaeidae), and Bracon mellitor, Cardiochiles nigriceps, and Cotesia marginiventris (Hymenoptera: Braconidae) in cotton. J Econ Entomol 93(6):1638–1643 Torres JB, Ruberson JR (2005) Canopy-and ground-dwelling predatory arthropods in commercial Bt and non-Bt cotton fields: patterns and mechanisms. Environ Entomol 34(5):1242–1256 Tunca H, Doğru A, Köçkar F et al (2020) Evaluation of Azadirachtin on Arthrospira plantensis Gomont growth parameters and antioxidant enzymes. Ann Limnol Int J Limnol 56:8 Tunca H, Hödük K, Köçkar F et al (2021) Effects of two synthetic pyrethroids on Arthrospira platensis Gomont growth and antioxidant parameters. Acta Botanica Croatica 80(2):117–124 Ullah MI, Altaf N et al (2019) Effects of Entomopathogenic Fungi on the Biology of Spodoptera litura (Lepidoptera: Noctuidae) and its Reduviid Predator, Rhynocoris marginatus (Heteroptera: Reduviidae). Int J Insect Sci 11:1179543319867116 Van De Veire M, Klein M, Tirry L (2002) Residual activity of abamectin and spinosad against the predatory bug Orius laevigatus. Phytoparasitica 30(5):525–528 Vasanthakumar D, Babu A, Shanmugapriyan R et al (2013) Impact of Azter (azadirachtin 0.15% EC), a neem-based pesticide, against tea red spider mite, Oligonychus coffeae Neitner (Acarina: Tetranychidae), and its natural enemies. Int J Acarol 39(2):140–145 Vinuela E, Adán A, Smagghe G et al (2000) Laboratory effects of ingestion of azadirachtin by two pests (Ceratitis capitata and Spodoptera exigua) and three natural enemies (Chrysoperla carnea, Opius concolor and Podisus maculiventris). Biocontrol Sci Tech 10(2):165–177 Wang Q, Ju X, Huang J (2014) Predation functional responses of Coccinella septempunctata Linnaeus to Toxoptera aurantii (Boyer) and security evaluation of several botanical biopesticides. Acta Agric Univ Jiangxiensis 36(6):1247–1252 Xu HX, Zheng XS, Yang YJ (2015) Methyl eugenol bioactivities as a new potential botanical insecticide against major insect pests and their natural enemies on rice (Oriza sativa). Crop Prot 72:144–149 Yang B, Ouyang F, Parajulee M et al (2015) Crop dominance exerts specific effects on foliage dwelling arthropods in Bacillus thuringiensis cotton. Agric For Entomol 17:225–238 Zhang GF, Wan FH, Murphy ST et al (2008) Reproductive biology of two nontarget insect species, Aphis gossypii (Homoptera: Aphididae) and Orius sauteri (Hemiptera: Anthocoridae), on Bt and non-Bt cotton cultivars. Environ Entomol 37(4):1035–1042 Zhao Y, Zhang S, Luo JY et al (2016) Bt proteins Cry1Ah and Cry2Ab do not affect cotton aphid Aphis gossypii and ladybeetle Propylea japonica. Sci Rep 6(1):1–10

Index

A Abaris basistriata, 30 Abax ater, 450 Abdomen, 74, 75, 77–79, 81, 82, 84–86, 88–90, 92, 118, 129, 151, 152, 159, 167, 174, 191–194, 211, 212, 222, 223, 231, 232, 234, 242, 257, 308, 311, 314, 319, 325, 339, 357, 360, 363, 367–369, 392, 400, 401, 432, 548, 549 Acanthaspis, 102, 338, 339, 400 Acanthaspis cincticrus, 400, 424 Acanthaspis pedestris, 51, 52, 101–103, 358, 360, 400, 401, 593, 594, 614, 617, 618, 634, 635, 674, 676, 747, 749, 789 Acanthaspis petax, 488 Acanthaspis quinquespinosa, 52, 101, 102, 104, 488 Acanthaspis rama, 108 Acanthaspis sexguttata, 108 Acanthaspis siva, 51, 102, 104, 108, 231, 360 Acanthaspis tergemina, 108 Acanthaspis trimaculata, 108 Acanthaspis vitticollis, 488 Acanthoscelides obtectus, 481 Acanthspis pedestris, 488 Acclimatization, 276 Acetamiprid, 618, 741, 744, 745, 747, 751, 752, 763, 764, 768 Acetic acid, 14, 126, 539, 541, 545 Acetylcholinesterase, 804 Achaea janata, 344, 494, 495, 597, 634, 680 Acherontia styx, 198 Acheta domestica, 451 Acheta domesticus, 427, 435, 437, 466, 481, 581 Acid phosphatase, 422, 423, 428 Acisomapa norpoides, 37 Acorus calamus, 804

Acroster numhilare, 281 Acupalpus meridianus, 29 Acyrthosiphon cyparissiane, 454 Acyrthosiphon kondoii, 317, 478 Acyrthosiphon pisum, 106, 132, 154, 187, 203, 260, 274, 278, 317, 344, 454, 455, 463, 464, 478, 483, 486, 563, 651, 751 Adalia bipunctata, 27, 31, 32, 43, 106, 141, 173, 225, 261, 283, 322, 335–338, 371, 372, 452, 611, 612, 652, 705, 713, 718, 755, 765 Adalia decempunctata, 31, 106, 141, 173, 188, 189, 338, 452 Adalia fasciatopunctata revelieri, 452 Adalia tetraspilota, 7, 694 Adelges tsugae, 563, 642–644, 658, 666, 713 Adelgidae, 642, 685 Adenanthera pavonina, 48 Adonia variegate, 31, 397 Adonis vernalis, 386, 389 Adoxo phyesorana granulovirus (AoGV), 788 Adult, 6, 30, 72, 98, 116, 150, 184, 222, 252, 300, 357, 379, 410, 460, 529, 581, 629, 666, 706, 737, 786 Adult longevity, 159, 164, 177, 186, 191, 198, 201, 202, 204, 210, 211, 232, 253, 255, 261, 368, 401, 537, 561, 598, 601, 659, 745, 762, 765–768, 789, 800 Aedia leucomelas, 466 Aegle marmelos, 48 Aeolothripidae, 3, 6, 31, 82, 124, 174, 670 Aeolothrips intermedius, 716, 723 Aero-micropylar, 120 Aeschnidae, 3 Aeshnidae, 37 Aethriamanta brevipennis, 37 Agallia constricta, 591 Agar, 206, 555, 566, 567, 600

# The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2023 K. Sahayaraj, E. Hassan, Worldwide Predatory Insects in Agroecosystems, https://doi.org/10.1007/978-981-99-1000-7

821

822 Age age-specific fecundities, 189, 264 age-specific life expectancy, 189 age-specific net fecundity, 189 age-specific reproductive value, 189 age-specific survival rates, 167, 189, 202 age-stage specific fecundities, 189 Ageneotettix deorum, 470 Aggressive behaviour, 240 Aggressive predator, 151 Agonoscena pistaciae, 334 Agonum dorsalis, 27, 306, 450 Agonum mülleri, 29 Agriocnemis femina, 28, 31, 78, 674 Agriocnemis lacteola, 38 Agriocnemis pygmaea, 28, 31, 34, 37 Agriosphodrus dohrni, 130, 323 Agroforests, 39 Agromyzid fly, 6 Agropyron desertorum, 611 Agropyron repens, 14 Agrotis ipsilon, 449 Ahasverus advena, 481, 491 Ailanthus, 49, 53 Aiolocaria hexaspilota, 7 Alabama argillacea, 466, 477, 774 Alate, 101–104, 108, 109, 230, 231, 235, 498 Alate forms, 108 Alcaeorrhynchus, 156 Alcohol, 2, 14, 341, 401 Aldehydes, 341, 384, 432 Aleochara bilineata, 28, 44, 346, 448, 717, 794, 796 Aleurodothrips fasciapennis, 124, 716, 723 Aleurolobus barodensis, 669 Aleyrodidae, 265–266, 273, 315, 585, 667, 684, 685 Alfalfa, 2, 10, 26, 28, 32, 42–51, 54, 85, 134, 185–187, 207, 338, 344, 397 Alfalfa weevil larva and caterpillar, 495 Aliphatic esters, 384 Aliphatic glucosinolates, 387, 389 Alkaline phosphatase, 422, 423 Alkaloids, 185, 343, 345, 385, 389, 411, 425, 427, 434, 764, 802, 804 Alkanes, 227, 372, 384 Allantus rufocinctus, 157, 158 Allelochemicals, 125, 190, 385, 614, 806 Allaeocranum biannulipes, 203, 280, 488, 599, 609–611 Alleocranum quadrisignatus, 159, 160 Allicin, 804 Allium sativum, 57, 804, 810

Index Alloeocranum biannulipes, 203, 280, 488, 599, 609–611 Alloeorhynchus reinhardi, 98, 233 Alloeorhynchus sp., 479 Allograpta exotica, 167, 193, 471 Allograpta fly, 193 Allograpta obliqua, 193, 194, 470, 687 Aloconota gregagia, 28 Alphitobius diaperinus, 50, 322, 489 Alphitobius venator, 50 Alsophila pometaria, 316 Alternative food, 637, 654 Amara aenea, 27, 450 Amara anthobia, 385 Amara eurynota, 27 Amara familiarisa, 27 Amara fulva, 44 Amaranthus retroflexus, 14 Amara ovate, 27 Amauromorpha accepta, 29 Amblyseius cucumeris, 649 Amblyseius swirskii, 632, 678, 792 Ambrosia artemisiifolia, 18 Ambush bugs, 87, 314, 477 Ametabolous, 71 Amino acids, 229, 421, 427, 429, 527, 529, 540, 544, 552, 565, 568 Aminopeptidase N, 428 Amischa analis, 28 Amischa bifoveotata, 28 Amphiareus constrictus, 165, 166, 742, 809 Amphibolus venator, 50, 481, 488, 609, 678 Amphorophora rubi, 457 Ampulex compressa, 116, 426 Amrasca biguttula, 670, 799 Amritodus atkinsoni, 490 Amsacta albistriga, 476, 493 Amylase, 277, 422, 423 Amyotea malabaricus, 476 Anacardium occidentale, 200, 596 Anagasta kuehniella, 50, 137, 206, 270, 480, 488–490, 497 Anagrus atomus, 718 Anagyrus dactylopii, 718 Anagyrus fusciventris, 718 Anagyrus pseudococci, 721 Anaphes iole, 346, 718 Anasa tristis, 486 Anastrepha fraterculus, 466 Anastrepha ludens, 466 Anastrepha suspensa, 466, 473 Anax guttatus, 37 Anaximma culifrons, 37

Index Anax imperator, 171 Anaxipha longipennis, 28, 31, 460 Anax strenuus, 78 Anchomenus dorsalis, 259–260, 385, 397 Andrallus spinidens, 157, 423, 478, 616, 617, 640, 769 Anegleis cardoni, 32 Anisochrysa basalis, 465 Anisodactylus merula, 689 Anisodactylus sanctaecrucis, 450 Anisolabella marginalis, 475 Anisolabididae, 3, 16, 105, 171, 251, 252, 307, 473, 797 Anisolabis maritima, 105, 395 Anisoptera, 37, 79, 235, 382 Ankylopteryx octopunctata, 669 Annonaceae, 804 Annona squamosa, 803, 804 Annonin, 803, 804 Anomis flava, 492–494, 634 Anoplolepis longipes, 102 Anotilus inustus, 449 Anotylus rugosus, 29 Ant assemblages, 96 Antennae, 74, 75, 77, 80, 81, 84–86, 90–93, 158, 167, 174, 175, 192, 193, 212, 227, 243, 305, 309–311, 318–320, 322, 339, 344, 359–362, 368, 371, 373, 543, 547– 549 Antennifer, 159 Anthecorid, 48, 633 Antheraea paphia, 532 Anthicidae, 3, 16, 41, 459, 669 Anthicus cervinus, 19 Anthicus ruficallis, 669 Anthicus unicolor, 459 Anthocorid, 2, 9, 13, 33, 48, 49, 85, 86, 104, 105, 124, 128, 343, 532, 533, 585, 631, 633, 677, 679, 706, 713, 714 Anthocoridae, 3, 27, 84, 98, 128, 164, 235, 256, 321, 364, 385, 480, 530, 581, 630, 669, 710, 741, 789 Anthocoris confusus, 49, 134, 321 Anthocoris flavipes, 49 Anthocoris minki, 48, 49, 256, 364 Anthocoris muraleedharani, 49, 584, 585 Anthocoris nemoralis, 47, 104, 105, 134, 321, 364, 365, 385, 389, 480, 677–679, 705, 741 Anthonomus grandis, 264, 317, 425, 473, 478 Anthonomus grandis grandis, 171 Anticarsia gemmatalis, 425, 428, 437, 463, 476, 477, 479, 484, 486, 738, 791

823 Anticarsia irrorata, 477 Anti-collision behaviour, 378 Antifeedant, 14, 158, 764, 804, 805 Antilochus conqueberti, 12 Antioxidant, 176, 213, 275–277, 689, 802, 814 Anti-predatory acts, 339–340 Ant-mimicking, 117 Ants, 3, 30, 74, 96, 124, 154, 223 Aonidiella aurantii, 345, 464, 686 Aonidiella orientalis, 463 Apateticus, 156 Aphaenogaster obsidiana, 467 Aphelinus abdominalis, 719 Aphelinus asychis, 719 Aphelinus mali, 40, 718 Aphelinus varipes, 718 Aphididae, 27, 155, 166, 173, 176, 189, 192, 261, 334, 344, 372, 562, 653, 685 Aphidius colemani, 719 Aphidius ervi, 719, 766 Aphidius gifuensis, 719 Aphidius matricariae, 721 Aphidius transcaspinus, 719 Aphidius urticae, 718 Aphid lions, 5 Aphidoletes aphidimyza, 27, 38, 129, 131, 138, 139, 335, 336, 344, 470, 643, 644, 652, 678, 705, 706, 709, 713, 714, 722 Aphidophagous predator, 26, 188, 324, 335, 652 Aphids, 2, 29, 85, 96, 117, 150, 184, 228, 251, 303, 366, 380, 453, 532, 590, 627, 666, 705, 736, 787 Aphid wolves, 5 Aphis acetosae, 453, 456 Aphis affinis, 452–457 Aphis craccivora, 135, 136, 152, 153, 165–169, 184–186, 194, 207, 211, 290, 291, 453, 454, 456, 464–466, 471, 472, 474, 492, 494, 566, 594, 596, 601, 602, 673, 674, 676, 687, 694 Aphis euphorbiae, 453, 454, 457 Aphis fabae, 132, 169, 173, 188, 189, 210, 211, 259, 333, 335, 452–457, 463–465, 471, 472, 562, 605, 606, 681, 686 Aphis fabae subsp. cirsiiacanthoidis, 453 Aphis fabae subsp. mordvilkoi, 457 Aphis frangulae, 453–455 Aphis glycines, 27, 454, 455, 470, 643–645, 773, 807, 811 Aphis gossypii, 135–137, 139, 152, 155, 166, 173, 184, 185, 187–190, 211, 261, 267, 269, 274, 282, 289, 290, 326, 328, 330,

824 334, 336, 344, 449, 453–455, 457, 458, 464, 465, 471, 482, 486, 493, 494, 497, 534, 562, 594, 595, 600, 602, 604, 635, 644, 645, 651, 670, 675, 678, 686, 690, 745 Aphis grossulariae, 472 Aphis illinoisensis, 449 Aphis maidis, 458 Aphis nerii, 452–454, 456, 458, 464, 486 Aphis pomi, 169, 452, 456, 457, 464, 473, 694 Aphis punicae, 210, 211, 452, 457, 465 Aphis ruborum, 453, 457 Aphis sambuci, 457, 471 Aphis spiraecola, 170, 366, 464 Aphytis diaspidis, 718 Apiomerus crassipes, 122, 489 Apiomerus duckei, 489 Apiomerus luctuosus, 489 Apiomerus pilipes, 489 Apis mellifera, 489, 798 Apis mellifera pupae, 567 Aposematic, 72, 236, 371, 385, 390 Apple, 2, 17, 27, 28, 39–41, 44, 45, 55, 101, 124, 133, 196, 287, 316, 346, 450–452, 455, 458, 471, 474, 476, 478, 481, 484, 486, 497, 603, 652, 681, 694, 695, 736, 755, 812 maggot, 450–452 orchards, 17, 27, 28, 55, 101, 196, 287, 472, 476, 478, 484, 486, 497, 681, 694, 736 Approaching time, 229, 257, 542, 544, 545, 547–551 Apricot trees, 17 Apristus latens, 689 Aproaerema modicella, 493 Apterous, 75, 87, 108, 109, 222, 228, 235, 242 Aptery, 222 Aquatic, 76, 96–98, 170, 365, 381, 629, 736 Arabidopsis thaliana, 387 Arachis hypogaea, 14, 128 Arachnids, 10, 29, 315, 689 Arachnocoris albomaculatus, 98, 154 Araecerus fasciculatus, 491 Araneids, 689, 693 Archimantis latistyla, 12, 35 Arge ochropus, 157 Arge pagana, 157, 158 Argidae, 157, 158 Arilus cristatus, 54, 489, 634, 650, 657 Arilus gallus, 122, 489 Arion lusitanicus, 451 Armyworms, 4, 88, 138, 153, 154, 201, 208, 263, 307, 329, 590, 599, 753, 787

Index Arnulphus, 46 Arousal, 257, 309, 311, 314, 318, 358, 360– 363, 547 Artemia franciscana, 532, 560, 671 Artemia urmiana, 174, 190 Arthropod predators, 20, 29, 95, 98, 106, 627, 641, 642, 644, 689, 756, 758 Artificial diet holidic, 527 liquid, 538, 541, 557 meridic, 527, 528, 531 natural diet, 557, 559, 561 oligidic diets, 529, 531, 532, 538, 540, 541, 545–547, 549–551, 673 semiliquid, 538 solid diet, 528, 557 Ascalaphidae, 3, 27, 120, 151, 433 Ascalaphids, 152 Ascalobyas, 152 Ascaloptynx, 152 Ascia monuste orseis, 307, 474 Ascorbic acid, 539, 540 Asiagomphus pryeri, 171 Asian lady beetle, 477, 499, 559 Asopinae, 98, 156, 157, 234, 236, 318, 340, 363, 402, 530 Asparagus beetle, 89, 713, 803 Aspartic proteinases, 421 Aspergillus niger, 439 Aspidiotus nerii, 463, 564 Aspirator, 2, 7–9, 16, 18, 19 Assassin bugs, 4, 12, 29, 86, 87, 89, 101, 104, 107, 200, 201, 230, 242, 312, 315, 323, 339, 400, 410, 411, 413, 414, 422, 429, 433, 435, 436, 440, 498, 587, 593, 597, 673, 675, 747 Astata occidentalis, 309, 344 Asterolecaniidae, 685 Athalia lugens proxima, 476 Atheta coriaria, 448, 450, 705, 709, 713, 714 Atheta (=Xenota) mucronata, 449 Atopozelus opsimus, 394, 489 Atractomorpha crenulata, 202, 204, 492–494, 674, 676, 694 Atropine, 805 Attack, 2, 29, 86, 98, 156, 188, 234, 251, 300, 366, 378, 415, 533, 580, 641, 665, 749, 793 Attack rates (a), 251, 259, 274, 580, 586, 600, 615, 647 Attack ratio (y/x), 596, 598 Auditory organs, 76, 77 Augmentation, 210, 666, 716

Index Aulacaspis citri, 462, 463 Aulacaspis tegalensis, 669, 686 Austrosimulium immatures, 470 A-vac beating tray, 8 Axinidris acholli, 468 Axinoscymnus cardilobus, 265, 266, 273 Azadirachta indica, 764, 804, 806, 813 Azadirachtin, 742, 743, 751, 757, 764, 771, 773, 804, 806–811 Azalea lace bug, 653 Azalea plant bug, 653 B Bacillus firmus, 787 Bacillus thuringiensis (Bt), 33, 57, 175, 688, 689, 739, 755, 786, 797 Bacillus thuringiensis israelensis, 787 Bacillus thuringiensis kurstaki, 787, 789 Bacillus thuringiensis sphaericus, 433, 787 Bacillus thuringiensis tenebrionis, 755, 787 Bacterial symbionts, 433 Bactericera cockerelli, 483, 638–640, 671 Bagrada hilaris, 317 Baited pitfall traps, 2 Banana, 34, 36, 40, 41, 43, 567, 674, 675, 708 Bathycoelia thalassina, 490 Bauhinia purpurea, 49 Bean, 6, 50, 99, 274, 315, 333, 389, 583, 612, 613, 655, 812 flower thrips, 583 leaf beetle, 803 pods, 532 Beat bucket, 19 Beating tray, 2, 7, 8, 10–11, 18 Beat sampling, 8, 17 Beauveria bassiana, 278, 786–789, 792, 793, 797 Beauveria brongniartii, 787 Beef liver, 539, 540, 544, 545, 568–571 Beet armyworm, 138, 153, 154, 208, 329, 590 Beetle, 3, 27, 72, 99, 117, 157, 184, 223, 259, 301, 370, 380, 421, 477, 559, 587, 629, 680, 704, 736, 788 Beet leafhopper, 481 Belenois solilucis, 466 Belonochus rufipennis, 448 Bembidion femoratum, 28, 44 Bembidion guttula, 27, 450 Bembidion lampros, 27, 43, 450 Bembidion obtusum, 27 Bembidion properans, 44 Bembidion quadrimaculatum, 109

825 Bemisia argentifolii, 645, 646, 648 Bemisia tabaci, 131, 205, 206, 265, 273, 282, 326, 336, 454, 458, 464, 483–485, 487, 489, 584, 585, 590, 592, 601, 602, 605, 607, 615, 638–640, 647, 648, 667, 672, 675, 679, 760, 761, 768, 799, 800 Bendhi, 49 Berlese funnel, 2 Berothidae, 151, 432, 465 Berry borer, 604, 787 Berytidae, 2, 3, 84 β-caryophyllene, 373 Beta carotene, 567 Big-eyed bug, 8, 19, 84, 124, 153, 176, 207, 209, 268, 326, 329, 364, 688 Bigger is better, 150 Bilia castanea, 480 Bio-intensive integrated pest management (BIPM), 1, 291, 693–695, 786, 802, 803, 806, 815 Biological control agents (BCAs), 18, 27, 46, 54, 92, 97, 106, 110, 124, 135, 150, 169, 188, 199, 203, 252, 254, 263, 333, 410, 525, 526, 530, 551, 558, 566, 580, 583, 591, 592, 595, 597, 600, 606, 614, 615, 629, 632, 643, 644, 658, 665, 667, 668, 674, 677, 683, 686, 736, 739, 751, 756, 757, 769, 774, 807, 809, 815 Biology, 113–142, 149–177, 183–214, 261– 275, 355–373 Biprorulus bibax, 281, 493 Black ant, 304 Black earwigs, 3 Blackfly, 470, 629, 630 Black scale, 603 Blaptostethus kumbi, 49 Blaptostethus pallescens, 2, 49, 137, 206, 242, 480, 532, 584, 585, 633, 634, 742, 809 Blatta orientalis, 317 Blattids, 103, 109 Bledius filipes, 31 Blow fly, 498 Blueberry maggot, 452 Blue grama, 132 Bombyx mori, 166, 487, 539, 543, 798 Bombyx mori pupae, 567 Borniochrysa squamosus, 462 Botanicals anti-feeding, 807 chemistry, 125 oviposition, 803–805 phytochemistry, xxi Botanochara sedecimpustulata, 157, 234

826 Botrytis cinerea, 439 Bowl traps, 7, 11 Brachiaria decumbens, 16 Brachinus longipalpis, 12 Brachinus sclopeta, 385 Brachinus sexmaculeatus, 12 Brachycaudus cardui, 453, 456 Brachycaudus divaricatae, 452 Brachycaudus schwartzi, 452 Brachydiplax sobrina, 37 Brachygastra lecheguana, 466 Brachymyrmex incisus, 467 Brachyptery, 89, 223, 227, 233 Brachythemis contaminate, 31, 37, 38 Bradley, 460 Bradycellus harpalinusa, 27 Bradysia coprophila, 448 Brassica chinensis, 55 Brassica napus, 125, 689, 692 Brassica oleracea, 55, 346, 641 Brassica rapa, 636, 654, 680 Brassica rapa oleifera, 131 Brassivola hystrix, 121 Brevicoryne brassicae, 136, 137, 139, 152, 154, 168, 169, 188, 194, 211, 273, 387, 389, 448, 454, 471, 472, 486, 601–603, 694, 811 Bromus inermis, 14 Bromus japonicus, 14 Brontispa longissima, 473, 687 Brontocoris, 156, 417 Brontocoris tabidus, 417, 423, 476, 719 Brooding, 130, 394 Brown lacewing, 3, 209, 556, 657, 695, 714 Brown plant hopper (BPH), 213, 592, 652 Brumoides suturalis, 19, 42, 559, 693, 813 Brumus suturalis, 32 Brutfürsorge, 396 Bryozoa, 98, 151 Bt crops Bt corn, 33, 56, 57, 799 Bt cotton, 19, 56, 176, 214, 326–332, 688, 689, 800–802 Bt maize, 33, 56, 328–330, 332, 799, 801 Bt rice, 327 Bt toxins Bt (Cry1Ab), 33, 57, 326–330, 690, 795 Bt (Cry1Ac), 56, 176, 213, 214, 325, 326, 328–332, 558, 559, 688–691, 693, 798– 801 Bt (Cry3), 691 Bt (Cry3A), 691 Bt (Cry3Aa), 691

Index Bt (Cry3Bb1), 690 Bt (Cy1Ac), 690 Bt (VIP3A + Cry1Ab), 690 Buccula, 4, 87 Bufadienolides, 388 Buffalo grass, 132 Burkholderia spp., 787 Butea monosperma, 48 B vitamins, 557 C Cabbage, 28, 38, 42, 44, 49, 51, 55, 90, 133, 138, 139, 195, 286, 306, 343, 389, 450, 471, 472, 475, 486, 554, 601, 603, 604, 618, 641, 642, 644, 646, 648, 667, 694, 695, 757, 774, 802, 811 aphid, 139, 601, 603, 646, 694, 695, 811 root fly, 450 Cacopsylla (Psylla) pyri, 480, 677–679 Cacopsylla pyricola, 788 Cadra cautella, 608 Caesalpinia pulcherrima, 48 Cage experiment, 636, 640, 652 Calamol, 804 Calathus fuscipes, 385 Calathus melanocephalus, 28, 44 Calciferol (D), 527 Cales noacki, 718 Caliothrips phaseoli, 137, 474 Calleida decora, 650, 657 Callosobruchus chinensis, 467 Callosobruchus maculatus, 481 Calocoris anguslatus, 486 Calocoris norvegicus, 317, 478 Calopterygidae, 171 Calopteryx maculate, 61 Calopteryx splendens, 19 Calosoma calidum, 14, 450 Calosoma granulatum, 30 Calosoma sayi, 689 Calosoma sycophanta, 450, 569 Calosoma vagans, 18 Calvia punctata, 7 Camellia sinensis, 128, 595 Camouflage decamouflaging, 401 Campaniform, 87 Campodeiform larva, 151 Camponotine ants, 103 Camponotus compressus, 30, 40, 101, 104 Camponotus crassus, 467 Camponotus festinatus, 378

Index Camponotus melanoticus, 467 Camponotus novogranadensis, 467 Camponotus oasium, 467 Camponotus pilicornis, 467 Camponotus rufipes, 467, 469 Camponotus sericeus, 30 Camponotus xerxes, 467 Campylomma nicolasi, 484 Campyloneuropsis infumatus, 485, 552, 638, 639 Cannabis sativa, 14 Cannibalism, 73, 99, 106–108, 139–141, 199, 239, 240, 270, 314, 323, 337, 357, 359, 361–363, 380, 393, 394, 526, 536, 558, 559, 563, 652, 655, 712, 715, 772 Canola, 55 Canthecona bhoutanica, 381 Canthecona furcellata, 11, 381 Capsaicin, 804 Capsicum annuum, 53, 256, 631, 804 Capture, 9, 10, 12, 14, 17–19, 73, 75, 87, 135, 199, 264, 301, 302, 305, 306, 309, 311, 312, 318, 321, 322, 325, 340, 378, 379, 399, 400, 413, 416, 424, 433, 435, 440, 547, 580, 596, 610, 612, 691, 693, 766, 803 Capture/attack rate, 580 Carabidae, 3, 5, 8, 10, 12–14, 16, 26–28, 30–32, 41, 43, 44, 55, 58, 72, 99–101, 107, 286, 305, 380, 396, 397, 414, 415, 433–435, 450–452, 499, 569, 656, 669, 690, 736, 754, 772, 773, 801 Carabus meander, 14 Carabus nemoralis, 452 Carabus violaceous, 27 Carbohydrates, 44, 155, 276, 307, 397, 428, 430, 437, 527, 528, 535, 540, 812 Carbon and nitrogen isotopes, isotopes C13, 107 Carcinops pumilio, 569, 717 Cardamom, 49, 51 Cardenolides, 386, 388, 389 Cardiac glycosides, 384 Cardiastethus, 32, 48, 86 Cardiastethus affinis, 48, 585 Cardiastethus exiguus, 86, 480, 677, 706 Cardiocladius africanus, 470 Cardiocladius australiensis, 470 Cardiocladius oliffi, 470 Cardiocondyla elegans, 467 Cardiocondyla stambouloffi, 467 Carpelimus corticinus, 29 Carpophilus dimidiatus, 481

827 Carpophilus hemipterus, 448 Carpophilus mutilatus, 491 Carterus calydonius, 396 Cashew, 29, 48, 191, 200, 312, 495, 496, 595, 596 Casnoidea indica, 28, 31 Casnoidea ishii ishii, 28 Cassia javanica, 48 Cassida rubiginosa, 466, 618 Cassida vittata, 449 Cassidine beetles, 157, 234 Cataglyphis albicans var. auratus, 467 Cataglyphis auratus, 467 Cataglyphis lividus, 467 Cataglyphis nodus var. drusa, 467 Cataglyphis semitonsus, 467 Catamiarus brevipennis, 51, 52, 411, 416, 417, 422, 489, 634, 635 Catana parcesetosa, 669 Caterpillars, 30, 74, 88, 92, 93, 100, 102, 174, 201, 254, 282, 301, 303, 310, 312, 318, 320, 323, 331, 336, 339, 384, 385, 388, 399, 428, 466, 474, 477, 495, 499, 534, 588, 589, 595, 640, 646–648, 677, 711– 715, 738, 757, 787, 796, 803, 811 Cathepsin B, 420 Catopsilia pyranthe, 476, 477 Caura rufiventris, 491 Cavariella, 470, 471 Cecidomyiidae, 3, 13, 27, 32, 38, 131, 138, 335, 344, 470, 667, 678, 706, 709 Cecidomyiid midge, 3 Celastraceae, 804 Celastrus angulatus, 804 Cellular differentiation, 421 Cellular respiration, 805 Cephalonomia stephanoderis, 719 Ceraeochrysa, 152, 225 Ceraeochrysa cincta, 31, 44, 762, 763 Ceraeochrysa claveri, 31, 45, 462 Ceraeochrysa cubana, 31, 45, 385, 389, 462 Ceraeochrysa elegans, 44 Ceraeochrysa everes, 45 Ceraeochrysa smithi, 432 Ceraeochrys avalida, 44 Cerasus avium, 17, 41 Ceratagallia agricola, 479, 591 Ceratitis capitata, 449, 451, 452, 466, 473 Ceratovacuna lanigera, 28, 669, 683 Cerci, 75, 76, 81–83 Ceriagrioncerino rubellum, 37 Ceriagrioncoro mandelianum, 37 Cermatulus nasalis, 281

828 Ceroplastes sinensis, 463 Cerotainiops abdominalis, 470 Cevadine, 805 Chaetoleon pusillus, 80 Chaetopsis massyla, 597 Chaitophorus leucomelas, 457 Chamaemelum mixtum, 630 Chauliognathus marginatus, 489 Cheilomenes lunata, 9, 43 Cheilomenes propinqua, 459 Cheilomenes sexmaculata, 19, 32, 43, 173, 185, 186, 225, 226, 238, 284, 285, 287, 372, 454, 458, 693, 706, 798, 812 Cheilomenes undecimpunctata, 186 Chelisoches morio, 171, 195, 196, 369, 473, 475, 687 Chelisochidae, 3, 16, 196, 369, 473, 475, 669, 687 Chemical control, 677 Chemical ecology, 341–346, 372–373 Chemical pesticides, 656, 665, 689, 745, 758, 785, 786, 806, 813 Chemosterilant, 804 Cherry orchards, 41 Chewing, 72, 74–76, 108, 402, 787, 803 Chicken liver (Cl), 560–562, 568, 569, 571 Chillies, 49, 53 Chilli thrips, 632 Chilochorus bipustulatus, 32 Chilocorus baileyi, 717 Chilocorus bipustulatus, 31, 453, 458, 564, 685, 717, 759 Chilocorus cacti, 41, 685 Chilocorus circumdatus, 17, 669, 717 Chilocorus kuwanae, 41, 454, 685 Chilocorus nigritus, 41, 173, 454, 669, 685, 717 Chilocorus rubidus, 17 Chilomenus propinquaisis, 189, 190, 565 Chilo partellus, 49 Chilo suppressalis, 138, 213, 316, 460, 461, 616, 769 Chinch bugs, 479, 713 Chironomus thummi, 470 Chlaenius nigricans, 12 Chlaenius pictus, 12 Chlaenius velutinus, 385 Chloral hydrate, 14 Chlorobalius leucoviridis, 309, 461 Chlorochroa ligata, 497 Chlorochroa sayi, 495 Chlorogenic acid, 385, 389 Chlorogomphidae, 171 Chloropidae, 3 Chloropid flies, 3

Index Chlorpyrifos, 616, 617, 751–753, 755, 768, 769, 794–797, 809, 814 Chlosynelacinia saundersii, 466 Chlysopa carnea, 556 Chlysopa formosa, 556 Chlysopa perla, 556 Choice test, 268, 545, 619 Cholesterol, 541, 545 Choline, 557 Chorionated follicles, 553 Choristoneura houstonana, 497 Choristoneura rosaceana, 602, 603 Chromaphis juglandicola, 456 Chromobacterium subtsugae, 787 Chrotogonus trachypterus, 492, 674, 676, 694 Chrysanthemum, 630 Chrysanthemum cinerariifolium, 804 Chrysodeixis chalcites, 477, 588, 589 Chrysomelidae, 18, 98, 157, 234, 236, 398, 459, 605, 636, 654, 656, 680, 690, 772 Chrysomphalus aonidum, 462–464 Chrysomya albiceps, 470 Chrysopa carnea, 45, 465, 645, 646, 648–651, 657, 714 Chrysopa claveri, 462 Chrysopa cubana, 462 Chrysopa formosa, 283, 462 Chrysopa lacciperda, 465, 747 Chrysopa nigricornis, 44, 346, 385, 389, 462, 646, 648 Chrysopa oculata, 344, 346, 462, 464, 581 Chrysopa pallens, 211, 342, 464, 558, 559, 602–604, 684, 690, 768 Chrysopa phyllochroma, 126, 127 Chrysopa rufilabris, 465, 650, 657 Chrysopa sanchezi, 462 Chrysopa scelestes, 462 Chrysopa vulgaris aeqyptiaca, 464 Chrysoperla (=Chrysopa) carnea, 224 Chrysoperla sp., 28, 32, 126, 152, 209, 214, 224, 556–558, 653, 681, 703, 704, 712– 715, 722, 800, 813, 815 C. bicarnea, 462 C. carnea, 28, 97, 126, 153, 213, 224, 286, 328, 386, 463, 556, 600, 643, 667, 704, 737, 796 C. comanche, 45, 657, 682, 695 C. congrua, 9, 46 C. downesi, 556, 683 C. exotera, 44 C. externa, 44, 45, 224, 463, 556, 602, 604, 722, 763, 764, 792, 793 C. harrisii, 463 C. lucasina, 133 C. mutata, 133

Index C. nipponensis, 463, 464, 556, 600–602, 757, 758, 767, 769, 770 C. orestes, 463 C. pallid, 133 C. plorabunda, 26, 45, 224, 225, 346, 464, 556 C. pudica, 464 C. rufilabris, 19, 44, 126, 127, 133, 379, 380, 464, 465, 528, 556, 571, 645, 648, 650, 653, 657, 677, 682, 683, 695, 704, 711, 713, 722, 760, 798, 811 C. sinica, 214, 556, 693 C. zastrowi arabica, 152 C. zastrowi sillemi, 120, 135, 152, 153, 211, 464, 647, 648, 706, 799, 800 Chrysopid, 58, 151, 308, 337, 646, 694, 706, 794 Chrysopidae, 5, 27, 81, 96, 135, 151, 209, 286, 328, 355, 380, 432, 463, 556, 600, 646, 666, 711, 751, 792 Chrysopodes collaris, 464 Chrysoteuchia topiaria, 449 Chrysotoxum bactrianum, 39 Chysoperla carnea, 133, 135 Cicada prey, 461 Cicindela flexuosa, 12 Cicindela melancholia, 28 Cicindela undulate, 28 Cicindelines, 689 Cinara cedri, 455 Cinara cedri Mimeur, 456 Cinerin, 804 Circalifer tenellus, 497 Citrus, 6, 26, 38, 40, 44, 49, 53, 165, 195, 274, 284, 345, 366, 446, 629, 630, 656, 657, 666, 686, 708–712, 761, 810 Citrus orchard, 44, 449, 452, 467, 473, 676, 677, 683 Cladomorphus phyllinus, 402, 477, 490 Claw, 92, 151, 382, 383 Climate change, 95, 250, 253, 254, 288, 290, 291 Clitostethus arcuatus, 717 Clitostethus oculatus, 41, 684, 685 Clivina fossor, 27, 30 Clomiphene citrate, 554 Clothianidin, 618 Clusters, 98, 114, 117, 120–124, 130, 133, 161–163, 194, 385, 424, 562 Cluysopa septempunctata, 556 Clypeus, 234 Cnaphalocrocis medinalis, 138, 308, 316, 460 Coccidae, 603, 685 Coccidencyrtus ochraceipes, 718 Coccidoxenoides perminutus, 720

829 Coccinella maculata, 13 Coccinella octomaculata, 693 Coccinella quinquepunctata, 32, 101, 458 Coccinella septempunctata, 7, 13, 19, 27, 28, 32, 41–43, 54, 57, 106, 134, 137, 173, 184, 186, 187, 259, 283, 284, 286, 287, 326, 337, 346, 372, 390, 397, 453, 454, 458, 560, 562, 566, 567, 604, 605, 617, 651, 656, 685, 694, 695, 717, 754–756, 807, 812, 813 Coccinella sexmaculata, 807 Coccinella transversalis, 10, 17, 28, 30, 32, 42, 108, 184, 454, 669, 670, 693 Coccinella transversoguttata, 26, 134, 396 Coccinella undeceimpunctata, 184 Coccinellidae, 5, 26, 72, 96, 117, 173, 225, 260, 309, 370, 380, 452, 559, 585, 642, 666, 705, 736, 789 Coccinellinae, 5, 9, 16, 72, 379 Coccinula quatuordecimpustulata, 32, 454 Coccivora, 46 Coccoidea, 685 Coccophagus cowperi, 718 Coccus hesperidum, 464 Cochliomyia macellaria, 470 Cocinella septempunctata, 458 Cockroaches, 198, 317, 426, 427 Coco, 49, 52, 484 Coconut, 26, 38, 48, 49, 52, 195, 197, 473, 475, 480, 491, 673–675, 677 Coconut palms, 195 Cocoon, 81, 116, 174, 212 Codling moth, 316, 450–452, 788 Coenagrionidae, 3, 12, 27, 31, 37, 240, 498 Coenosia attenuata, 722 Coleomegilla maculata, 26, 43, 55, 57, 127, 213, 283, 309, 337, 379, 397, 499, 559, 560, 565, 614, 642, 644, 650, 651, 654, 655, 690, 693, 789 Coleomegilla maculatalengi, 126 Coleomegilla maculate, 42, 605, 644, 737 Coleoptera, 3, 25, 72, 106, 117, 157, 185, 223, 260, 304, 370, 396, 410, 448, 530, 585, 636, 666, 708, 756, 786 Coleoptery, 223 Collection methods, vii, 1–20, 415–416 Collection net/bag, 13 Collops quadrimaculatus, 581 Collops vittatus, 478, 481, 486, 495, 497 Colorado potato beetle, 89, 117, 397, 398, 479, 587, 605, 636, 658, 680, 692, 711, 713, 787, 803 Commercial fields, 529 Comperiella bifasciata, 718 Competition

830 Competition (cont.) dominates, 332–333 Compound eyes, 75, 77–81, 91, 93, 159, 301 Compsobiella, 46 Congregate, 365, 396 Coniopterygid, 150, 151, 321 Coniopterygidae, 3, 26, 44, 46, 151, 152 Coniopterygid larvae, 44, 150, 321 Coniopteryx josephus, 46 Coniopteryx tineiformis, 722 Connexivum, 89, 90, 159, 229, 319 Conocephalus, 12, 461 Conocephalus longipennis, 28, 213, 461 Conocephalus maculatus, 31, 391, 461 Conservation, 7, 39, 54, 71, 95, 98, 99, 338, 396, 559, 586, 607, 653, 668, 716, 742, 745, 787 Consistency, 411, 528, 538–540, 543, 568 Consumption (Ne), 580 Contact poison continuous method, 417, 418 discontinuous method, 416, 417 electric stimulation, 415, 417, 418 Conventional cotton, 328, 693 Conventional crops, 54–55 Conwentzia psociformis, 722 Coperam arginipes, 34 Copestylum caudatum, 472 Coptopteryx viridis, 116 Coptosoma crubrarria, 486 Coptotermes heimi, 488 Copulation, 108, 175, 189, 192, 284, 285, 355, 358–365, 367–370 Coranus C. aegpus, 490 C. africana, 489, 675 C. fuscipennis, 595, 599 C. nodulosus, 51, 101, 159 C. obscurus, 121 C. siva, 489 C. spiniscutis, 12, 52, 53, 121, 159, 163, 312, 358, 359, 490 C. vitellinus, 159, 164, 312, 358 Corcyra cephalonica, 49, 135–137, 152, 153, 159, 161, 162, 164, 165, 198–202, 204, 211, 418, 464, 465, 488–491, 493–497, 500, 525, 530, 532, 546, 548–551, 595, 598, 599, 601, 706 Cordalia obscura, 449 Cordulegastridae, 171 Cordyline australis, 132 Corium, 22, 90, 92 Corn

Index earworm, 88, 154, 208, 590, 788, 803 leaf aphid, 185, 187, 654 meal, 798 Correlation, 54, 106, 285, 287, 370, 385, 544, 596, 655 Cosmoclopius nigroannulatus, 490 Cosmolestes picticeps, 10, 52 Cosmolestes sp., 19, 490 Cosmopolites sordidus, 473 Cost-effective, 321, 531 Cotton, 2, 28, 85, 101, 131, 153, 185, 264, 310, 382, 465, 528, 581, 628, 665, 706, 741, 798 aphid, 155, 185, 190, 282, 289, 327, 330, 456, 534, 595, 600, 659, 675, 686 boll weevil, 171, 264, 473 Courtship, 156, 355–357, 363, 365, 368, 370– 372, 384, 386 Cowpea, 49, 52, 136, 168, 213, 287, 290, 389, 554, 583, 687, 693 Cowpea trypsin inhibitor (CpTI), 213, 691 Coxa, 73 Crematogaster C. antaris, 467 C. punctulata, 648, 649 C. scutellaris, 469 C. subdentata, 467 C. warburgi, 467 Crepuscular, 101–104, 209 Crickets, 76, 77, 176, 192, 213, 427, 435, 803 Crocidolomia pavonana, 123, 161, 201, 202, 495 Crocothemis servilia, 28, 37 Cropping seasons, 638, 689, 801 Crosopedophorus elegans, 11, 12 Crowding, 136, 401, 526, 559 Cry proteins Cry1Ab, 33, 56, 326–331, 690, 795 Cry1Ac, 56, 176, 213, 214, 325, 326, 328– 332, 558, 559, 688–691, 693, 798–801 Cry2Aa, 34 Cry2Ab, 56, 213, 214, 558, 559, 691, 693, 798–800 Cry3A, 691 Cryptognatha nodiceps, 685 Cryptolaemus, 713 Cryptolaemus montrouzieri, 32, 40, 273, 370, 454, 560, 563, 684, 685, 705, 706, 708, 713, 714, 718, 747, 759, 760, 772, 789 Cryptolestes pusillus, 80, 481 Cryptophlebia leucotreta, 492, 788 Crytolestes pusillus, 80, 481 Cucujoidea, 72

Index Cucumber beetle, 803 leaf folder, 596 Cucumis sativus, 6 Cultivars, 27, 33, 190, 326, 586, 604, 614, 616, 632, 689, 800 Cunctochrysa baetica, 133 Curculionidae, 171, 264, 317, 425, 691 Curinus coeruleus, 41, 379, 380, 685, 706, 772 Cursorial legs, 75 Cyanocitta cristatata, 388 Cybocephalus aegyptiacus, 459 Cybocephalus micans, 459, 604 Cybocephalus nigriceps, 459 Cybocephalus nipponicus, 717 Cybocephalus pullus, 459 Cycloneda munda, 26, 43 Cycloneda sanguinea limbifer, 454 Cydia pomonella, 788, 794 Cydia pomonella granulovirus (CpGV), 788, 794 Cydnocoris gilvus, 29, 52, 122, 159, 197, 312, 358, 360, 490, 594, 596 Cypermethrin, 614, 615, 617, 689, 747, 749, 750, 752, 754, 756, 766, 789, 796, 797 Cyrtacanthacris ranacea, 311 Cyrtopeltis (=Nesidiocoris) tenuis, 482 Cyrtopeltis (=Engyatus) varians, 485 Cyrtorhinus lividipennis, 28, 31, 46, 138, 205, 253, 316, 592, 652, 658, 659, 748, 808, 812 Cystatin, 420 Cysteine, 421, 423 Cysteine protease inhibitor, 689 Cytochrome P450, 213, 424 D Dachylopius nipae, 486 Dacnusa sibirica, 720 Dacus curcurbitae, 475 Dalbulus maidis, 474, 497 Dalotia coriaria, 448, 713 Dalotia (Atheta) coriaria, 709, 713, 717 Damsel bug, 8, 20, 33, 88, 154, 155, 233, 264– 265, 317, 801 Damsel flies, 3, 19, 28, 31, 49, 78, 79, 235, 240, 380, 477, 498, 674 Danaus plexippu, 388, 644, 645 Dance flies, 3, 17, 302, 325, 399 Date palm bostrichid, 459 Datura stramonium, 805 Deacetylsalannin, 804

831 Debitterized annona oil, 804 Defense, 710 Defoliation, 636, 680, 810 Delonix regia, 48 Delphastus catalinae, 454, 605, 607, 705, 708, 713, 714, 718, 761, 763, 813 Delphastus pallidus, 605, 607 Delphastus pusillus, 708, 718 Demetrias atricapillusa, 27 Demographic parameters doubling time (in days), 187, 205, 210, 255, 261, 553 finite rate of increase (λ), 165, 202, 261, 266 generation time, 150, 151, 188, 197, 202, 205, 210, 255, 261, 266, 284, 533, 534, 553 gross reproductive rate (GRR), 205, 210, 534 innate capacity for increase (rc), 210 net reproductive rate (NRR) R0, 187, 196, 202, 205, 207, 283, 534, 768 Deraeocoris lutescens, 482 Deraeocoris nebulosus, 158, 205, 316, 424, 435 Dermaptera, 3, 4, 8, 12–14, 16, 17, 26–29, 31, 32, 39, 41, 72, 74–75, 105–106, 171– 173, 195–196, 251–252, 262–264, 307– 309, 317, 334, 368–370, 395–397, 432, 473, 669, 687, 752, 764, 794, 797, 814 Dermapterans, 473, 687, 689, 693, 752–754 Deroceras reticulatum, 451, 689 Derris chinensis, 803, 805 Derris elliptica, 803, 805 Detoxication, 213 Diabolocatantops pinguis, 202–204, 493 Diacrisia obliqua, 494 Diamondback moth, 154, 533, 604, 618 Diapause syndrome, 208 Diaphania indicus, 594, 596 Diaphorina citri, 656, 657, 687, 695 Diaplacodes trivialis, 37 Diaspididae, 685, 686 Diaspidiotus perniciosus, 463, 686 Diatraea saccharalis, 307, 463, 466, 473–475 Diazinon, 616, 617, 754, 768 Dichochrysa alcestes, 210 Dichochrysa flavifrons, 210 Dichochrysa formosana, 210 Dichochrysa prasina, 210, 464 Dichochrysa tacta, 210, 465 Dichochrysa ussuriensis, 210 Dichochrysa zelleri, 210 Dicranolaius bellulus, 10

832 Dictyophara europea, 481 Dictyopterans, 115 Dicyphus bolivari, 252, 253, 277, 637 Dicyphus cerastii, 107, 336, 337, 592 Dicyphus eckerleini, 252, 253, 277 Dicyphus errans, 107, 138, 253, 277, 482, 552, 671, 718 Dicyphus flavoviridis, 252, 253, 277 Dicyphus hesperus, 253, 482, 637–639, 671, 679, 714, 718, 720 Dicyphus maroccanus, 158, 205, 482, 485 Dicyphus tamaninii, 480, 482, 484, 552, 569, 570, 671 Dielocroce chobauti, 98 Diet composition parafilm, 528, 530, 535, 561, 569–571 Diet ingredients dried aphids, 560, 561 eggs diet, 530 egg yolk, 531, 538, 539, 541, 544, 558, 560, 566, 569–571 elements, 536 frozen aphids, 560, 561, 566 fructose, 554, 557 glycerine, 533, 565 holidic diet, 526–529, 537 honey, 189, 210, 282, 531, 536, 539, 541, 544, 554, 557, 563, 566, 751 inositol, 557 insect whole body, 539 iron-coated brine shrimp eggs, 534 KCl, 539 vitamins, 539, 540 water, 541, 545, 557, 567 Wesson’s salt, 539 yeast, 531, 533, 540, 541, 557, 565, 566 Diet promotion, 538 Dietrick vacuum insect net, 12, 13 Digestion, 213, 315, 331, 421, 422, 424, 435, 739 Digger wasp, 3, 116 Digging, 76 Digitalis purpurea, 386, 389 Diglyphus begini, 720 Diglyphus isaea, 720 Dilaridae, 79, 321 Dinoderus porcellus, 203, 481, 488, 599, 609– 611 Dinotefuran, 618, 743, 744, 752, 768 Dinothenarus badipes, 448 Diomus hennesseyi, 41, 685 Diomus pumilio, 657, 695

Index Diomus sp., 717 Dipha aphidovora, 28 Diplacodes lefebvrii, 37 Diplocheila polita, 12 Diplocodes trivialis, 31, 37 Diptera, 3, 4, 7, 8, 12, 13, 16, 17, 25–29, 31, 32, 38, 39, 72, 78, 93, 97, 106, 116, 129, 131–132, 138, 166, 167, 197, 201, 227, 237, 302, 324–325, 335, 343, 344, 372, 399, 410, 411, 414, 425, 431–432, 466, 470, 497, 531, 555, 589, 594, 597, 667, 669, 670, 678, 687, 706, 716–722, 759, 764, 794 Discovery time, 587 Distantiella theobroma, 482, 490 Diuraphis noxia, 455, 459, 611, 647, 648 Diurnal, 11, 109, 342 DNA, 107, 229, 238, 337, 338, 395, 688, 800, 809 Dolichopodidae, 3, 28, 29, 38, 39, 498 Dome-shaped curve, 615, 749 Dortus primarius, 482, 592, 659 Doru lineare, 4, 195, 307, 395, 474, 475 Doru luteipes, 172, 307, 474, 752 Doru taeniatum, 263, 307, 432, 474, 794, 797, 814 Dorylus labiatus, 469 Dorymyrmex brunneus, 468 Dotriacontane, 344 Dragon-fly, 3, 31, 78, 79, 477, 498 Drepanosiphum, 471 Dried fish, 189, 190 Dropping, 141, 379, 548, 581 Drosicha stebbingi, 462 Drosophila melanogaster, 186 Dung beetle, 12, 395 Dungfly, 12, 395 Dustywings, 3, 16 D-vac UC-vac methods, 8, 18 D-VAC vacumn method, 8, 13, 18 Dysaphis plantaginea, 27, 452, 455, 456, 474, 481, 652 Dyschirius globulosus, 14 Dysdercus D. cingulatus, 257, 258, 344, 476, 489, 490, 492–495, 546–551, 594, 595, 598, 620, 634, 635, 672, 673 D. koenigii, 202, 318–320, 476, 488, 493, 499, 500, 594, 598 D. laetus, 488 D. superstitiosus, 233, 491

Index E (E)-2-Alkenals, 384 Earias biplaga, 492 Earias inornatus, 496 Earias insulana, 197, 492, 530 Earias vitella, 136, 490, 530 Earwigs, 3, 4, 8, 10, 13, 14, 29, 33, 39, 74, 75, 105, 263, 264, 287, 307, 308, 395–397, 432, 653, 752, 753, 794, 795, 797, 814 Easchistus, 497 (E)-β-farnesene (EβF), 125–127, 386, 554 Echinothrips americanus, 256, 481, 482 Echis carinatus, 103, 104 Eclosion, 115, 168–170, 195, 238, 266, 279, 284, 312, 327, 567 Ecology, vii, 17, 26, 33, 95–110, 184, 234, 237, 264, 275, 341–345, 372–373, 772 Ecosystem services, 26, 655 Ectatomma brunneum, 468, 469 Ectomocoris E. cardiger, 102 E. tibialis, 51, 101, 103, 129, 230, 231, 358, 490, 634, 635, 673 E. ululans, 12 E. xavierei, 101, 158 Ectrychotes pilicornis, 129 Edocla slateri, 101, 163 Egg aggregation, 140 attendance, 396 cluster, 130 dumping, 141 egg-case, 115, 132 egg-laying, 114, 119, 120, 125–141, 184, 196, 198, 203, 209, 213, 238, 264, 275, 356, 398, 534, 561, 562 egg-watching, 396 mortality, 136, 284, 393, 737 of nitidulids, 450 plant, 44, 481 UV-irradiated eggs, 532 Eicosane, 344 Elaeidobius kamerunicus, 304, 312, 466 Elasmopalpos lignosellus, 466 Elatophilus, 46, 105, 236, 267 Elatophilus nigricornis, 105, 267 Elattoneura nigerrima, 34 Elaunon bipartitus, 12 Elytra, 73, 184, 223, 226, 238, 259, 306, 370, 564 Emergence, 116, 117, 119, 158, 159, 161, 163, 167, 184, 187, 189, 200, 208, 224, 229, 241, 261, 265, 275, 281, 283, 289, 322,

833 360, 387, 393, 529, 540, 555, 557, 562, 565, 566, 647, 765, 767, 773, 786, 800, 814 Emesinae, 12, 86, 87, 121, 122, 228, 314, 338, 340 Empoasca kerri, 458 Empusidae, 32, 34 Enallagma cyathigerum, 34 Encarsia citrina, 718 Encarsia guadeloupae, 718 Encarsia hispida, 718 Encarsia protransvena, 718 Encarsia tricolor, 719 Encyrtus infelix, 719 Encyrtus lecaniorum, 719 Endochus africanus, 490 Endochus albomaculatus, 30, 137, 163, 199, 313, 496 Endochus cinqalensis, 121 Endochus inornatus, 490, 706 Endochus umbrinus, 159, 160 Endomychidae, 72 Endopterygote, 79 Engytatus nicotianae, 483, 639, 640 Engytatus varians, 316, 484, 485, 552, 638– 640, 671, 672 Ensiform, 76 Entatomidae, 3, 12, 32, 46–50, 90–91, 98, 99, 124, 138, 156–158, 234–236, 254–255, 280, 288, 301, 309, 318, 329, 340, 343, 362–364, 381, 384, 387, 388, 397, 402, 423, 425, 428, 435, 501, 552, 570, 571, 581, 587, 589, 616, 617, 636, 640, 654, 669, 670, 679, 680, 711, 738, 769, 774, 791–792, 810 Entomopathogens, 175, 786, 789, 790, 793 Entomophagbous insects, 2 Entomosuccivorous, 101–104 Enzyme catalase, 276 cytolysis protease, 420 enzymatic casein hydrolysate, 557 enzymatic soya hydrolysate, 557 esterase, 213, 278, 421–423 esterase (EST) detoxification, 213, 332, 386 glutathione S-transferase (GST), 213, 276, 277 histidine phosphatase, 420 inositol-phosphate phosphatase, 420 kinase, 420 phospholipase, 421–423 phospholipase A, 428, 433

834 Eocanthecona furcellata, 12, 32, 157, 254, 255, 381, 476, 589, 669, 670, 770 Eocanthecona robusta, 609, 670 (E)-2-octenyl acetate, 373, 384 Ephestia cautella, 50, 481, 489, 534 Ephestia kuehniella, 155, 165, 173, 174, 188– 190, 205, 207, 210, 259, 268, 269, 282, 336, 452, 455, 456, 463, 465, 476, 478, 482–485, 487, 530, 534, 535, 560, 562, 565, 566, 592, 668, 671, 769 Ephestiasula pictipes, 191 Epicauta maclini, 670 Epicauta waterhousei, 670 Epicuticular waxes, 133, 383–384 Epidaus bicolor, 29, 51, 137, 163, 199, 313, 496, 594, 595 Epilachna varivestis, 317, 477, 478, 588, 589 Epilachninae, 5, 72 Epimeron, 91 Episcopomantis sp., 9, 36 Episternum, 91 Episyrphus balteatus, 27, 38, 39, 97, 127, 131, 132, 137, 139, 140, 166, 168–170, 194, 195, 290, 338, 470–472, 554, 555, 652, 722 Episyrphus confrater, 170, 554 Epitheca princeps, 61 Eremochrysa punctinervis, 44 Eretmocerus corni, 720 Eretmocerus eremicus, 719, 720 Eretmocerus mundus, 720 Eretmocerus warrae, 720 Erigone atra, 653 Erigonidae, 653 Eriococcidae, 685 Eriopis connexa, 455, 605, 606, 756, 758, 763, 764 Eriosoma lanigerum, 316, 472, 474 Erythroneura elegantula, 681, 682 Erythroneura variabilis, 681, 682 Escape reactions, 321 Escherichia coli, 439 Essigella californica, 459 Ethylene glycol, 14 Euagoras plagiatus, 29, 137, 163, 199, 312, 361, 496 Euborellia annulata, 195, 473 Euborellia moesta, 473 Euborellia stali, 28 Eucalyptus benthamii, 484, 486 Eucolliuris fuscipennis, 31 Eugubinus annulatus, 490 Eulissus chalybaeus, 30

Index Eumeninae, 74 Eumerinae, 97 Euparagiinae, 74 Eupeodes americanus, 133, 193, 256, 470 Eupeodes bucculatus, 554, 555 Eupeodes corolla, 27, 39, 97, 166–169, 194, 195, 472 Eupeodes frequens, 137, 139, 169 Eupeodes lundbeckii, 27 Eupeodes luniger, 131 Eupeodes (=Eupeodes) nuba, 39 Eupeodes volucris, 193, 346, 470 Euprepocnemis alacris alacris, 311 Euproctis mollifera, 493 Eurema hecabe, 476 European corn borer, 33, 354, 803 European earwig, 3, 173, 196, 307, 317, 432, 753 Eurytopic, 100 Euschistus biformis, 384 Euschistus heros, 387–389 Eutectona machaeralis, 493, 594, 596 Euthyrhynchus, 156, 309, 322, 499 Eutolmus rufibarbis, 413, 431 Euwallacea interjectus, 473 Euxesta annonae, 597, 598 Euxesta eluta, 597, 598 Euxesta stigmatias, 597, 598 Exochomus laeviusculus, 717 Exochomus nigripennis, 273, 459 Exochomus nigromaculatus, 32, 455 Exochomus quadripustulatus, 397, 455, 563, 717 Extatosoma tiaratum, 392 Extension of rostrum, 199 External ovipositor, 77 Extraguild prey (EGP), 332, 333, 335, 629 F Facultative carnivorous, 2 Fagaceae, 379 Farmland, 39, 757 Fatty acid, 109, 190, 276, 341, 555, 813 Fecundity, 34, 120, 125, 128, 129, 135, 136, 141, 150, 153, 162, 183, 185, 186, 188– 191, 194, 198, 199, 202, 204, 206, 207, 209–211, 214, 224, 226, 229, 231, 238, 250, 255, 261–265, 268–270, 272, 273, 280, 282–285, 289, 327, 330, 332, 371, 530, 533–537, 541, 544, 553–556, 559, 560, 562, 567, 583, 587, 589, 590, 595, 598, 601, 693, 738, 742, 744, 746, 747,

Index 754–757, 760, 761, 763, 765, 768, 774, 795, 798, 800, 809 Feeding approach, 318, 548 deterrent, 791, 805 filter, 580 glass olfactometer, 545 group, 301, 304, 323, 347, 542 group ambushing, 322, 325 group hunting, 301, 322 post-feeding, 318, 549 potential, 206, 211, 533, 584 preference, 151, 264, 545, 592 Feltiella acarisuga, 705, 710, 714, 722 Feltiella acarisuga (=Therodiplosis persicae), 705, 710, 714, 722 Female emergence, 275 Femora, 76, 88, 89, 92, 158, 174, 175, 242, 370, 371 Femur, 76, 82, 84, 87, 90, 162, 194, 200, 212, 382, 581 Fenitrothion, 616, 617, 754, 768 Ferrisia dasyrilii, 260, 265 Ferrisia virgata, 49, 463 Fertilization, 114, 135, 355, 356, 656, 659, 678 Ficus benjamina, 632, 633 Ficus retusa, 49, 480 Field evaluations, 687 Field releases, 18, 279, 529, 530, 556, 667, 673, 678, 680 Filiform, 75, 76, 91, 227 Finite rate of increase, 165, 202, 261, 266 Fiorinia theae, 462–464 Flagellomeres, 87 Flea beetle, 387, 803 Flies, 3, 7, 11, 16, 17, 78, 80, 97, 177, 186, 192, 194, 195, 228, 302, 324, 325, 380, 399, 410, 411, 413, 414, 431, 433, 458, 465, 466, 554, 594, 597, 598, 654, 657, 693, 706, 709, 713, 714, 717, 787, 803 Flight intercepts traps, 2 Flooding of eggs, 136 Flour moth, 174, 671 Flower beetles, 3 Flying insects, 6, 8, 15, 16, 174, 325, 366, 380 Foliage, 1, 10, 26, 76, 128, 153, 163, 209, 239, 288, 317, 380, 675, 693, 759 Foliar damage, 603, 632 Food chain, 34, 288, 332, 798, 802 composition, 276, 527 consumption, 305, 534, 547, 552, 689, 692, 765, 800 lures, 2 Forage crops, 645

835 Foraging behaviour, 125, 172, 253, 259, 300, 303, 320, 328, 341, 586, 614, 766 Forests, 5, 10, 17, 30, 39, 61, 96, 100–104, 199, 236, 307, 810 Forewing, 74, 75, 77, 83, 84, 88, 92, 166, 209, 211, 212, 319 Forficula auricularia, 11, 17, 19, 28, 32, 41, 173, 196, 262, 287, 307, 308, 317, 334, 368, 397, 432, 434, 466, 473, 687, 753 Forficula lurida, 17, 19, 41 Forficula pubescens, 11, 474 Forficula smyrnensis, 17, 18, 41 Forficulidae, 3, 4, 29, 31, 32, 40, 41, 101, 173, 195, 196, 307, 334, 368, 395, 473–475, 687, 752, 794, 797 Forficulina Arixeniina, 75 Forficulinae, 4 Formica fusca, 467, 468 Formica glauca, 467 Formica subpolita, 366 Formicidae, 3, 4, 13, 14, 18, 19, 26–28, 30, 31, 39, 40, 93, 303, 304, 317, 321, 343, 380, 466–469 Formicimus braminus, 669 Frankliniella F. occidentalis, 206, 329, 336, 344, 448, 464, 465, 472, 480–482, 487, 602–604, 613, 630–633, 649, 676–678, 684, 695, 743, 744, 792 F. schultzei, 484, 585, 592, 659 F. williamsi, 669 Franklinothrips, 6, 124 F. atlas, 124 F. basset, 124 F. brunneicornis, 124 F. fulgidus, 125 F. lineatus, 125 F. megalops, 716 F. megalops (=myrmicaeformis), 124, 716 F. orizabensis, 124 F. rarosae, 125 F. tenuicornis, 329, 330 F. variegatus, 124 F. vespiformis, 124, 174, 175, 211, 212, 667, 670, 716, 723 French beans, 40, 47, 582 Fringed wings, 6, 82 Fruit crops, 49, 52 Fruit flies, 186, 654 Functional response approaching time, 229, 542, 544, 545, 547– 551 handling times (Th), 259, 580, 585, 590, 592, 600, 616–618, 647, 659, 769 highest prey density (x), 581, 596, 597, 602

836 Functional response (cont.) Holling’s disc equation, 583, 585, 588, 595 maximum prey consumed, 600 predatory rate, 618 prey attack ration, 589 prey density (N0), 209, 580, 581, 601, 615 prey–predator ratio, 261, 284, 378, 388, 569, 582, 589, 596, 617, 641 resting time, 602 Fungicide, 617, 741, 754, 771–773, 775 Fungivores, 6 G Gabrius pennatus, 29 Galchana, 46 Galepsus lenticularis, 115, 116 Galerucella nymphaeae, 641, 644 Galerucella pusilla, 641 Galleria mellonella, 138, 191, 199, 398, 421, 423, 430, 437, 476, 496, 531–532, 599 Gall midges, 3, 643 Galosoma obsoletum, 14 Garlic, 55, 810 Gauropterus fulgidus, 449 General feeders, 6, 82 Generalist predator, 4, 5, 30, 34, 72, 134, 155, 157, 186, 206, 213, 288, 312, 328, 333, 336, 387, 402, 435, 552, 590, 591, 597, 608, 616, 635, 637, 648, 651, 656, 657, 659, 665, 678, 679, 684, 695, 705, 740, 757, 758, 792, 798, 801, 802, 807 Generalized predators, 1, 2, 397 Genetically engineered crops, vii, 57 Genitalia, 79, 85, 86, 93, 102, 114, 228, 229, 233, 358–360, 362, 367 Geocoris G. amabilis, 486 G. atricolor, 269, 335 G. bullatus, 479, 658 G. erythrocephalus, 154, 207, 209, 486 G. floridanu, 19, 153, 154, 207, 208, 486, 590, 801 G. ochropterus, 166, 486, 534 G. pallens, 8, 269, 346, 385, 389, 649, 650, 657 G. pallidipennis, 176, 209, 214, 282, 536, 688, 689 G. pollens, 207, 336 G. proteus, 269, 270, 487 G. punctipes, 8, 19, 20, 46, 138, 153, 176, 207, 268–270, 328–332, 336, 346, 373, 486, 487, 528, 535, 569, 581, 590, 591,

Index 648–651, 657, 670, 691, 722, 741, 742, 761, 789, 794, 798–801 G. punctipes sonoraensis, 486 G. superbus, 154, 207–209, 487 G. uliginosus, 19, 487, 650, 657 G. varius, 269, 270, 487, 535, 536, 740 Giant Asian Mantis, 132 Gitona perspicax, 669 GK-12 (Bt), 326, 799 Glands accessory, 82, 115, 356, 411–413, 422, 432, 434 anterior main, 422 Dufour’s glands, 413 labial, 413 maxillary, 414, 433 posterior main, 411, 412, 436 sex-accessory, 411, 412, 420–422, 424, 435, 436 thoracic, 413, 414, 431 venom, 411, 413, 414, 422, 433, 435, 440 Glandular trichomes, 383, 384, 614 Glucosinolates, 125, 386, 387, 389 Glycaspis brimblecombei, 489 Glycine javanica, 128 Glycine max, 18, 287, 346 Gomphidae, 27, 37, 171 Goniozus legneri, 720 Gonocoxae, 91 Gossyparia spuria, 273 Gossypol, 190 Grasping, 87, 89, 564 Grasshoppers, 12, 28, 76, 77, 102, 198, 243, 310, 311, 385, 391, 593, 673, 694, 787, 793 Grasslands, 5, 10, 58, 59, 100, 109, 399, 470 Grazed oil palm, 18 Greater wax moth, 599 Green coverworm, 479 Green foxtail, 132 Greenhouse, 2, 6, 101, 107, 130, 155, 158, 209, 211, 238, 290, 321, 333, 335, 480, 481, 530, 536, 588, 592, 605, 607, 612, 618, 630–632, 637–641, 649, 655, 657, 671, 672, 677–679, 681, 683, 689, 695, 704, 706, 708–713, 744, 757, 792, 812 Green pepper, 649 Gregarious behavior, 156 Gressitt design, 16 Gressitt trap, 15 Grooming, 87, 311, 322, 359, 396, 549 Ground beetle, 3, 5, 13, 28, 30, 44, 55, 61, 72, 73, 99, 107, 118, 396, 397, 433, 656

Index Ground-dwelling groundling dragonfly, 3 predator, 8, 689 Groundnut, 50, 101, 104, 135, 136, 157, 481, 492, 597, 673, 674, 676, 694, 706 Growth regulator, 805, 807 Gryllacridids, 649 Gryllidae, 26, 27, 54, 77, 176, 192, 460, 581 Grylloblatta campodeiformis, 309 Gryllotalpa gryllotalpa, 54 Gryllus campestris, 54 Gryllus desertus, 54 Gryllus pennsylvanicus, 460 Guaiacol, 126, 127 Guanchia hincksi, 17, 19, 41 Guionius nigripennis, 129 Gynaikothrips ficorum, 480, 679 Gynaikothrips uzeli, 480, 486, 632, 633 Gyranusoidea litura, 719 H Haematobia irritans, 449 Haematochares obscuripennis, 490 Haematorrhophus marginatus, 108 Haemocytes granulocyte (GC), 438 plasmatocyte (PL), 438, 439 Halmus chalybeus, 41, 685 Halotydeus destructor, 307, 473 Halyomorpha halys, 46, 306, 317, 344, 466, 478, 492, 495, 581 Halyziats chitscherini, 7 Hand picking, 8, 9, 16, 19 Haploglenius, 152 Haplothrips, 6 Haplothrips brevitubus, 6 Haplothrips subtilissimus, 6 Haplothrips victoriensis, 6 Harmonia axyridis, 11, 13, 19, 27, 32, 33, 41, 43, 117, 126, 127, 186, 187, 226, 238, 259–261, 274, 285, 289, 290, 337, 338, 346, 372, 373, 379, 380, 387, 396, 455, 474, 499, 562, 563, 569, 581, 582, 585, 605, 607, 642–646, 648, 654, 666, 685, 686, 691, 692, 717, 718, 761, 763, 772, 789, 802, 807 Harmonia conformis, 714, 807 Harmonia dimidiata, 42, 187, 188, 274, 370, 371 Harmonia octomaculata, 28, 30, 31, 455, 670, 714 Harmonia oxivridis, 565

837 Harmonia quadripunctata, 32, 455 Harpactor angulosus, 242, 490 Harpactor costalis, 490 Harpactorinae, 12, 87, 102, 103, 108, 120, 122, 197–199, 242, 301, 309, 310, 313–315, 323, 338, 340, 358, 359, 362, 363, 394, 411, 434, 587, 596 Harpactorine, 102, 103, 122, 129, 198, 199, 309, 313, 323, 340, 358, 359, 394, 411, 434, 587 Harpactor nilqiriensis, 121 Harpactor pyqmaeus, 121 Harpalus aeneus, 452 Harpalus affinis, 259, 260, 451 Harpalus distinguendus, 259, 452 Harpalus gravis, 689 Harpalus pennsylvanicus, 452, 651, 689 Harpalus rufipes, 27, 28, 44, 109, 306, 451 Hatchability, 136, 185, 199, 204, 214, 255, 271, 280, 285, 537, 595, 745, 755 Head, 4, 16, 72, 75, 76, 78, 80, 81, 83–87, 89– 91, 93, 116, 129, 137, 159, 166, 167, 174, 191, 207, 209, 212, 234, 236–238, 242, 243, 313, 314, 319, 320, 322, 340, 357, 359, 360, 366, 370, 378, 381, 394, 401, 413, 414, 431, 548 Head roll, 322 Hedgerows, 30, 59, 396 Hediocoris tibialis, 497 Helicoverpa armigera, 48, 136, 157, 176, 187, 208, 209, 214, 269, 288, 326, 328, 332, 343, 344, 418, 430, 438, 439, 451, 463, 465, 476, 477, 483, 486–495, 500, 530, 536, 600–603, 634, 635, 669, 673–676, 688, 694, 773, 787, 788, 807 Helicoverpa armigera nucleopolyhedrovirus (HearNPV), 788, 798 Helicoverpa punctigera, 788 Heliothis virescens, 465, 486, 593, 594, 645, 646, 648, 650, 788 Heliothis zea, 153, 154, 208, 465, 479, 486, 590, 645, 648, 742, 788 Heliothrips haemorrhoidalis, 6 Helopeltis antonii, 199, 313, 490, 496, 599 Helopeltis theivora, 594, 595 Hemelytra, 84, 85, 88, 89, 91, 92 Hemerobiid, 151 Hemerobiidae, 3, 32, 44, 46, 101, 209, 225, 433, 465, 556, 657, 667, 683, 695, 712, 773 Hemerobiiformia, 79 Hemerobius bolivari, 465 Hemiberlesia lataniae, 48, 464

838 Hemileuca oliviae, 648, 649 Hemimerina, 75 Hemimetabolous, 71, 75 Hemiptera, 2, 26, 72, 105, 116, 154, 189, 233, 252, 302, 363, 380, 411, 497, 532, 581, 636, 667, 706, 742, 789 Hemlock, 563, 644, 666 Hemlock woolly adelgid, 642, 643, 658, 713 Heptacosane, 344 Herbicides, 55, 736, 753, 754, 758 Herbivore-induced plant volatiles (HIPVs), 125, 341–344, 372, 386 Herbivorous, 4, 41, 54, 141, 150, 176, 278, 284, 290, 317, 330–332, 641, 666, 688, 695, 798 Heringia calcarata, 133, 472 Hermetia illucens, 531 Heteracris, 466 Heteromurus nitidus, 450, 451 Heteroptera, 2, 25, 83, 98, 128, 157, 222, 253, 318, 362, 384, 410, 476, 530, 585, 630, 677, 722, 736, 792 Hexacosane, 126, 127 Hibernation, 105, 225, 226, 254, 267, 380, 396–398 Hierodula membranacea, 35, 36, 132 Himacerus apterus, 98 Hind legs, 73, 76, 77, 82, 129, 311, 319, 325, 359, 360, 370, 402 Hippodamia axyridis, 455 Hippodamia convergens, 19, 41, 42, 44, 96, 278, 284, 286, 455, 581, 644, 650, 685, 691, 713, 717, 754, 799–801, 811 Hippodamia undecimnotata, 32 Hippodamia variegata, 32, 135, 188, 189, 225, 273, 274, 277, 333, 335, 453, 455, 458, 694, 717 Hippodamia (Adonia) variegate, 7 Hippodomia convergens, 26 Hippodomia parenthesis, 26 Hippodomia tredecimpunctata tibialis, 26 Hirsutella thompsonii, 789 Holding, 8, 79, 310, 314, 322, 357, 360, 361, 379, 609 Hologastric abdomen, 92 Holometabolous, 93, 238, 302 Holoptilinae, 108, 122, 338 Holoptilus melanospilus, 101 Homopterans, 37, 93, 151, 265, 668 Honeybees honeycomb, 120 honeydew, 5, 44, 126, 150, 192, 344, 386, 554, 758

Index Hoplistoscelis deceptivus, 479, 650, 657 Hoplistoseeis deeeptivus adult, 479 Hornets, 73, 74, 237 Horticultural crops, 2, 336, 592, 640 Horvathiolus superbus, 386, 389 Host plant resistance, 190 Host/prey location, 58, 126, 341 Hoverfly, 16, 96, 97, 132, 139, 140, 167, 168, 192, 290, 687 Hunterbugs, 3 HvCPI-1, 692 Hyadaphis coriandri, 454 Hyadaphis tataricae, 452 Hyaline wings, 80, 225 Hyalopterus, 471 Hyalopterus amygdali, 452, 453, 455 Hyalopterus pruni, 452–454, 456, 645, 646, 648 Hyaluronidase, 422, 423, 428, 435 Hyblaea puera, 476 Hybotidae, 3, 17, 38, 325 Hydrellia philippina, 213, 461 Hydroquinones, 432 Hylemya brassicae, 448 Hylesia paulex, 490 Hymenopodidae, 34, 399 Hymenoptera, 3, 25, 72, 106, 116, 151, 227, 289, 300, 413, 618, 653, 670, 764 Hymenopus coronatus, 399 Hyperaspis femorata, 455 Hyperaspis maindroni, 32, 455 Hyperaspis notata, 685 Hyperaspis polita, 605–607 Hyperaspis quadrimaculata, 456 Hyperaspis repensis, 456 Hyperaspis trilaneata, 669 Hypermetamorphic, 151 Hyphantria apterus, 478 Hypodamia tridecimpunctata, 189, 190, 565 Hypodermal pigments, 174, 175 Hypoponera eduardi, 467 Hypothenemus hampei, 489, 602, 604 Hypothetical F2 females, 165 Hyppodamia convergens, 56 I Icerya aegyptiaca, 464, 684, 686 Icerya pattersoni, 464 Icerya purchasi, 464, 563, 684, 686 Ictinogomphu srapax, 37 Idioscopus clypealis, 490 Idioscopus nitidulus, 490

Index Idioscopus niveosparsus, 490 Illeis cincta, 30, 32, 43 Imidacloprid, 615–618, 736, 739–742, 744, 745, 747, 751, 754, 755, 757, 766–771, 773, 789 Immature stages, 4, 81, 89, 149–177, 275, 283, 400, 450, 604, 651, 658, 675, 746 Immune response, 237 Incubation period, 123, 136, 137, 139, 156, 159, 161, 163, 171, 175, 185, 192, 195, 199, 203, 211, 265, 269, 271, 280, 439, 562, 745, 789, 790, 793 Index of dispersion, 597 Infestations, 195, 555, 560, 630, 634, 635, 647, 667, 668, 671, 675, 677–680, 766, 800 Insect growth regulators (IGRs), 741, 751, 756, 759–764, 805, 807 Insecticidal, 33, 56, 425, 427, 428, 433, 436, 765, 766, 802, 804–806, 813 Insemination, 355, 364, 369, 370 Integrated pest management (IPM), 1, 11, 33, 190, 210, 261, 556, 583, 591, 608, 615, 616, 634, 647, 648, 665, 672, 684, 693, 706, 708–711, 738, 742, 743, 746, 761, 762, 768, 770, 786, 797, 798, 808, 810 Inter-guild predation, 140, 335 Inter-specific competition, 642 Intra-guild interactions, 333, 651 Intra-guild predation (IGP), 106, 107, 139–141, 274, 317, 332–333, 336, 347, 591, 619, 629, 632, 643, 651, 653, 654, 678, 792 Intra-specific competition, 586, 587, 641 Intrinsic rate of increase, 135, 165, 202, 205– 207, 211, 534, 553, 554, 583, 746 Inula viscosa, 140 Invertase, 277, 422, 423 Invertebrate predators, 93, 98, 618 1-Iododecane, 344 Ippodamia (Adonia) variegate, 7 Irantha armipes, 29, 51, 102, 122, 137, 163, 199, 200, 313, 340, 362, 490, 496 IRGM, 56 Iridoid glycosides, 384, 389 Iridomyrmex humilis, 343 Irrigated condition, 635 Isaria fumosorosea, 787, 791 Ischiodon scutellaris, 32, 38, 97, 131, 166, 168–170, 194, 195, 365 Ischnura aurora, 31, 34, 37 Ischnura elegans, 34, 235 Ischnura nursei, 34 Ischnura senegalensis, 34, 171 Ischocyttarus flavitarsis, 466 Ischyropalpus nitidulus, 459 Ishidona egyptius, 38

839 Isoscelipteron okamotonis, 465 Isotomurus palustris, 451 Isyndus heros, 490, 706 Ithotze fusca, 151 Iwao's patchiness regression, 597 J Jackknife, 188 Jasmolin, 804 Jaw set, 322 Jetropa, 38, 40, 45, 48, 49, 53 Johnson-Taylor suction trap, 12 Joppeicus paradoxus, 481 Juvenile hormone (JH), 222, 243, 753, 759, 760, 770 Juvenile stages, 9, 153, 222, 255, 265, 270, 762 Juvocimene, 805 K Kairomones, 125, 341, 344, 346 Karanjin, 805 Karnyothrips melaleucus, 716, 723 Katydids, 76, 77, 240, 309, 391 Kerosene, 14 Kets, 28 Kharif season, 11, 272, 673 Kleptoparasite, 74 L Labidura riparia, 41, 263, 279, 307, 395, 475 Labidura truncata, 41, 475 Labiduridae, 3, 4, 14, 28, 41, 307, 475 Labium, 4, 78, 81, 83, 87, 90, 91, 242, 381, 382, 413 Lablab purpureus, 645 Laboratory hosts, 134–139, 150, 164, 198, 201, 538, 545 Labrum, 4, 81, 91, 152, 381 Lacewing brown, 3, 209, 556, 657, 695, 714 green, 5, 8, 16, 19, 27, 32, 152, 209, 224, 225, 289, 328, 333, 382, 556, 557, 600, 603, 604, 645, 647, 650, 651, 653, 657, 667, 668, 681, 683, 695, 714, 737, 752, 757, 765, 792, 793, 796, 808 Lactuca sativa, 55 Lactucas cariola, 14 Lady beetle, 3, 5, 19, 54, 100, 108, 117, 134, 173, 186–188, 396, 397, 462, 477, 498, 499, 559, 611, 641, 643, 655–657, 736, 737, 747, 758, 761, 774, 788, 812, 813 Laemostenus complanatus, 317

840 Lagenaria siceraria, 483 Landscape, 17, 30, 55, 59–61, 96, 100, 108, 158, 274, 338, 396, 653, 656 Lantana camara, 132 Laphygma exiguae, 488 Larger grain borer, 610 Laricobius nigrinus, 642, 644 Laricobius osakensis, 658 Larva, 6, 72, 74, 92, 116, 139, 150–153, 158, 167, 169, 170, 193, 198, 203, 212, 226, 283, 289, 303, 305, 308, 309, 318, 320, 322, 330, 385, 401, 417, 428, 430, 433, 436, 495, 499, 562, 582, 583, 606, 640, 646, 651–654, 684, 747, 799, 802 Larval stadium, 167 Lasioderma serricorne, 50, 481 Lasius alienus, 40, 467 Lasius fuliginosus, 467 Lasius grandis, 467 Lasius neglectus, 467 Lassius niger, 467 Latheticus oryzae, 49, 489 Lathrobium longulum, 29 Leaf galls, 632, 679 Leafminers, 154, 315, 619 Leaf vacuum, 10 Lecanicillium lecanii, 787 Lecanium nigrofasciatum, 316 Legume field, 646 Lepidophallus hesperius, 449 Lepidoptera, 4, 32, 72, 93, 106, 154, 155, 157, 158, 161, 162, 165, 166, 173, 175, 191, 196–198, 200, 201, 205, 211, 214, 251, 252, 254, 260, 288, 289, 303, 306, 307, 315, 316, 321, 329, 336, 343, 369, 395, 417, 434, 449, 466, 479, 530–533, 535, 546, 552–554, 559, 562, 589, 594, 595, 599, 601, 602, 605, 608, 609, 613, 616, 618, 638, 646, 679, 683, 690–692, 694, 738, 742, 751, 764, 767, 769, 774, 786, 787, 791, 794, 795, 797, 809 Lepidopteran larvae, 2, 5, 90, 199, 316, 331, 332, 634, 684, 688, 798 Lepidopterous rice pests, 460 Lepidosaphes beckii, 462, 464 Lepidosaphes tapleyi, 463 Lepisiota frauenfeldi subsp. Karavievi, 467 Lepisiota karavievi, 468 Leptinotarsa decemlineata, 397, 398, 562, 587, 589, 605, 658, 680, 692, 772, 788 Leptocorisa acuta, 490 Leptocorisa oratorius, 213, 461 Leptocorisa vericornis, 490

Index Leptomastidea abnormis, 720 Leptomastix dactylopii, 720 Leptomastix epona, 719 Leptomastix histrio, 719 Leptothrips mali, 37, 124 Lerida pugnax, 491 Lestesthor acicus, 34 Lestesum brinus, 34 Lestes viridulus, 34 Lettuce, 27, 47, 55, 169, 773 Leucochrysa, 152 Leucochrysa americana, 44 Leucochrysa arizonica, 46 Leucopis glyphinivora, 38, 470 Leucoptera coffeella, 466 Libellulidae, 3, 12, 27, 31, 37, 498, 614 Life-history, 98, 106, 157, 173, 183, 221, 224, 241, 264, 275, 277, 282, 330, 331, 534, 689, 746 Life span, 152, 162, 175, 184, 185, 195, 197, 199, 203, 213, 368 Life table, 34, 135, 165, 183, 184, 187–189, 196, 202, 205, 208, 210, 214, 255, 260, 266, 270, 275, 284, 329, 526, 531, 537, 552, 563, 583, 586, 598, 614, 742, 762, 791 Light, 2, 11, 12, 15, 20, 99, 100, 106, 107, 115, 117, 119, 120, 123, 132, 140, 150, 154, 162, 184, 193, 194, 203, 205, 226, 230, 233, 236, 239, 251, 252, 254, 259, 267, 269, 275, 305, 317, 320, 415, 552, 592, 628, 683, 693, 786 Light traps, 2, 7, 8, 11–12, 16, 109 Liliaceae, 97, 804 Limonene, 126, 127, 803, 804 Linalool, 126, 127, 803, 804 Linepithema humile, 467, 468 Lipaphis erysimi, 106, 135, 173, 184–186, 189, 458, 464, 480, 604, 605, 647, 648 Lipase, 420, 422, 423 Lipids, 114, 138, 155, 191, 261, 276, 307, 397, 420, 421, 428, 437, 527, 528, 535, 540, 552, 590 Lipophilic materials, 383 Liquid, 83, 168, 304, 331, 413, 419, 421, 424, 432, 433, 528, 532, 538, 541, 557 Listeria grayi, 439 Litter-dwelling prey, 450 Liturgusidae, 34 Live aphids, 176, 269, 560, 561, 566 Lixophaga diatraeae, 720 Lixus junci, 449 Lloyd’s patchiness indices of dispersion, 597

Index Locusts, 76, 77, 793 Lomamyia laiipennis, 152 Lomamyia longicollis, 152 Longevity, 135, 156, 159, 161, 164, 175, 177, 185, 186, 191, 198–204, 206, 207, 210, 211, 224, 229, 232, 253, 255, 261, 270– 273, 279, 285, 330, 356, 368, 396, 401, 532, 537, 561, 566, 569–571, 585, 595, 598, 601, 659, 738, 745, 747, 750, 754, 757, 762, 765–768, 774, 789, 799, 800 Long-horned grasshopper, 3, 12, 28 Long-legged flies, 3 Lophocephala gueriniis, 102 Lophyra intermedia, 17, 108 Lophyrotoma zonalis, 157 Loricera pilicornis, 27, 451 L-tryptophan, 126, 127 Lubber grasshoppers, 385 Lucerne, 10, 49, 52, 317 Lycosa pseudoannulata, 652 Lyctocoris campestris, 104 Lydella minense, 720 Lygaeidae, 3, 27, 84–85, 98, 99, 137, 138, 207– 209, 222, 268, 336, 346, 391, 398, 435, 486, 530, 569, 590–591, 691, 740–741, 794 Lygaeus equestris, 386, 389 Lygocoris communis, 476, 478, 486, 497 Lygus hesperus, 268, 478, 528, 649, 657 Lygus lineolaris, 478, 484, 486, 497, 528, 591 Lymantria dispar, 450, 487, 788 Lymantria dispar multiple nucleopolyhedrovirus (LdMNPV), 788 Lyophilised artificial diet, 174, 190 Lysiphlebus fabarum, 722 Lysiphlebus testaceipes, 718 Lythrum salicaria, 641 M Machimus arthriticus, 431 Maconellicoccus hirsutus, 210, 462, 464, 465, 560 Macracanthopsis nodipes, 121 Macroilleis (=Halyzia) hauseri, 7 Macrolophus basicornis, 316, 418, 485, 552, 638, 639 Macrolophus caliginosus, 140, 483, 484, 570, 667, 705, 723, 809 Macrolophus praeclarus, 267, 268, 485, 672 Macrolophus pygmaeus, 107, 140, 252, 253, 267, 277, 333, 335, 336, 480, 483–485,

841 552, 617, 618, 637, 640, 671, 678, 713, 714, 716, 722, 769, 792 Macromiidae, 171 Macropterous, 89, 93, 222, 223, 227, 231–233, 236, 242 Macrosiphoniella artemisiae, 457 Macrosiphum euphorbiae, 267, 455, 481, 484, 605, 606 Macrosiphum miscanthi, 458 Macrosiphum rosae, 169, 188, 210, 211, 453, 458, 464, 465, 470, 471, 605, 608 Macrotrachelia, 46 Magnetic resonance imaging (MRI), 411 Maize, 7, 18, 30, 33, 37, 41–46, 49, 52, 54, 56, 57, 99, 153, 172, 185, 195, 207, 208, 265, 307, 308, 328, 338, 395, 474, 475, 480, 533, 590, 597, 652, 653, 690, 703, 706, 752, 794, 795, 797, 802, 812 Malaise, 2, 7, 8, 11, 14–18 Mallada basalis, 106, 107, 465, 668–670 Mallada boniensis, 32, 706 Mallada boninensis, 209, 211, 465 Mallada desjardinsi, 210, 465, 647, 808 Mallada signata, 328, 712, 714, 722 Malus domestica, 346, 448, 603 Mamestra brussicae, 306, 450, 451 Mandibles, 72, 80, 81, 83, 89, 91, 99, 152, 304, 306, 309, 313, 321, 322, 368, 379–382, 413, 414, 431, 433 Mandible stylets, 313 Manduca sexta, 385, 389 Mango, 26, 31, 34, 36, 40, 44–48, 287 Mantidae, 27, 29, 32, 34, 115–116, 132, 357, 384 Mantoidea, 3 Manual milking, 415, 417, 418 Marasmia patnalis, 213, 460, 461 Margasus afzelii, 490 Marigolds, 200, 632, 633 Masarinae, 74, 92 Mating success, 150 Maxillae, 80, 81, 83, 89, 91, 152, 340, 381, 382, 413, 414, 433 Mealybug, 28, 48, 49, 165, 173, 209, 260, 265, 284, 310, 487, 499, 500, 532, 559, 584, 585, 598, 601, 606, 607, 618, 668, 684, 686, 705, 706, 708, 709, 712–714, 722, 751, 772, 787, 789, 799 Mealybug destroyer, 789 Mediterranean crops, 168, 633 Megacephala affinis, 459

842 Megacephala carolina, 689, 693 Megaloprepus coerulatus, 78 Megalurothrips sjostedti, 480, 582–584 Megalurothrips usitatus, 164, 165, 583, 584 Megoura viciae, 132 Melanitis leda ismene, 460, 476 Melanocoris, 46 Melanoplus femurrubrum, 581 Melanostoma melenium, 39 Melanostoma scalare, 472 Melanotus communis, 475 Melanthiaceae, 805 Melridae, 3, 41 Membranous wings, 74, 75, 77, 80, 84 Menochilus sexmaculatus, 225, 226, 290, 291, 334, 370, 453, 560, 561, 688, 755, 813 Mercury vapour lamp, 12 Meridic diet, 527–529, 531, 532, 538–544, 565 Mesocoxae, 89 Mesolecanium nigrofasciatum, 462 Mesopleural, 91, 325 Mesopleuron, 92 Mesoscutum, 91 Mesosternum, 92 Mesothorax, 75, 174 Mesovelia vittigera, 28 Messor alexandri, 467 Messor barbarus, 467 Messor bouvieri, 467 Messor darianus, 468 Messor denticulatus, 468 Messor medioruber, 468 Messor structor, 467 Messor sultanus, 468 Metabolic disruption, 804 Metabolites, 125, 213, 288, 425, 437, 440, 764, 771, 786, 793–798, 802, 803 Metamorphosis, 82, 83, 168, 171, 807 Metaphycus flavus, 720 Metaphycus helvolus, 720 Metaphycus lounsburyi, 720 Metaphycus spp. Peru, 720 Metaphycus stanleyi, 719 Metaphycus swirskii, 719 Metapterini, 87, 228, 229 Metarhizium anisopliae, 615, 786, 787, 789, 790, 792, 793, 797 Metarhizium brunneum, 787, 792 Metasyrphus confrater, 471 Metatetranvchusulmi, 478 Metathorax, 174, 212 Meteorological, 14, 287 Meteorus gyrator, 719

Index Methyl salicylate, 126, 127, 345, 346, 385, 389 Metioche bicolor, 460 Metioche vittaticollis, 28, 31, 213, 460 Metisa plana, 496 Metopolophium dirhodum, 450, 451 Mexican bean beetle, 117, 588 Micraspis crocea, 28, 31, 453 Micraspis discolor, 28, 30, 42, 668–670 Micraspis inops, 28 Micraspis nr. crocea, 453 Micraspis sp., 29, 453 Microclimate, 18, 100, 693 Microenvironmental cages (MECs), 537 Microhabitats, 15, 86, 102–104, 108, 129, 150, 237, 251, 307, 397 Microlophium carnosum, 131, 471 Micromus angulatus, 282, 714, 722, 804 Micromus igorotus, 28, 683 Micromus tasmaniae, 556, 722 Micromus timidus, 465 Micropterous, 87, 96, 103 Microptery, 89, 222, 223, 233 Micropyle, 114, 120 Microterys, 469 Microterys lunatus, 471 Microtheca ochroloma, 636, 654, 680, 681 Microvelia douglasi, 28, 31 Milk powder, 531, 539–541 Milkweed bugs, 3, 207, 391 Mimicry aggressive, 224, 301 Batesian, 224, 228, 306 Mertensian, 224 Miillerian, 224 Wasmannian, 224 Mimics, 6, 223, 224, 227, 228, 234, 237, 243, 301, 314, 392, 498, 528, 759, 760, 764 Minerals, 189, 190, 307, 540, 568, 656, 794 Mineral salts, 540, 557, 772 Minute pirate bugs, 3, 583, 705, 742 Miomantis caffra, 116, 132 Miomantis paykulli, 116 Miridae, 2, 3, 27, 28, 31, 46–50, 84, 91–92, 98, 99, 101, 138, 140, 158, 199, 205–206, 252–254, 267–268, 277, 315–316, 333, 335, 340, 346, 372, 381, 398, 423, 424, 435, 482, 530, 552, 569, 570, 592, 595, 596, 617, 618, 637–640, 667, 671–672, 706, 711, 767, 809–810 Mirid bugs, 2, 637, 639, 640, 659, 705 Mischocyttarus flavitarsus, 343 Mites, 2, 5, 6, 30, 33, 44, 48, 85, 88, 136, 150, 151, 164, 187, 273, 307, 316, 321, 379,

Index 381, 487, 526, 532, 566, 583, 584, 604, 605, 618, 629, 633, 637, 655, 656, 668, 671, 677, 678, 684, 705, 712–714, 716, 718, 722, 737, 741, 751, 756, 771, 772, 787, 792, 803 Mite, spider, 4, 6, 82, 187, 274, 275, 289, 315, 329, 331, 336, 487, 533, 583–585, 600, 633, 647, 668, 671, 705, 710–715, 787, 797 Mntispidae, 80 Mobility, 56, 117, 192, 301, 305, 421, 586, 592, 614, 766, 793, 800, 803 Moericke traps, 11 Monalonion bondari, 497 Monarch butterfly, 388, 644 Monoctenus sanchezi, 472, 497 Monoculture, 16, 49, 55, 97, 99, 655 Monolepta signata, 459 Monomorium areniphilum, 468 Monomorium ergatogyna, 317 Monomorium minimum, 469 Monomorium pharaonis, 468 Monomorium venustum, 468 Monophlebidae, 684–686 Montandoniola confusa, 2, 632, 633, 679 Montandoniola indica, 480, 585 Montandoniola moraguesi, 480, 632, 679 Mortality, 18, 106, 128, 150, 186, 226, 262, 329, 378, 421, 534, 580, 641, 677, 715, 737, 786 Mosquitoes, 129, 380, 787 Mouthparts, 73–76, 78, 80–82, 86, 90, 152, 170, 303, 357, 367, 371, 380–382, 395, 410–413, 428, 432, 433, 435, 538, 543 MTI-2, 692 Multiple mating, 356, 366, 367, 369 Multiple predators, 618, 648–657 Multivitamins, 190, 541, 554 Multivoltine, 101–105, 150, 191, 265, 397 Musca domestica, 448 Muscidifurax raptorellus, 721 Muscidifurax zaraptor, 720 Mustard, 49, 52, 125, 173, 189, 290 Mycene, 805 Mycetophagids, 82 Myiodactylus, 151 Mylabris indica, 344, 492, 495, 547, 587, 597, 673, 674 Mylabris phalerata, 670 Mylabris purtulata, 488 Mylloceous viridanus, 486 Myopopone castanea, 304, 468 Myrmeleon mobilis, 322

843 Myrmeleontid, 151 Myrmeleontidae, 28, 80, 120, 321, 432, 433 Myrmeleontiformia, 79 Myrosinase, 387 Myrothecium verrucaria, 787 Myrrha octodecimguttata, 32, 456 Myrrneleontoid, 151 Mythimna separata, 276, 492, 493, 670 Myzus, 473 Myzus cerasi, 452, 453, 456 Myzus persica, 125, 131, 139, 166, 184, 186, 188, 195, 259, 261, 267, 274, 452, 454, 457, 464, 472, 478, 483, 561, 592, 605, 606, 612, 647, 648, 657, 670, 691, 713 N Nabicula flavomarginata, 50, 397 Nabicula subcoleoptrata, 98, 317, 478 Nabidae, 2, 3, 9, 10, 13, 19, 27, 31, 32, 46–50, 84, 88–89, 98–99, 101, 123, 130, 133, 154–156, 203–205, 233, 262, 282, 286, 316–317, 336, 372, 398, 478, 530, 534, 591, 656 Nabis, 8–16, 26, 31, 49, 50, 88, 89, 99, 233, 316, 317, 326, 331, 336, 591, 651, 688 Nabis alternatus, 20, 154, 203, 336, 478, 658, 761, 801 Nabis americoferus, 20, 50, 99, 133, 264, 317, 336, 478, 479, 649, 650, 657, 801 Nabis capsiformis, 130, 155, 203, 205, 264, 265, 581, 789 Nabis consimilis, 156, 265 Nabis ferus, 50, 478 Nabis kinbergii, 154, 203, 317, 478, 711 Nabis major, 50, 397 Nabis pseudoferus, 31, 50, 106, 154, 155, 205, 282, 372, 534, 591, 615, 619 Nabis pseudoferus ibericus, 723 Nabis roseipennis, 20, 50, 130, 317, 478, 479, 650, 801 Nabis spp., 8–10, 26, 50, 265, 316, 336, 479, 656 NaCl, 539, 541 Nagusta goedelii, 51, 492 Nala lividipes, 195, 475 Nams roseipennis, 20, 50, 130, 317, 478, 479, 801 Naringenin, 126, 127 Nasonia vitripennis, 720, 765 Nasonovia ribisnigri, 465, 773 Nasutitermes corniger, 400 Nasutitermes exitiosus, 496

844 Natural diet, 557, 559, 561 Nearctaphis bakeri, 316 Nearctic vespine, 466 Nebria brevicollis, 27, 451, 691, 773 Necrobia rufipes, 491 Necrotic rings, 268, 639, 671, 672 Neem, 758, 790, 803, 806–809, 812–814 Negundoside, 805 Neivamyrmex nigrescens, 378 Nematicidal, 805, 806 Nematode, 30, 705, 785–787 Neochrysocharis formosa, 719 Neoconis szirakii, 46 Neodiprion abietis, 788 Neodiprion abietis nucleopolyhedrovirus (NeabNPV), 788 Neodiprion swainei, 466 Neohaematorrhophus therasii, 101, 103, 122, 232, 233 Neoleucinodes elegantalis, 316, 484 Neonicotinoid insecticides, 618, 745, 752 Neopterous, 79 Nephotettix virescens, 213, 461 Nephus arcuatus, 458 Nephus bipunctatus, 713 Nephus bisignatus, 283, 456, 563 Nephus includens, 283, 284, 456, 563, 717 Nephus nigricans, 453 Nephus reunioni, 456, 563, 717 Nesidiocoris tenuis, 107, 140, 252, 253, 268, 277, 315, 336, 337, 372, 482–485, 552, 617, 618, 637, 640, 672, 714, 716, 722, 767–769, 810–812 Nesidiocoris volucer, 158, 205, 206, 484 Nettle caterpillar, 201, 312 Neuroptera, 3, 26, 79, 97, 120, 150, 209, 224, 283, 302, 399, 410, 462, 556, 600, 645, 666, 711, 752, 792 Neurothemis fulvia, 31 Neurothemis tullia, 37 Neurotoxic, 302, 410, 420, 425, 429, 431, 435, 738, 740, 804 Neurotoxin, 420, 429, 435, 440, 794 Nevrorthiformia, 79 Nezara viridula, 481, 491, 670 Niche, 98–100, 107–110, 240 Nicotiana tabacum, 805 Nicotine, 805 Nilaparvata lugens, 34, 166, 213, 253, 316, 327, 330, 448, 450, 453, 460, 461, 592, 594, 598, 652, 658, 659 Nimbin, 804, 806 Nipaecoccus viridis, 458

Index Nocturnal, 10, 11, 14, 109, 209, 342 Non-Bt, 19, 35, 176, 326, 329–331, 689, 693, 799–801 maize, 329, 330, 799 maize rice, 799 Non-choice test, 365, 545 Non-target arthropods, 33, 58, 794 Non-tibiaroliate, 108 Non-venomous, 413, 424, 433 Non-volatile, 125–127, 242, 372 Notiophilus biguttatus, 27 Notiophilus semistriatus, 14 Notomus gravis, 451 Notonecta kirbyi, 614 Notoxus calcaratus, 8, 650 Notoxus monodon, 19 Nuclear polyhedrosis virus (NPV), 788–790, 799 Nuclease, 420, 435 NuCOTN 33B (Bt), 326, 799 Nudiscutella frontispina, 109 Number of prey (y), 185, 205, 250, 547, 580, 581, 593, 612, 616, 641, 769 Numerical response, 579, 586–587, 599, 658, 659 Nutrient regime, 57 Nutrients, 134, 183, 190, 191, 250, 281, 288, 356, 528, 538, 540, 568, 593 Nutritional composition, 134, 530 Nylanderia fulva, 467 Nymphal period, 154, 155, 157, 161, 172, 201, 231, 532, 742, 799 Nymphal stages, 77, 139, 153, 154, 157, 159, 162, 173, 192, 197, 213, 230, 252, 255, 256, 258, 265, 269, 280, 310, 391, 400– 402, 531, 533, 535, 542, 546, 549, 551, 584, 598, 658, 744, 746, 789, 790, 796 O Obligately feed, 93, 152 Occamus typicus, 200, 496 Ocelli, 75, 78, 80, 81, 87, 88, 212, 232, 431 Ocimin, 805 Ocimum sanctum, 805 O-coumaric acid, 126 Octacosane, 126, 127, 344 Octodecane, 344 Ocypus olens, 385, 449 Odentotermes obesus, 101, 102, 104 Odonata, 3, 7, 11, 12, 17, 26–29, 31, 34, 61, 66, 71, 78–79, 106, 170–171, 235, 239, 241, 380, 381, 614

Index Odontocheila nodicornis, 30 Odontogonus dimensis, 490 Odontomachus brunneus, 468 Odontoponera denticulata, 17, 108 Odontotermes brunneus, 595, 597 Odontotermes wallonensis, 488 Oebalus insularis, 495 Oechalia schellenbergii, 10, 281, 288 Oecophylla longinoda, 26, 40, 482, 483, 488, 490 Oecophylla smaragdina, 26, 157, 166, 303, 468, 487, 534, 666 Oedothorax apicatus, 653, 654 Oenopia conglobata, 7, 32, 174, 190, 334, 456, 560, 755 Oenopia kirbyi, 453, 605, 607, 608, 614 Oenopia lyncea, 32 Oenopia oncina, 456 Oil palm, 10, 16, 18, 19, 26, 54, 104, 200, 304, 312, 468, 469, 750 Oilseed rape (OSR), 259, 689, 692 Okra, 28–30, 136, 480, 634, 635, 706 Olfactometer, 107, 126, 545, 547 Olfactory stimuli, 301 Oliarces clara, 209 Oligidic diet, 529, 531, 532, 538, 540, 541, 545–547, 549–551, 673 Oligonychus coffeae, 187, 448, 647 Oligota kashmirica benefica, 448 Oligota oviformis, 448 Oligota pygmaea, 187, 448, 808 Olive oil, 565, 567 Olive orchards, 44, 45 Ommatissus lybicus, 459 Omnivorous, 2, 26, 76, 328, 629, 637, 679 Omnivorous predators, 56, 140, 158, 164, 172 Oncocephalus annulipus, 101 Oncocephalus sannulipes, 12 Oncopeltus fasciatus, 207, 391 Onitis falcutus, 12 Onthophagus landolti, 459 Onthophagus marginicollis, 459 Oocytes, 115, 141, 808 Ooencyrtus kuvanae, 721 Ooencyrtus pityocampae, 721 Oogenesis, 114, 115, 141, 570, 571 Oothecas, 15, 132, 191, 192 Ootstrap techniques, 188 Ophelosia crawfordi, 721 Ophionea nigrofasciata, 28, 30, 31, 451 Ophionia indica, 669 Ophraella communa, 18, 276 Ophyra aenescens, 722

845 Opilionids, 17 Opisina arenoosella, 480 Opisthorhynchous head, 91 Opius pallipes, 721 Orchard grass, 132 Orchards, 4, 6, 17, 18, 26–28, 33, 37, 38, 41– 45, 47, 49, 55, 100, 101, 124, 132, 133, 196, 287, 308, 397, 449, 452, 467, 472– 474, 476, 478, 484, 486, 497, 603, 629, 646, 648, 657, 677, 679, 681, 683, 686, 694, 736, 753, 755 Orchid bee, 489 Organic matter, 57 Organic vineyard, 49, 53 Orgilus obscurator, 721 Orius O. albidipennis, 31, 47, 48, 165, 207, 480, 582–586, 613, 614, 676, 723, 745, 792 O. amnesius, 48 O. arenosella, 48, 677 O. armatus, 723, 745 O. dravidiensis, 48 O. insidiosus, 13, 26, 128, 206, 242, 266, 321, 480, 531, 581, 630, 676, 710, 741, 789 O. laevigatus, 31, 47, 48, 140, 333, 344, 480, 481, 531, 533, 569, 584, 585, 630– 633, 678, 679, 704, 713, 714, 723, 741, 744–746, 792, 796 O. lindbergi, 49 O. majusculus, 33, 47, 266, 276, 332, 338, 480, 481, 530, 531, 630, 632, 633, 678, 679, 704, 723, 792, 799 O. maxidentex, 31, 48 O. minutus, 31, 47–49, 480, 481, 533, 583, 584, 723, 741, 745, 746 O. niger, 31, 47, 48, 338, 583, 584, 630–633 O. niger aegypitiacus, 48 O. nubilalis, 654, 690 O. persequens, 2, 669 O. sauteri, 48, 124, 128, 129, 164, 165, 206, 207, 242, 256, 481, 531, 582–586, 632, 633, 679, 744, 792, 799 O. shyamavarna, 48 O. strigicollis, 132, 136, 481, 531–533, 590, 631, 633, 649, 677, 679, 723 O. tantills, 487 O. tristicoler, 481 O. vicinus, 47, 378, 584, 585 Orthacris maindroni, 202, 204, 488, 493, 593, 594 Orthaga euadrusalis, 490 Orthaga exvinacea, 48

846 Orthetrumlu zonicum, 37 Orthetrumpruin osumneglectum, 37 Orthetrum sabina, 28, 31, 37 Orthetrumtaeni olatum, 37 Orthodera novaezealandiae, 132 Orthoptera, 3, 12, 17, 26–29, 31, 54, 76–77, 106, 176, 192, 202, 213, 214, 239, 240, 309, 317, 391, 460, 581, 593, 594, 668, 669 Oryctes boas, 491 Oryctes monoceros, 491 Oryctes rhinoceros, 304, 466, 491, 673 Oryzacystatin-1 (OC-1), 689, 692 Oryzaephilus mercator, 491 Oryzaephilus surinamensis, 481, 609, 610 Oryzopsis hymenoides, 611 O. sexpunctatus, 491 Osmylops, 151 Ostrinia furnacalis, 195, 308, 473, 475, 669 Ostrinia nubilalis, 33, 332, 480, 654 Oulema melanopus, 656 Ovarian maturation, 553 Ovarioles, 82, 553, 564–567, 759 Ovatus mentharius, 452–457 Ovicidal, 804, 805 Oviduct, 141, 564 Oviposition deterrent, 804, 805 period, 129, 153, 163, 164, 184, 185, 188, 189, 191, 203, 206, 254, 262, 272, 275, 744 sites, 115, 117, 119, 125, 128–131, 139– 141, 192, 193, 300, 341, 392 Ovipositional substrate, 532 Ovipositor, 74, 77, 79, 80, 82, 113, 129, 132, 193, 213 Owlflies, 3, 424, 432 Oxya nitidula, 202, 204, 311, 493 P Pacer beetles, 787 Pachycondyla, 435, 467 Pachycondyla goeldii, 425, 427, 437, 439 Pachycondyla striata, 428, 468 Pachycrepoideus vindemiae, 721 Paddy fields, 12, 28, 61 Paecilomite insects, 787 Paecilomyces fumosoroseus, 787 Paecilomyces lilacinus, 787 Paederus fuscipes, 28, 30, 34, 44, 448, 449, 652, 669, 802 Paederus littorarius, 449

Index Paederus memnonius, 449 Paederus riparius, 31 Pagasa fusca, 154 Pagasa sp., 479 Pagasa species, 479 Paleacrita vernata, 316 Palearctic syrphid, 96 Palmaspis phoenicis, 458 Palmipenna aeoleoprera, 465 Palmipenna cf. pilicornis, 465 Palps, 81, 83, 152, 305 Pamara mathias, 476 Panonychus ulmi, 316 Pantala flavescens, 12, 31, 37, 61 Panthous bimaculatus, 29, 53, 137, 159, 161, 163, 199, 313, 361, 362, 491, 496, 543, 548 Papillae, 120, 151 Paracercion malayanum, 34 Paracoccus marginatus, 584, 585, 602 Paragus albifrons, 39 Paragus compeditus, 39 Paragus haemorrhous, 470 Paragus quadrifasciatus, 39 Paragus serratus, 168, 170, 194, 365, 554 Paragus tibialis, 38, 365 Paramere, 89, 93, 102, 241, 242, 360 Parasaissetia nigra, 464 Parasitoids, 11, 13, 56, 93, 175, 288, 289, 300– 302, 332, 341, 355, 372, 383, 525, 526, 619, 635, 653, 656, 665, 678, 684, 689, 693, 705, 759, 766, 773, 774, 794, 798, 802, 806 Paratheresia claripalpis, 721 Paratrechina parvula, 469 Paratrechinica longicornis, 30 Paratriphleps laeviusculus, 2 Parental care guarding females, 394 guarding males, 108, 393 uniparental female care, 393 uniparental male care, 393 Parlatoria blanchardi, 463 Parlatoria cinerea, 463 Parlatoria date scale, 458 Parlatoria scale, 459 Parthenogenesis, 6, 116, 212 Parthenolecanium persicae, 463 Parthenolecanium pruinosum, 464 Pasimachus elongates, 14 Passive methods, 6, 7, 13 Pasture pests, 97 Patanga succincta, 670

Index Pathogens, 17, 278, 356, 790 Pea aphid, 54, 134, 154, 187, 479, 495, 563, 566, 651 Peanut, 14, 609, 673, 674, 802 Pear orchards, 17, 677 Pear thrips, 3 Pecan, 49, 52 Pectinophora gossypiella, 488, 492, 650 Pediobius foveolatus, 721 Pedunculate, 87 Peeling tree bark, 1 Pegomyia mixta, 449 Peiratinae, 12, 122, 359, 394 Pelopidas mathias, 308, 477 Pemphigus bursarius, 457 Pentacomia cupricollis, 30 Pentacosane, 126, 127, 227, 344 Pentatomidae, 3, 32, 90, 98, 124, 156, 234, 254, 301, 362, 381, 417, 499, 530, 581, 636, 669, 711, 738, 791 Pepper, 28, 36, 43, 252, 256, 267, 484, 612, 613, 630–633, 649, 742, 747, 757 Peprius nodulipes, 491 Peregrinator biannulipes, 50, 481 Peregrinus maidis, 454, 459, 465, 484 Pergidae, 157, 158 Pericallia ricini, 492, 673 Peridroma saucia, 477, 679 Perigrinator biannulipes, 50, 481, 491 Perillus, 156, 398 Perillus bioculatus, 254, 381, 397, 398, 528, 554, 570, 571, 587, 589, 636, 680, 692 Peristenus digoneutis, 721 Pestiferous insects, 39, 90 Pest suppression efficacy, 606 Petalurain gesntissima, 78 Phalaenopsis amabilis, 399 Phalaris arundinacea, 14 Pharoscymnus anchorago, 458 Pharoscymnus horni, 32, 458 Pharoscymnus numidicus, 305, 458 Pharoscymnus ovoideus, 458 Pharoscymnus pharoides, 458 Pharoscymnus setulosus, 458 Pharoscymnus varius, 458 Phaseolis spp., 6 Phaseolus vulgaris, 128, 582, 583, 655 Pheidole gertrude, 469 Pheidole magacephala, 469 Pheidole oxyops, 469 Pheidole pallidula, 467 Phenacoccus gossypii, 462

847 Phenacoccus madeirensis, 465 Phenacoccus manihoti, 455, 465 Phenacoccus solenopsis, 154, 202, 208, 493, 494, 584, 585, 594, 595, 598, 605–607, 634, 635, 799–800 Phenolics, 339, 384 Phenology, 28, 33, 54, 131–133, 208, 237, 250, 251, 255, 287, 586, 613, 646, 687 Pheromones aggregation, 342, 344, 345, 372 alarm, 290, 342, 372, 386 marking, 372 sex, 191, 242, 342, 344, 357, 365, 366, 368, 372, 373 traps, 2 Philanthus triangulum, 426 Philonthus flavolimbatus, 449 Philonthus hepaticus, 449 Philonthus longicornis, 449 Philonthus sericans, 449 Philonthus ventralis, 449 Philosamia cynthia ricini, 531 Phlaeothripidae, 3, 6, 31, 82, 124 Phloenomus minimus, 449 Phonoctonus fasciatus, 51, 120, 491 Phonoctonus lutescens, 120, 491 Phonoctonus nigrofasciatus, 51, 491 Phonoctonus spp., 491 Phonoctonus subimpictus, 51, 120, 233, 491 Phorodon humuli, 453, 474 Photinus pyralis, 44, 489 Photoperiod photophase, 154, 203, 270, 563 phototropic, 11 Phthorimaea operculella, 343, 475, 556 Phyla obtusa, 109 Phyllocnistis citrella, 463 Phyllocrania paradoxa, 399 Phyllotreta cruciferae, 460 Phylloxerae, 150 Phymata pennsylvanica, 484, 486 Phymatidae, 486 Physalis heterophylla, 14 Physokermes hellenicus, 467 Physopelta analis, 490 Phytophagous, 2, 4–6, 39, 58, 76, 82, 85, 90, 91, 93, 96, 97, 106, 156, 302, 363, 383, 387, 743, 762, 806 Picromerus bidens, 157, 158, 318, 320, 571, 723 Piercing, 74, 81–83, 309, 320, 380, 421 Piercing-sucking, 80, 83, 90, 309, 319, 320, 410

848 Pieris brassicae, 202, 492, 646–648, 658 Pieris rapae, 137, 159, 161, 201, 599, 641, 642, 644 Pigeon pea, 48, 49, 51, 596, 635, 673 Pigweed, 132, 138 Pineapple mealy bug, 458 Piperidine alkaloids, 425 Pirate bug, 2, 3, 8, 28, 531, 583, 678, 705, 706, 737, 742, 809 Pirates affinis, 101, 108, 434 Pisilus tipuliformis, 53, 394, 490, 492, 497 Pit clearing, 322 Pitfall traps, 2, 7, 8, 13–14, 16–19, 55, 321, 432, 649, 691, 801 Pithecellobium dulce, 394 Pittosporum tobira, 564 Plagiolepis maura, 468 Plagiolepis pygmaea, 467 Plagiotropic, 96 Planococcus citri, 165, 207, 260, 283, 284, 454, 456, 462, 465, 563, 684 Plant bug, 3, 88, 102, 315, 637, 653 growth factors, 19 pathogenic nematodes, 787 phenology, 54, 133, 251, 586 secretions, 382 shaking, 9, 19 trichomes, 382–383 varieties, 190 volatile chemicals, 125 Plastic cage, 647 Platerus pilcheri, 121 Platycheirus scutatus, 131, 166 Platymeris laevicollis, 491 Platymeris rhadamanthus, 429, 440, 491 Platynaspis luteorubra, 456 Platynus dorsalis, 452 Platystethus cornutus, 449 Platystethus nitens, 29 Plectrocnemia conspersa, 608 Plega signata, 80 Plesiochrysa ramburi, 209, 210, 465, 668, 669 Plodia interpunctella, 50, 489, 608–610 Plutella xylostella, 154, 159, 161, 173, 175, 196, 197, 201, 203, 251, 252, 264, 343, 386, 389, 428, 437, 473, 496, 530, 533, 599, 604, 605, 692, 774, 788, 791 Podabrus rugulosus, 489 Podisus bidens, 381 Podisus lewisi, 381 Podisus maculiventris, 138, 156, 236, 254, 281, 309, 329–331, 381, 385, 389, 423, 435,

Index 476, 477, 528, 552–554, 571, 587–589, 616, 617, 636, 650, 654, 655, 658, 679– 681, 691, 692, 711, 723, 737, 760, 774, 791, 811 Podisus nigrispinus, 156, 175, 254, 363, 387– 389, 417, 423, 425, 428, 434, 435, 437, 477, 571, 588–590, 613, 615, 616, 618, 723, 738–740, 774, 791, 792, 810–812 Podisus sagitta, 477, 528 Poecilus cupreus, 28, 44, 45, 55, 259, 260, 385, 451 Poecilus lucublandus, 452 Pogonomyrmex, 470 Pogonomyrmex naegelli, 469 Polididus, 12 Polididus armatissmus, 12 Polistes cf. olivaceus, 466 Polistes dominulus, 466, 618 Polistes versicolor, 466 Polistinae, 74 Polybia fastidiosuscula, 466 Polybia occidentalis, 466 Polybia paulista, 466 Polybia scutellaris, 466 Polybia (=Trichothorax) sericea, 466 Polybiinae, 151, 152 Polygonum convolvulus, 14 Polyhouse, 592, 627, 628, 630, 633, 634, 679 Polymorphic alate male, 230, 231, 498 brachypterous female, 230, 231, 234 brachypterous male, 230, 231 colour polymorphism, 105, 225, 226, 235, 236, 238, 391 entire red male, 231 entire violaceous black male, 231–233 morphs, 229 niger, 229, 230 nigrosanguineous, 229 sanguineous, 229, 230 violaceous black male with pale red corium, 231, 232 Polyphaga, 72 Polyphagous, 39, 82, 101–104, 131, 526, 528, 559, 592, 627–659 Polyphagous predators, 84, 128, 317, 333, 535, 562, 632, 639, 671 Polyrhachis lacteipennis, 468 Polystoechotids, 79 Polytoxus fuscoviftatus, 31 Polyunsaturated fatty acids, 527 Ponericins, 425, 427 Ponericins G, L, W, 425, 427

Index Pongamia pinnata, 803, 805 Pongamol, 803, 805 Pongapin, 803, 805 Population ecology, 184 Porcelloderes impenetrabilis, 400–402 Pork liver, 539, 541, 543, 565, 567, 570 Post-copulatory behaviour, 355, 359 Post-oviposition period, 156, 184 Post-predatory behaviour, 318, 548 Potato, 27, 31, 44–46, 49, 52, 97, 125, 129, 131, 139, 265, 284, 389, 467, 480, 563, 583, 584, 587, 590, 605, 667, 680, 688, 691, 692, 708, 713, 772 Potato leafhopper, 479 Potato psyllid, 638 Praon volucre, 721 Predaceous, 2, 30, 33, 82, 90, 98, 128, 156, 255, 281, 305, 307, 316, 342, 400, 429, 435, 590, 641 Predaceous mites, 33 Predator–prey interactions, 250, 286, 321, 384, 584, 600, 608, 762 Predator–prey-mutualist, 58 Predator size, 582, 606 Predator stage, 592, 606, 607 Predatory bug, 104, 154, 272, 290, 316, 331, 332, 344, 380, 385, 429, 530, 533, 552, 585, 586, 610, 614, 619, 640, 671, 676, 705, 706, 739, 761, 769 Predatory efficiency, 84, 608, 618, 653, 680, 809 Predatory midge, 678, 792 Pregnanes, 384 Pre-imaginal development, 203, 210, 284, 590 Pre-imaginal survival, 196 Prenolepis imparis, 321 Pre-oviposition, 116, 153–156, 184, 189, 197, 198, 203, 204, 206, 209, 232, 254, 275, 284, 285, 365, 392, 536, 537, 553, 554, 561, 562, 567, 595, 808 Pre-oviposition period, 153, 163, 187, 189– 191, 195, 198, 201, 203, 205, 208, 262, 553, 559, 590, 761, 767 Prerostichus melanarius, 27, 43, 55, 100, 119, 131, 227, 305, 306, 451, 452, 645, 656, 689, 692, 773 Preservation, 2, 210 Prey beating, 322 capture, 75, 199, 264, 306, 315, 325, 399, 413, 433, 435, 596, 766 clearing, 322 defensive secretions, 301

849 deprivation, 229, 417, 418, 552 immobilisation, 310 preferences, vii, 33, 591, 681 Preying mantids, 3, 9, 80, 191, 688 Priscibrumus uropygialis, 17 Pristhesancus plagipennis, 51–53, 271, 411, 421, 435, 436, 491, 536, 673 Proctacanthus milbertii, 470 Production cost, 530, 534, 535 Pronotum, 74, 76, 84, 85, 87, 90–92, 159, 184, 209, 234, 238, 370, 401 Propylea dessecta, 19 Propylea japonica, 11, 33, 139, 190, 214, 276, 289, 326, 327, 330, 372, 456, 693, 798, 802 Propylea quatuordecimpunctata, 32, 117, 186, 273, 274, 277, 337, 605, 606, 611 Propylene glycol, 14, 743 Proreus similans, 669 Prorops nasuta, 721 Prospaltella spp., 721 Prostemma sp., 480 Prostephanus truncatus, 609–611 Prosternum, 86, 88, 91 Protease, 277, 411, 420–424, 436, 440, 689 inhibition, 420 inhibitor, 440, 689 Proteinases, 421, 422 Protein family, 420 Protein X, 538–540 Proteolysis Fibrillin, 420 Proteolysis S1 protease, 420 Prothorax, 75, 76, 80, 82, 174, 212, 356, 361, 362, 390, 414 Prothrombin activator, 421 Prothyma sp., 12 Protoalkaloids, 804 Protocatechuic acid, 126 Protonectarina sylverae, 466 Protopulvinaria pyriformis, 463 Pselliopus latispina, 490 Pselliopus zebra, 497 Pseudagrion decorum, 34 Pseudagrion hypermelas, 34 Pseudagrion microcephalum, 34 Pseudagrion rubriceps, 37 Pseudaphycus angelicus, 721 Pseudaphycus flavidulus, 721 Pseudaphycus maculipennis, 721 Pseudoazya trinitatis, 41, 685 Pseudococcidae, 165, 260, 283, 284, 684, 685 Pseudococcus comstocki, 462, 464 Pseudococcus jackbeardsleyi, 210, 465, 669

850 Pseudococcus maritimus, 462, 463 Pseudococcus sp., 486 Pseudocreobotra wahlbergi, 399 Pseudodendro thrips mori, 6 Pseudodorus clavatus, 170, 195, 366 Pseudomallada prasinus, 133 Pseudomantis albofimbriata, 191 Pseudomonas aeruginosa, 277, 439 Pseudomyrmex gracilis, 427 Pseudomyrmex penetrator, 427 Pseudomyrmex termitarius, 427 Pseudopanax crassifolius, 132 Pseudophonus rufipes, 55, 385, 451 Pseudoplusia includen, 464, 466, 479, 487, 489, 491, 657 Pseudorhopus testaceus, 469 Pseudotheraptus wayi, 491 Psychopsidae, 151, 432 Psylla pyricola, 321 Psyllidae, 256, 334, 638, 685, 686 Psyllids, 4, 5, 188, 256, 308, 474, 638, 677, 686, 712–714, 719, 757 Psyllobora vigintiduopunctata, 32, 456 Psyllopsis repens, 256 Psyttalia concolor, 721 Pterochloroides persicae, 457 Pterohelaeus nr. darlingensis, 477 Pteroma pendula, 303, 466 Pterophorus lienigianus, 496, 594, 595 Pterostichus anthracinus, 29, 118, 119 Pterostichus cupreus, 27, 451, 772 Pterostichus melanarius, 27, 43, 55, 100, 119, 131, 227, 305, 451, 452, 645, 656, 689, 692, 773 Pterostichus mutus, 452 Pterostichus nigera, 27 Pterostichus oblongopunctatus, 380 Pterostichus strennusa, 27 Pterostigma, 78 Pterothorax, 174, 212, 359 Puccinia polysora, 172 Pulvinaria floccifera, 463 Pulvinaria horii, 463 Pulvinaria tenuivalvata, 463 Pulvinaria vitis, 463, 464 Pumpkin, 49, 53, 209, 560, 810 Pupa, 6, 18, 27, 72, 81, 82, 132, 151, 152, 166, 167, 173, 174, 187, 189, 197, 198, 201, 212, 225, 227, 264, 288, 303, 341, 343, 387, 430, 449–452, 468, 470, 473, 475, 531–535, 555, 556, 560, 567, 588, 591, 609–611, 641, 645, 713, 714, 737, 756– 758, 760, 763, 767

Index Pupation, 74, 81, 152, 168, 170, 238, 260, 290, 327, 566, 755, 799, 807 Purified water, 330, 567 Purple loosestrife, 641, 644 Pygolam pisfoeda, 12 Pygophoresetose, 91 Pyralidae, 16, 158, 162, 165, 177, 196, 200, 205, 280, 336, 343, 369, 530–532, 535, 560, 562, 599, 608, 616, 683, 769 Pyraris farinalis, 50 Pyrethrins, 746, 753, 803, 804, 810–812 Pyrethrum, 803 Pyrrhalta viburni, 477 Pyrrhocoridae, 12, 28, 318, 598 Pyrrolizidine alkaloids, 385, 389 Q Quercetin, 126, 127 Quercus glauca, 379 Quiescence, 322 Quiescent prey, 105 R Ragmus importunitus, 486 Ragweed, 18 Rain fed condition, 635 Rapeseed pollen, 567 Raphidiodean, 81 Rapismatids, 79 Rasping, 81, 87, 89, 564 Rastrococcus iceryoides, 463 Rastrococcus invadens, 462 Red clover, 591 Red date palm scale, 458 Red spider mite (RSM), 187, 585, 647, 714, 808 Redtop grass, 132 Reduviidae, 3, 27, 84, 98, 137, 158, 199, 242, 257, 301, 358, 393, 411, 488, 530, 593, 635, 669, 711, 746, 789 Reduviinae, 12, 123, 338, 339, 394, 400 Reduviolus americoferus, 478 Reduviolus roseipennis, 479, 650, 657 Relative density (RD), 26 Release-rate, 636, 638, 639, 645, 647, 654, 655, 671, 673, 680–682, 704 Repellent, 339, 384, 388, 804, 805 Reproductive response, 579, 613, 658 Resilin, 76 Resistant variety, 630 Resorcinol, 126 Reticulitermes flavipes, 386

Index Reticulitermes speratus, 465 Retreating, 369 Rhagoletis mendax, 452 Rhagoletis pomonella, 448, 450, 452, 466 Rhapftidosoma atkinsoni, 121 Rhinacloa forticornis, 315 Rhinocapsus vanduzeei, 653 Rhinocoris albopunctatus, 130, 340, 492 Rhinocoris bicolor, 492, 497 Rhinocoris carmelita, 440, 497 Rhinocoris fuscipes, 103 Rhinocoris kumarii, 102 Rhinocoris loratus, 497 Rhinocoris marginatus, 101 Rhinocoris tristis, 107 Rhinocoris tropicus, 492 Rhizobius lophantae, 564 Rhizopertha dominica, 481, 489 Rhopalosiphum, 471 Rhopalosiphum maidis, 96, 185, 187, 453, 456, 457, 464, 465, 536, 654, 690 Rhopalosiphum padi, 29, 186, 188, 328, 372, 450, 451, 457, 463, 482, 556, 646, 648, 655, 797 Rhynchophorus ferrugineus, 195, 475 Rhynocoris albopilosus, 492, 770, 771 Rhynocoris fuscipes, 29, 51–53, 87, 121, 258, 271, 272, 276–278, 280, 287, 313, 415– 418, 422, 426, 436, 438, 439, 477, 493, 494, 496, 500, 528, 543, 587, 594–598, 688, 694, 748, 749, 771 Rhynocoris iracundus, 419, 423, 430, 439, 492, 495 Rhynocoris kumarii, 51, 88, 121, 122, 198, 344, 358, 415, 416, 418, 477, 494, 499, 528, 536, 548, 552, 587, 594, 596, 597, 634, 672, 673, 675, 676, 693, 694, 748–750, 770, 789, 790 Rhynocoris longifrons, 51, 477, 494, 500, 594, 595, 634–637, 673, 706 Rhynocoris marginatus, 51, 88, 164, 198, 229, 257, 309, 394, 411, 492, 537, 587, 634, 673 Rhynocoris squamulosus, 495 Rhyothemis variegate, 37 Rhyparochromina sp., 479 Rhyzobius chrysomeloides, 717 Rhyzobius forestieri, 457, 717 Rhyzobius lophanthae, 41, 345, 458, 657, 684, 685, 695, 718 Rhyzobius (=Lindorus) lophanthae, 709 Rhyzopertha dominica, 609, 610 Rice leaffolder, 138 medium, 164, 198

851 pests, 28, 31, 34, 449, 460, 466 stem borer, 138, 316 Ricolla pallidinervis, 495 Rihirbus trochantericus, 29, 51, 120, 312, 495, 594, 595 Ringlegged earwig, 3 Ripple bug, 28 Riptortus clavatus, 490 Riptortus pedestris, 384 Rivula atimeta, 451, 453, 460 Robber flies, 7, 97, 380, 411, 413, 414, 431, 433, 462, 465, 498, 512 Rocconota tuberculigera, 495 Rodolia cardinalis, 41, 564, 684–686, 759 Rodolia pumila, 41, 686 Romalea microptera, 385, 389 Root grubs, 787 Ropalidia brevita, 466 Ropalidia fasciata, 30 Roses, 630, 678 Rostrum rostral probing, 257, 318, 547–549 rostrum (labium), 83 rostrum extension, 394 Rotenone, 803, 805, 810 Rove beetle, 3, 13, 19, 28, 30, 33, 44, 57, 72, 306, 652, 656 Royal jelly, 189, 190, 566 Ryania, 803, 805 Ryania speciosa, 803, 805 Ryanodine, 803, 805 S Sabadilla, 803 Saccharicoccus sacchari, 462, 669 Saccharopolyspora, 738, 787, 793, 794, 796 Saccharose, 557 Sahlbergella singularis, 492 Saicinae, 108 Saintpaulia, 630, 676, 792 Saissetia oleae, 463, 602, 603 Salannin, 804, 806 Salicylic acid, 126, 127 Saltatorial, 76 Salyavainae, 123 Salyavata variegata, 400 Sanguineous, 159, 229, 230 Saprophagous, 97 Saropogon combustus, 470 Saropogon pritchardi, 470 Satiation threshold, 580 Sawfly, 88, 157, 158 Sawing, 74 Scaevaalbo maculata, 39

852 Scaeva pyrastri, 27, 194 Scale insects, 4, 44, 93, 150, 321, 380, 454, 456, 457, 787 Scaphinotus elevates, 14 Scarabaeidae, 12, 304 Scarabaeiform, 151, 152 Scarites quadriceps, 14 Sceliphron madraspatanum, 30 Schacht trap, 15–16 Schizaphis graminum, 29, 96, 169, 195, 458, 586, 613, 614 Schizoneura, 471 Schizotetranychus brevisetosus, 379 Schoenocaulon officinale, 805 Scirpophaga incertulas, 213, 461 Scirpophaga innotata, 461 Scirtidae, 12 Scirtothrips dorsalis, 632, 633 Scolothrips longicornis, 6, 275 Scolothrips sexmaculatus, 124, 712, 716, 723 Scolothrips takahashii, 6, 274, 379, 617 Scotinophara coarctata, 453, 460, 461 Scrobipalpa ocellatella, 449 Scutellista caerulea (cyanea), 721 Scutellum, 83, 85, 87, 90–92, 184, 234, 319 Scymnus apetzi, 32, 42, 457 Scymnus apiciflavus, 669, 670 Scymnus bipunctata, 458 Scymnus bivulnerus, 32 Scymnus coccivora, 32, 706, 772 Scymnus coniferarum, 643, 644 Scymnus flagellisiphonatus, 32 Scymnus frontalis, 32, 457 Scymnus interruptus, 32, 42, 457 Scymnus levaillanti, 261, 457, 801 Scymnus mimulus, 457 Scymnus pallipediformis, 32, 457 Scymnus pictus, 458 Scymnus punetillum, 458 Scymnus quadriguttatus, 457 Scymnus rubromaculatus, 32, 42, 457, 717 Scymnus subvillosus, 32, 457 Searching ability, 587 Seed beetles, 609 Seed bugs, 3 Segmental bristles, 193 Selenaspidus articulatus, 463 Selenophorus palliatus, 689, 693 Selenophorus seriatoporus, 30 Selenothrips (=Heliothrips), 484 Selenothrips rubrocinctus, 484 Semi-aquatic, 97, 98 Semidalis boliviensis, 46 Seminal receptacle, 82

Index Semiothisa pervolgata, 476, 477 Semi-solid, 557 Serangium japonicum, 131, 133, 615, 617, 618, 768 Serangium parcesetosum, 458, 629, 630 Serine, 421 Serine proteases, 420, 421, 423, 425 Serological techniques, 107 Sesamia inferens, 138, 316, 476, 669 Sesquiterpene, 764, 804 Setora nitens, 466 Setothosea asigna, 201, 495 Sex lures, 2 Sex ratio, 154, 156, 162–164, 166, 173, 175, 176, 186, 191, 192, 195, 199, 205, 206, 210, 212, 229–231, 241, 262, 265, 268, 270, 275, 279, 366, 401, 537, 560, 562, 566, 641, 659, 766, 790, 809 Sexual dimorphism, 85, 86, 173, 191, 200, 209, 231, 232, 239–242 Shake bucket, 19 Short-term, 282, 529, 606 Sidnia kinbergi, 154, 317, 478 Sigmoid increase, 580 Silverleaf whitefly, 645, 679 Silvopastoral, 16, 97 Silvopasture, 16, 97 Simian 3 (non-Bt), 326, 799 Simulium damnosum, 470 Simulium pontina, 470 Sinapis alba, 125 Sinea diadema, 53, 317, 495, 537 Sinea sanguisuga, 495 Sinea spinipes, 122, 495 Sirthenea carinata, 12 Sisyridae, 152 Sitobion avenae, 186, 289, 290, 450, 471, 653, 656 Sitodiplosis mosellana, 452 Sitotroga cerealella, 154, 174, 190, 208, 464, 484, 486, 530–532, 559, 601–603 Slant-faced grasshopper, 593 Smooth brome, 132 Social wasps, 4, 92, 380, 466 Soft-winged flower beetle, 3 Sogatella furcifera, 213, 461 Soil-dwelling, 57, 73, 450, 789 Soil moisture, 57, 99 Solanaceae, 97, 804 Solanaceous crops, 638–640 Solanum lycopersicum, 639, 671, 679, 743, 813 Solanum melongena, 6, 595, 801 Solanum nigrum, 14, 125 Solanum tuberosum, 125, 691, 692

Index Solenopsins, 425, 427 Solenopsis geminata, 425, 427, 428, 437, 466, 469 Solenopsis invicta, 14, 55, 176, 416, 425, 427, 428, 437, 469, 799 Solenopsis saevissima, 468, 469 Solenopsis wolfi, 468 Solenopsis xyloni, 317 Solenostethium liligerum, 491 Solitary, 74, 92, 123, 397 Solitary wasps, 4, 92, 151, 152, 302 Sophora microphylla, 132 Sorghum, 2, 48, 96, 164, 195, 198, 454, 459, 465, 484, 653, 675, 743 Sorghum medium, 164, 198 Soybean, 9, 10, 16–19, 27, 30, 31, 37, 38, 40, 43, 44, 48–50, 55, 85, 99, 108, 128, 130, 132, 138, 195, 201, 207, 287, 321, 345, 464, 470, 477, 479, 487, 489, 588, 634, 643–645, 650, 657, 670, 706, 738, 802, 807, 811 Soy bean lecithin, 557 Space clearance rate, 582, 606, 612 Spalangia cameroni, 721 Spalangia endius, 721 Spalangia gemini, 721 Spalangia nigroaenea, 722 Spalgius epius, 28 Spanogonicus albofasciatus, 484, 650 Spartocera dentiventris, 490 Spathodea campanulata, 48 Spatial scales, 109, 629 Spatio-temporal, 39, 100, 108, 109, 277, 666 Specific predators, 1, 177, 274, 695 Spectroscopy techniques, 411 Spermatheca, 212, 369 Sphaerophoria contigua, 470 Sphaerophoria macrogaster, 472 Sphaerophoria rueppelli, 39, 168 Sphaerophoria scripta, 27, 31, 38, 39, 166 Sphaerophoria sulphuripes, 38, 194 Sphecidae, 3, 27, 28, 31, 39, 40, 309, 344, 670 Sphedanolestes aterrimus, 103 Sphedanolestes himalayensis, 123, 162, 198, 496 Sphedanolestes leucocephalus, 497 Sphedanolestes minusculus, 122, 123, 162, 198, 496 Sphedanolestes pubinotum, 123, 162, 198 Sphedanolestes signatus, 123, 137, 162, 163, 198–200, 312, 340, 496 Sphedanolestes variabilis, 123, 162, 198, 309, 358, 496, 593, 594

853 Spherophoria scripta, 39 Spherophoria turkmenica, 39 Sphex viduatus, 670 Sphincter valves, 412 Sphingomyelinase C, 433 Spiders, 4, 6, 9, 17, 18, 31, 33, 55, 82, 87, 93, 98, 109, 151, 152, 187, 239, 274, 275, 289, 303, 315, 317, 322, 329, 331, 336, 338–340, 367, 489, 533, 563, 580, 583– 585, 600, 633, 647, 649, 652, 653, 656, 668, 671, 674, 688, 691, 705, 710–715, 747, 787, 797, 801, 802, 808, 812 Spilonota ocellana, 316 Spined soldier bug, 618, 636, 654, 679, 774 Spiracle, 82, 89, 90, 92, 167, 168 Spitting of saliva, 310, 323 Spodoptera exigua, 153, 154, 176, 208, 254, 328, 329, 476–478, 480, 486, 590, 613, 691, 788, 810 Spodoptera exiguan nucleopolyhedrovirus (SeNPV), 788 Spodoptera frugiperda, 4, 138, 157, 172, 263, 307, 395, 428, 448, 454, 455, 466, 474– 477, 487, 490, 491, 496, 497, 584–586, 589, 599, 605, 607, 616, 652, 653, 753, 795, 797, 811 Spodoptera littoralis, 175, 205, 328, 451, 463, 489, 690, 786, 788, 795, 797, 812 Spodoptera littoralis nucleopolyhedrovirus (SpliNPV), 788–790 Spodoptera litura, 136, 137, 159, 161, 164, 201, 202, 254, 255, 257, 258, 260, 320, 330, 418, 427, 429, 430, 436–439, 466, 476, 483, 488–495, 500, 530, 531, 539, 545, 546, 548–551, 581, 594, 597, 634, 635, 673, 674, 676, 694, 788–790 Spodoptera litura nucleopolyhedrovirus (SLNPV), 788 Spoviruses, 174 Springbok mantis, 116 Spumaline secretions, 121 Squamocin, 803, 804 Square beating trays, 7, 8, 10–11 Stable isotopes, 107, 109, 338 Stagmatoptera, 115 Standardised rearing, 526, 703 Staphylinidae, 3, 8, 10, 13, 16, 26–28, 30, 31, 33, 41, 44, 72, 86, 101, 224, 305, 346, 397, 448–450, 656, 669, 690, 709, 736, 794, 796 Staphylinids, 14, 18, 29, 30, 55, 186, 187, 653, 689, 693, 704, 705, 794 Staphylinoidy, 223

854 Starvation, 229, 323, 363, 417, 544–552, 765 Statilia maculata, 11, 36 Stegobium paniceum, 50, 481 Stemborer, 461 Stenogastrinae, 74, 92 Stenolophus comma, 452 Stenolophus ochropezus, 693 Stenopodinae, 12, 108, 359 Stenotopic, 100 Stephanitis pyrioides, 653 Stereptophenicol, 189, 190 Sterilant, 805 Sterilised eggs, 136, 152, 153 Sternite, 86, 212, 232, 242, 401, 563 Sternum, 85, 92, 381, 548 Sterols, 134, 527 Steropus gallega, 55 Steropus melas italicus, 385 Stethorus, 448 Stethorus gilvifrons, 32, 457, 566 Stethorus punctillum, 32, 690, 713, 718 Sticky traps, 7, 11, 13, 16, 315 Stilt bugs, 3 Stimulation method, 417 Stinging, 74, 92, 228, 304, 394 Stiretrus, 156 Stiretrus anchoago, 650 Stiretrus decemguttatus, 157, 234, 236 Stizus continuus, 469 Stomach poison, 805 Storage pests, 2, 203, 497, 605, 608–611, 613, 619, 678, 814 Strawberries, 18, 49, 480, 535, 591, 630, 631, 633, 678, 691, 711, 746, 792, 803 Streptomycin, 539–541 Stressed, 150, 254, 276, 277, 279, 415, 416, 739, 764, 765, 773, 774, 809 Stridulation, 76, 306, 360 Striped earwigs, 3, 4, 263, 279 Subcoccinella vigintiquattuorpunctata, 32 Submacroptery, 223 Submergence, 322 Sucking insects, 173, 787 Sucking pests, 135, 282, 313, 534, 596, 615, 647 Sucrose, 189, 190, 380, 527, 531, 539, 541, 554, 561, 565–567 Suction discs, 152 Suction sampling, 8, 11, 17 Suction trap, 7, 12–13 Sugar beet, 55, 681 Sugarcane, 35, 49, 51, 54, 195, 307, 308, 467– 469, 669, 686, 758

Index Sunflower, 48, 49, 53, 207, 743 Superoxide dismutase (SOD), 276, 277 Supputius, 156 Supputius cincticeps, 402, 477 Survival costs, 393 Survivorship, 807 Sustainable agriculture, 559, 603, 716 Swarming, 366 Sweep-netting, 1, 7–10, 16–19, 649 Sweet pepper, 2, 48, 268, 333, 389, 588, 611, 613, 614, 630–633, 649, 671, 678, 711, 792 Sweet potato, 49, 52, 131, 467, 590, 667, 708, 713, 760 Sweetpotato whitefly, 638, 640 Sycanus affinis, 197, 198, 495 Sycanus annulicornis, 123, 161, 201, 202, 495 Sycanus ater, 103 Sycanus aurantiacus, 496 Sycanus collaris, 121, 200, 201, 495, 496, 552, 594, 595, 669, 670 Sycanus dichotomus, 10, 104, 161, 162, 199, 200, 312, 496, 750, 751 Sycanus falleni, 137, 159, 199, 201 Sycanus galbanus, 29, 200, 360, 496 Sycanus indagator, 51, 496, 599 Sycanus reclinatus, 161, 197, 496 Sycanus sichuanensis, 200 Sycanus versicolor, 197 Sylepta derogata, 492 Sympherobius barberi, 695, 712 Sympherobius fallax, 722 Sympherobius subcostalis, 46 Symphrasinae, 151, 152 Synacra paupera, 719 Synoeca cyanea, 466 Synonycha grandis, 669 Syrphidae, 3, 7, 13, 16, 27, 28, 31, 32, 38, 96, 97, 101, 126, 132, 166–170, 227, 237, 302, 343, 346, 365, 380, 385, 389, 413, 470–472, 499, 555, 670, 687, 757, 811 Syrphid fly, 119, 133, 227, 228, 343, 380, 495, 752, 755 Syrphus balteatus, 670 Syrphus opinator, 194 Syrphus rectus, 133, 470 Syrphus ribesii, 166, 471 Systematics, 6, 71, 558, 712 T Tachyporus chrysomelinusa, 27 Tachyporus hypnorum, 27, 186, 306, 450

Index Taeniolate marshhawk, 37 Taeniopoda eques, 389 Tagetes eracta, 632 Tagetone, 805 Tagetes erecta, 205, 805 Tamarinadus indica, 102–104 Tamarixia radiata, 656, 657 Tapinoma carininotum, 468 Tapinoma festae, 468 Tapinoma karavievi, 468 Tapinoma simrothi subsp. karavievi, 468 Tarsomere, 75, 92 Tarsus, 75, 92, 151, 159, 174, 175 Tasgius (=Paratasgius) ater, 449 Taxifolin, 126 Taxonomic, 33, 34, 71, 85, 151, 240, 333, 388 Taylor’s power law, 597 Tea, 2, 6, 129, 187, 200, 274, 595, 647 Teak skeletonizer, 201, 596 Technomyrmex moerens, 468 Tecoma stans, 48 Tegea atropicta, 496 Tegmina, 75, 77 Telenomus remus, 720 Teleogryllus commodus, 475 Temnostethus, 46, 236 Temnostethus paradoxus, 49 Temnostethus pusillus, 104 Temperature, 20, 99, 119, 150, 187, 225, 250, 305, 363, 391, 526, 581, 627, 674, 751 Temporal scales, 629 Tenebrio, 497 Tenebrio molitor, 123, 138, 156, 157, 161, 162, 176, 199–202, 255, 476, 477, 491, 495, 496, 530, 536, 537, 792 Tenebrio molitor pupae, 567 Tenodera aridifolia, 390 Tenodera aridifolia sinensis, 384, 391 Tenuisvalvae notata, 260, 265 Terebrantia, 82 Teretrius nigrescens, 609, 610 Teretrius pulex, 459 Tergite, 89, 92, 564 Termatophylidea opaca, 484 Termatophylidea pilosa, 484 Termites, 78, 96, 102, 152, 198, 306, 307, 322, 338, 386, 400, 488, 597, 787 Terpenes, 372, 384, 411, 764, 802 Terrestrial arthropods, 1, 13, 25 Tetrabrachinae, 72 Tetracha brasiliensis, 30 Tetracnemoidea brevicornis, 719 Tetracycline, 212

855 Tetradecane, 387, 389 Tetramorium bicarinatum, 467 Tetramorium caespitum, 467, 468 Tetramorium taurocaucasicum, 468 Tetranychidae, 315, 329, 533, 583, 585, 600, 633, 746 Tetranychus cinnabarinus, 632 Tetranychus evansi, 455, 605, 606 Tetranychus parvispinus, 205, 206 Tetranychus urticae, 6, 49, 267, 275, 328, 329, 448, 453, 480, 482, 483, 487, 490, 533, 583–585, 592, 600, 602, 633, 655, 671, 690, 746, 792, 797 Tetraphleps abdulghanii, 104, 105 Tetraphleps gracilis, 104 Tetraponera rufonigra, 30, 40 Tetrastichus coeruleus (asparagi), 719 Tettigoniidae, 3, 27, 76, 77, 240, 309, 461 Thasus acutangulus, 384 Thaumastocoridae, 486 Thaumastocoris peregrinus, 465, 484, 486 Thermal tolerance, 253, 277, 278 Thiacidas postica, 476 Thiamethoxam, 617, 618, 739, 740, 744, 745, 753, 756, 770, 773, 789 Thinopinus pictus, 33 Tholymis tillarga, 31, 37 Thorax, 73, 76, 78, 80, 82–85, 87, 89–91, 129, 151, 191, 193, 234, 236, 242, 310, 319, 367, 428, 548 Thosea cervina, 476 Thripidae, 3, 6, 124, 164, 315, 329, 344, 345, 583, 603, 630–632, 712 Thripobius semiluteus, 720 Thrips, 2, 44, 81, 124, 164, 212, 256, 315, 381, 480, 532, 583, 630, 667, 705, 741, 792 Thrips and aphids, 333, 632, 633, 678 Thrips hawaiiensis, 315 Thrips orientalis, 315 Thrips palmi, 582–584, 631, 633, 649, 679 Thrips parvispinus, 484 Thrips setosus, 481, 667, 668 Thrips tabaci, 472, 583, 584, 631, 633, 670, 813 Thunderflies, 3 Thysanoptera, 3, 6–8, 25, 26, 81–82, 106, 124– 125, 164, 166, 174–175, 315, 321, 329, 344, 381, 583, 602, 603, 630–632, 670, 712, 723, 764 Tibia, 77, 151, 158, 174, 311, 360, 370, 371, 382, 549 Tibiaroliate, 108 Tibiotarsus, 174, 212

856 Ticks, 803 Tirathaba rufivena, 196, 369 Tobacco, 49, 52, 130, 140, 382, 484, 485, 490, 597, 615, 645, 650, 651, 711, 714, 810, 813 Tocopheryl acetate, 554 Tomato leafminer, 154, 619 Tomato pinworm, 155, 166, 268, 618, 742, 767, 809 Tomato-potato psyllid (TPP), 640 Tornosvaryella oryzaetora, 31 Toxoderidae, 34 Toxomerus marginatus, 194, 470 Toxomerus occidentalis, 194 Toxoptera citricida, 170, 366 Trameabasilaris burmeisteri, 37 Trameaonusta trans C11, 425 trans C13, 425 trans C13:1, 425 trans C17:1, 425 Transgenic rice, 33 Transmembrane protein, 388 Trechus quadristriatus, 27, 44 Tree canopy, 10 Trehalase, 422, 423, 770 Trialeurodes vaporariorum, 267, 336, 454, 482–484, 671 Triatominae, 123, 400 Tribelocephalinae, 108, 242 Tribolium castaneum, 50, 481, 489, 491, 492, 495, 609, 610 Tribolium confusum, 50, 481, 488, 489, 609, 678 Trichobothria, 81, 313 Trichogramma brassicae, 155, 619 Trichomalopsis oryzae, 29 Trichome, 54, 117, 129, 131, 133, 139, 315, 382–384, 400–402, 586, 614 Trichoplusia ni, 138, 213, 475, 553, 554, 693, 810 Trichoptera, 72, 106, 579, 608 Tricosane, 126, 127 Tridecane, 344, 387, 389 Trifolium pratense, 479, 591 Trifolium repens, 128 Trilophidia annulata, 202, 204, 493 Trithemis aurora, 37 Trithemis pallidinervis, 37 Triticum aestivum, 29, 655 Tri-trophic interactions, 175–177 Tritrophic system, 290, 342, 689 Trochanter, 73, 92 Trochosa ruricola, 109 Trogoderma granarium, 49, 481, 489

Index Trombidiformes, 633 Tropiconabis capsiformis, 20, 154, 203, 479, 650, 657, 801 Tropieonabis eapsiformis, 479 Trypsin, 213, 423, 693 Trypsin inhibitor, 213, 693 Tsuga canadensis, 643, 644, 658 Tubulifera, 82 Tupiocoris cucurbitaceus, 484, 592 Tuta absoluta, 154, 155, 166, 268, 270, 316, 482–485, 552, 592, 615, 618, 638, 639, 672, 714, 742, 767, 769, 809 Two-spotted spider mite, 275, 329, 331, 533, 583, 584, 600, 633, 671, 705, 713 Tyloprobidus variecornis, 597 Tylospilus, 156 Tylotropidus variecornis, 488, 593 Type I, 580–582, 584, 609–613 Type II, 252, 256, 260, 274, 324, 580–588, 590, 592–607, 609–613, 615–618, 647, 653, 749, 769, 770 Type III, 251, 252, 254, 256, 324, 580, 581, 584–587, 590, 592, 600–602, 604–606, 609–611, 613, 614, 619 Type IV, 615, 749 Typhaea stericorea, 481 Tytthaspis sedecimpunctata, 32 Tytthus chinensis, 253, 254 Tytthus mundulus, 158 Tytthus parviceps, 484 U Umbelliferae, 97 Umbonia crassicornis, 395 Unaspis yanonensis, 463 Unfertilized, 116, 162, 197 Urbanization, 58–60 Uresiphita reversalis, 343 Urolepis rufipes, 720 Uroleucon compositae, 136, 152–153, 211, 458 Uroleucon (=Uromelan) jaceae, 456, 457 Urothemis signata, 37 Urtica dioica, 131 Utetheisa ornatrix, 385, 389 Utethesia pulchella, 476 UV-radiation, 236 V Vachiria natolica, 497 Vacuum net method, 10 Various polyhouse, 633 Vegetable crop pests, 155, 252, 482, 667, 679 Vegetable ecosystem, 39, 47

Index Velvetleaf, 132, 138 Venom channel, 413, 424 collection, 417 family, 419, 420 peptidase S10, 420 peptide Pr8a, 420 peptides, 431 proteins, 420, 428, 431, 436 Ptu1, 420, 425, 430, 435, 440 redulysin, 420 RmIT-1, 425, 430 RmIT-2, 425, 430 RmIT-3, 425, 430 serpin, 420 transferrin, 420 triabin 1, 420 Vip3Aa20, 56 Ventrites, 92 Veratridine, 805 Verticillium lecanii, 786, 787, 789 Vesbius sanguinosus, 497 Vespa cincta, 30 Vespa velutina, 466 Vespidae, 3, 4, 27, 28, 30, 39, 40, 73, 74, 92–93, 96, 237, 343, 380, 435, 466, 618 Vespiform thrips, 3 Vespinae, 74 Vespula gennanica, 466 Vespula maculifrons, 466 Vespula rufa consobrina, 466 Vespula vulgaris, 390, 466 Vestula lineaticeps, 497 Vicia glabrescens, 128 Vicia sativa, 128 Vicia villosa, 128 Vigna angularis, 128, 129 Vigna radiate, 128 Vineyards, 14, 49, 53, 59, 681–683, 708 Violaceous, 90, 103, 230–233 Viola philippica, 128 Visual observations, 18 Visual stimuli, 310 Vitellogenesis, 553, 571 Vitellogenic, 553 Vitellogenic follicles, 553 Vitexin, 805 Vitex negundo, 693, 694, 805 Vitex trifolia, 805 Volatile, 125, 126, 242, 290, 341–345, 372, 373, 386, 387, 394, 766, 798, 814 Volatile organic compounds (VOCs), 125–127, 342, 345, 386

857 W Walnut, 7 Wasp, 3, 4, 8, 16, 19, 39, 57, 73, 74, 92, 96, 108, 116, 151, 152, 192, 223, 227, 228, 243, 300, 302, 303, 309, 310, 341, 343, 344, 368, 380, 382, 390, 393, 395, 414, 415, 426, 466, 498, 526, 674, 773, 774, 803 Watermelon, 42, 49, 487, 645, 648 Weight gain, 229, 257, 330, 544, 547, 549–552, 566, 642, 689, 692, 765, 791 Western flower thrips (WFT), 331, 604, 613, 630, 632, 714, 715, 743, 744 Wheat bran, 533 flour, 533, 678 medium, 164, 198 Whitefly, 5, 6, 44, 131, 133, 151, 158, 164, 252, 315, 336, 480, 482–484, 585, 590–592, 601, 607, 613, 616, 618, 629, 630, 637, 638, 640, 645–647, 660, 667, 671, 672, 675, 679, 684, 686, 705, 706, 708, 710–718, 720, 722, 723, 751, 756, 760, 761, 767, 787, 799 Whole gland extraction, 415 Wind velocity, 150, 251, 285–287, 291 Wing buds, 77, 82, 174, 212 Wingless, 74–77, 82, 89, 243, 264 Wing polymorphism aptery, 222 brachyptery, 223, 227, 233 coleoptery, 223 macroptery, 223, 227 microptery, 222–223, 233 staphylinoidy, 223 submacroptery, 223 Wing venation, 79 Wolf spider, 109, 652, 656 Wollastoniella parvicuneis, 670 Wollastoniella rotunda, 669, 670 Woolly apple aphids, 133, 316, 474 X Xylocoris afer, 49 Xylocoris ampoli, 49 Xylocoris clarus, 49 Xylocoris confusus, 49 Xylocoris flavipes, 49, 206, 481, 608–610, 678 Y Yam, 203, 481, 488, 610

858 Yellow-margined, 636, 654, 680 Yellow-pan trap, 16 Yellow stemborer, 461 Yellow sticky traps, 7, 16 Z Zatrephina lineata, 157, 234 Zelus exsanguis, 52, 136 Zelus longipes, 497, 594, 597, 598 Zelus obscuridorsis, 497

Index Zelus pedestris, 497 Zelus renardii, 49, 51, 53, 258, 271, 336, 398, 421, 422, 497, 593, 594, 634, 651, 711 Zelus socius, 52, 497 Zelus tetracanthus, 497 Zicrona caerulea, 477 Zonocerus uariegatus, 497 Zoogeographical regions, 2 Zoophytophagy, 155 Zyxomma petiolatum, 37