Ethnobotany of the Mountain Regions of Mexico 3030993566, 9783030993566

Research in recent years has increasingly shifted away from purely academic research, and into applied aspects of the di

211 63 73MB

English Pages 1580 [1581] Year 2023

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Preface
Acknowledgements
Contents
About the Editors
Contributors
Part I: Regions
Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction
Introduction
The Biocultural Diversity of the Mountains of Mexico
The Book
References
Ethnobotanical Knowledge in Mexico: Use, Management, and Other Interactions Between People and Plants
Introduction
Research Strategy
BADEPLAM
Field Studies
Ethnobotanical Diversity of Mexico
Diversity of Management Forms
Diversity of Domestication Processes
Perspectives
Appendix 1
References
Agroforestry Complexes in the Mountain Regions of Mexico
Introduction
Mexican Mountains, Biocultural Diversity, and Agroforestry Complexes
The Mountains, Their People, and History
Case Studies
Living in the Cold Mountains of Mexico Through Agrosilviculture
Agroforestry and Rituality in Ayuujk Montains
Homegardens and Ornamentality in Cuanajo, Michoacn
Milpas of the Mountain as Agroforestry Complex
Agrosilviculturization in Urban Contexts: Community and Collective Gardens in Peri-Urban Mexico City
Subhumid and Semiarid Mountains of Mexico and Agrosilviculture
Tempered-Subhumid Transition and Agroforestry Management
Agrosylviculture of Palms: Landscape Traditional Management
Interactionships Between Habitants and Diversity in Semiarid Agroforestry Complexes
Metepantles: Agroforestry Terraces and Semiterrace of a Template-Semiarid Zone of Tlaxcala, Mexico
Concluding Remarks
Appendix 1 (IPNI)
Appendix 2
References
Wild, Weedy and Domesticated Plants for Food Security and Sovereignty
Introduction
General Research Methods
La Montaña de Guerrero Region
The Mazahua of the Monarch Butterfly Biosphere Reserve
The Rarmuri of Cuiteco
Peoples of the Tehuacn-Cuicatln Biosphere Reserve
Discussion
Biocultural Diversity and Diversity of Food Patterns
References
Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers, Conduits, and Generators of Plant-People Interactions and...
Introduction
Rarmuri: Guardians of the Sierra Tarahumara
Aspects of Interactions and Relationships Between the Rarmuri and Their Plant-world Environment
Benefits of Ethnobotanical Research in the Sierra Tarahumara
Conclusion
Appendix 1
References
Rarmuri Ethnobotany: Peasant Subsistence and Biodiversity Conservation at Local Scale
Introduction
The Sierra Tarahumara
The Ejido Cuiteco
Research Methods
Plant Species Used in Cuiteco
Environmental Units
Rarmuri Economy and Plant Resources
Timber Commercial Extraction
Firewood
Agriculture
Noncrop Edible Plants
Medicinal
Conclusion
References
Wild Food and Traditional Knowledge of the Kumiai from Baja California
Introduction: Wild Food as Cultural Heritage
From the High Sierra to the Coast: The Wild Food Corridor
Discovering Landscape and Culture Through Wild Foods: Experiences with Plants and Animals
Utensils
From the Sierra: Piñón (Pine Forest)
Jojoba
From the Valleys: Bellota (Acorns)
Wild Leaves
Agaves and Cacti
Uses in Contemporary Food
A Landscape of Opportunities and Diversity to Be Discovered
References
Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre Occidental in Mexico
Introduction: The Wixaritari or Huichol Culture
The Wixaritari Territory
Methodology
Relations Between the Wixaritari and the Plants
Craft Plants
Ceremonial, Sacred, or Religious Use Plants
Edible Plants
Timber Plants
Medicinal Plants
Other Uses
Perspectives
References
Traditional Management and Diversity of Opuntia: General Panorama in Mexico and a Case Study in the Meridional Central Plateau
Introduction
Methods
Literature Review
Site of the Case Study
Ethnobotany of Opuntia and the Nopalera System in the Case Study
Participatory Workshop
Results
Opuntia in Mexican Cultures
Management of Spatial Areas of Opuntia
The Monte or Nopalera
Home Gardens
Traditional Products of Opuntia Species
Prickly Pears and Sweet Products
Queso De Tuna and Melcocha: Opuntia Sweets
Colonche: A Traditional Fermented Juice
Traditional Varieties of Opuntia, Classification, and Attributes
Conclusion
References
Ethnobotanical Knowledge Within the Sierra Gorda, Querétaro, Mexico
Introduction
The Sierra Gorda
Environment
Biological Diversity
Historical and Cultural Context
Anthropogenic Ecosystems: Agroecosystems
Ethnobotanical Categories
Ethnobotanical Knowledge
Plant Uses
Relevant Plant Uses
Plants from Natural Vegetation and Agroecosystems
Discussion and Projections
Appendix
References
Ethnobotany of the Sierra Norte de Puebla
Introduction: The Territory and the People
Ethnobotanical Exploration in the Sierra Norte de Puebla
The Useful Flora and Non-timber Forest Products
Non-timber Forest Products (NTFPs)
Agricultural Systems and Traditional Agriculture
The Milpa
Permanent Parcels (without fallow period)
Milpa in Acahuales or ``Coamil´´ (short, 2 to 5 years fallow periods)
Milpa with Ilite (Alnus spp.) (short, 4 to 5 years fallow)
Slash-and-Burn Milpas (long, 20 to 25 years fallow period)
The Frijolares
The Squashes
The Chilares (chili peppers plots)
The Homegardens
The Coffee Plantations or Cafetales
Perspectives
References
Preserving Healthy Eating Habits: Quelites in the Food System of a Nahua Mountain Community, Mexico
Introduction
Methods
Study Area
Fieldwork
Socioeconomic Data
Data Analysis
Results
General Observations on the Food System
Quelite Species and Sources
Preparation
Frequency and Quantity of Consumption
Relationship Between Sociodemographic Factors and Quelite Knowledge and Use
Quelite Preferences and Knowledge Among Children and Young People
Discussion
Quelite Species
Sources of Quelites
Preparation
Amounts Consumed
Socioeconomic and Sociodemographic Factors and Quelite Knowledge and Consumption
Conclusions and Perspectives
References
Ethnobotany of the Nahua People: Plant Use and Management in the Sierra Negra, Puebla, Mexico
Introduction
Study Site
Interviews, Vegetation, and Ethnobotanical Sampling
Ethnobotany of the Sierra Negra
Importance of Ornamental Plants
Medicinal Plants
Edible Plants and Their Importance in Local Food Patterns
Plants Used as Fuel
Ecological Complementarity and Multiple Use of Ecosystems in Plant Use
Ecological and Sociocultural Strategies Influencing Plant Management
Conclusion
Appendix 1 Plants used by local people in Sierra Negra, Puebla, Mexico
References
Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Ppalo, Oaxaca
Introduction
The Cuicatec
The Community Studied: San Lorenzo Ppalo
Research Methods
Results
Inventory of Useful Plants
Extraction Rates of Plant Resources
Distribution and Abundance of the Main Plant Resources
Discussion and Conclusions
References
Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding the Tehuacán Valley
Introduction
The Ixcatec and Their Territory
A Subsistence Strategy Based on Diversity
The Essential Plants for Livelihood
Nourishment and Edible Plants
Plants and Healthcare
Ceremonial Plants
Ornamental and Luxury Plants
The Feeding and Care of Domestic Animals
Firewood
Housing
Utensils and Other Uses
The Palm: From Mats, Baskets, and Hats to Artisanal Contemporaneous Designs
Maguey as Sources of Sacred and Spirituous Drinks
Conclusion
Appendix 1
References
Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of Craft Traditions and Biocultural Landscapes
Introduction
Theoretical and Methodological Bases
Textiles Representing Nature and Human Linkages
Biocultural Landscapes in Mountainous Regions
Methodological Underpinning for a Participatory and Dialogical Ethnobotany
Landscape Dynamics in the Sierra de Zongolica
Main Historical and Socioeconomic Conditions
Tree Cover Changes as Part of Regional Land Use Changes
Weaving as Part of the Biocultural Landscape
Women Artisans: Weavers of Life
Wool Textile Production Process
Dye Plants
Sheep Wool
Weaving Knowledges
Changes, Innovations, and Losses
Perspectives
References
Ethnobotanical Knowledge and the Patterns of Plant Use and Management in the Sierra de Huautla Biosphere Reserve and the Chich...
Introduction
Sources of Information for the Study of the Ethnobotany of the Mountain Zones of Morelos, Mexico
Ethnobotany of the Sierra de Huautla Biosphere Reserve
Ecological Context
Sociocultural Context of the Peoples in the Sierra de Huautla Biosphere Reserve
Patterns of Use and Management of the Flora of the REBIOSH
Useful Flora
Provenance of Useful Species
Growth Habit
Plant Use Categories
Ecological Status
Management Categories
Use of Medicinal Plants in the REBIOSH: Current Issues and Strategies for Their Sustainable Management
The Traditional Management of Copal: A Case of Sustainable Appropriation of the Tropical Deciduous Forests in the REBIOSH
Ethnobotany of the Chichinautzin Biological Corridor
Ecological Context
Rural Settlements
Ethnobotany of the COBIO
Use of Medicinal Plants in the Chichinautzin Biological Corridor
Threats to the Traditional Knowledge About Plants in the COBIO
Conclusions
References
Mazahua Ethnobotany: Traditional Ecological Knowledge, Management, and Local People Subsistence
Introduction
Materials and Methods
Study Site
Data Collection
Mapping of Environmental Units
Distribution and Abundance of Useful Plants
Temporal Availability of Useful Plants
Results
Environmental Units
Distribution of Useful Plants in the Environmental Units
Temporal Availability of the Main Useful Plants
Conclusions and Perspectives
Appendix 1 Useful plants of the Mazahua village of Francisco Serrato. Habitats: 1 = ruderal; 2 = agricultural areas as weeds; ...
References
Risk Management of Availability of Plant and Fungi Resources Among the Purépecha in Michoacn, Central-Western Mexico
Introduction
Methods
Study Area
Ethnobotanical Studies
Data Analyses
Results
Local Ecological Knowledge of Wild Plants and Mushrooms
Perception on Distribution, Abundance, Extraction, and Motivations for Management
Management of Wild Plants and Mushrooms
Ecological Risk of Wild Plants and Mushrooms
Distribution, Regulations, and Access to Wild Plants and Mushrooms
Discussion
Management and Management Intensity
Rules of Access to Resources as a Factor Motivating Management
Conclusion
References
Traditional Ecological Knowledge and Biodiversity Conservation in the Tierra Caliente Region of Michoacn
Introduction
Methods
Delimitation of the Tierra Caliente Region in Michoacn
Review of Literature, Reports, Databases, and Herbarium Specimens
Case Studies
The Tierra Caliente of Michoacn
Environmental Description
Plant Diversity
The Population of the Tierra Caliente Region in Michoacn
The State of Ethnobotanical Documentation in the Tierra Caliente of Michoacn
The Useful Plant Species of the Tierra Caliente of Michoacn Region
The COINBIO Project and Valuing of Traditional Ecological Knowledge for Conservation
The Trees of the Ejido Huacana: A Collaborative Exercise
The Guacamayas Calentanas A.C. Self-Management Initiative for Biocultural Conservation
Reflections, Perspectives, and Concluding Remarks
References
Ethnobotanical Science in the Clouds: Useful Plants of Northeastern Oaxaca, Mexico
Introduction
Methods
Study Area
Ethnobotanical Information
Species Determination of Plant Specimens
Data Analysis
Results
Geographical and Ecological Aspects of Ethnofloristic Approach Sampling
Taxonomic Aspects of Useful Plants
Traditional Knowledge
Distribution of Traditional Knowledge Among Ethnic Groups
Discussion
Geographical and Ecological Aspects of the Ethno-Floristic Sampling Approach
Taxonomic Aspects of Useful Plants
Traditional Knowledge
References
Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca
Introduction
Ethnobotanical Background for Oaxaca, Sierra Norte, and Sierra Sur
Floristic Diversity in the Mountains of Oaxaca
Outline of Oaxaca´s Cultural Diversity
Biocultural Ethnobotany of the Zapotec Mountains
Zapotec Ethno-Agroecology: Milpas, Home Gardens, and Coffee Plantations in Zoogochí, Sierra Norte
The Milpa
Homegardens
The Coffee Plantation
Gathering and Food Consumption
Complementarity of Mountain Agroecosystems: Las Delicias, Sierra Norte
Agroecosystems
Categories of Plant Use
Mountains Are Living Pharmacies: Coatln, Sierra Sur
Final Comments
Dedication
References
Nopal de Monte: Cacti Named and Used by a Mixtec Community in Mountainous Oaxaca
Introduction
Materials and Methods
Results
Socioeconomic Characteristics of the Collaborators
Traditional Knowledge
Uses and Use Value Index
Discussion
Perspectives
References
Floristic Diversity on Rubber Plantations and Their Importance for Subsistence at Foothill Landscapes of Mexico
Introduction
The Rubber Plantations in the Uxpanapa Valley, Veracruz, Mexico
Methods
Study Area
Vegetation Sampling
Characterization of Rubber Plantation Management
Results
Vegetation and Structure in Rubber Plantations of Uxpanapa
Useful Species in Rubber Plantations
General Characteristics of the Rubber Plantations
Characteristics of Management Practices
Discussion
Perspectives
References
Ethnobotany of the Highlands of Chiapas
Introduction
Background of Ethnobotanical Studies in the Highlands of Chiapas
Forest Management
Useful Plants in the Highlands
Medicinal Ethnobotany in the Highlands of Chiapas
Plants and Worldview
Final Thoughts
Referens
Perspectives of the Ethnobotanical Research in Mexico
Abstract
Ethnobotany in Mexico
Contemporary Theoretical Challenges for Ethnobiological Sciences
Methodological Approaches
Final Comment
References
Part II: Plant Profiles
Agastache spp. Lamiaceae. Important Species of Hyssop in Mexico
Synonyms
Local Names
Botany and Ecology
Phytochemistry
Local Medicinal Uses
Pharmacological Studies
Horticulture and Other Uses
References
Agave americana L. Agave angustifolia Haw. Agave atrovirens Karw. ex Salm-Dyck. Agave asperrima Jacobi. Agave bovicornuta Gent...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Local Fermented Beverage Uses
Local Distilled Beverage Uses
Local Handicraft and Other Uses
Local Construction Uses
Local Fiber and Handicrafts
Appendix 1
References
Amaranthus crassipes Schltdl Amaranthus cruentus L. Amaranthus dubius Mart. ex Thell Amaranthus fimbriatus (Torr.) Benth. ex S...
Synonyms
Local Names
Botany and Ecology
Amaranthus crassipes Schltdl
Amaranthus cruentus L.
Amaranthus dubius Mart. ex Thellung
Amaranthus fimbriatus (Torr.) Benth. ex S. Watson
Amaranthus graecizans L.
Amaranthus hypochondriacus L.
Amaranthus palmeri S. Watson
Amaranthus polygonoides L.
Amaranthus powellii S. Wats
Amaranthus retroflexus L.
Amaranthus scariosus Benth
Amaranthus spinosus L.
Amaranthus viridis L.
Amaranthus watsonii Standley
Local Food Uses
Local Medicinal Uses
Local Handicraft and Other Uses
References
Amphipterygium adstringens (Schltdl.) Standl. Amphipterygium glaucum (Hemsl. & Rose) Hemsl. & Rose Amphipterygium molle (Hemsl...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Socioecological Impact
References
Aristolochia spp. Aristolochiaceae
Aristolochia Species
Synonyms
Local Names*
Botany and Ecology
Aristolochia L.
Aristolochia subg. Aristolochia*
Aristolochia subsection Pentandrae Duch
Aristolochia ser. Hexandrae F. Gonzlez
Aristolochia subser. Hexandrae F. Gonzlez
Aristolochia subser. Anthocaulicae F. Gonzlez
Aristolochia ser. Thyrsicae F. Gonzlez
Aristolochia subg. Siphisia (Duch.) O.C. Schmidt*
Local Medicinal Uses
Local Handicraft and Other Uses
The Phytochemical Paradox of Aristolochia Consumption
Consulted Herbaria
References
Brahea aculeata (Brandegee) H. E. Moore Brahea armata S. Watson Brahea brandegeei (Purpus) H. E. Moore Brahea calcarea Liebm. ...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Biocultural Importance
Local Handicraft and Other Uses
References
Bursera bipinnata (Moc. y Sessé ex DC.) Engl.Bursera copallifera (Sessé & Moc. Ex DC.) BullockBursera fagaroides (Kunth) Engl....
Synonyms
Local Names
Botany and Ecology
Bursera bipinnata
Bursera copallifera
Bursera fagaroides
Bursera glabrifolia
Bursera linanoe
Bursera morelensis
Bursera simaruba
Local Medicinal Uses
Local Food Uses
Local Handicraft and Other Uses
References
Capsicum annuum L. var. annuum Capsicum annuum L. var. glabriusculum (Dunal) Heiser & Pickersgill Capsicum chinense Jacq. Caps...
Synonyms
Local Names
Botany and Ecology
The Genus Capsicum
Local Medicinal Uses
Local Food Uses
Local Handicraft and Other Uses
References
Chamaedorea Willd. Chamaedorea alternans H. Wendl. Chamaedorea arenbergiana H. Wendl. Chamaedorea cataractarum Mart. Chamaedor...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Local Handicraft Uses
Local Ornamental and Marketable Uses
Local Ceremonial Uses
References
Crescentia alata Kunth Crescentia cujete L. Bignoniaceae
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Local Handicraft and Other Uses
References
Cucurbita argyrosperma Huber ssp. argyrosperma Cucurbita argyrosperma Huber ssp. sororia (L.H. Bailey) Merrick and Bates Cucur...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Other Uses
References
Escontria chiotilla (F. A. C. Weber ex K. Schum.) Rose Cactaceae
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Management and Domestication
References
Gaultheria acuminata Schltdl. and Cham. Ericaceae
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Ritual Uses
Management
Socioeconomic Importance
References
Gossypium aridum (Rose y Standley ex Rose) Skovsted Gossypium armourianum KearneyGossypium barbadense Linnaeus Gossypium david...
Synonyms
Local Names
Botany and Ecology
Subgenus Houzingenia: American Diploid Wild Species
Section Houzingenia
Subsection Houzingenia
Subsection Integrifolia
Subsection Caducibracteolata
Section Erioxylum
Subsection Erioxylum
Subsection Selera
Subgenus Karpas: American Tetraploid Species
Local Uses
Subgenus Houzingenia
American Domesticated Cottons
Gossypium barbadense
Gossypium hirsutum
Conservation
Final Remarks
References
Hintonia latiflora (Sessé et Mociño ex DC.) Bullock Hintonia octomera (Hemsl.) Bullock Hintonia standleyana Bullock Rubiaceae
Synonyms
Local Names
Botany and Ecology
Phytochemistry
Local Medicinal Uses
Local Handicraft and Other Uses
Socioecological Impact
References
Litsea glaucescens Kunth Lauraceae
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Biocultural Importance
Local Handicrafts and Other Uses
References
Magnolia mexicana DC. Magnoliaceae
Synonyms
Local Names
Botany and Ecology
Conservation
Local Medicinal Uses
Local Food Uses
Local Handicraft and Other
Economic Aspects
References
Melothria pendula L. Melothria pringlei (S.Watson) Mart.Crov. Melothria trilobata Cogn. Cucurbitaceae
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Local Handicraft and Other Uses
Management
References
Myrtillocactus Console Myrtillocactus cochal (Orcutt) Britton & Rose Myrtillocactus eichlamii Britton & Rose Myrtillocactus ge...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Local Handicraft and Other Uses
Management and Domestication
References
Nicotiana alata Link & Otto Nicotiana attenuata Torr. ex S. Watson Nicotiana clevelandii A. Gray Nicotiana glauca Graham Nicot...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Handicraft and Other Uses
References
Phaseolus acutifolius A. Gray Phaseolus coccineus L. Phaseolus dumosus Macfad. Phaseolus filiformis Benth. Phaseolus glabellus...
Synonyms
Local Names
Botany and Ecology
Phaseolus acutifolius A. Gray
Phaseolus coccineus L.
Phaseolus dumosus Macfadyen
Phaseolus filiformis Benth.
Phaseolus glabellus Piper
Phaseolus leptostachyus Benth.
Phaseolus lunatus L.
Phaseolus maculatus Scheele
Phaseolus ritensis M. E. Jones
Phaseolus salicifolius Piper
Phaseolus vulgaris L.
Local Food Uses
Local Medicinal Uses
Local Handicraft and Other Uses
References
Physalis philadelphica Lamarck Physalis angulata L. Physalis chenopodifolia Lamarck Physalis cinerascens (Dunal) C. L. Hitchk....
Synonyms
Local Names
Botany and Ecology
Physalis philadelphica Lamarck
Physalis angulata L.
Physalis chenopodifolia Lamarck
Physalis cinerascens (Dunal) C. L. Hitchck. var. cinerascens
Physalis pubescens L.
Physalis acutifolia (Miers) Sandith
Physalis coztomatl Moc. and Sessé ex Dunal
Local Medicinal Uses
Local Food Uses
Local Handicraft and Other Uses
References
Polaskia chichipe (Rol.-Goss.) Backeb. Polaskia chende (Gosselin) A.C. Gibson and K.E. Horak Cactaceae
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Management and Domestication
References
Sabal mauritiiformis (H. Kars.) Griseb. & H. Wendl. Sabal mexicana Mart. Sabal pumos (Kunth) Burret Sabal uresana Trel. Sabal ...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Local Construction, Handicraft and Other Uses
References
Sechium chinantlense Lira & Chiang. Sechium compositum (J.D. Smith) C. Jeffrey Sechium edule ssp. edule Sechium edule ssp. syl...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Other Uses
References
Solanum americanum Mill. Solanum cardiophyllum Lindl. Solanum demissum Lindl. Solanum douglasii Dunal Solanum elae agnifolium ...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Other Uses
References
Spondias mombin L. Spondias purpurea L. Spondias radlkoferi J. D. Smith Anacardiaceae
Synonyms
Local Names
Botany and Ecology
Spondias mombin L.
Spondias purpurea L.
Spondias radlkoferi J.D. Smith
Local Medicinal Uses
Spondias mombin L.
Spondias purpurea L.
Spondias radlkoferi J.D. Smith
Local Food Uses
Spondias mombin L.
Spondias purpurea L.
Spondias radlkoferi J.D. Smith
Biocultural Importance
Spondias spp.
Spondias mombin L.
Spondias purpurea L.
Spondias radlkoferi J.D. Smith
References
Stenocereus (A. Berger) Riccob. Cactaceae Stenocereus alamosensis (J.M. Coult.) A.C. Gibson and K.E. Horak Stenocereus benecke...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Management and Domestication
References
Vanilla cribbiana Soto Arenas Vanilla hartii Rolfe Vanilla helleri A.D. Hawkes Vanilla inodora Schiede Vanilla insignis Ames V...
Synonyms
Local Names
Botany and Ecology
Local Medicinal Uses
Local Food Uses
Other Local Uses
Management
Conservation
Conclusions
References
Zea diploperennis Iltis, Doebley, n Zea perennis (Hitchc.) Reeves & Mangelsd. Zea luxurians (Durieu & Asch.) R.M. Bird Zea nic...
Synonyms
Local Names
Botany and Ecology
Uses
References
Recommend Papers

Ethnobotany of the Mountain Regions of Mexico
 3030993566, 9783030993566

  • 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

Ethnobotany of Mountain Regions Series Editors: Rainer W. Bussmann Narel Y. Paniagua-Zambrana

Alejandro Casas José Juan Blancas Vázquez Editors

Ethnobotany of the Mountain Regions of Mexico

Ethnobotany of Mountain Regions Series Editors Rainer W. Bussmann, Department of Ethnobotany, Institute of Botany and Bakuriani Alpine Botanical Garden, Ilia State University, Tbilisi, Georgia Narel Y. Paniagua-Zambrana, Herbario Nacional de Bolivia, Saving Knowledge, Universidad Mayor de San Andres, La Paz, Bolivia

Ethnobotanical research in recent years has increasingly shifted into applied aspects of the discipline, including climate change research, conservation, and sustainable development. It has by now widely been recognized that “traditional” knowledge is always in flux and adapting to a quickly changing environment. Trends of globalization, especially the globalization of plant markets, have greatly influenced how plant resources are managed nowadays. While ethnobotanical studies are now available from many regions of the world, no comprehensive encyclopedic series focusing on the worlds mountain regions is available in the market. Scholars in plant sciences worldwide will be interested in this website and its dynamic content. The field (and thus the market) of ethnobotany and ethnopharmacology has grown considerably in recent years. Student interest is on the rise, attendance at professional conferences has grown steadily, and the number of professionals calling themselves ethnobotanists has increased significantly (the various societies—Society for Economic Botany, International Society of Ethnopharmacology, Society of Ethnobiology, International Society for Ethnobiology, and many regional and national societies in the field currently have thousands of members). Growth has been most robust in BRIC countries. The objective of this new series on Ethnobotany of Mountain Regions is to take advantage of the increasing international interest and scholarship in the field of mountain research. We anticipate including the best and latest research on a full range of descriptive, methodological, theoretical, and applied research on the most important plants for each region. Each contribution will be scientifically rigorous and contribute to the overall field of study.

Alejandro Casas • Jose´ Juan Blancas Vázquez Editors

Ethnobotany of the Mountain Regions of Mexico With 527 Figures and 97 Tables

Editors Alejandro Casas Instituto de Investigaciones en Ecosistemas y Sustentabilidad Universidad Nacional Autónoma de México Morelia, Michoacán, Mexico

José Juan Blancas Vázquez Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos Cuernavaca, Morelos, Mexico

ISSN 2523-7489 ISSN 2523-7497 (electronic) Ethnobotany of Mountain Regions ISBN 978-3-030-99356-6 ISBN 978-3-030-99357-3 (eBook) https://doi.org/10.1007/978-3-030-99357-3 © Springer Nature Switzerland AG 2023 This work is subject to copyright. All rights are reserved 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 Switzerland AG. The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

To the Indigenous and mestizo peoples of rural Mexico, whose knowledge and experience are the bases of the studies contained in this work. To those who initiated the studies of ethnobotany in Mexico, especially Efraim Hernández-Xolocotzi, Arturo Gómez-Pompa, Víctor M. Toledo, Miguel Ángel Martínez-Alfaro, Montserrat Gispert-Cruells, Salvador Flores-Guido, Javier Caballero-Nieto, Robert Bye-Boettler, and José Sarukhán-Kermez. They established the bases of different research approaches and are crucial pillars that performed and inspired the studies reported in this volume. Our deepest recognition and gratitude to Javier Caballero-Nieto, who started with us the general plan of this work. Let this volume be a tribute to his memory. Alejandro Casas and José Blancas

Preface

Presence of humans in the Americas was recently discovered to be much older than previously considered. Several studies in North and South America have accumulated evidence that Homo sapiens was present in this region more than 30,000 years ago, and that several waves of migration between the late Pleistocene and early Holocene took place. The earliest human occupation of the territory that today is called Mexico appears to have occurred about 25,000 years ago, starting interactions with local environments, exploring the attributes of their components, and using them to satisfy their subsistence needs. Such interactions are the roots of the extraordinary knowledge about the diverse Mexican flora, which, through such a long history, has left its mark in the current ethnobotanical knowledge occurring in the high diverse biocultural settings that compose Mexico. In 2019 we were invited by Dr. Rainer Bussmann, coordinator of the series Ethnobotany of the Mountain Regions, to be editors of a compendium about the ethnobotany of the mountains of Mexico. We, together with Dr. Javier Caballero, considered the proposal as a pertinent way to complement the previous work Ethnobotany of Mexico. Interactions of People and Plants in Mesoamerica that we edited in 2016 together with Dr. Rafael Lira. We dedicated three years to this project, with the unfortunate death of our partner Javier Caballero, and in the middle of the dramatic COVID-19 pandemic. We accepted the challenge to complete the work in this adverse time, convinced that the series offered the opportunity to continue showing to the world the ethnobotanical richness of Mexico, an eminent mountainous region of the world. Ethnobotanists working in Mexico have recorded nearly 8,000 species of plants Mexican cultures interact with, and we estimate that the universe could be more than 11,000, a figure that illustrates the high biocultural diversity of this country. It is an extensive work, summarized in two volumes resulting from a hard effort of numerous researchers. It aspires to provide a general panorama of the cultures and regions that have been investigated, a sample of representative plant groups that have influenced the cultures of this country, and reflections about the perspectives of ethnobotany in connection with other disciplines and the theoretical construction that are in progress and those that are needed. We organized the volumes in two parts, one dedicated to general aspects of ethnobotanical research and case studies of particular regions and cultural groups, vii

viii

Preface

the other dedicated to plant groups meaningful for Mexican cultures. However, the current volumes and that published in 2016 continue to be a fraction of what is known and that that is to be investigated about the ethnobotany of Mexico. We therefore consider this as a new step of several that are needed to complete the work. Ethnobotany of the Mountain Regions of Mexico is a work that provides information and reflections. We aspire to contribute to specialized and non-specialized readers interesting issues to value the enormous biocultural heritage, routes for continuing studying it, and some directions for its conservation. Alejandro Casas José Juan Blancas Vázquez Editors

Acknowledgements

The editors of this work thank to Rainer W. Bussmann and Narel Y. PaniaguaZambrana, general coordinators of the series Ethnobotany of the Mountain Regions. He invited us to undertake this challenge, helped very much in delimiting Mexico, a vast biocultural setting, as the universe of the studies reported. Rainer and his team made careful review of the manuscripts composing the opus and provided valuable suggestions. We thank the collaboration of many persons in reviewing and editing the innumerable details of edition and images. The collaboration of Itzel Abad-Fitz in reviewing the format of bibliography of a number of chapters, as well as the translation and copy editing work of texts by Sergio Zárate and Mathew Rose, was invaluable. Other collaborations were possible thanks to the PAPIIT-DGAPA of the National Autonomous University of Mexico (UNAM, IN206520 and IN224023) and the National Council of Science and Technology of Mexico (CONACYT, A1-S-14306) We thank the National Commission for the Knowledge and Use of Biodiversity of Mexico (CONABIO), FAO GEF (Project ID 9380) for the support to the project Mexican Agrobiodiversity. We especially thank for support to the project “Manejo y domesticación de agrobiodiversidad en Mesoamérica: Bases para la soberanía alimentaria sustentable” (RG023). Our gratitude to the editorial team of Springer, especially Sylvia Blago and Jawahar Babu, for their continual support throughout the long production process of this work.

ix

Contents

Volume 1 Part I

Regions

...........................................

1

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Alejandro Casas and José Juan Blancas Vázquez

3

Ethnobotanical Knowledge in Mexico: Use, Management, and Other Interactions Between People and Plants . . . . . . . . . . . . . . . . . . . . . . . . . Javier Caballero, Laura Cortés-Zárraga, Cristina Mapes, José Juan Blancas Vázquez, Selene Rangel-Landa, Ignacio Torres-García, Berenice Farfán-Heredia, Andrea Martínez-Ballesté, and Alejandro Casas Agroforestry Complexes in the Mountain Regions of Mexico . . . . . . . . Ana Isabel Moreno-Calles, Gerardo Hernández-Cendejas, Wilfrido López-Martínez, Alexis Daniela Rivero-Romero, Yessica Angélica Romero-Bautista, Karla Guzmán-Fernández, Ana Mitzi García-Leal, Ernesto Gutiérrez-Coatecatl, Cloe X. Pérez-Valladares, Ana Rojas-Rosas, Ignacio Torres-García, and Selene Rangel-Landa Wild, Weedy and Domesticated Plants for Food Security and Sovereignty . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Alejandro Casas, Berenice Farfán-Heredia, Andrés Camou-Guerrero, Ignacio Torres-García, José Juan Blancas Vázquez, and Selene Rangel-Landa Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers, Conduits, and Generators of Plant-People Interactions and Relationships . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Robert Bye and Edelmira Linares

25

65

97

129

xi

xii

Contents

Rarámuri Ethnobotany: Peasant Subsistence and Biodiversity Conservation at Local Scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Andrés Camou-Guerrero, Juan Vega, María Teresa Guerrero-Olivares, and Alejandro Casas Wild Food and Traditional Knowledge of the Kumiai from Baja California . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Carolina Gutiérrez-Sánchez and Claudia Leyva-Aguilera Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre Occidental in Mexico . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Martha Cedano-Maldonado, Luis Villaseñor-Ibarra, and Mara Ximena Haro-Luna

149

169

197

Traditional Management and Diversity of Opuntia: General Panorama in Mexico and a Case Study in the Meridional Central Plateau . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Amaranta Paz-Navarro, César I. Ojeda-Linares, Gonzalo D. Álvarez-Ríos, Mariana Vallejo, and Alejandro Casas

233

Ethnobotanical Knowledge Within the Sierra Gorda, Querétaro, Mexico . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Luis Hernández-Sandoval and Hugo Castillo-Gómez

265

Ethnobotany of the Sierra Norte de Puebla . . . . . . . . . . . . . . . . . . . . . . Francisco Basurto, Cristina Mapes, Tania Escobar, and Juan Carlos González Preserving Healthy Eating Habits: Quelites in the Food System of a Nahua Mountain Community, Mexico . . . . . . . . . . . . . . . . . . . . . . Claudia Sánchez-Ramos, Heike Vibrans, María Rivas-Guevara, Edelmira Linares, Edmundo García-Moya, and Alfredo Saynes-Vásquez Ethnobotany of the Nahua People: Plant Use and Management in the Sierra Negra, Puebla, Mexico . . . . . . . . . . . . . . . . . . . . . . . . . . . . José Juan Blancas Vázquez, Alejandro Casas, Hilda Ramírez-Monjaraz, Andrea Martínez-Ballesté, Ignacio Torres-García, Itzel Abad-Fitz, Leonardo Beltrán-Rodríguez, Carolina Larios, Aketzalli Olvera-Espinosa, Myriam A. Miranda-Gamboa, Elisa Lotero, and Mariana Vallejo

401

431

453

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pápalo, Oaxaca . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Leonor Solís and Alejandro Casas

517

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding the Tehuacán Valley . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Selene Rangel-Landa and Alejandro Casas

555

Contents

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of Craft Traditions and Biocultural Landscapes . . . . . . . . . . . . . . . . . . Citlalli López-Binnqüist, Belinda Contreras-Jaimes, Fortunata Panzo-Panzo, and Edward A. Ellis Ethnobotanical Knowledge and the Patterns of Plant Use and Management in the Sierra de Huautla Biosphere Reserve and the Chichinautzin Biological Corridor in Morelos, Mexico . . . . . . . . . . José Juan Blancas Vázquez, Araceli Tegoma-Coloreano, Itzel Abad-Fitz, Leonardo Beltrán-Rodríguez, Belinda Maldonado-Almanza, María Idalia Villalpando-Toledo, Fabiola Mena, Angélica Alemán, and Amanda Ortiz-Sánchez

xiii

593

623

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management, and Local People Subsistence . . . . . . . . . . . . . . . . . . . . . Berenice Farfán-Heredia and Alejandro Casas

663

Risk Management of Availability of Plant and Fungi Resources Among the Purépecha in Michoacán, Central-Western Mexico . . . . . . Berenice Farfán-Heredia and Alejandro Casas

693

Traditional Ecological Knowledge and Biodiversity Conservation in the Tierra Caliente Region of Michoacán . . . . . . . . . . . . . . . . . . . . . Selene Rangel-Landa, María Elizabeth Saucedo-Gudiño, Erandi Lizbeth Guzmán-Gómez, Maria Fernanda Salazar-Ramirez, Arnulfo Blanco-García, Delia Caldera-Cano, Aglaen Lucero Carbajal-Navarro, Rosendo Caro-Gómez, Andrea Ponce-Rangel, José Isabel Texta-Hernández, and Xavier Madrigal-Sánchez Ethnobotanical Science in the Clouds: Useful Plants of Northeastern Oaxaca, Mexico . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Beatriz Rendón-Aguilar, David Bravo-Avilez, Luis Alberto Bernal-Ramírez, Abisaí García-Mendoza, Adolfo Espejo-Serna, Ana Rosa López-Ferrari, Carlos Durán-Espinosa, David S. Gernandt, Francisco Lorea-Hernández, Guillermo Ibarra-Manríquez, Jaime Jiménez-Ramírez, Jesús Ricardo de Santiago-Gómez, Jorge Santana-Carrillo, José Luis Villaseñor, Laura Yáñez-Espinosa, Lucio Lozada-Pérez, Marie-Stéphanie Samain, Susana Valencia-Ávalos, Rosa María Fonseca-Juárez, and Salvador Arias-Montes Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca . . . . . . Marco Antonio Vásquez-Dávila, Gladys I. Manzanero-Medina, Adonicam Santiago-Martínez, and Sunem Pascual-Mendoza Nopal de Monte: Cacti Named and Used by a Mixtec Community in Mountainous Oaxaca . . . . . . . . . . . . . . . . . . . . . . . . . . . Luis E. Ortiz-Martínez, Gladys I. Manzanero-Medina, Jordan Golubov, Marco Antonio Vásquez-Dávila, and María C. Mandujano

721

755

871

899

xiv

Contents

Floristic Diversity on Rubber Plantations and Their Importance for Subsistence at Foothill Landscapes of Mexico . . . . . . . . . . . . . . . . . Juan Carlos López-Acosta, Emmanuel Ismael Pantoja-Aparicio, Jorge Antonio Gómez-Díaz, Maite Lascurain-Rangel, and Ina Falfán

909

Ethnobotany of the Highlands of Chiapas . . . . . . . . . . . . . . . . . . . . . . . Felipe Ruan-Soto, Fausto Bolom-Ton, Eréndira J. Cano-Contreras, Lucia Domínguez-Torres, Fernando Guerrero-Martínez, and Ramón Mariaca-Méndez

929

Perspectives of the Ethnobotanical Research in Mexico . . . . . . . . . . . . Alejandro Casas, José Juan Blancas Vázquez, and Heike Vibrans

953

Volume 2 Part II

Plant Profiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Agastache spp. LAMIACEAE. Important species of Hyssop in Mexico . . . Guadalupe Carrillo-Galván and Robert Bye Agave americana L. Agave angustifolia Haw. Agave atrovirens Karw. ex Salm-Dyck. Agave asperrima Jacobi. Agave bovicornuta Gentry. Agave cupreata Trel. & A. Berger. Agave hookeri Jacobi. Agave inaequidens K. Koch. Agave karwinskii Zucc. Agave kerchovei Lem. Agave lechuguilla Torr. Agave mapisaga Trel. Agave marmorata Roezl Agave maximiliana Baker Agave montana Villareal Agave potatorum Zucc. Agave rhodacantha Trel. Agave salmiana Otto ex Salm-Dyck Agave scaposa Gentry Agave tequilana F.A.C. Weber Agave victoriae-reginae A. Berger ASPARAGACEAE . . . . . . Ignacio Torres-García, América Minerva Delgado-Lemus, Alejandro Casas, Gonzalo D. Álvarez-Ríos, Selene Rangel-Landa, Raymundo Martínez-Jiménez, Carmen Julia Figueredo-Urbina, Ofelia Vargas-Ponce, Guadalupe Casarrubias-Hernández, Oassis Huerta-Galván, Dánae Cabrera-Toledo, and Nancy Vázquez-Pérez

981 983

997

Amaranthus crassipes Schltdl Amaranthus cruentus L. Amaranthus dubius Mart. ex Thell Amaranthus fimbriatus (Torr.) Benth. ex S. Watson Amaranthus graecizans L. Amaranthus hybridus L. Amaranthus hypochondriacus L. Amaranthus palmeri S. Watson Amaranthus polygonoides L. Amaranthus powellii S. Amaranthus retroflexus L. Amaranthus spinosus L. Amaranthus viridis L. Amaranthus watsonii Standl AMARANTHACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1049 Cristina Mapes, Ángel Mujica-Sánchez, and Laura Cortés-Zárraga

Contents

xv

Amphipterygium adstringens (Schltdl.) Standl. Amphipterygium glaucum (Hemsl. & Rose) Hemsl. & Rose Amphipterygium molle (Hemsl.) Hemsl. & Rose Amphipterygium simplicifolium (Standl.) Cuev.-Fig. ANACARDIACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1067 Leonardo Beltrán-Rodríguez and Robert Bye Aristolochia spp. ARISTOLOCHIACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1081 Anna Paizanni-Guillén and David Douterlungne-Rotsaert Brahea aculeata (Brandegee) H. E. Moore Brahea armata S. Watson Brahea brandegeei (Purpus) H. E. Moore Brahea calcarea Liebm. Brahea decumbens Rzed. Brahea dulcis (Kunth) Mart. Brahea edulis H. Wendl. ex S. Watson Brahea moorei L. H. Bailey ex H. E. Moore Brahea pimo Becc. ARECACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1139 Viviana Andrade-Erazo, Cloe X. Pérez-Valladares, and Selene Rangel-Landa Bursera bipinnata (Moc. y Sessé ex DC.) Engl. Bursera copallifera (Sessé & Moc. Ex DC.) Bullock Bursera fagaroides (Kunth) Engl. Bursera glabrifolia (Kunth) Engl. Bursera linanoe (La Llave) Rzed., Calderón, and Medina Bursera morelensis Ramírez Bursera simaruba (L.) Sarg. BURSERACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1155 José Juan Blancas Vázquez, Itzel Abad-Fitz, Leonardo Beltrán-Rodríguez, Alejandro Casas, Belinda Maldonado-Almanza, José Antonio Sierra-Huelsz, Feliciano García-Lara, Laura Cortés-Zárraga, Fabiola Mena, and María Inés Ayala-Enríquez Capsicum annuum L. var. annuum Capsicum annuum L. var. glabriusculum (Dunal) Heiser & Pickersgill Capsicum chinense Jacq. Capsicum frutescens L. Capsicum lanceolatum (Greenm.) C.V. Morton & Standley Capsicum pubescens Ruiz & Pav. Capsicum rhomboideum (Dunal) Kuntze SOLANACEAE . . . . . . . . . . . . . . . . . . . . . . . 1179 Araceli Aguilar-Meléndez, Esther Katz, Marco Antonio Vásquez-Dávila, and Gloria E. Barboza

xvi

Contents

Chamaedorea Willd. Chamaedorea alternans H. Wendl. Chamaedorea arenbergiana H. Wendl. Chamaedorea cataractarum Mart. Chamaedorea elatior Mart. Chamaedorea elegans Mart. Chamaedorea ernesti-augusti H. Wendl. Chamaedorea geonomiformis H. Wendl. Chamaedorea graminifolia H. Wendl. Chamaedorea hooperiana Hodel Chamaedorea liebmannii Mart. Chamaedorea metallica O.F.Cook ex H.E.Moore Chamaedorea oblongata Mart. Chamaedorea pinnatifrons (Jacq.) Oerst. Chamaedorea pochutlensis Liebm. Chamaedorea radicalis Mart. Chamaedorea sartorii Liebm. Chamaedorea schiedeana Mart. Chamaedorea seifrizii Burret Chamaedorea tepejilote Liebm. ARECACEAE . . . . . . . . . . . . . . . . . . . . . . . 1197 Viviana Andrade-Erazo and Myriam A. Miranda-Gamboa Crescentia alata Kunth Crescentia cujete L. BIGNONIACEAE . . . . . . . . . . . 1225 Xitlali Aguirre-Dugua and Alejandro Casas Cucurbita argyrosperma Huber ssp. argyrosperma Cucurbita argyrosperma Huber ssp. sororia (L.H. Bailey) Merrick and Bates Cucurbita ficifolia Bouché Cucurbita moschata (Duchesne ex Lam.) Duchesne ex Poir. Cucurbita pepo L. ssp. pepo Cucurbita pepo L. ssp. fraterna (Bailey) Andres CUCURBITACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1235 Rafel Lira, Luis E. Eguiarte, Salvador Montes-Hernández, and Isela Rodríguez-Arévalo Escontria chiotilla (F. A. C. Weber ex K. Schum.) Rose CACTACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1253 Alejandro Casas Gaultheria acuminata Schltdl. and Cham. ERICACEAE . . . . . . . . . . . . . . 1267 A. Carolina Elizondo-Salas, M. Fernanda Morán-Tzópitl, and David Jimeno-Sevilla Gossypium aridum (Rose y Standley ex Rose) Skovsted Gossypium armourianum Kearney Gossypium barbadense Linnaeus Gossypium davidsonii Kellogg Gossypium gossypioides (Ulbrich) Standley Gossypium harknessii Brandegee Gossypium hirsutum Linnaeus Gossypium laxum Phillips Gossypium lobatum Gentry Gossypium schwendimanii Fryxell and S. Koch Gossypium thurberi Todaro Gossypium trilobum Skovsted Gossypium turneri Fryxell MALVACEAE . . . . . . . . . . . . . . . . . . . 1273 Valeria Alavez, Melania Vega, Alejandra Gutiérrez-Cedillo, Rodrigo Hernández-Pacheco, and Ana Wegier

Contents

xvii

Hintonia latiflora (Sessé et Mociño ex DC.) Bullock Hintonia octomera (Hemsl.) Bullock Hintonia standleyana Bullock RUBIACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1305 Leonardo Beltrán-Rodríguez, Sol Cristians, and Robert Bye Litsea glaucescens Kunth LAURACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . 1317 Daniela Ortega-Meza, María Teresa Pulido-Silva, José Juan Blancas Vázquez, and Carolina J. Da Silva Magnolia mexicana DC. MAGNOLIACEAE . . . . . . . . . . . . . . . . . . . . . . . . . 1327 David Jimeno-Sevilla and A. Carolina Elizondo-Salas Melothria pendula L. Melothria pringlei (S.Watson) Mart.Crov. Melothria trilobata Cogn. CUCURBITACEAE . . . . . . . . . . . . . . . . . . . . . . . . 1335 Paulina Guerrero-Torres, Luis Hernández-Sandoval, and Alejandro Casas Myrtillocactus Console Myrtillocactus cochal (Orcutt) Britton & Rose Myrtillocactus eichlamii Britton & Rose Myrtillocactus geometrizans (Mart. ex Pfeiff.) Console Myrtillocactus schenckii (J.A. Purpus) Britton & Rose CACTACEAE . . . . . . . . . . . . . . . . . . . . . . . . 1349 José Juan Blancas Vázquez and Alejandro Casas Nicotiana alata Link & Otto Nicotiana attenuata Torr. ex S. Watson Nicotiana clevelandii A. Gray Nicotiana glauca Graham Nicotiana longiflora Cav. Nicotiana obtusifolia M. Martens & Galeotti Nicotiana plumbaginifolia Viv. Nicotiana rustica L. Nicotiana tabacum L. SOLANACEAE . . . . . . . . . . . . . 1369 Marco Antonio Vásquez-Dávila, Gladys I. Manzanero-Medina, P. Pelayo-Delgado, and Araceli Aguilar-Meléndez Phaseolus acutifolius A. Gray Phaseolus coccineus L. Phaseolus dumosus Macfad. Phaseolus filiformis Benth. Phaseolus glabellus Piper Phaseolus leptostachyus Benth. Phaseolus lunatus L. Phaseolus maculatus Scheele Phaseolus ritensis M. E. Jones Phaseolus salicifolius Piper Phaseolus vulgaris L. LEGUMINOSAE/FABACEAE . . . . . . . . . . . . . . . . . . . . . 1383 Alfonso Delgado-Salinas and Leticia Torres-Colín Physalis philadelphica Lamarck Physalis angulata L. Physalis chenopodifolia Lamarck Physalis cinerascens (Dunal) C. L. Hitchk. Physalis pubescens L. Physalis acutifolia (Miers) Sandwith Physalis coztomatl Dunal SOLANACEAE . . . . . . . . . . . . 1407 Mahinda Martínez, Ofelia Vargas-Ponce, and Pilar Zamora-Tavares Polaskia chichipe (Rol.-Goss.) Backeb. Polaskia chende (Gosselin) A.C. Gibson and K.E. Horak CACTACEAE . . . . . . . . . . . . . . . . . . . . . . . . 1419 Alejandro Casas

xviii

Contents

Sabal mauritiiformis (H. Kars.) Griseb. & H. Wendl. Sabal mexicana Mart. Sabal pumos (Kunth) Burret Sabal uresana Trel. Sabal yapa C. Wright ex Becc. ARECACEAE . . . . . . . 1437 Andrea Martínez-Ballesté, María Teresa Pulido-Silva, and Laura Cortés-Zárraga Sechium chinantlense Lira & Chiang. Sechium compositum (J.D. Smith) C. Jeffrey Sechium edule ssp. edule Sechium edule ssp. sylvestre Lira & Castrejón Sechium hintonii (P.G. Wilson) C. Jeffrey CUCURBITACEAE . . . . . . . . . . . . . . . . . . . . . . . . . 1447 Rafel Lira, Luis E. Eguiarte, Salvador Montes-Hernández, and Isela Rodríguez-Arévalo Solanum americanum Mill. Solanum cardiophyllum Lindl. Solanum demissum Lindl. Solanum douglasii Dunal Solanum elaeagnifolium Cav. Solanum ehrenbergii (Bitter) Rydb. Solanum houstonii Martyn Solanum lycopersicum L. Solanum nigrescens M. Martens & Galeotti Solanum pubigerum Dunal Solanum rostratum Dunal Solanum rudepannum Dunal Solanum stoloniferum Schltdl. Solanum umbellatum Mill. Solanum verrucosum Schltdl. SOLANACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1461 Ignacio Torres-García, América Minerva Delgado-Lemus, and Alejandro Casas Spondias mombin L. Spondias purpurea L. Spondias radlkoferi J. D. Smith ANACARDIACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1483 María del Rocío Ruenes-Morales, Patricia Irene Montañez-Escalante, Miriam Monserrat Ferrer, Juan José Jiménez-Osornio, Emiliano González Iturbe-Ruenes, and José Antonio González Iturbe-Ahumada

Contents

xix

Stenocereus (A. Berger) Riccob. Stenocereus alamosensis (J.M. Coult.) A.C. Gibson and K.E. Horak Stenocereus beneckei (Ehrenb.) A. Berger and Buxb. Stenocereus chacalapensis (Bravo & T. MacDoug.) Buxb. Stenocereus chrysocarpus Sánchez-Mej. Stenocereus dumortieri (Scheidw.) Buxb. Stenocereus fricii Sánchez-Mej. Stenocereus griseus (Haw.) Buxb. Stenocereus gummosus (Engelm.) A. Gibson and K.E. Horak Stenocereus heptagonus (L.) Mottram Stenocereus huastecorum Alvarado-Sizzo, Arreola-Nava y Terrazas Stenocereus laevigatus (Salm-Dyck) Buxb. Stenocereus littoralis (K. Brandegee) L.W. Lenz Stenocereus martinezii (J.G. Ortega) Buxb. Stenocereus montanus (Britton & Rose) Buxb. Stenocereus pruinosus (Otto ex Pfeiff.) Buxb. Stenocereus queretaroensis (F.A.C. Weber ex Mathes.) Buxb. Stenocereus quevedonis (J. G. Ortega) Buxb. Stenocereus standleyi (J.G. Ortega) Buxb. Stenocereus stellatus (Pfeiff.) Riccob. Stenocereus thurberi (Engelm.) Buxb. Stenocereus treleasei (Rose) Backeb. CACTACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1501 Hernán Alvarado-Sizzo and Alejandro Casas Vanilla cribbiana Soto Arenas Vanilla hartii Rolfe Vanilla helleri A.D. Hawkes Vanilla inodora Schiede Vanilla insignis Ames Vanilla odorata C. Presl Vanilla phaeantha Rchb. f. Vanilla planifolia Jacks ex. Andrews Vanilla pompona Schiede ORCHIDACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1547 Melania Vega, Lislie Solís-Montero, Valeria Alavez, Pamela Rodríguez-Juárez, Manuel Gutiérrez-Alejo, and Ana Wegier Zea diploperennis Iltis, Doebley, & R. Guzmán Zea perennis (Hitchc.) Reeves & Mangelsd. Zea luxurians (Durieu & Asch.) R.M. Bird Zea nicaraguensis Iltis & B.F. Benz Zea vespertilio Gómez-Laur. Zea mays subsp. huehuetenangensis (Iltis & Doebley) Doebley Zea mays subsp. mexicana (Schrad.) Iltis Zea mays subsp. parviglumis Iltis & Doebley Zea mays L. subsp. Mays POACEAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1569 Cecilio Mota-Cruz, Ramón Cuevas-Guzmán, Cesar Azurdia, and Francisco Palemón-Alberto

About the Editors

Alejandro Casas is a Mexican researcher, with bachelor and master’s degrees in Biological Sciences from the Faculty of Sciences of the National Autonomous University of Mexico (UNAM) and a doctoral degree in Plant Sciences from the University of Reading, UK. Since 1997 he is fulltime Researcher at the Instituto de Investigaciones en Ecosistemas y Sustentabilidad (Institute of Research on Ecosystems and Sustainability), at UNAM, campus Morelia, in the state of Michoacán. He is member of the National System of Researchers, with the highest level (level 3). Dr. Casas has conducted research in the following fields: (1) ecology, culture, and evolution of biodiversity under processes of domestication, (2) management of ecosystems and landscape domestication, (3) ecology for the sustainable management of biotic resources and ecosystems, (4) in situ management of genetic resources, and (5) ethnoecology and biocultural heritage. His research group has made theoretical contributions on the evolutionary mechanisms operating on current processes of domestication, as well as on the origins of food production, agriculture, and domestication in the Neotropics. Also, his group has made contributions to understand biocultural processes generating agrobiodiversity and the bases for their conservation, models of sustainable use of non-timber forest products and agroforestry systems. His work has been mainly conducted in regions of Mexico and Peru and has collaborated in studies in Brazil. The results of his research have been published in nearly 200 peer reviewed articles, more than 100 book chapters, and 13 books that have been authored or edited. These and other works can be consulted and downloaded from ResearchGate (https://www.researchgate.net/profile/ Alejandro-Casas/research). xxi

xxii

About the Editors

Dr. Casas is section editor of the journals Botanical Sciences Genetic Resources and Crop Evolution and Ethnobotany Research and Applications and member of the Mexican, Latin American, and international societies of botany, ecology, and ethnobiology. He was promoter of the Biosphere Reserve Tehuacán-Cuicatlán, Director of the Centro de Investigaciones en Ecosistemas (CIEco), and promoter of the creation of the Instituto de Investigaciones en Ecosistemas y Sustentabilidad (IIES) and the Escuela Nacional de Estudios Superiores (ENES), UNAM-Morelia. José Blancas is a Mexican researcher, with bachelor and master’s degrees in biological sciences from the Faculty of Sciences of the National Autonomous University of Mexico (UNAM) and a doctoral degree from the Centro de Investigaciones en Ecosistemas (Ecosystem Research Center, UNAM). Since 2015 he is fulltime Researcher at the Centro de Investigación en Biodiversidad y Conservación (Biodiversity and Conservation Research Center), at the Autonomous University of the State of Morelos, in the city of Cuernavaca. He is member of the National System of Researchers (level 2). Dr. Blancas has conducted research in the following fields: (1) ethnobotany, (2) management strategies of wild plant resources by traditional human communities, (3) ecological consequences of plant management, and (4) non-timber forest products. He has taught courses on the epistemology of ethnobiology, as well as on quantitative methods used in ethnobiological studies. He has published 35 articles, 4 books, and 20 book chapters and currently is the editor-in-chief of Etnobiología, the journal of the Mexican Ethnobiological Association A.C. (AEM), and section editor of Botanical Sciences. His works can be consulted and downloaded from ResearchGate (https://www.researchgate.net/profile/ Jose-Vazquez-26/research).

Contributors

Itzel Abad-Fitz Centro de Investigación en Biodiversidad y Conservación (CIByC) Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico Araceli Aguilar-Meléndez Centro de Investigaciones Tropicales, Universidad Veracruzana, Xalapa, Veracruz, Mexico Xitlali Aguirre-Dugua Posgrado en Botánica, Colegio de Postgraduados, Texcoco, Estado de México, México Valeria Alavez Laboratorio de Genética de la Conservación, Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Posgrado en Ciencias Biológicas, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Angélica Alemán Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico Gonzalo D. Álvarez-Ríos MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico Laboratorio Manejo y Evolución de Recursos Genéticos, Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Hernán Alvarado-Sizzo Laboratorio de Biogeografía y Sistemática, Departamento de Biología Evolutiva, Facultad de Ciencias, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Viviana Andrade-Erazo Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Salvador Arias-Montes Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, México xxiii

xxiv

Contributors

María Inés Ayala-Enríquez Centro de Investigaciones Biológicas, Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico Cesar Azurdia Consejo Nacional de Áreas Protegidas, Guatemala, Guatemala Gloria E. Barboza Instituto Multidisciplinario de Biología Vegetal (IMBIV)CONICET, Córdoba, Argentina Francisco Basurto Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Coyoacán, Ciudad de México, Mexico Leonardo Beltrán-Rodríguez Laboratorio de Etnobotánica Ecológica, Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Luis Alberto Bernal-Ramírez Departamento de Biología, Universidad Autónoma Metropolitana Iztapalapa, Ciudad de México, Mexico José Juan Blancas Vázquez Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico Arnulfo Blanco-García Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Mexico Fausto Bolom-Ton Centro de Investigaciones Multidisciplinarias sobre Chiapas y la Frontera Sur, Universidad Nacional Autónoma de México, San Cristóbal de Las Casas, Mexico David Bravo-Avilez Departamento de Biología, Metropolitana Iztapalapa, Ciudad de México, Mexico

Universidad

Autónoma

Robert Bye Laboratorio de Etnobotánica, Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Javier Caballero Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Coyoacán, Ciudad de México, Mexico Dánae Cabrera-Toledo MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico Laboratorio Nacional de Identificación y Caracterización Vegetal (LaniVegCONACYT), Departamento de Botánica y Zoología, Centro Universitario de Ciencias Biológicas y Agropecuarias, Universidad de Guadalajara, Zapopan, Jalisco, Mexico Delia Caldera-Cano Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Mexico Javier Caballero: deceased.

Contributors

xxv

Andrés Camou-Guerrero Escuela Nacional de Estudios Superiores (ENES), Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Eréndira J. Cano-Contreras San Cristóbal de Las Casas, Mexico Aglaen Lucero Carbajal-Navarro Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Mexico Rosendo Caro-Gómez Independent Services Provider, Morelia, Mexico Guadalupe Carrillo-Galván Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Guadalupe Casarrubias-Hernández Facultad de Ciencias Químico-Biológicas, Universidad Autónoma de Guerrero, Chilpancingo, Guerrero, Mexico Alejandro Casas Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico Hugo Castillo-Gómez Facultad de Ciencias Naturales, Universidad Autónoma de Querétaro, Querétaro, Mexico Martha Cedano-Maldonado Instituto de Botánica, Departamento Botánica y Zoología, Centro Universitario de Ciencias Biológicas y Agropecuarias (CUCBA), Universidad de Guadalajara, Zapopan, Jalisco, Mexico Belinda Contreras-Jaimes Doctorado en Ciencias de la Sustentabilidad, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Laura Cortés-Zárraga Jardín Botánico-Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Sol Cristians Laboratorio de Etnobotánica Integrativa, Jardín Botánico, Instituto de Biología, Ciudad de México, Mexico Ramón Cuevas-Guzmán Laboratorio de Botánica, Departamento de Ecología y Recursos Naturales, Centro Universitario de la Costa Sur, Universidad de Guadalajara, Autlán de Navarro, Jalisco, Mexico Carolina J. Da Silva Centro de Limnologia, Biodiversidade e Etnobiologia do Pantanal, Universidade do Estado de Mato Grosso, Cuiabá, Mato Grosso, Brazil América Minerva Delgado-Lemus MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico Alfonso Delgado-Salinas Departamento de Botánica, Instituto de Biología, Universidad Nacional Autónoma de México, CDMX, México Jesús Ricardo de Santiago-Gómez Facultad de Ciencias, Universidad Nacional Autónoma de, Ciudad de México, México Lucia Domínguez-Torres San Cristóbal de Las Casas, Mexico

xxvi

Contributors

Carlos Durán-Espinosa Red de Biodiversidad y Sistemática, Instituto de Ecología, Mexico City, Mexico Luis E. Eguiarte Instituto de Ecología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico A. Carolina Elizondo-Salas ZON Herbarium, Forest Engineering Academy, Instituto Tecnológico Superior de Zongolica, Zongolica, Veracruz, Mexico Edward A. Ellis Centro de Investigaciones Tropicales, Universidad Veracruzana, Veracruz, Mexico Tania Escobar Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Coyoacán, Ciudad de México, Mexico Adolfo Espejo-Serna Departamento de Biología, Universidad Autónoma Metropolitana Iztapalapa, Ciudad de México, Mexico Ina Falfán Red Ambiente y Sustentabilidad, Instituto de Ecología, A.C, Xalapa, Mexico Berenice Farfán-Heredia Universidad Intercultural Indígena de Michoacán (UIIM), Pichátaro, Michoacán, Mexico Miriam Monserrat Ferrer Departamento de Manejo y Conservación de Recursos Naturales Tropicales, Campus de Ciencias Biológicas y Agropecuarias-Facultad de Medicina Veterinaria y Zootecnia, Universidad Autónoma de Yucatán. Xmatkuil, Mérida, Yucatán, México Carmen Julia Figueredo-Urbina Cátedra CONACYT. Centro de Investigaciones Biológicas, Instituto de Ciencias Básicas e Ingeniería, Universidad Autónoma del Estado de Hidalgo, Mineral de la Reforma, Hidalgo, Mexico Rosa María Fonseca-Juárez Facultad de Ciencias, Universidad Nacional Autónoma de México, Ciudad de México, México Ana Mitzi García-Leal Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Feliciano García-Lara Escuela de Estudios Superiores Totolapan, Universidad Autónoma del Estado de Morelos, Totolapan, Morelos, Mexico Abisaí García-Mendoza Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, México Edmundo García-Moya Posgrado en Botánica, Colegio de Postgraduados. Montecillo, Texcoco, Estado de México, Mexico David S. Gernandt Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, México Jordan Golubov Universidad Autónoma Metropolitana Xochimilco, Ciudad de México, Mexico

Contributors

xxvii

Jorge Antonio Gómez-Díaz Centro de Investigaciones Tropicales, Universidad Veracruzana, Xalapa, Veracruz, Mexico Juan Carlos González Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Coyoacán, Ciudad de México, Mexico José Antonio González Iturbe-Ahumada Académico Licenciatura Diseño del Hábitat, Facultad de Arquitectura., Universidad Autónoma de Yucatán y Maestría de Arquitectura de Paisaje-Universidad Marista de Mérida, Mérida, Yucatán, México Emiliano González Iturbe-Ruenes Consultor Ambiental independiente (flora y fauna), Hacienda de Xcanatún, Mérida, Yucatán, México Fernando Guerrero-Martínez Centro de Investigaciones Multidisciplinarias sobre Chiapas y la Frontera Sur, Universidad Nacional Autónoma de México, San Cristóbal de Las Casas, Mexico María Teresa Guerrero-Olivares Consultoría Técnica Comunitaria A.C., Chihuahua, Chihuahua, México Paulina Guerrero-Torres Facultad de Ciencias Naturales, Universidad Autónoma de Querétaro, Querétaro, Mexico Anna Paizanni-Guillén Department of Environmental Sciences, Instituto Potosino de Investigación Científica y Tecnológica, A.C. (IPICyT), San Luis Potosí, Mexico Alejandra Gutiérrez-Cedillo Genética de la Conservación, Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Mexico City, Mexico Escuela Nacional de Antropología e Historia, Instituto Nacional de Antropología e Historia, Mexico City, Mexico Facultad de Ciencias, Universidad Nacional Autónoma de México, Mexico City, Mexico Ernesto Gutiérrez-Coatecatl Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Carolina Gutiérrez-Sánchez Universidad Autónoma de Baja California, Ensenada, Mexico Manuel Gutiérrez-Alejo Facultad de Ciencias, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Karla Guzmán-Fernández Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Erandi Lizbeth Guzmán-Gómez Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Campus Morelia, Morelia, Mexico Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Mexico

xxviii

Contributors

Mara Ximena Haro-Luna Instituto de Botánica, Departamento Botánica y Zoología, Centro Universitario de Ciencias Biológicas y Agropecuarias (CUCBA), Universidad de Guadalajara, Zapopan, Jalisco, Mexico Gerardo Hernández-Cendejas Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Luis Hernández-Sandoval Facultad de Ciencias Naturales, Universidad Autónoma de Querétaro, Querétaro, Mexico Oassis Huerta-Galván MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico Maestría en Biosistemática y Manejo de Recursos Forestales y Agrícolas, Centro Universitario de Ciencias Biológicas y Agropecuarias, Universidad de Guadalajara, Zapopan, Jalisco, Mexico Guillermo Ibarra-Manríquez Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de, Ciudad de México, México Juan José Jiménez-Osornio Departamento de Manejo y Conservación de Recursos Naturales Tropicales, Campus de Ciencias Biológicas y AgropecuariasFacultad de Medicina Veterinaria y Zootecnia, Universidad Autónoma de Yucatán. Xmatkuil, Mérida, Yucatán, México Jaime Jiménez-Ramírez Facultad de Ciencias, Universidad Nacional Autónoma de México, Ciudad de México, México David Jimeno-Sevilla ZON Herbarium, Forest Engineering Academy, Instituto Tecnológico Superior de Zongolica, Zongolica, Veracruz, Mexico Esther Katz UMR 208 PALOC IRD/MNHN (Joint Team 208 “Local heritage, environment and globalization” Institute of Research for Development /National Museum of Natural History), Institute of Research for Development (IRD), Paris, France Juan Carlos López-Acosta Centro de Investigaciones Tropicales, Universidad Veracruzana, Xalapa, Veracruz, Mexico Citlalli López-Binnqüist Centro de Investigaciones Tropicales, Universidad Veracruzana, Xalapa, Veracruz, Mexico Ana Rosa López-Ferrari Departamento de Biología, Universidad Autónoma Metropolitana Iztapalapa, Ciudad de México, Mexico Wilfrido López-Martínez Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Carolina Larios Instituto de Investigaciones en Ecosistemas y Sustentabilidad – Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico

Contributors

xxix

Maite Lascurain-Rangel Red Ambiente y Sustentabilidad, Instituto de Ecología, A.C, Xalapa, Veracruz, Mexico Claudia Leyva-Aguilera Department of Sciences, Environmental Education Specialty, Universidad Autónoma de Baja California, Ensenada, Mexico Edelmira Linares Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Rafel Lira Unidad de Biotecnología y Prototipos (UBIPRO), Facultad de Estudios Superiores Iztacala, Universidad Nacional Autónoma de México, Tlalnepantla, Estado de México, Mexico Francisco Lorea-Hernández Red de Biodiversidad y Sistemática, Instituto de Ecología, Mexico City, Mexico Elisa Lotero Jardín Botánico, Instituto de Biología – Universidad Nacional Autónoma de México, Ciudad de México, Mexico Lucio Lozada-Pérez Facultad de Ciencias, Universidad Nacional Autónoma de México, Ciudad de México, México Xavier Madrigal-Sánchez Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Michoacán, Mexico Belinda Maldonado-Almanza Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico María C. Mandujano Instituto de Ecología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Gladys I. Manzanero-Medina Instituto Politécnico Nacional, Centro Interdisciplinario de Investigación para el Desarrollo Integral Regional, Unidad Oaxaca, Oaxaca, Mexico Cristina Mapes Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Coyoacán, Ciudad de México, Mexico Ramón Mariaca-Méndez El Colegio de la Frontera Sur, San Cristóbal de Las Casas, Mexico Mahinda Martínez Universidad Autónoma de Querétaro, Querétaro, Mexico Andrea Martínez-Ballesté Jardín Botánico-Instituto de Biología, Universidad Nacional Autónoma de México, Coyoacán, Ciudad de México, Mexico Raymundo Martínez-Jiménez Investigador independiente, Villa Sola de Vega, Oaxaca, Mexico Fabiola Mena Centro de Investigación en Biodiversidad y Conservación (CIByC) Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico

xxx

Contributors

Myriam A. Miranda-Gamboa Jardín Botánico, Instituto de Biología – Universidad Nacional Autónoma de México, Ciudad de México, Mexico Patricia Irene Montañez-Escalante Departamento de Manejo y Conservación de Recursos Naturales Tropicales, Campus de Ciencias Biológicas y AgropecuariasFacultad de Medicina Veterinaria y Zootecnia, Universidad Autónoma de Yucatán. Xmatkuil, Mérida, Yucatán, México Salvador Montes-Hernández Instituto Nacional de Investigaciones Forestales, Agrícolas y Pecuarias, Ciudad de México, México M. Fernanda Morán-Tzópitl ZON Herbarium Forest Engineerin Academy, Instituto Tecnológico Superior de Zongolica, Zongolica, Veracruz, Mexico Ana Isabel Moreno-Calles Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Cecilio Mota-Cruz Laboratorio en Manejo de Recursos Genéticos, Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Ángel Mujica-Sánchez Facultad de Ciencias Agrarias, Universidad Nacional del Altiplano, Puno, Perú César I. Ojeda-Linares Laboratorio Manejo y Evolución de Recursos Genéticos, Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Aketzalli Olvera-Espinosa Instituto de Investigaciones en Ecosistemas y Sustentabilidad – Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Daniela Ortega-Meza Programa Educativo de Turismo, Universidad Tecnológica del Valle del Mezquital, Ixmiquilpan, Hidalgo, Mexico Luis E. Ortiz-Martínez Instituto Politécnico Nacional, Centro Interdisciplinario de Investigación para el Desarrollo Integral Regional, Unidad Oaxaca, Mexico Amanda Ortiz-Sánchez Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico Rodrigo Hernández-Pacheco Facultad de Filosofía y Letras, Universidad Nacional Autónoma de México, Mexico City, Mexico Francisco Palemón-Alberto Facultad de Ciencias Agropecuarias y Ambientales, Universidad Autónoma de Guerrero, Iguala de la Independencia, Guerrero, Mexico Emmanuel Ismael Pantoja-Aparicio Centro de Investigaciones Tropicales, Universidad Veracruzana, Xalapa, Veracruz, Mexico

Contributors

xxxi

Fortunata Panzo-Panzo Campus Atlahuilco, Sierra de Zongolica and People and Plants International, Universidades para el Bienestar “Benito Juárez García”, Atlahuilco, Veracruz, Mexico Sunem Pascual-Mendoza Instituto Politécnico Nacional, Centro Interdisciplinario de Investigación para el Desarrollo Integral Regional, Unidad Oaxaca, Oaxaca, Mexico Amaranta Paz-Navarro Laboratorio Manejo y Evolución de Recursos Genéticos, Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico P. Pelayo-Delgado Investigadora independiente, Tepetlixpa, Estado de México, Mexico Cloe X. Pérez-Valladares Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Andrea Ponce-Rangel Guacamayas Calentanas A.C., Churumuco, Michoacán, Mexico María Teresa Pulido-Silva Centro de Investigaciones Biológicas, Universidad Autónoma del Estado de Hidalgo, Pachuca de Soto, Hidalgo, Mexico Hilda Ramírez-Monjaraz Facultad de Estudios Superiores Iztacala – Universidad Nacional Autónoma de México, Tlalnepantla, Estado de México, Mexico Selene Rangel-Landa Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Consejo Nacional de Ciencia y Tecnología (CONACYT) – Escuela Nacional de Antropología e Historia (ENAH), Mexico City, Mexico MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico Beatriz Rendón-Aguilar Departamento de Biología, Universidad Autónoma Metropolitana Iztapalapa, Ciudad de México, Mexico María Rivas-Guevara Universidad Autónoma Chapingo, Texcoco, Estado de México, Mexico Alexis Daniela Rivero-Romero Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Isela Rodríguez-Arévalo Unidad de Biotecnología y Prototipos (UBIPRO), Facultad de Estudios Superiores Iztacala, Universidad Nacional Autónoma de México, Tlalnepantla, Estado de México, Mexico Pamela Rodríguez-Juárez Universidad Simón Bolívar, Ciudad de México, Mexico

xxxii

Contributors

Ana Rojas-Rosas Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Yessica Angélica Romero-Bautista Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico David Douterlungne-Rotsaert CONACYT Research Fellow, Department of Environmental Sciences, Instituto Potosino de Investigación Científica y Tecnológica, A. C. (IPICyT), San Luis Potosí, Mexico Felipe Ruan-Soto Instituto de Ciencias Biológicas, Universidad de Ciencias y Artes de Chiapas, Tuxtla Gutiérrez, Chiapas, Mexico María del Rocío Ruenes-Morales Departamento de Manejo y Conservación de Recursos Naturales Tropicales, Campus de Ciencias Biológicas y AgropecuariasFacultad de Medicina Veterinaria y Zootecnia, Universidad Autónoma de Yucatán. Xmatkuil, Mérida, Yucatán, México Maria Fernanda Salazar-Ramirez Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Campus Morelia, Morelia, Michoacán, Mexico Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Michoacán, Mexico Marie-Stéphanie Samain Red de Diversidad Biológica del Occidente, Instituto de Ecología, Xalapa, Veracruz, México Claudia Sánchez-Ramos Posgrado en Botánica, Colegio de Postgraduados. Montecillo, Texcoco, Estado de México, Mexico Jorge Santana-Carrillo Departamento de Biología, Universidad Autónoma Metropolitana Iztapalapa, Ciudad de México, Mexico Adonicam Santiago-Martínez Instituto Politécnico Nacional, Centro Interdisciplinario de Investigación para el Desarrollo Integral Regional, Unidad Oaxaca, Oaxaca, Mexico María Elizabeth Saucedo-Gudiño Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Campus Morelia, Morelia, Michoacán, Mexico Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Michoacán, Mexico Alfredo Saynes-Vásquez Posgrado en Botánica, Colegio de Postgraduados. Montecillo, Texcoco, Estado de México, Mexico José Antonio Sierra-Huelsz People and Plants International, Bristol, VT, USA Leonor Solís Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico

Contributors

xxxiii

Lislie Solís-Montero CONACYT-ECOSUR. Departamento de Agricultura, Sociedad y Ambiente. Tapachula, Chiapas, Mexico Araceli Tegoma-Coloreano Centro de Investigaciones Biológicas (CIB), Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico José Isabel Texta-Hernández Guacamayas Calentanas A.C., Churumuco, Michoacán, Mexico Leticia Torres-Colín Departamento de Botánica, Instituto Universidad Nacional Autónoma de México, CDMX, México

de

Biología,

Ignacio Torres-García Escuela Nacional de Estudios Superiores (ENES), Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico Susana Valencia-Ávalos Facultad de Ciencias, Universidad Nacional Autónoma de México, Ciudad de México, México Mariana Vallejo Jardín Botánico, Instituto de Biología – Universidad Nacional Autónoma de México, Ciudad de México, Mexico Ofelia Vargas-Ponce Laboratorio Nacional de Identificación y Caracterización Vegetal (LaniVeg-CONACYT), Departamento de Botánica y Zoología, Centro Universitario de Ciencias Biológicas y Agropecuarias, Universidad de Guadalajara, Zapopan, Jalisco, Mexico Marco Antonio Vásquez-Dávila Instituto Tecnológico del Valle de Oaxaca, Oaxaca, Mexico Tecnológico Nacional de México, Campus Valle de Oaxaca, Xoxocotlán, Oaxaca, Mexico Nancy Vázquez-Pérez Centro de Investigación en Biodiversidad y Conservación, Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico Juan Vega Consultoría Técnica Comunitaria A.C., Chihuahua, Chihuahua, México Melania Vega Genética de la Conservación, Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Mexico City, Mexico Posgrado en Ciencias Biológicas, Universidad Nacional Autónoma de México, Mexico City, Mexico Heike Vibrans Posgrado en Botánica, Colegio de Postgraduados. Montecillo, Texcoco, Estado de México, Mexico

Juan Vega: deceased.

xxxiv

Contributors

María Idalia Villalpando-Toledo Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico José Luis Villaseñor Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, México Luis Villaseñor-Ibarra Instituto de Botánica, Departamento Botánica y Zoología, Centro Universitario de Ciencias Biológicas y Agropecuarias (CUCBA), Universidad de Guadalajara, Zapopan, Jalisco, Mexico Ana Wegier Laboratorio de Genética de la Conservación, Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico Laura Yáñez-Espinosa Instituto de Investigación de Zonas Desérticas, Universidad Autónoma de San Luis Potosí, San Luis Potosí, México Pilar Zamora-Tavares Universidad de Guadalajara, Zapopan, Jalisco, Mexico

Part I Regions

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction Alejandro Casas and Jose´ Juan Blancas Vázquez

Abstract

This chapter introduces the whole book. It is dedicated to analyze interactions between peoples and plants in the mountain regions of an eminently mountainous country: Mexico. Most of the Mexican territory are mountains that for thousands of years have harbored an extraordinarily high diversity of plants, human cultures, and ecosystems. Richness of native vascular plants has been documented to be more than 23,000 species, which, together with the introduced plants conform a diversity that has been estimated to be nearly 30,000 plant species. Humans have occupied the territory that currently is Mexico since nearly 24,000 years ago. Such long history, the influence of several waves of human occupation, and the predominant mountainous ecosystems favored a high cultural diversification, which is currently represented by peoples speakers of nearly 290 languages. The interaction of these biological and cultural diversities has configured a rich ethnobotanical knowledge, which has maintained valuable expressions disseminated throughout the Mexican territory. The biocultural richness of Mexico is especially high in the mountain regions, which have been refuges of both native peoples and biodiversity. This book gathers studies from different regions representing the mosaic of ecosystems and cultures of Mexico. The first part shows case studies at community and regional levels documenting the role of plants in people’s subsistence, their knowledge, and management forms. The second part is dedicated to synthesizing taxonomic, ethnobotanical, and A. Casas (*) Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] J. J. Blancas Vázquez Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_1

3

4

A. Casas and J. J. Blancas Vázquez

ecological information of a selected group of plant genera that are rooted in Mexican cultures. The two parts of the book aspire to provide a piece of the vast universe of experiences developed by humans interacting with plants in the mountain regions of Mexico.

Introduction Mountains are earth’s abrupt elevations, usually more than 300 m of altitude with respect the surrounding areas, with surface commonly smaller than plateaus and larger than hills (Gerrard 1990). Microenvironments in these areas are highly diverse since in short distances temperature, rainfall, and atmospheric pressure may change drastically, as well as soils and vegetation (Gerrard 1990; Funnell and Parish 2001; Funnell and Price 2003). Mountains and highlands usually harbor high biodiversity, with a marked richness of species assembling in a variety of vegetation types and forming highly heterogenous ecosystems. In fact, mountains are considered hotspots of biodiversity of the world (Körner 2004; Payne et al. 2017; Körner and Spehn 2020). But, in addition, hundreds of millions of people currently inhabit these areas of the planet. These people and their ancestors have learned very much about the peculiarities of such variable environments, and developed a high diversity of interactions with the ecosystems and their components as well as techniques and strategies to live in them (Stepp et al. 2005; Payne et al. 2017). Throughout human history the mountainous landscapes have favored isolations of human groups and, consequently, their cultural and linguistic diversification. Therefore, mountains are scenarios and reservoirs of an extraordinary biocultural diversity (Stepp et al. 2005). Mexico is the setting of a complex geological history. It is a mountainous country with a marked and predominant rugged relief. Nearly 15% of the terrestrial territory of the country is constituted by narrow coastal plains which, together with the Yucatán Peninsula, or Yucatán Platform, are the main plain lowlands; but the remaining area is elevated. About 70% of the Mexican territory has 800 m of altitude or more (Ferrusquía-Villafranca 1994). Going from the coastal plains into the inland, soon appear complex mountain systems dominating the landscape from north to south. The mountain chains are followed by plateaus of considerable elevation and variable extents, traversed and surrounded by more mountains (Fig. 1). The mountain chains of the Eastern and Western Sierras Madre and the Transvolcanic Belt are expressions of an extraordinary scenario of tectonic activity and vulcanism that have configured the geographic relief characterizing this mountainous country (Ferrusquía-Villafranca 1994). But Southern Mexico is also plenty of complex mountainous systems. In the west the Sierra Madre del Sur is the most extended, which joins with smaller systems, among them the Sierra de Juárez (Fig. 1). In the east, the Sierra de Santa Martha in Southern Veracruz is one of the most outstanding, and then in South Central Mexico the system called Sierra Madre de Chiapas gives a peculiar configuration to mountainous landscapes with elevated intermountain valleys and plateaus (Fig. 1). The altitudinal gradients and topographic diversity

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction

5

Fig. 1 General orography of Mexico showing the main mountain systems in yellow brackets and the main plainlands indicated in green labels. The Central Plateau is divided into northern and southern areas by the Transversal Sierras. The Yucatán Peninsula and the Coastal Plainlands of the Gulf of Mexico and the Pacific Ocean are the real lowland plan areas of Mexico, which are nearly 15% of the Mexican territory, the rest are mountainous systems and elevated plateaus. About 70% of the terrestrial surface of Mexico is at elevations higher than 800 m

associated to the relief, the extraordinary edaphic and climatic variation, as well as the latitudinal gradient and the confluence of the Nearctic and Neotropical biogeographic regions have influenced the high diversity of ecosystems that occur in Mexico, which makes it one of the most biodiverse countries of the world (Graham 1994; Rzedowski 1994; Toledo and Ordóñez 1994; Fig. 2). The history of humans in this setting has been ancient. The recent finding of Ardelean et al. (2020) in the Chiquihuite cave in the state of Zacatecas provides evidence to support the argument that humans have been present in the Mexican territory for 24,000 to probably 30,000 years. As documented by several archaeological studies, several waves of humans arrived into the Americas and some occupied parts of the territory of Mexico at different times (Madsen 2015; Williams and Madsen 2020; Shillito et al. 2020; Ardelean et al. 2020; Becerra-Valdivia and Higham 2020). Languages and cultures have therefore different origins and histories and then established in Mexico continued diversifying. The diversification was favored by the diffusion and occupation of the also diverse ecosystems, as well as progressive local adaptations and isolation (Buchanan et al. 2016; Hamilton et al. 2019; Crevels and Muysken 2020). The result of such history is the current occurrence of nearly 290 languages (Eberhard et al. 2022) that

6

A. Casas and J. J. Blancas Vázquez

Fig. 2 A general aspect of the biodiversity that can be found in the mountain regions of Mexico. In the case illustrated, the semiarid mountains surrounding the Tehuacán-Cuicatlán Valley, Central Mexico. (Photos: Francisco Javier Rendón-Sandoval (IIES, UNAM))

derived from several main trunks of linguistic families. The current diversity of languages and cultures is extraordinarily high, one of the highest in the Americas. But the mosaic of cultures existing before the arrival of the European conquerors was even more

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction

7

complex. It has been estimated that the depopulation of Mexico after the Spanish conquest was from 80–90% in several regions, which determined the extirpation of a significant portion of cultures and languages existing in this country in pre-Columbian times (Howell 2002).

The Biocultural Diversity of the Mountains of Mexico Diversity of ecosystems, biodiversity and diversity of cultures, and a long time of cultural history, such is the context of one of the regions with the highest biocultural diversity of the Americas. What ethnobotanical and, in general, ethnobiological studies currently document are biocultural expressions derived from such history of interactions between peoples and ecosystems. Knowledge on the relationships of the ecosystem components, people’s views about the meaning of these interactions and their influence on human life, as well as the techniques mediating the interactions between humans and ecosystems are all elements of the cultures conformed throughout time. It is the history of interactions among peoples and nature and among peoples themselves. Thousands of years of history of the interactions among peoples of the called New World, and then centuries of interaction with peoples coming from the Old World. Throughout such long history it has been configured the biocultural diversity existing in Mexico, in highlands and lowlands, in mountains and plains (Figs. 3 and 4). Most Mexican peoples have lived and currently live in the highlands. The past and present settlements are a constellation of towns occupying the largest area of the territory: mountain regions. The influence of mountains has therefore been crucial in

Fig. 3 Nahua people interact with mountainous rainforest ecosystems in the lowlands of the Sierra Negra, Puebla. The boy is collecting the edible inflorescences of the palm Chamaedorea tepejilote. (Photo: Francisco Javier Rendón-Sandoval (IIES, UNAM))

8

A. Casas and J. J. Blancas Vázquez

Fig. 4 People interacting with mountain ecosystems of temperate forests. An aspect of mushrooms collecting by a P’uhrépecha woman in pineoak forests of the mountains surrounding the Pátzcuaro Lake, in Michoacán. (Photo: Berenice Farfán-Heredia (Universidad Intercultural Indígena de Michoacán, UIIM))

Mexican people’s life, even for those living in coastal plains and altiplanos. The most common patterns of settlements are establishing of villages in the middle of the cold highlands and the warm lowlands. Peoples living in the coastal plains use to migrate and interchange products with people of the highlands and vice versa. Separating peoples of mountains, piedmont, and valleys is difficult since they all have been historically interconnected, interacting with highland and lowland ecosystems directly, seasonally migrating or interchanging products from one area to another. Therefore, peoples of the mountainous regions are partly peoples of the lowlands or the plains and vice versa. Mexico is a mountainous country, and their ecosystems and human cultures are eminently mountainous (Fig. 5).

The Book This book examines interactions between peoples and plants of mountainous ecosystems and their connections with the plains and lowlands (Fig. 6). It is in part the continuation of the book Ethnobotany of Mexico published previously (Lira et al. 2016), and it is also in part one of the windows of the series Ethnobotany of the Mountain Regions to which this book belongs (Paniagua-Zambrana and Bussmann 2020; Batsatsashvili et al. 2020a, b; Abbasi and Bussmann 2021; Kunwar et al. 2021; Franco 2021). We have gathered studies that represent the human cultures and particular plant groups representing the mountain regions of Mexico. The book comprises two main parts. The first one is dedicated to examining ethnobotanical aspects of communities and regions harboring human cultures of the mountains. The second part is dedicated to analyzing plant groups representative of mountainous ecosystems and that have been in interaction with peoples for long time, becoming important components of their culture.

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction

9

Fig. 5 Ixcatec people, inhabiting the mountainous areas of Oaxaca. Men of Santa María Ixcatlán weaving handcrafts with the palm Brahea dulcis. (Photo: Selene Rangel-Landa (IIES, UNAM))

The first part starts with this introductory chapter and then it is followed by the chapter ▶ “Ethnobotanical Knowledge in Mexico: Use, Management, and Other Interactions Between People and Plants” written by Caballero et al., which shows a general panorama of the ethnobotanical information documented for Mexico. It is based on the construction of a research strategy designed 40 years ago to carry on communitarian and regional ethnobotanical studies and systematize them in the first database of ethnobotanical information constructed in Mexico. The authors summarize information on use and management of nearly 8000 plant species, estimating that the total could be more than 11,000 species. In addition, the authors identify the regions and cultural groups that have been more studied and those that need more research efforts, emphasizing the need to dedicate more attention to aspects of plant management. Chapter ▶ “Wild, Weedy and Domesticated Plants for Food Security and Sovereignty” analyzes the role of wild, weedy, and domesticated plants in diet among rural people from different regions. The authors (Casas et al.) gather several case studies showing that after thousands of years of practicing horticulture and agriculture and more than a century of industrialization, rural communities continue to practice gathering of wild and weedy plants complementing their diet based on agricultural products. These authors found cases in which wild and weedy plants may be 8–24% of the components of the annual diet, the variation being influenced by multiple social, economic, and cultural factors, mainly changes in land use,

10

A. Casas and J. J. Blancas Vázquez

Fig. 6 Women and mountains. People of the Sierra Mixe in Tahuitoltepec, Oaxaca, a region of outstanding biocultural diversity. (Photo: Ana Isabel Moreno-Calles (Escuela Nacional de Estudios Superiores-Morelia, UNAM))

migration, monetarization of their economy, and outstandingly, cultural discrimination. The loss of wild and weedy edible products is accompanied by the inclusion of fast food and health problems, therefore the authors put into perspective the importance of conserving and recovering traditional food as central strategy for food sovereignty. In chapter ▶ “Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre Occidental in Mexico,” Cedano-Maldonado et al. review ethnobotanical information among the Wixaritari or Huichol people in the Western Sierra Madre of Mexico. The Wixaritari have maintained their ethnic identity and ancestral social, political, educational, and religious structure, for over hundreds of years. The authors documented their traditional knowledge and worldview over the useful plants they interact with. In addition, the authors provide a regional inventory of nearly 600 plant species used by these people and identify factors influencing loss of the Wixaritari culture and priority areas to conduct ethnobotanical research. In the fifth and sixth chapters Bye and Linares (▶ “Ethnobotany in the Sierra Tarahumara, Mexico: Mountains as Barriers, Conduits, and Generators of PlantPeople Interactions and Relationships”) and Camou-Guerrero et al. (▶ “Rarámuri Ethnobotany: Peasant Subsistence and Biodiversity Conservation at Local Scale”) provide regional and communitarian perspectives, respectively, of ethnobotanical

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction

11

knowledge and plant management practices of the Rarámuri or Tarahumara people in the Western Sierra Madre of Northern Mexico. Both studies analyze at different scales the role of plants in the Rarámuri subsistence, processes that have affected their lives and deterioration of their ecosystems. Also, the communitarian responses and initiatives to maintain their culture and territory, and the role of ethnobotanical research to support such efforts. Chapter ▶ “Wild Food and Traditional Knowledge of the Kumiai from Baja California” is dedicated to show the spectrum of wild food and traditional knowledge of the Kumiai from Baja California, a cultural group with few studies available. The authors (Gutiérrez-Sánchez and Leyva-Aguilera) carried out a case study about food, edible resources, and customs from an ethnographic approach. They illustrate the relationship between resources of the mountain and the coast through ancestral corridors that at present people recognize to be part of their cultural landscape, and the importance of recovering it for their food sovereignty. Chapters ▶ “Mazahua Ethnobotany: Traditional Ecological Knowledge, Management, and Local People Subsistence,” ▶ “Risk Management of Availability of Plant and Fungi Resources Among the Purépecha in Michoacán, Central-Western Mexico,” ▶ “Agroforestry Complexes in the Mountain Regions of Mexico,” and ▶ “Traditional Ecological Knowledge and Biodiversity Conservation in the Tierra Caliente Region of Michoacán,” comprise studies of the mountainous regions of the state of Michoacán in Central Western Mexico. The scenarios are complex mountainous areas where the Western Sierra Madre and the Transversal Volcanic Belt are imbricated. Chapters ▶ “Mazahua Ethnobotany: Traditional Ecological Knowledge, Management, and Local People Subsistence” and ▶ “Risk Management of Availability of Plant and Fungi Resources Among the Purépecha in Michoacán, CentralWestern Mexico,” prepared by Farfán-Heredia and Casas, provide case studies of ethnobotany among the Mazahua and P’uhrépecha people. The study of the Mazahua was conducted in the Monarch Butterfly Biosphere Reserve, an important area for conservation of world natural heritage where these people are the main stewards. Then, these authors analyze aspects of the ecological knowledge in relation to management of plants and fungi. They describe and examine the relation of plant management as an expression of risk management that is conducted by the P’uhurépecha of the Pátzcuaro Lake region. In chapter ▶ “Agroforestry Complexes in the Mountain Regions of Mexico” Moreno-Calles et al. show a panorama of their studies on agroforestry systems in different mountainous regions of Mexico. The authors emphasize the importance of these systems as reservoirs of biodiversity, maintained, managed, and domesticated there. Also, they analyze agrobiodiversity and food production systems as keystones for food sovereignty in the regions studied as well as in the whole country. Chapter ▶ “Traditional Ecological Knowledge and Biodiversity Conservation in the Tierra Caliente Region of Michoacán” is dedicated to analyzing the traditional ecological knowledge and its role in biodiversity conservation in the Tierra Caliente region of Michoacán. The region studied by Rangel-Landa et al. is part of the Biosphere Reserve Zicuirán-Infiernillo, an important area of the Balsas River region managed by mestizo people, which is reservoir of diverse tropical dry forest and

12

A. Casas and J. J. Blancas Vázquez

other vegetation types. The authors discuss current and potential actions considered by local people to contribute to the conservation and restoration of natural areas. In chapter ▶ “Preserving Healthy Eating Habits: Quelites in the Food System of a Nahua Mountain Community, Mexico,” Sánchez-Ramos et al. report their study about the importance of the traditional green vegetables of Mexico called quelites (a term derived from the Nahuatl word “quilitl”) in food systems. They identify their healthy properties and the importance to promote the maintenance of this diverse group of plants. They analyze the case of a Nahua community, but their analysis, reflections, and conclusions are extendable to numerous communities of Mexico that traditionally consume quelites. Chapter ▶ “Ethnobotany of the Sierra Norte de Puebla” summarizes ethnobotanical information of the Sierra Norte de Puebla. Basurto et al. provide a panorama of studies conducted in the region for more than 30 years. It is one of the most studied regions from ethnobotanical approaches and the results and methods have influenced the ethnobotany of Mexico. It is a valuable work based upon an inventory of more than 1000 species of useful plants. In the neighboring region, the Sierra de Zongolica, which as the Sierra Norte de Puebla is part of de Eastern Sierra Madre, López Binnqüist et al. (▶ “Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of Craft Traditions and Biocultural Landscapes”) show the results of their study on wool textiles, emphasizing the role of plants used in the craft traditions and the perspectives for reshaping the activities, based on an approach of biocultural landscapes. In chapter ▶ “Floristic Diversity on Rubber Plantations and their Importance for Subsistence at Foothill Landscapes of Mexico” López-Acosta et al. report their study on the floristic diversity occurring in rubber plantations of foothill landscapes in Uxpanapa, Veracruz. This area, originally with tropical rain forest, was transformed (nearly 80% of the original forest cover was cleared in the last 50 years) for establishing rubber plantations, orange plantations, and grasslands. The authors analyzed the vegetation associated to rubber plantations, finding that an important diversity of native plant species is maintained in those systems. They found the presence of 197 plant species which are managed by local people because of their utility for subsistence, which in turn favors conservation of other species and the restoration of the area. In chapter ▶ “Ethnobotanical Science in the Clouds: Useful Plants of Northeastern Oaxaca, Mexico” Rendón-Aguilar et al. communicate a general panorama of the ethnofloristic richness of Northeastern Oaxaca. The state of Oaxaca is one of the most biocultural diverse areas of Mexico. The authors systematized information from previous studies and directly investigated sites of the northeast. They gathered ethnobotanical information from the mountainous regions surrounding the Tehuacán-Cuicatlán Valley, the Sierra Mixe and the Selva Zoque, where eight indigenous peoples have lived in the area for thousands of years, outstandingly the Zapotec, Mixe, Mazatec, and Mixtec. The ethnofloristic inventory is nearly 800 species, mainly medicinal, edible, and ornamental plants. This is one of the most important regional inventories of ethnobotanical knowledge of the mountains of

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction

13

Mexico, and the authors emphasize the role of plants on local people’s life and the role of people as local safeguards of biodiversity. In chapter ▶ “Ethnobotany of the Highlands of Chiapas” Ruan-Soto et al. report a synthesis of their studies on ethnobotany of the Highlands of Chiapas. This region is a biocultural mosaic of ecosystems and cultures in the mountains of Southern Mexico. The region is covered by different vegetation types dominated by pine and oak species. There, the Maya-related Tsotsil, Tseltal, and Tojol-ab’al indigenous groups have developed a deep knowledge on the regional biodiversity. The authors show a general panorama of the diversity of wild and domesticated, native, and exotic plant species used by these people to satisfy their basic needs of food, fuel, medicine, ornamental, and other purposes (Fig. 7). Chapter ▶ “Ethnobotanical Knowledge and the Patterns of Plant Use and Management in the Sierra de Huautla Biosphere Reserve and the Chichinautzin Biological Corridor in Morelos, Mexico” is dedicated to summarizing the ethnobotany of the Sierra de Huautla and the Chichinautzin Biological Corridor in the state of Morelos (Fig. 8). These are natural protected areas constituting reservoirs of tropical dry forest and other vegetation types and are also the territory of Nahua and Mestizo communities. In this chapter, Blancas et al. focus their attention on documenting aspects of knowledge, use, and management of the regional flora. The state of Morelos is in part in the Transvolcanic Belt and in part in the Balsas River Basin, one of the areas with the greatest biocultural wealth in Mexico, with a high number of endemic species. Although the European invasion produced great changes in the indigenous forms of cultivation and on the landscape of the region, great civilizations such as the Olmec, Tlahuica, and Mexica that flourished there maintained their cultural heritage. Plants use and management of the current human communities are expressions of those cultures and ethnobotany may significantly contribute to conserve their biocultural memory. Chapter ▶ “Ethnobotany of the Nahua People: Plant Use and Management in the Sierra Negra, Puebla, Mexico,” is dedicated to the Sierra Negra in the state of Puebla, which is a mountainous region that forms part of the Sierra Madre del Sur. The Sierra Negra harbors a great diversity of ecosystems and vegetation types like pine, pineoak, cloud forest, rain forest, tropical dry forest, and xerophytic scrub. In addition, in this region the Nahua, Mazatec, Popoloca, and mestizos people coexist and interact. These peoples have remained relatively isolated from other regions and the multiple and diverse management of the ecosystems are part of their strategies to deal with such isolation. In this chapter, Blancas et al. document a wide spectrum of forms of use and management of plants, considering the region as an important biocultural refuge (Fig. 9). In chapter ▶ “Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding the Tehuacán Valley” Rangel-Landa and Casas synthesize information of nearly 20 years of studies of the Ixcatec ethnobotany. The Ixcatec people live in Santa María Ixcatlán, in the mountains surrounding the Tehuacán-Cuicatlán Valley. The Ixcatec is an Otomanguean language in process of extinction, with only 12 speakers alive. This situation has motivated efforts from several scholars and institution to contribute to conserve the people’s biocultural memory and language. The chapter

14

A. Casas and J. J. Blancas Vázquez

Fig. 7 People and plants in the Highlands of Chiapas, where the Tsotsil and Tseltal have developed an extraordinary ethnobotanical knowledge. (Photos: Felipe Ruán-Soto (Universidad de Ciencias y Artes de Chiapas))

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction

15

Fig. 8 Landscapes in the Chichinautzin Biological Corridor (a), and the Sierra de Huautla in the state of Morelos, México (b). These are natural protected areas constituting reservoirs of pine-oak forest and tropical dry forest that are the territory of Nahua and Mestizo communities. (Photos (a) Araceli Tegoma Coloreano and (b) Luis Sánchez Méndez)

summarizes information on plant knowledge, nomenclature, classification, use, and management of plants, analyzing the role of wild and cultivated species in people’s subsistence. Also, the authors discuss the role of ethnobotanical studies and the ways

16

A. Casas and J. J. Blancas Vázquez

Fig. 9 General aspect of the landscape in the Sierra Negra, where Nahua, Mazatec, Popoloca, and mestizo people have interacted with a high diversity of ecosystems and plant species throughout long time. (Photo: Javier Rendón-Sandoval (IIES, UNAM))

these can interact with other disciplines to contribute to conserving the Ixcatec biocultural memory. In chapter (▶ “Ethnobotanical Knowledge Within the Sierra Gorda, Querétaro, Mexico”) Hernández-Sandoval and Castillo-Gómez gathered information from different studies including their own work for decades in this region. The Sierra Gorda is a portion of the Eastern Sierra Madre, with tropical, temperate, and semiarid vegetation used and managed by several cultures, outstandingly the Nya’xu (Northern Pame), Teenek (Huastec), Ximpeces, Uzá’ (Chichimeco-jonaz), and Hñöñho (Semidesert Otomí). The authors report 739 useful plant species from 19 vegetation types and five types of agroecosystems, providing an extraordinary panorama of the ethnobotany of that mountainous region. As in other regions, the authors found that most useful pants are obtained from tropical deciduous forest, oak forest, and piedmont scrub, but they also recorded that numerous species are gathered from secondary vegetation. In chapter ▶ “Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca” Vásquez-Dávila et al. analyze the ethnobotany of the mountain landscape of Oaxaca, the state of Mexico with the highest biocultural diversity. This and other chapters of this book provide complementary views of such a fascinating kaleidoscope that Oaxaca constitutes. Following Alcorn (1991), the authors confirm that most biodiversity is harbored in the lands managed by traditional marginal human groups. They illustrate this fact through case studies in the mountains of Oaxaca. The

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction

17

chapter describes the sacrality of the mountains and illustrates it through ethnographic information on agriculture and use of medicinal plants in two communities of the Sierra, analyzing the tensions between local people and external agents. In chapter ▶ “Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pápalo, Oaxaca” Solís-Rojas and Casas report their ethnobotanical study with the Cuicatec of San Lorenzo Pápalo, Oaxaca. The Cuicatec are cultural and linguistically related with the Mixtec but formed a separate branch some thousands of years ago. Recently, the authors (Solis-Rojas and Casas 2019) reported ethnobiological Cuicatec information on fauna, and the current chapter reports the corresponding part of ethnobotanical studies. The role of plants in people’s subsistence is analyzed, contextualizing their views, nomenclature, and classification system in relation to the management practices and other interactions between local people and ecosystems. The role of wild, weedy, and domesticated plant species in the Cuicatec subsistence is also reviewed. In several chapters referred to above, the traditional markets are outstanding places to document biocultural diversity. These are areas where products from highlands and lowlands are obtained by peoples, making possible an important ecological complementarity (Fig. 10). The second part of the book is dedicated to review groups of plants representative of the Mexican cultures. These chapters are organized by genus in alphabetical order. We start this part with a group of aromatic plants of the genus Agastache, contributed by Carrillo-Galván and Bye (▶ “Agastache spp. Lamiaceae. Important Species of Hyssop in Mexico”); these plants are widely used for medicinal purposes, some of them with signs of domestication. Then, we continue with a review of the genus Agave, elaborated by Torres-García et al. (“Agave spp. Agavaceae”). This is an important group of plants with overwhelming diversity and presence in Mexican landscapes, with long history of interaction with humans and rooted in their cultures. Dozens of species have been used as food, for their fiber, their sap to prepare fermented beverages, and, more recently, their stems, which are cooked to prepare mescal spirits. Most species used are wild, extracted from forests and their overuse in some areas has determined important dangers over populations. Nowadays, conservation of several types of forests in Mexico are linked to correct planning of using agaves. The chapter provides a general panorama of these plants, their ethnobotany and ecology, and the challenges for their conservation. Then, Mapes et al. review the useful species of genus Amaranthus, a group of plants providing edible grains and vegetables since ancient times. Some species are domesticated and others are weedy plants providing valuable food. Cristina Mapes has conducted important studies on the management and domestication of Amaranthus species of Mexico for decades. Therefore, the chapter offers an interesting window to this interesting group of plants. Amphipterygium is a genus of the family Anacardiaceae, whose four species growing in tropical dry forests of Mexico have outstanding medicinal properties. Particularly noteworthy is the cuachalalate (A. adstringens), since its bark is widely used for gastric ulcers and is sold in markets of Mexico. Beltrán-Rodríguez and Bye describe this species and some aspects of its use, ecology, and needs for its conservation. Paizanni-Guillén and Douterlungne (▶ “Aristolochia spp. ARISTOLOCHIACEAE”)

18

A. Casas and J. J. Blancas Vázquez

Fig. 10 Traditional markets, impressive reservoirs of biocultural diversity, spaces where people from lowlands and highlands interchange their products and allow ecological complementarity. In the top, aspect of a seller in Cuetzalan, the main town of the Northern Sierra of Puebla. In the bottom, aspect of the variety of products interchanged in the market of Chilapa, in the Mountain of Guerrero region. (Photos: Ignacio Torres-García (ENES-Morelia, UNAM))

continue the series analyzing the genus Aristolochia, a group of plants characterized by the presence of a high diversity of secondary compounds associated to traditional medicine, now explored by scientists for its great pharmaceutical potential. Palms of the genus Brahea spp. are among the most ancient plants associated to human remains, as suggested by archaeological discoveries by Ardelean et al. (2020).

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction

19

Its diversity in Mexico is outstanding and the fiber, edible fruit, and fibrous stems of the species forming the genus have been widely used by Mexican cultures This genus is reviewed by Andrade-Erazo et al. Then, Blancas et al. provide a panorama of the species of the important genus Bursera. Species of this genus provide multiple benefits to people, wood, firewood, medicine, and, outstandingly, the resin called copal, which is deeply linked to the ceremonial life of Mesoamerican and Aridoamerican peoples from ancient times to the present. Capsicum is a group of plants iconic of the Mexican culture. Aguilar-Meléndez et al. provide a valuable review of the taxonomy, ethnobotany, and evolutionary aspects related to the domestication of these plants. The authors of this chapter have been working in an ambitious project of a series of books gathering aspects on the natural and cultural history of the chili peppers, two of them (Aguilar Meléndez et al. 2018; Vásquez-Dávila et al. 2022) were recently published. Thus, the chapter is an extraordinary link to those important works. Andrade-Erazo and Miranda Gamboa reviewed another important genus of palms: Chamaedorea. This genus provides several benefits to humans, the inflorescences of some species are edible (Fig. 3), very much appreciated, and, in general, these palms are valued worldwide for ornamental purposes. It is a genus with extraordinary diversity in the Mexican territory, deeply rooted in several cultures and nowadays threatened for their intense harvest from forests. This is a reason why these palms have been important systems of population ecology studies motivated to identify sustainable thresholds of harvest rates (Hernández-Barrios et al. 2015; Jansen et al. 2018). Bowls have been crucial artifacts in human cultures, some of them preceding the ceramic. Among the most important are the Lagenaria siceraria fruits which were early introduced from the Old World, species of the genus Cucurbita and the native gourd trees of the genus Crescentia. In the Mexican territory, Crescentia alata and C. cujete are the species of this genus that are available and have been important part of Mexican cultures. Aguirre-Dugua and Casas give a review of taxonomic and ethnobotanical information of these species which is linked with their studies on domestication and phylogeography of these plants in the neotropics (Aguirre-Dugua et al. 2012, 2013, 2018). The genus Cucurbita has an especial place in the cultures of the Americas. Species of this genus have been used as bowls, food, medicine, and are important components of the iconic milpa systems associated with maize and beans. Lira et al. prepared a review of taxonomic, ethnobotanical, and evolutionary studies conducted by their group for decades. It is a valuable contribution that allows connecting the reader with taxonomic challenges and questions related to the origins and domestication of these important plants. Escontria chiotilla is an arborescent candelabriform cactus whose fruits are widely appreciated by cultures of the semiarid areas of Central Mexico. Escontria is a monospecific genus and in this chapter Casas (▶ “Escontria chiotilla (F. A. C. Weber ex K. Schum.) Rose CACTACEAE”) provides a general panorama of ethnobotanical and ecological information related to the interactions between people and

20

A. Casas and J. J. Blancas Vázquez

these plants. Also, Casas provides information about the economic importance of this species, challenges for conservation, and the relevance of the local experience of management for this purpose. Elizondo-Salas et al. (▶ “Gaultheria acuminata Schltdl. and Cham. ERICACEAE”) present an overview of the use and management of Gaultheria acuminata Schltdl. and Cham. ERICACEAE, an important medicinal plant from the Sierra de Zongolica (the central area of the Eastern Sierra Madre is called this way). G. acuminata is a key resource in many communities, since it is an important medicinal plant with commercial value at regional scale, but also has a prominent role in the cosmovision and symbolism of the Nahua people. The Mesoamerican cotton Gossypium hirsutum has been extraordinarily important for its fiber to Mesoamerican peoples from pre-Columbian times to the present. Currently this species of cotton is one of the most important crops worldwide. Several species of the genus occur in Mexico, maintaining interactions with cultivated cotton. The research group of Dr. Ana Wegier, leading this chapter, has made relevant contributions to understand the genetic interactions among these species and their role in the process of domestication. In this chapter Alavez et al. review taxonomic, ethnobotanical, ecological, and evolutionary information of 13 species of the genus, including the crop species G. hirsutum and G. barbadense, which allow the readers a general panorama of the importance of these plants, and it is a connection with the fascinating studies carried out by the authors. Beltrán-Rodríguez et al. elaborated a general overview of the genus Hintonia in Mexico. These are species whose bark has medicinal uses, mainly as a febrifuge and for treating other ailments. The case of H. latiflora stands out since it constitutes an important nontimber forest product of tropical deciduous forest. The American wild or false laurel Litsea glaucescens Kunth is one outstanding nontimber forest product. It is a spice very much appreciated in Mexico and other countries of the world. Commercializing its leaves is a profitable activity that enhances unsustainable harvesting practices that endanger entire populations in some areas. Ortega-Meza et al. (▶ “Litsea glaucescens Kunth LAURACEAE”) examine use and management strategies of this species in different areas, bringing ethnobotanical, ecological, and economic information. The chapter allows the readers general insights about the species, the interactions, the problems, and views for attending need for the species conservation. Jimeno-Sevilla and Elizondo-Salas (▶ “Magnolia mexicana DC. MAGNOLIACEAE”) address the case of the Magnolia mexicana DC. MAGNOLIACEAE, regionally known as “yoloxóchitl,” which is used for several purposes, mainly medicinal. It is a tree that inhabits cloud forest and rainforests in the mountainous area of Veracruz, Mexico. This species is classified as vulnerable in the IUCN list, therefore some conservation strategies are documented and discussed, both in wild vegetation and in some agroecosystems. Melothria is a genus of the Cucurbitaceae widely spread in the Americas. In Mexico species of this genus are appreciated as food, harvested from wild and weedy

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction

21

areas. Fruits are consumed raw (with cucumber-like flavor) or cooked (sometimes substituting green tomatoes in sauces). Guerrero-Torres et al. summarize information that provides a general panorama of the ethnobotany of species of this genus. Myrtillocactus is a genus of candelabriform shrubby or arborescent cacti. All species produce edible fruit and flowers, which are collected from wild populations, but some species receive different forms of management, and at least two species (M. geometrizans and M. schenckii) have signs of incipient domestication. Blancas and Casas show in the chapter dedicated to this genus a review of ethnobotanical and ecological information on use and management of these plants. Vásquez-Dávila et al. (2022) address the genus Nicotiana, which is made up of species of annual or perennial plants, widely used in Mexico for multiple purposes, highlighting medicinal, ritual, and ornamental uses. An overview of nine species is presented, emphasizing the case of N. tabacum. Delgado-Salinas and Torres-Colín address the case of the genus Phaseolus, important food resources of various cultures in Mexico. The use of dry seeds and pods stands out, which, together with corn, are part of the basic diet of the Mexican population. The authors provide information about the use of wild relatives and several varieties of P. vulgaris. Martínez et al. address taxonomic and ethnobotanical information of the genus Physalis, whose complexity and diversity are part of different cultures of Mexico. The green tomato P. philadelphica is the best known and studied species of this genus, but other species are in the background of several indigenous cultures. Most species described have edible uses and a vast culinary complex throughout Mexico, but some species have in addition other uses, mainly medicinal. Polaskia chichipe and P. chende are candelabriform cacti used and managed by people of the Tehuacan Valley. In this chapter Casas reviews information generated by his research team about these species in relation to their main use for their edible fruit, and the management practices including gathering of fruit from the wild, the silvicultural management of populations in agroforestry system, and, in the case of P. chichipe, its cultivation in plantations in regional homegardens. Ethnobotanical, morphological and ecological information is provided to analyze the incipient levels of domestication identified in the managed populations. Martínez-Ballesté et al. address the genus Sabal, palms of the Arecaceae family, whose leaves are widely used as construction material from pre-Hispanic times to the present. Some species of the genus Sabal are outstanding plant resources in regional cultures, especially for the Maya of the Yucatán Peninsula. Information is provided on past and present uses, as well as the different ways in which they are managed in wild vegetation and in agroecosystems. Lira et al. review the genus Sechium, whose species are important food resources in mountainous areas of Mexico since pre-Hispanic times. Sechium edule is one of the species that has the greatest morphological diversity, which in part has been

22

A. Casas and J. J. Blancas Vázquez

promoted by human management. In addition to edible uses, these vines have uses as veterinary medicine, as well as fodder for livestock. Torres-García et al. address the Solanum genus, which is one of the most complex and diverse plant group of the Solanaceae family. The authors emphasize the medicinal and edible uses, as well as the management strategies that have been documented among various peoples and cultures that inhabit mountainous areas in Mexico. Ruenes-Morales et al. (▶ “Spondias mombin L. Spondias purpurea L. Spondias radlkoferi J. D. Smith ANACARDIACEAE”) evince the ethnobotanical importance of the Spondias genus, whose species are mainly used for their fruits. In addition, they have a complex traditional nomenclature that includes the great diversity of shapes, colors, flavors, and ways of preparation. The implications of the diversity of forms of management are also addressed, which have led to domestication processes from the incipient to the most intense management and selection. Alvarado Sizzo and Casas review ethnobotanical and ecological information of the genus Stenocereus, which is formed by shrubby and arborescent cacti widely used for their edible fruits. The genus is naturally distributed from the Southern USA to the arid and semiarid regions of Colombia and Venezuela. Mexico is the main area of diversity of the genus and species have been part of several cultures inhabiting xerophytic vegetation and tropical dry forest areas. Fruits of practically all species are appreciated and gathered from wild populations but plants of some species receive silvicultural management in agroforestry systems or in live fences. Species like S. griseus, S. pruinosus, S. stellatus, S. fricii, and S. queretaroensis show signs of domestication and the authors summarize the information available to document this aspect. Vega et al. review the genus Vanilla, whose best-known species is V. planifolia, with which various edible products are made and it is one of the most important crops of the world. However, in this section other lesser-known species are considered, all of them locally important genetic resources with economic and biocultural relevance. Finally, Mota-Cruz et al. provide an overview of the ethnobotanical importance of the Zea genus, whose wild, weedy, ruderal, and domesticated species have forged the identity of various Mesoamerican cultures, especially those distributed in Mexico. The case of Zea mays ssp. mays is emblematic, not only because Mexico is the center of origin of the species, but also because of the diversity and complexity of interactions that humans have established with it throughout its evolutionary history. In addition, in the context of climate change, knowledge of the various species is crucial for the conservation of the genetic diversity of the Zea genus. Acknowledgments The authors thank CONACYT for financial support to the project A1S-14306, as well as support from the GEF Project ID 9380 CONABIO-GEF-FAO/RG023 “Manejo y domesticación de agrobiodiversidad en Mesoamérica: Bases para la soberanía alimentaria sustentable,” and PAPIT, UNAM (project IN206520 and IN224023).

Human Cultures and Plant Diversity in the Mountains of Mexico: An Introduction

23

References Abbasi AM, Bussmann RW, editors. Ethnobiology of mountain communities in Asia. Cham, Switzerland: Cham, Switzerland: Springer International Publishing; 2021. Aguilar-Meléndez A, Vásquez-Dávila MA, Katz E, Hernández-Coronado MR. Los chiles que le dan sabor al mundo. Universidad Veracruzana, Institut de Recherche pou le Dévelopment. Xalapa, Veracruz, México, 2018. Aguirre-Dugua X, González A, Eguiarte L, Casas A. Round and large: morphological and genetic consequences of artificial selection on the gourd tree Crescentia cujete by the Maya from the Yucatán Peninsula, Mexico. Ann Bot. 2012;109:1307–16. Aguirre-Dugua X, Casas A, Pérez-Negrón E. Phenotypic differentiation between wild and domesticated varieties of Crescentia cujete and culturally relevant uses of fruits as bowls in the Yucatan Peninsula, Mexico. J Ethnobiol Ethnomed. 2013;9:76. Aguirre-Dugua X, Llanderal-Mendoza J, González-Rodríguez A, Eguiarte L, Casas A. Anthropogenic dispersion of selected germplasm creates a new phylogeographic mosaic of Crescentia cujete populations in Mesoamerica. Tree Genet Genomes. 2018;14:18. Alcorn JB. Ethics, economies and conservation. In: Oldfield ML, Alcorn JB, editors. Biodiversity. Culture, conservation and ecodevelopment. Oxford, UK: West View Press; 1991. p. 317–49. Ardelean CF, Becerra-Valdivia L, Pedersen MW, Schwenninger JL, Oviatt CG, Macías-Quintero JI, Arroyo-Cabrales J, Sikora M, Ocampo-Díaz YZE, Rubio-Cisneros II, Watling JG, Medeiros VB, Oliveira PE, Barba-Pingarón J, Ortiz-Butrón A, Blancas-Vázquez J, Rivera-González I, Solís-Rosales C, Rodríguez-Ceja M, Gandy DA, Navarro-Gutierrez Z, de La Rosa-Díaz JJ, Huerta-Arellano V, Marroquín-Fernández MB, Martínez-Riojas LM, López-Jiménez A, Higham T, Willerslev E. Evidence of human occupation in Mexico around the Last Glacial Maximum. Nature. 2020;584:87–92. Batsatsashvili K, Kikvidze Z, Bussmann RW. Ethnobotany of the mountain regions of far Eastern Europe: Ural, Northern Caucasus, Turkey, and Iran. Cham, Switzerland: Springer International Publishing; 2020a. Batsatsashvili K, Kikvidze Z, Bussmann RW. Ethnobotany of the mountain regions of Central Asia and Altai. Cham, Switzerland: Springer International Publishing; 2020b. Becerra-Valdivia L, Higham T. The timing and effect of the earliest human arrivals in North America. Nature. 2020;584:93–7. Buchanan B, Hamilton MJ, Kilby D, Gingerich JAM. Lithic networks reveal early regionalization in late Pleistocene North America. J Archaeol Sci. 2016;65:114–21. Crevels M, Muysken P, editors. Language dispersal, diversification, and contact. Oxford, UK: Oxford University Press; 2020. Eberhard DM, Simons GF, Fennig CD, editors. Ethnologue: languages of the world. 25th ed. Dallas: SIL International; 2022. Online version: http://www.ethnologue.com Ferrusquía-Villafranca I. Geology of Mexico: a synopsis. In: Ramamoorthy TR, Bye R, Lot A, Fa J, editors. Biological diversity of Mexico: origins and distribution. Oxford, UK: Oxford University Press; 1994. p. 3–108. Franco FM, editor. Ethnobotany of the mountain regions of Southeast Asia. Cham, Switzerland: Springer International Publishing; 2021. Funnell DC, Parish R. Mountain environments and communities. London: Routledge; 2001. Funnell DC, Price MF. Mountain geography: a review. Geogr J. 2003;169(13):183–90. Gerrard JA. Mountain environments: an examination of the physical geography of mountain. Cambridge, MA: MIT Press; 1990. Graham A. Historical factors and biological diversity in Mexico. In: Ramamoorthy TR, Bye R, Lot A, Fa J, editors. Biological diversity of Mexico: origins and distribution. Oxford, UK: Oxford University Press; 1994. p. 109–28. Hamilton MJ, Buchanan B, Walker RS. Spatiotemporal diversification of projectile point types in western North America over 13,000 years. J Archaeol Sci Rep. 2019;24:486–95.

24

A. Casas and J. J. Blancas Vázquez

Hernández-Barrios JC, Anten NPR, Martínez-Ramos M. Sustainable harvesting of non-timber forest products based on ecological and economic criteria. J Appl Ecol. 2015;52(2):389–401. Howell KW. In the wake of conquest: a global perspective on the depopulation of indigenous peoples of Latin America. Diálogos Latinoamericanos. 2002;5:58–72. Jansen M, Anten N, Bongers F, Martínez-Ramos M, Zuidema P. Towards smarter harvesting from natural populations by sparing the most important individuals in an exploited palm species. J Appl Ecol. 2018;55(4):1682–91. Körner C. Mountain biodiversity, its causes and function. AMBIO J Hum Environ. 2004;33 (sp13):11–7. Körner C, Spehn EM. Mountain biodiversity. A global assessment. New York: Routledge; 2020. Kunwar RM, Sher H, Bussmann RW. Ethnobotany of the Himalayas. Springer International Publishing; 2021. Lira R, Casas A, Blancas J. Ethnobotany of Mexico. Interactions of people and plants in Mesoamerica. New York: Springer Nature; 2016. Madsen DB. A framework for the initial occupation of the Americas. PaleoAmerica. 2015;1(3): 217–50. Paniagua-Zambrana NY, Bussmann RW, editors. Ethnobotany of the Andes. Cham, Switzerland: Springer International Publishing; 2020. Payne D, Spehn EM, Snethlage M, Fischer M. Opportunities for research on mountain biodiversity under global change. Curr Opin Environ Sustain. 2017;29:40–7. Rzedowski J. Diversity and origins of the phanerogamic flora of Mexico. In: Ramamoorthy TR, Bye R, Lot A, Fa J, editors. Biological diversity of Mexico: origins and distribution. Oxford, UK: Oxford University Press; 1994. p. 129–44. Shillito LM, Whelton HL, Blong JC, Jenkins DL, Connolly TJ, Bull ID. Pre-Clovis occupation ofthe Americas identified by human fecal biomarkers in coprolites from Paisley Caves, Oregon. Sci Adv. 2020;6(29):eaba6404. Solís-Rojas L, Casas A. Cuicatec ethnozoology: traditional knowledge, use and management of fauna by people of San Lorenzo Pápalo, Oaxaca, Mexico. J Ethnobiol Ethnomed. 2019;15:58. Stepp JR, Castaneda H, Cervone S. Mountains and biocultural diversity. Mount Res Dev. 2005;25(3):223–7. Toledo VM, Ordóñez MJ. The biodiversity scenario of Mexico: a review of terrestrial habitats. In: Ramamoorthy TR, Bye R, Lot A, Fa J, editors. Biological diversity of Mexico: origins and distribution. Oxford, UK: Oxford University Press; 1994. p. 757–77. Vásquez-Dávila MA, Aguilar-Meléndez A, Katz E, Manzanero-Medina GI. Chiles en México. Historias, culturas y ambientes. Municipio de la Ciudad de Oaxaca, Universidad Veracruzana, Instituto Politécnico Nacional CIDIIR, Oaxaca, Tecnológico Nacional de México campus Valle de Oaxaca. Xalapa, Veracruz, México, 2022. Williams TJ, Madsen DB. The upper palaeolithic of the Americas. PaleoAmerica. 2020;6(1):4–22.

Ethnobotanical Knowledge in Mexico: Use, Management, and Other Interactions Between People and Plants Javier Caballero, Laura Corte´s-Zárraga, Cristina Mapes, Jose´ Juan Blancas Vázquez, Selene Rangel-Landa, Ignacio Torres-García, Berenice Farfa´n-Heredia, Andrea Martínez-Balleste´, and Alejandro Casas Abstract

This chapter shows a general panorama of ethnobotanical research and information generated during the twentieth and twenty-first centuries among Mexican cultures, according to the database Base de Datos Etnobotánicos de Plantas Mexicanas (BADEPLAM) of the Botanical Garden at the Institute of Biology, UNAM. This is the most complete database with ethnobotanical information in Mexico, whose Javier Caballero: deceased. J. Caballero · C. Mapes Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Coyoacán, Ciudad de México, Mexico e-mail: [email protected]; [email protected] L. Cortés-Zárraga Jardín Botánico-Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico e-mail: [email protected] J. J. Blancas Vázquez Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico e-mail: [email protected] S. Rangel-Landa Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Consejo Nacional de Ciencia y Tecnología (CONACYT) – Escuela Nacional de Antropología e Historia (ENAH), Mexico City, Mexico e-mail: [email protected] I. Torres-García Escuela Nacional de Estudios Superiores (ENES), Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_2

25

26

J. Caballero et al.

construction started nearly 40 years ago. It was a pioneer effort to systematize biocultural information in this country, which has continued until now and has stored nearly 60,000 records on plants used and managed by different cultural groups in different ecosystems of Mexico. It includes information on nearly 7823 useful plant species, which is approximately one-third of the total native vascular flora of the country. Through different approaches, it is estimated that the real number could be more than 11,500 species, which gives an idea of the effort still required to complete the inventory. The current listing has information from numerous Mestizo people communities, but only 32 of the 68 main linguistic groups of Mexico; not all the states of Mexico have been studied, and ethnobotanical research has concentrated in half of the states composing Mexico. All this information indicates that although BADEPLAM is probably the oldest project of biocultural informatics in Latin America, there is a long way to complete the task of inventorying the ethnobotanical knowledge of the country. BADEPLAM has records for 4222 medicinal plant species, 2265 ornamental, 2051 edible, 1974 used as fodder, and 975 for fuelwood, among other uses. Most species (nearly 64%) are wild and weedy plants collected from forests, mainly tropical dry forests (1995 species), tropical rain forests (1928 species), temperate forests (1440 species) and xerophytic vegetation (1361 species), grasslands, and agricultural areas. However, nearly 3000 species are managed through one or more forms, some of them showing incipient or intermediate signs of domestication. Nearly 500 species are fully domesticated crops, approximately one-half of them (251 species) being native to the Mesoamerican region. Plant families contributing with the highest richness of useful plants are Fabaceae (752 species), Asteraceae (727), Poaceae (476), Cactaceae (474), Euphorbiaceae (233), Malvaceae (198), and Solanaceae (195). Associated with BADEPLAM, several research groups have articulated our work coordinating different approaches to generate inventories of knowledge, management techniques, and different forms of interactions between people and plants. These inventories have been performed at rural community (more than 150 communities) and regional levels (17 main biocultural regions of Mexico) feeding the database while constructing theoretical frameworks on traditional classification and worldviews, use, management and domestication, and bases for sustainable use of plants and ecosystems. Several approaches have enhanced our

B. Farfán-Heredia Universidad Intercultural Indígena de Michoacán (UIIM), Pichátaro, Michoacán, Mexico A. Martínez-Ballesté Jardín Botánico-Instituto de Biología, Universidad Nacional Autónoma de México, Coyoacán, Ciudad de México, Mexico e-mail: [email protected] A. Casas (*) Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected]

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

27

studies, but plant management and domestication have been some of the most important issues. We understand that management is a crucial expression of interactions between people and plants, reflecting their knowledge and worldviews, and it is a topic that allows connecting ethnobotany with social, cultural, and economic topics. In addition, studying plant management allows establishing socioecological bases for sustainable management and studies on evolution of plants through domestication at populations and landscape levels. In this chapter, we show general insights of the research approaches developed by our teams. Most of our studies have been conducted in mountainous regions since Mexico is an eminent mountainous country. Therefore, this text provides general perspectives of the ethnobotanical knowledge of Mexico, as well as methodological approaches that are helpful to contextualize the entire volume of this book.

Introduction Mexico and the Mesoamerican region in the neighboring countries of Central America is one of the areas with the highest biocultural diversity of the world (Maffi 2005; Toledo and Barrera-Bassols 2008; Boege 2008). This region includes more than 300 native languages (284 only in Mexico, according to Ethnologue; Eberhard et al. 2022) and more than 39,300 species of vascular plants (Hanelt 2001), which is nearly one-third of the flora of the Americas (Ulloa et al. 2017), as well as a high diversity of vertebrate and arthropod species. Such diversity has a notable expression in the ethnobiological knowledge and the systems of management and domestication of plants, animals, mushrooms, and microorganisms, as well as the regional ecosystems, their components and functions, and landscapes. As recently reported by Clement et al. (2021), nearly 6500 native plant species of Mexico and the Mesoamerican area, belonging to 265 families, have been recorded to have one or more uses by the Indigenous cultures and other rural people of the region. Among the main families providing plant resources are Fabaceae (699 species), Asteraceae (571), Cactaceae (438), Poaceae (335), Euphorbiaceae (205), Malvaceae (171), Solanaceae (162), Rubiaceae (159), Asparagaceae (143), Apocynaceae (133), and Lamiaceae (133). Compared with information from the Andean region of Peru and the Amazonia and lowlands of Brazil, these numbers are outstanding, not only because of the high biocultural diversity, but also due to the active and long tradition of ethnobotanical research conducted in the area and, importantly, because of the extraordinary efforts to systematize the information in databases (Clement et al. 2021). Most native plant species in Mexico and the Mesoamerican area are medicinal (3478 species), edible (1810), fodder (1637) and used for construction (1224) and as fuelwood (883). Interactions between people and plants are mostly through gathering, since nearly 6000 species are obtained this way from forests. However, 1555 species receive some form of management: (i) tolerance or let standing of plants in areas cleared for different purposes, (ii) enhancing or promotion actions directed to increase abundance of desirable plants, (iii) special protection and care of plants against herbivores, competitors, frosts, or for procuring water, shade, or sunlight, or (iv) their cultivation

28

J. Caballero et al.

by planting seeds, vegetative propagules, and/or (v) transplanting of complete individuals with the purpose of cultivating or relocating them. These forms of management may involve selection on particular phenotypes and have determined that at least 727 species have incipient signs of domestication. Other 170 species can be considered semidomesticated, and 251 species are fully domesticated plants, with clear signs of domestication syndrome, phenotypic divergence from wild populations, and marked dependence on humans for survival and reproduction (Clement et al. 2021). These native species were complemented with others introduced from different regions of the Americas throughout history, and then, after European colonization, numerous wild, weedy, and domesticated plant species from the Old World were introduced and adopted by the human cultures of the Americas (Corona et al. 2021). Scholars studying management and domestication of plants have recognized that Mexico and the neighboring Mesoamerican area are one of the most ancient and dynamic scenarios where management and domestication can be documented in the Americas (Vavilov 1992; Harlan 1975; Hawkes 1983; Smith 2006). But also, because these are ongoing processes. Studying how and why they are initiated, maintained, and innovated may contribute to understanding how and why these processes occurred in the past. The region called Mesoamerica was originally proposed by Paul Kirchhoff (1943) as a cultural area with special features that distinguish it from other regions of the Americas. According to Kirchhoff (1943), Matos-Moctezuma (1994), and others, in Mesoamerica flourished human cultures with distinctive settlements and buildings, agricultural systems and techniques, food patterns, and numerous other cultural aspects compared with the neighboring northern arid region of Mexico, called Aridoamerica, and other cultures further North America, as well as those of the Andean, Amazonian, or Patagonian regions in South America. The human cultural features considered by Kirchhoff (1943) have been partly confirmed or refuted by several archaeological and anthropological studies conducted for decades in the region and the whole American Continent. However, the term continues being used and it is still a helpful reference to studies of both cultural and biological diversity. According to Matos-Moctezuma (1994), the Mesoamerican region comprised the southern half of Mexico until the north-western area of the current Costa Rica, but he and other authors have discussed the dynamic limits of this region throughout time. In addition, it is pertinent to say that cultural elements and products of the regional biodiversity from the Aridoamerica and from South America arrived at Mesoamerica continually since prehistory (MacNeish 1967, 1992; Piperno and Pearsall 1993; López-Austin and López-Luján 2002; Clement et al. 2021; Corona et al. 2021). This illustrates that the frontiers, if these really existed, were not only dynamic but also with high porosity. The early presence of maize in the Andean region (Piperno and Pearsall 1993; Clement et al. 2021) and the ancient presence of cacao (Theobroma cacao L.), manioc (Manihot esculenta Crantz), peanuts (Arachis hypogaea L.), sweet potato (Ipomoea batatas (L.) Lam.), and other South American crops in Mesoamerica are indicators of the antiquity and intensity of technological interactions and interchange of crop species and varieties among regions (Pease et al. 2016; Zarrillo et al. 2018; Kistler et al. 2020; Corona et al. 2021). Archaeological remains and ethnohistorical sources, as well as studies from anthropology, ecology, population genetics,

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

29

phylogeography, and genomic approaches, have been progressively clarifying the biocultural history of the region and will continue doing it with new research tools and approaches. The information available now appears to suggest that the discontinuities proposed by Vavilov and other scholars among the biocultural regions are hypothetical and deserve more research to be confirmed or modified. In this chapter, we will summarize information of the ethnobotanical knowledge documented among peoples from the Mesoamerican and Aridoamerican regions of Mexico, including the current scenario of native and introduced species that became adopted by human cultures occupying the area. This is part of the information that requires to be analyzed to contribute to reconstructing the biocultural history of the area. Cultures of the Mexican Mesoamerica started developing techniques to manage biotic resources and ecosystems that led to early domestication of plants and food production systems, approximately 9000–10,000 years ago (MacNeish 1967, 1992; Benz 2006; Smith 1997; Piperno et al. 2009) and have continued doing it until the present (Casas et al. 1997, 2017; Parra et al. 2010; Clement et al. 2021). Cultures of Aridoamerica, apparently, adopted in some areas these experiences of management and domestication and initiated their own processes (Nabhan 1985). Different practices like gathering, interchange of products, and incipient management and domestication have been reconstructed based on archaeological information and strongly supported by ethnobotanical and ethnographic studies of how current cultures perform activities that configure these processes (Alcorn 1984; Zizumbo and Colunga 1982; Casas et al. 1994, 1996, 1997; Blancas et al. 2010; Rangel-Landa et al. 2016). In addition, since much of these practices are still carried out, important details about the perception of variation, targets of selection, mechanisms to put it in practice, and their evolutionary consequences can be documented through ethnobiological, ecological, and evolutionary biology approaches (Casas et al. 1997, 2007; Blancas et al. 2010, 2013; Aguirre-Dugua et al. 2012, 2013, 2018; Rangel-Landa et al. 2016; Moreira et al. 2017; Clement et al. 2021; Arévalo-Marín et al. 2021). Although the current social and ecological contexts are different to those occurring in the past, the current processes are valuable empirical bases that can be used as models to understand the motivations that enhanced people to manage plants in the past and ways that could have operated (Casas et al. 1997, 2007; Parra-Rondinel et al. 2021; Rangel-Landa et al. 2016; Clement et al. 2021; Arévalo-Marín et al. 2021). Our research groups have conducted studies on plant management and domestication in several regions of Mexico, in different ecological contexts, different cultural groups, and different groups of plants, including annual herbaceous, shrubby, small trees, agaves, cacti, and long-lived perennials. These studies could provide information and theoretical frameworks to support an interpretation of what happened in the past. Our groups have systematized ethnobotanical information for the whole Mexican territory for nearly 40 years, through the Base de Datos Etnobotánicos de Plantas Mexicanas (BADEPLAM, Database of Ethnobotanical Information of Mexican Plants, in English), of the Botanical Garden at the Institute of Biology, UNAM. Both the database and field studies in ethnobotany, ecology, and evolutionary biology related to management and domestication allow identifying general patterns of the processes analyzed and how and why these are currently occurring. The

30

J. Caballero et al.

information generated is now being useful not only to analyze plant management in Mexico, but also may be helpful to our colleagues working in the Amazonian and Andean regions, which are exceptional areas related to the human culture of plant management. Our research groups started their work with the coordination by the first author of this chapter and then developed their own profiles but maintained most of the original purposes, among them, to: (1) systematize the ethnobotanical information generated among peoples and plants of Mexico, (2) analyze the information on uses, management, traditional nomenclature and classification, habitats, and ecological information of plants Mexican cultures interact with, (3) identify general patterns on the groups of plants mostly incorporated in human subsistence by cultural groups of Mexico, (4) identify the factors motivating people to practice plant management, the different types of practices, and those involving processes of domestication, (5) develop views on sustainable management of nontimber forest products at population and ecosystem levels, (6) document the general trends of morpho-physiological, reproductive, and genetic changes in plants associated to domestication, (7) analyze how landscape domestication influences domestication processes on particular species and vice versa, and (8) analyze how different evolutionary forces operate to influence domestication. To address these issues, our research groups have combined ethnobiological, ecological, and evolutionary approaches. Our ethnobotanical studies have inventoried the diversity of forms of use and management of plants, and we have systematized our own research, as well as that published in the scientific literature and that registered in herbarium specimens. Those are the basic sources of information stored in BADEPLAM. In the field, we have worked in these inventories at community and regional levels, while the information of BADEPLAM allows a general panorama of the state of ethnobotanical knowledge for the whole country. During decades, most ethnobotanical studies in Mexico have emphasized collecting information on use of plants; therefore, since the 1990s our research has emphasized studying cultural, ecological, and evolutionary aspects related to plant management. We have documented the diversity of plant management forms in forests (silviculture), agricultural systems (horticulture and agriculture), agroforestry systems (agro-silviculture), and livestock-raising systems (plant management associated with pastoralism, free raising of goats and cattle, and agro-silvo-pastoralism). These studies look for understanding the different management techniques and the social and ecological factors motivating and influencing the way the management practices are. More particularly, how the need to ensure the availability of desirable products, esthetic purposes, curiosity, and other factors move people’s inventiveness, their interest in innovation, how they transmit their experiences to others, and how they adopt techniques developed by others. We are especially interested in understanding why and how these mechanisms enhance decision-making, as well as the consequences of management and domestication on different sociocultural, ecological, and evolutionary aspects. These are topics that could help to analyze how domestication and food production started and changed the human ways of life and, also, to the understanding of current processes of technological innovation, adoption, and diffusion in traditional rural contexts.

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

31

After the “Origins of Species” (Darwin 1859), by the end of the nineteenth and throughout the twentieth centuries, several studies explored the areas of origin of domestication. Among the most outstanding works are those by De Candolle (1882) and Vavilov (1992), who collected information from botanical, geographical, anthropological, linguistic, ecological, and genetic fields to suggest some regions of the world that were supposed to be the areas of origin of cultivated plants. The regions proposed were valuable hypotheses that led archaeologists to investigate remains to test the suppositions and to support information about the process of domestication of the most important crops. Then, after several classic archaeological and genetic studies appeared the proposals by Harlan (1975), Bruce Smith (1989, 2006), Zeder (2008, 2011, 2012, 2015, 2017), and other scholars that have had important dissertations about the origins and causes of domestication of plants and animals, which are still in debate, particularly about the questions of where, when, how, and why management and domestication started and developed. Periods of scarcity of resources associated with climate change or demographic growth of humans determining pressures on ecosystems are among some of the explanations, while for other authors environmental pressures and technological innovation should be integrally analyzed (Flannery 1986; Harris and Hillman 1989; Harris 1996). The research groups of the authors of this chapter consider that both management and domestication of plants are ongoing processes and, therefore, studying and understanding them provide elements to analyze the past, with reasonable bases for the interpretation of archaeological data. Looking for answers to the general questions referred to in the previous paragraph has, therefore, theoretical value, particularly in relation to how knowledge, management, and domestication of plants interact with social-cultural needs and the ecological conditions of the organisms used to satisfy them, as well as in relation to evolutionary-ecology issues. Management and domestication of plants and the systems where these are performed progressively configured a valuable biocultural heritage of the Mexican cultures from both Mesoamerican and Aridoamerican regions. This heritage is now a valuable experience to contribute to construct a general repertoire and catalogue of management techniques that are highly important to understand the past, but, at present, to construct strategies of sustainable management that several sectors of Mexico are interested in. Our research groups have conducted studies in more than 150 communities (see Appendix 1, Fig. 1) of Ch’ol, Cuicatec, Ixcatec, Lacandon, Mazatec, Mixtec, Mayan groups, Mazahua, Nahua, Popoloca, P’urhépecha, Rarámuri, Huastec or Teenek, Tlapanec, Tlahuica, Zapotec, and mestizo people in different regions of México (Fig. 1, see Caballero 1994; Caballero and Mapes 1985; Caballero et al. 1998; Caballero and Cortés 2001; Mapes et al. 1981, 1996, 1997; Casas et al. 1994, 2001, 2007, 2014, 2017; Pérez-Negrón and Casas 2007; Camou-Guerrero et al. 2008; Lira et al. 2009; Blancas et al. 2010, 2013; Cano et al. 2012; Torres 2004; Torres-García et al. 2013, 2015a, b, 2019, 2020; Martínez-Ballesté et al. 2005, 2006; Bunge-Vivier and Martínez-Ballesté 2017; Cuevas et al. 2021; Lotero-Velásquez et al. 2022; Farfán et al. 2007, 2018a, b; Ubiergo-Corvalán et al. 2019, 2020, 2021).

32

J. Caballero et al.

Fig. 1 Regions of Mexico where the more than 150 rural communities studied by our research groups are located. From north to south, Peninsula of Baja California, Sierra Tarahumara, Cuatrociénegas Valley, Huasteca, Northern Sierra of Puebla, Mountains of Northern Michoacán, Morelia Region, Monarch Butterfly Biosphere Reserve, Pátzcuaro Lake Basin, Tierra Caliente of Michoacán, Highlands of the state of Morelos, Balsas River Basin of the state of Morelos and Guerrero, Mountain of Guerrero, Tehuacán-Cuicatlán Valley, Central Valleys of Oaxaca, Highlands of Chiapas, and Yucatán Peninsula. See details in Appendix 1

We have promoted similar research with colleagues of the Andean region (Velásquez-Milla et al. 2011; Torres-Guevara et al. 2019; Pancorbo et al. 2020; Parra-Rondinel et al. 2021) and the Brazilian lowlands, especially the Caatinga (Lucena et al. 2014, 2016, 2017; Lins Neto et al. 2014; Trindade et al. 2015; Lima-Nascimento et al. 2021), Amazonia, and Mata Atlantica (Clement et al. 2021). One of the first attempts to show systematized ethnobotanical information on use and management in these regions was recently published (Clement et al. 2021). To achieve it, we put in practice a process of interaction involving conceptual and methodological interchange, looking for constructing databases with similar format compatible for further analyses and comparisons. In this process, we also started to include the information available for the neighboring Mesoamerican region of Central America. These activities are configuring a new stage in the systematization process that will require several years of effort while local and regional studies must continue, especially in those cultural and ecological areas with scarce or no studies available.

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

33

Research Strategy BADEPLAM The initiative of constructing BADEPLAM started in 1982 as part of an institutional project at the Botanical Garden of the Institute of Biology, UNAM. The work was initiated by designing the format of the database in a time when the technological tools for storing information were still limited and the personal computers were restricted. In fact, the earliest systems established failed to recover the stored information, and the format and storing system had to change from time to time. It was until the 1990s when BADEPLAM became more operative and dynamic for storing and providing services. In the first stage of construction, BADEPLAM had the name of Banco de Información Etnobotánica sobre Recursos Genéticos (BIERGEN). It was part of an ambitious project to integrate the Botanical Garden at UNAM as part of a research unit called Unidad de Recursos Genéticos (UNIRGEN), which was conceived and enhanced by Dr. José Sarukhán, who was the director of the Institute of Biology, and who some years later founded the National Commission for the Knowledge and Use of Biodiversity (CONABIO), the most important governmental institution in Mexico systematizing information on biodiversity. The project UNIRGEN started with the collaboration of several scholars, and BIERGEN was the main responsibility of the first author of this chapter (Javier Caballero). UNIRGEN aspired to carry out multidisciplinary work, in which ethnobotany was considered to be the direct source of information from the field about genetic resources for food and other purposes, emphasizing the documentation of information about management and domestication by different ethnic groups of Mexico. The project aspired to know the diversity of genetic resources, mainly used as food, and identify some of them with high potential to attend problems in Mexico. The general team of UNIRGEN included ethnobotanists, geneticists, taxonomists, and plant physiologists specialized in in vitro propagation (Caballero 1984; Caballero et al. 1985). The starting group of ethnobotanists was formed by Cristina Mapes, José Arellano, Javier Caballero, and Robert Bye, who conducted regional studies in the Isthmus of Tehuantepec, Oaxaca, the P’urhépecha region of Michoacán, the Yucatán Peninsula, and the Sierra Tarahumara, respectively. Later, Carmen Vázquez, Juan Luis Viveros, and Alejandro Casas were included in the team, investigating in different areas of the Balsas River Basin region and then in the Tehuacán-Cuicatlán Valley. In addition, the group of Miguel Angel Martínez-Alfaro, Francisco Basurto, and Alberto Villa in the Sierra Norte of the state of Puebla. Their regional approach included fieldwork, bibliographic compilation of inventories of useful plants, as well as studies on some plant groups (Amaranthaceae, Arecaceae, Fabaceae, and Solanaceae, among others) which were followed more deeply by collecting information from herbarium specimens at MEXU. The database was designed and coordinated by Javier Caballero and implemented by a mathematician (Juan Antonio Toledo), and a biologist (Laura Cortés), who captured and curated the entries of information. Juan Toledo elaborated an algorithm in algol language, which allowed the following: (1) loading all programs to manage

34

J. Caballero et al.

the database, (2) uniting text files captured in a PC for “5/4 floppy disks” which were then carried to a terminal of the supercomputer Burroughs B-7800, and (3) storing the files in magnetic tapes for later use through the database program. The ethnobotanical information was coded, and a dictionary of translated information was used. Such complex systems and processes made it difficult to manage the database and obtain results. The database included different fields which were discussed and carefully selected by the team of ethnobotanists. The information systematized included taxonomic information (plant family, species, subspecies, and other intraspecific categories), ecological information of the specimens recorded (location, vegetation type, elevation, climate, and soil), and the use and management types, among the most important. In the 1990s, each field of information was reconsidered, which allowed decreasing the previous huge structure of the database, adjusting it to the real information captured until that time, when information of important ethnobotanical works was already collected. Among the studies whose information was then captured, we counted those by Alcorn (1984) with the Huastec, Berlin, Breedlove, and Raven (1974) with the Tzeltal, several works by Maximino Martínez for the whole country (Martínez 1959, 1979, 1989), Felger and Moser (1985) with the Seri, Barrera-Marín et al. (1976) with the Maya, Pennington (1963) and Bye (1976) with the Tarahumara, and Pennington (1969) with the Tepehuan, among others, as well as information of useful species of the families Fabaceae, Arecaceae, and Amaranthaceae, among others. Also, several theses by students and collaborators of the project, and other unpublished works. Another important decision was to change the name BIERGEN into Base de Datos Etnobotánicos de Plantas Mexicanas (BADEPLAM). BADEPLAM is an application created and developed by Laura Cortés in Access 2016 Microsoft, Office. At present it has 59,487 records, from a total of 361 bibliographic sources of information, as well as information from herbaria and data from field studies by our research groups and others. It is a database with good curatorial work, with tables of relational reference which complement and help to minimize capture errors. There is in addition a good collection of works that are being captured. An aspect in progress is the restructuring of BADEPLAM according to international standards. Although it is important to maintain an original version of the information captured, it is continually necessary adjusting and updating taxonomic nomenclature and content of the information fields, which should be widely reviewed by ethnobotanists, taxonomists, geographers, and anthropologists with experience in bioinformatics. But importantly, the changes should not increase the basic structure of BADEPLAM but adjust some pertinent underlying concepts in the information fields. BADEPLAM is currently coordinated by Dr. Andrea Martínez-Ballesté who is also responsible of managing and using the information. The use has lacked specific operative rules, but these should be constructed in the near future to prevent misuse and misappropriation of information. BADEPLAM was a pioneer project in biodiversity informatics, with the vision of being a source of valuable information on nontimber forest products for academic and conservation programs. The experience has been adopted by several research groups, some of them have allowed the correct compatibility to feed the main database and strengthen its capacities, and this should be a process to enhance ahead.

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

35

Field Studies Our field studies have included regional and communitarian scales, including three main dimensions: one of them sociocultural, in which the main approach is ethnobiology and ethnobotany. It has been directed to inventory and document cultural information on plants, their economic value, and mechanisms of interchange at local and regional levels, the regulations existing in communities, municipalities, and regions, among other topics. The second dimension is ecological. In part, but not only we have analyzed the consequences of management on individuals, populations, biotic communities, and ecosystems; but we also have studied the influence of these aspects on the decision of people to manage populations, communities, and ecosystems. The third dimension is directed to analyze evolutionary processes associated with management, including domestication (Fig. 2). The ethnobiological studies look for understanding the cultural bases of use and management of plants, and we aspire to complement them with social research on institutions regulating access to resources; these regulations and commercialization of products in markets reflect the importance of resources to people. Ecological studies have had in principle the purpose of documenting the distribution and abundance of the most important plant resources, their diversity, biomass, and spatial and seasonal availability (Pérez-Negrón and Casas 2007; Blancas et al. 2013; Rangel-Landa et al. 2016). This information diagnosed in the vegetation types and anthropogenic areas of communitarian territories allows identifying possible use

Fig. 2 Processes studied by our research groups. In the intersection of these processes, the management of plants is a main issue of our research, which expresses knowledge, practices, and worldviews of people on plants they interact with. Management is influenced by the sociocultural context, including social and economic relations, forms of social organization and regulations constructed about the interactions among households and other communities and among these social units and the environment, and the technological aspects available for the interaction, among other relevant issues. In addition, management is markedly influenced by the contexts of ecosystems where it occurs, and in turn the management influences and drives changes in ecosystems, their components, and functions. Management influences evolutionary processes through domestication, which in turn is influenced by the natural evolutionary processes occurring in plant populations

36

J. Caballero et al.

patterns that may endanger the permanence of particular plant populations. But in addition, it allows documenting the ecological complementarity of the environmental units in people’s subsistence (Lotero-Velásquez et al. 2022). Also, ecological studies allow identifying the biotic interactions (pollinators, seed dispersers, facilitation, and other mutualist interactions) that should be maintained when planning use of forest products (Casas et al. 1999; Otero-Arnaiz et al. 2003; Torres-García et al. 2013, 2015a; León-Jacinto and Torres 2015; Rangel-Landa et al. 2015; Cuevas et al. 2021). Therefore, ecological studies together with ethnobotanical information allow constructing proposals on sustainable management at the community assemblages in territories. Another important approach developed by our research groups has been conducted at the population level. We have worked with populations of species particularly endangered or that may be endangered due to human activities. From this perspective, we have studied species of palms (Sabal spp. and Brahea dulcis (Kunth) Mart.; Martínez-Ballesté et al. 2005, 2006, 2008; Martínez-Ballesté and Martorell 2015; Martínez-Ballesté and Caballero 2016; Pulido and Caballero 2006; Rangel-Landa et al. 2014), several species of Agave (Torres-García et al. 2013, 2015a, 2019, 2020) mainly those extracted from forests to produce mescal, as well as some trees intensively used in some communities (Ceiba aesculifolia (Kunth) Britten & Baker f., by Arellanes et al. 2018) and Bursera bipinnata (Moc. & Sessé ex DC.) Engl. by Abad-Fitz et al. 2020). We are interested in documenting the effect of management on survival and reproduction of individuals that are under management and the populations they form part of. This information aspires to identify thresholds that are able to maintain or collapse the populations used and, therefore, develop ecological criteria to define sustainable rates of harvesting the useful products. These studies allow identifying how ecological processes influence plant management and the impact of management on ecosystems (Blancas et al. 2013). Through these studies, we explore hypotheses related to the influence of the scarcity or uncertain availability of resources of high cultural value on the people’s decisions to manage them. But in addition, the information allows analyzing the conditions for sustainable management at population and community levels. From this approach, aspects such as life form, length of life cycle, part or parts used, the type of reproduction system, distribution, abundance, and phenology are all relevant issues. The third dimension of our research is studying evolutionary processes associated with management: domestication, which involves adjustments in morphology and physiology of plants according to human purposes. Domesticates commonly diverge from wild and weedy plants, whose fitness is high in those environments while domesticates are successful in managed environments only through human assistance. Divergence between wild or weedy and the domesticates is not binary but may include a continuum of intermediate conditions, depending on the purposes of humans and the level of intensity of human selection (Casas et al. 1996, 1997, 2007). It is not unidirectional since multiple features, not only one, may motivate humans to practice selection. And complete domestication is not the unique destiny of management. In Mexico, hundreds of plants remain in a state of low divergence with respect to their wild or weedy relatives and may remain in that state for a long

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

37

time since with that intensity of management and selection the desired actual sociocultural benefits are obtained. At ecosystem level, we are particularly interested in how the management of populations or groups of populations influence changes at landscape level. And, in turn, how changes deliberately performed at landscape level influence changes of plant populations (Casas et al. 1997; Parra et al. 2015; Clement et al. 2021).

Ethnobotanical Diversity of Mexico México harbors a high biocultural diversity, and this is especially expressed in the ethnobotanical knowledge. The Mexican territory is inhabited by Indigenous peoples representing diverse cultures that speak nearly 291 languages (Eberhard et al. 2022). According to Ethnologue 284 are Indigenous languages, 84 are developing, 74 are vigorous, 88 are in trouble, 44 are dying, and 6 languages are extinct (Eberhard et al. 2022). The existing languages are catalogued by the Instituto Nacional de Lenguas Indígenas (INALI 2008) in 68 linguistic groups, each one with different linguistic variants that totalize the 284 Indigenous languages referred to above. It has been estimated that after the European invasion and conquest, nearly one half of languages became extinct because of wars, diseases, and extermination of a high percentage of people living in this country. All those cultures were configured for thousands of years. The recent discoveries of Ardelean et al. (2020) in the Chiquihuite Cave at the state of Zacatecas reveal that humans have been present in the territory that currently is Mexico since about 25,000 and possibly around 30,000 years ago. The diversity of flora is also high, the inventory of the native vascular plants, according to Villaseñor (2016), is 23,314 species, and according to Toledo and Ordóñez (1993), the native and introduced flora occurring in Mexico is nearly 30,000 species. And this is the setting of biocultural processes that molded what ethnobotanists working in Mexico have documented since the early twentieth century. The most recent information from BADEPLAM indicates that peoples of Mexico interact, know, use, and manage nearly 7823 plant species. This is an inventory documented among 32 of the 68 main linguistic groups of Mexico. Table 1 indicates the number of records by cultural group, which in turn indicates that nearly half of the cultural groups have been studied, some of them scarcely, and those with no ethnobotanical records would be one of the priorities to enhance studies about. Table 2 shows the number of ethnobotanical records by state, indicating that most studies have been conducted in the states of Puebla, Veracruz, Oaxaca, Chiapas, and others that are bioculturally diverse, whereas more efforts are required in the states of Querétaro, Colima, Baja California, and Zacatecas, among others. Similarly, Fig. 6 indicates tropical and temperate forests are the most studied vegetation types, whereas the xerophytic vegetation, grasslands, and cloud forests require more research effort. Figures 3, 4, and 5 show a general panorama of the inventory of use types, plant families, and life forms providing more plant resources, respectively. Figure 6 shows

38 Table 1 Ethnic groups and the number of records about use and management of plants in BADEPLAM. (Records are the references found in the different sources of information, among them literature reference, collection number referred to in documents or herbarium specimens)

J. Caballero et al.

Ethnic group Maya Tarahumara Náhuatl Teenek Totonaco Mixteco Mayo (yoremes) Zoque Seri Zapoteco Tzotzil Otomí Tzeltal Chinanteco Cuicateco Ixcateco Guarijío Chontal (Tabasco) Purépecha Tepehuanes Lacandón Kikapú Tepehuanes (Durango) Mazateco Mazahua Mixe Popoluca Pápago Huave Chol Pima Cora Huicholes Mestizos (in Spanish) Others

Number of records 4356 1918 1824 1572 1432 1144 1084 1046 1007 892 859 847 669 484 464 372 324 309 284 258 181 168 168 166 154 138 111 83 79 74 20 8 8 35,378 1606

the general panorama of vegetation types providing useful plant species. Figure 7 shows a panorama of the types of plant management recorded while Fig. 8 shows the panorama of the types of interaction in the main groups of plants under management: edible and medicinal plants. However, information from regional studies suggests that the inventory in BADEPLAM could be substantially increased, especially when the total flora is compared with the flora reported with use. For instance, in the Tehuacán-

Ethnobotanical Knowledge in Mexico: Use, Management, and. . . Table 2 Number of records about use and management of plants per state in BADEPLAM. (Records are the references found in the different sources of information (literature reference, collection number referred to in documents or herbarium specimens))

State Puebla Veracruz Oaxaca Chiapas Nuevo León Quintana Roo Morelos Chihuahua Guerrero Sonora Yucatán San Luis Potosí México City Hidalgo Michoacán Tabasco State of México Tamaulipas Guanajuato Coahuila Sinaloa Aguascalientes Nayarit Campeche Jalisco Durango Tlaxcala Zacatecas Baja California Baja California Sur Colima Querétaro Total

39 Number of records 10,575 4848 4544 3332 3025 2622 2621 2530 2374 2305 2186 1914 1745 1676 1597 1549 1463 1381 905 826 782 617 582 474 425 382 305 248 163 156 154 79 58,385

Cuicatlán Valley we have found that while the total flora is about 3000 plant species, the useful plants is nearly 2000 plant species (Casas et al. 2017), which is nearly 60% of the whole flora. Similar comparisons in other regions allow estimating that on average about 38.9% of the plant species in a region may be expected to have one or more uses (Table 3). This cipher, compared with the general inventory of native vascular plant species in Mexico, and 23,314 species according to Villaseñor (2016) would lead to expect 9069 useful native plant species. However, considering both native and introduced species, according to Toledo and Ordóñez (1993) nearly 30,000 plant species, in Mexico, the

40

J. Caballero et al.

Fig. 3 Number of plant species used with different purposes by the different cultures of Mexico according to the Base de Datos Etnobotánicos de Plantas Mexicanas BADEPLAM of the Botanical Garden, Institute of Biology, UNAM

Fig. 4 The plant families providing more useful species in Mexico, according to BADEPLAM

estimation allows expecting 11,670 useful plant species, native and non-native, disseminated throughout the Mexican territory. Through another approach, based on the number of species accumulated in the sources consulted for constructing BADEPLAM, and projecting the number of species that could potentially be included in the database, the curve of Fig. 9 was

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

41

Fig. 5 Life forms of the plant species used in Mexico by peoples of different cultures according to the Base de Datos Etnobotánicos de México BADEPLAM of the Botanical Garden, Institute of Biology, UNAM

Fig. 6 Vegetation types providing the highest richness of useful plants

obtained. This approach suggests that the number of useful species in Mexico could be about 11,500, a number similar to the estimation referred to above, which is reasonable since BADEPLAM stores information on native and non-native species.

42

J. Caballero et al.

Fig. 7 Number of species under different management types. The category incipient-management includes 570 species tolerated or let standing, 417 plant species promoted or enhanced, and 186 species under special protection by local people, according to the management categories by BADEPLAM

Fig. 8 Number of medicinal and edible species that are collected in the wild, managed incipiently, and those that are cultivated. (Information from the Base de Datos Etnobotánicos de Plantas Mexicanas BADEPLAM of the Botanical Garden at the Instituto de Biología, UNAM)

Diversity of Management Forms Different authors have proposed that food production like horticulture, agriculture, and pastoralism arose as strategies to decrease uncertainty in the availability of food and other products (Flannery 1986; Harris 1996). However, for thousands of years,

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

43

Table 3 Total number of plant species recorded in different regions of Mexico, compared with the general plant species richness recorded in those regions Region Tehuacán-Cuicatlán Valleya Sierra de Manantlánb,c Sierra Norte de Pueblad Selva Lacandonae Los Tuxtlasf Tuxtepece Uxpanapae Península de Yucatáng Sian Ka’ane Montaña de Guerreroh Huastec regioni Sierra Huicholaa Sierra del Abra Tanchipaj Sierra de Huautlak Tierra Caliente de Michoacánl Average (%) Mexico (native) Mexico (native and introduced)

Total spp. >3000 2983 1730 1660 2697 737 800 2900 558 800 1113 1652 427 1018 2634

Useful spp. >2000 650 720 415 730 296 336 1000 316 430 445 532 116 649 616

23,314 30,000

9139 11,760

% 66.7 21.8 41.6 24.9 27.1 40.2 40.6 23.4 56.6 53.8 40.0 32.2 27.2 63.8 23.4 38.9

a

Based on Casas et al. (2017, updated in this study); bSantana-Michel and CONABIO (2021); cBenz et al. (1994); dMartínez-Alfaro et al. (1995); eToledo et al. (1995); fCONANP/SEMARNAT (2006); g Flores (1999); hCasas et al. (1994); iAlcorn (1984); jDe-Nova et al. (2019); kBlancas et al. (chapter “▶ Ethnobotanical Knowledge and the Patterns of Plant Use and Management in the Sierra de Huautla Biosphere Reserve and the Chichinautzin Biological Corridor in Morelos, Mexico”); l Rangel-Landa et al. (chapter “▶ Traditional Ecological Knowledge and Biodiversity Conservation in the Tierra Caliente Region of Michoacán”).

Fig. 9 Estimation of the useful flora of Mexico based on the cumulative records of information sources and its projection. (According to data from BADEPLAM)

44

J. Caballero et al.

and even at present, the rural communities continue practicing gathering and extraction of forest products, including gathering of plants, hunting, and fishing together with agriculture livestock and horticulture in homegardens and other systems. Gathering, according to Casas et al. (1996, 1997), González-Insuasti and Caballero (2007), Blancas et al. (2010), Rangel-Landa et al. (2016), and Farfán et al. (2018a, b) may vary in complexity in the management strategies, differential investment of energy, complexity of tools, social agreements, and involving human selection with different levels of intensity and other evolutionary forces associated to management. It may be systematic, circumstantial, at random or following a plan, manual or involving tools and machines, generalist, or selective. For all these reasons, gathering should be considered a form of management. Currently, numerous plant resources are under management forms that are neither gathering nor agriculture, and that we have considered as “incipient” since they are less complex than agriculture (Casas et al. 1996, 1997, 2007, 2017; Clement et al. 2021). Among these management forms, the strategies of systematic, planned, and selective forms of gathering should be included, also, the tolerance of desirable plants when disturbing forests or when practicing weeding. It is also the case of induction or enhancement of abundance of desirable plants by sowing seeds, plating their vegetative propagules, or transplanting complete individuals, and also, the cases of plants protected through special ways to ensure their survival and reproduction as referred to above, including those from the wild, introduced deliberately to anthropogenic areas. All these forms of management are under different levels of intensity, and such intensity is related to the balance between the cultural and/or economic value, on the one hand, and their availability, on the other, which is commonly influenced by distribution, abundance, seasonal availability of products, vulnerability before interannual climate changes, pests, and natural or human-caused catastrophes, among other ecological aspects (Blancas et al. 2010, 2013; Rangel-Landa et al. 2016, 2017; Farfán et al. 2018a, b). Also, these are related to the resilience of individuals, communities, and ecosystems affected by human actions to use their products, depending on the plant part used and other biological aspects of the plants related to life cycle duration, reproductive systems, among others. Considering all these variables, it is possible to appreciate that plants used and managed by humans are under a continuous gradient of cultural/economic motivation of use, and ecological/biological aspects determining risk to ensure the availability of their products. Therefore, the management intensity is also expected to have a continuous expression of states. Through studies in different communities of the Tehuacán-Cuicatlán Valley (González-Insuasti and Caballero 2007; González-Insuasti et al. 2008, 2011; Blancas et al. 2013; Larios et al. 2013; Rangel-Landa et al. 2016, 2017), we analyzed the spectrum of forms and intensity of management of plants, mainly edible and ornamental plants in different rural communities. These studies show the broad spectrum of conditions of risk to ensure their availability, their relation to multiple ecological and social factors, and the responses to such risk. Likewise, the high relationship between risk conditions through the intensity of driving is highlighted.

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

45

Diversity of Domestication Processes Domestication is a consequence of management. Not all plants managed are domesticated nor eventually become domesticated, but all domesticated plants involve management. Through domestication, humans adjust forms, functions, and behavior of organisms according to human context (material and immaterial needs, values, esthetic purposes, and curiosities). Among the most important needs are food, most domesticated plants are edible, and humans select favoring quantity (e.g., number, size) and quality (flavor, color, texture, general aspect, and qualities associated with preparation, among others) of the edible products. However, a number of medicinal, ornamental and aromatic plants have been domesticated in Mexico. Most commonly, humans select in favor of several attributes of one or several plant traits, the processes producing a high diversification of the domesticated species. In addition to selection, people may drive gene flow and manipulate the reproduction system of plants and determine contexts for the propitious action of other evolutionary forces like inbreeding and genetic drift in small populations, bottleneck, and founder effects. The mechanisms and criteria associated with domestication are profusely linked to human culture; therefore, domestication is a biocultural expression. It is therefore important to document in studies of domestication the diversity of life forms of organisms under domestication, the diversity of attributes that people distinguish and value, and the diversity of mechanisms through which phenotypes are favored or unfavored and the action of other evolutionary forces. Domestication is an evolutionary process and therefore involves diversification. Darwin (1859) analyzed this process and adopted it in the first chapter of the “Origins of species” as a model to explain the origin of species in nature though developing the concept of natural selection. Domestication maintains and continually develops new varieties and in addition includes variation developed in different biocultural contexts through interchange of techniques, seeds and other propagules. It is a continuous process and therefore currently observable, which offers the possibility to document how it operates and provides to evolutionary biology and archaeology bases for interpreting what has happened in the natural evolutionary processes and ancient human-guided domestication. Through documenting domestication, it has been possible to describe and the broad spectrum of forms of plant management and ways through which human selection operates. This information is extraordinarily helpful to establish bases for sustainable management of genetic resources, particularly, to design strategies of in situ conservation. We have hypothesized that higher management intensity has caused higher divergence between managed and unmanaged organisms. Therefore, the silvicultural management is expected to determine lower differentiation with respect to wild populations than horticultural or agricultural management. For testing such hypotheses, we conducted several case studies. In all cases, we documented how variation in populations is perceived by people, how they value the variations and if they manage it differently. Ethnobotanical studies make possible documenting these aspects, as well as the mechanisms through which such variation is managed. The next step is evaluating the divergences (morphological, physiological, reproductive,

46

J. Caballero et al.

and genetic) and to test or reject the hypothesis. And we have analyzed all these aspects in annual plants, including some quelites like quintoniles Amaranthus spp. (Mapes et al. 1996, 1997), “alaches” (Anoda cristata (L.) Schltdl.), “chipiles” (Crotalaria pumila Ortega), some trees like “guajes” (Leucaena esculenta (Moc. & Sessé ex DC.) Benth; Casas et al. 1997, 2007; Casas and Caballero 1996; Zárate et al. 2005), tempesquistle (Sideroxylon palmeri (Rose) T.D.Penn.; GonzálezSoberanis and Casas 2004), and Ceiba aesculifolia (Kunth) Britten & Baker f. (Avendaño et al. 2006, 2009; Arellanes et al. 2013, 2018), which are widely appreciated and commercialized in the Tehuacán Valley and Oaxaca. All these trees are important since remains of them were found by archaeologists associated with humans in strata from prehistoric times of the Tehuacán Valley (Smith 1967). Other important trees that we have studied are the gourd trees Crescentia alata Kunth and C. cujete L. (Aguirre-Dugua et al. 2012, 2013, 2018) and the guava Psidium guajava L. (Arévalo-Marín et al. 2021). These and some columnar cacti species of the genus Stenocereus have allowed exploring questions related to the origin and diffusion of their domestication. We have found that C. alata appears to be native to Mexico, and its domestication has occurred in several areas of the territory. C. cujete has both native and introduced populations, but those with the clearest signs of domestication are genetically differentiated from the native populations, even where they coexist. We have not identified the area where these genotypes originated, but we have hypothesized that most probably such an area is in Central America, somewhere in Honduras or Nicaragua (Aguirre-Dugua et al. 2012, 2018; Moreira et al. 2017). The case of guava is a different story. Phylogenetic studies suggest that the genus Psidium originated and diversified in South America, and Psidium guajava is also from South America. It is a species with life history traits that make it able to spread and colonize wide areas. Arévalo-Marín et al. (2021) hypothesized several scenarios of its origin and diffusion, and one of the most probable is that the species arrived to Mexico thousands of years before the occupation of the territory by humans. However, the archaeological evidence indicates that the oldest remains are in South America, and its presence in Mesoamerica is relatively late. There are still several questions that are analyzed to clarify events of diffusion and domestications, and as in the cases of Crescentia, the phylogeographic approaches are particularly helpful. It is early to arrive at conclusions about this story, but the methodological approaches provided by molecular genetics, ecology, ethnohistory, and archaeology are keystones to reconstruct the natural and biocultural history of these species. The case of Stenocereus involved the analysis of a complex of related species grouped in the Stenocereus griseus (Haw.) Buxb. complex. Our study started exploring the origin and diffusion of S. pruinosus (Otto ex Pfeiff.) Buxb, a clearly domesticated species in central Mexico but with a wide distribution in Mexico. We soon found that what was identified as Stenocereus pruinosus were several species, including S. griseus. However, after a first step of our research we found that S. griseus is a South American species and that what was considered to be this species in Mexico was a well-differentiated new species, which was named Stenocereus huastecorum Alvarado-Sizzo, Arreola-Nava, and Terrazas (2018).

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

47

We have centered our attention in two additional systems: agaves and columnar cacti. In the case of agaves, we have documented the forest management of several wild populations (Agave potatorum Zucc., A. cupreata Trel. & A. Berger, and A. inaequidens K. Koch, A. angustifolia Haw.; Casarrubias-Hernández 2019; Delgado-Lemus et al. 2014; Illsley et al. 2018; Torres-García et al. 2013, 2015a, b, 2020) and states and changes associated to domestication in some of them and others (Agave inaequidens, A. hookeri Jacobi, A. salmiana Otto ex Salm-Dyck, A. mapisaga Trel., and A. americana; L. Colunga-GarcíaMarín et al. 2017; Figueredo-Urbina et al. 2017, 2018; Álvarez-Ríos et al. 2020). Studies of forest management have included the documentation of current rates of extraction, the effect of demand on it, and the effect of it on population dynamics and population genetics. In all cases, we have identified that the increasing demand of mescal has caused the extirpation of numerous populations of the target species. The extraction of adult individuals occurs just before producing flowers which cancel the sexual reproduction, the only form of reproduction in part of the species studied (A. potatorum, A. cupreata, and A. inaequidens). This fact directly affects the recovering capacity of the populations, which become extirpated progressively as the remaining individuals reach the extraction stage (Torres-García et al. 2015a, 2019). Some individuals escape to the extraction, but there is an effect densitydependent influencing the visits of bats to flowers to cause pollination. Several studies have identified that at least 30% of individuals at reproductive stage should be maintained in a population to allow pollinators visiting flowers; below such threshold, bats rarely visit a population (Torres-García et al. 2013). Species that produce vegetative propagules are able to recover their populations but reducing genetic diversity and therefore increasing their vulnerability to several factors. Several demographic models developed by our research group have identified the stages whose maintenance and growth are crucial for ensuring an appropriate growth population rate. These aspects may vary from population to population and among species. But proposals for actions have been possible. In all cases, ensuring pollination is crucial for preventing loss of genetic diversity, and in some species like Agave potatorum facilitation interactions with some species of shrubs are equally important to ensure the establishing of seedlings (Rangel-Landa et al. 2015). Studies on domestication were conducted documenting morphological, genetic, and phytochemical (saponin content) divergence between wild and domesticated populations, as specified for the general research strategy. In the case of Agave hookeri, the wild relative is unknown, but we performed a comparison with wild and domesticated populations of A. inaequidens, which has been proposed to be the wild relative (Figueredo-Urbina et al. 2015, 2017, 2018). In the case of A. americana, the study is still in progress, the wild subspecies (A. americana subsp. protoamericana) is being compared with the domesticated subspecies (A. americana subsp. americana), and divergences between varieties of the latter subspecies are being analyzed. Something similar was performed with varieties of A. salmiana, A. americana, and A. mapisaga whose varieties are managed together, some of them possibly being hybrids (Álvarez-Ríos et al. 2020).

48

J. Caballero et al.

The system that has been studied with more detail is that of the columnar cacti, which are important plant resources in several regions of Mexico. This system allows analyzing in one region several species in a gradient of management intensity. We have characterized such intensity in relation to the energy invested in managing plant populations versus the productivity of the managed system. This balance is influenced by the viability of management, which is very much influenced by the growth rate of the plants and the viability of managing vegetative propagules. Species like Escontria chiotilla (F.A.C. Weber ex K. Schum.) Rose and Polaskia chende (Rol.-Goss) (A.C. Gibson & K.E. Horak) have slow growth and are difficult to cultivate; others like Cephalocereus tetetzo (F.A.C. Weber ex J.M. Coult.) Diguet and Pachycereus weberi (J.K. Coult.) (Backeb) produce tasty fruits, seeds, and flower buds very much appreciated by people, but their growth is even slower than E. chiotilla. These species are let standing, protected, or transplanted (young plants) in agroforestry systems. Other species like Stenocereus pruinosus, S. stellatus Riccob., Lemaireocereus hollianus (F.A.C. Weber ex J.M. Coult.), and Britton & Rose are intensively cultivated in homegardens, live fences, and borders of agricultural plots. These species grow faster, and selection is easier than in the other species mentioned, and they show clear signs of domestication (Casas et al. 2007; Parra et al. 2010; Rodríguez-Arévalo et al. 2006). Studies of population genetics through neutral markers have showed that genetic differences among populations with different management are difficult to be visualized. Further studies with markers associated to traits could be more informative in this respect. For the moment, it has been found that silvicultural managed and cultivated populations have less genetic variation than wild populations, a pattern generally expected. However, in cultivated populations of Stenocereus stellatus, S. pruinosus, and in some cultivated agaves, we found higher genetic variation than in the wild. This is an interesting pattern that we have discussed considering the high gene flow among wild and managed populations, as well as the active movement of propagules from different communities and regions, carried out by people. To analyze this pattern, we have explored the provenance of materials from silvicultural managed and cultivated populations through interviews and molecular markers (Parra et al. 2010, 2012; Cruse-Sanders et al. 2013); this information allows identifying wild and managed populations that are sources of cultivated material within the territory of a community or among regions (the Tehuacan Valley and La Mixteca Baja region). These data illustrate the great capacity of traditional people to continually introduce and replace diversity in their management systems, and their crucial role in conserving and increasing the diversity these contain.

Perspectives Our studies have documented different types of interactions between people and plants in Mexico. These interactions are motivated by the cultural value of the products used by people, as well as their ecological attributes and biological features that make viable or not their management. Use, management, and ecological knowledge are closely interconnected, and therefore their systematization is extraordinarily

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

49

important to document the current state of biocultural diversity and for designing innovations based on what we currently know. In addition, these data and relations have high importance to construct and test hypotheses about the past processes that motivated people to manage plants. Documenting and systematizing ethnobotanical knowledge continues to be an important task. This is especially necessary among human cultures and ecosystems poorly or not studied, as identified in this diagnosis. Field work efforts are important in relation to plant use, but studies of management require to be emphasized, especially in relation to factors motivating management, innovation techniques, and domestication. The current state of information allows visualizing that there are hundreds of case studies yet to be analyzed to understand the context and patterns of management and efforts to catalogue the management techniques. It is important to mention that nowadays numerous scholars have worked in local or regional databases throughout the country, and the National Commission for the Knowledge and Use of Biodiversity (CONABIO) has enhanced an important project for systematizing the information on use and management of biodiversity. The effort by CONABIO dedicated to document and systematize information from a project on agrobiodiversity is outstanding. All these efforts spread throughout the country could be coordinated and shared. It requires leading institutions and clear rules to operate, but such a project is possible and necessary. Analyzing ecological and evolutionary consequences of management is a relevant avenue of research, the first one to develop strategies of sustainable management of forest and agroforestry systems, the second to understand the evolution of managed plants, and both related with cultural and social changes associated with it. Morphometric, physiological, reproductive systems and population genetics are important tools to analyze them, and the new methods related to the genomic approaches are extraordinary opportunities to clarify the history of the processes. Now it is also relevant to consider the inter-scalar influence of domestication at population and landscape levels, and such influence should be studied in depth. The collaboration of ethnobotanists using similar methods for studying different regions and cultural groups is relevant to produce comparable data to identify general biocultural patterns and contributing to construct theoretical frameworks. In addition, the complexity of the biocultural and social-ecological issues related to ethnobotany should enhance ethnobotanists to carry out interdisciplinary research, while the bridge that ethnobotanical research may construct between traditional ecological knowledge and the academy and other sectors indicates the extraordinary role ethnobotany may play to construct transdisciplinary approaches for constructing sustainability science. Acknowledgments The authors thank the Instituto de Biología, IIES and ENES-Morelia, UNAM, the CIBYC at the Universidad Autónoma del Estado de Morelos, the UIIM-Michoacán and CONACYT for support through the program Investigadoras e Investigadores por México, and financial support to the project A1S-14306. Also, we thank support from the GEF Project ID 9380 CONABIO-GEF-FAO/RG023 “Manejo y domesticación de agrobiodiversidad en Mesoamérica: Bases para la soberanía alimentaria sustentable,” and PAPIT, UNAM (project IN206520 and IN224023).

50

J. Caballero et al.

Appendix 1 Communities, regions, and cultural groups where our research groups have conducted studies, which are referred to in maps of Fig. 1. Communities Comondú

Region Baja California Península

El Pescadero

Baja California Península

La Purísima

Baja California Península

Mulegé

Baja California Península

San Ignacio

Baja California Península

San Isidro

Baja California Península

San Javier

Baja California Península

Santa Gertrudis

Baja California Península

Todos Santos

Baja California Península

Lacanjá Chansayab Tumbalá Ejido Cuiteco

Montes Azules

Antiguos Mineros del Norte Boquillas La Vega San Lorenzo Xichú Axaxacualco

North mountains Tarahumara mountains Cuatrociénegas Valley Cuatrociénegas Valley Cuatrociénegas Valley Cuatrociénegas Valley Sierra Gorda Balsas basin

State Baja California Sur Baja California Sur Baja California Sur Baja California Sur Baja California Sur Baja California Sur Baja California Sur Baja California Sur Baja California Sur Chiapas

Municipality Comondú

Ethnic groups Mestizo

Los Cabos

Mestizo

Comondú

Mestizo

Mulegé

Mestizo

Mulegé

Mestizo

Comondú

Mestizo

Comondú

Mestizo

San Quintín

Mestizo

La Paz

Mestizo

Bonampak

Lacandón

Chiapas Chihuahua

Tumbalá Urique

Ch’ol Raramuri

Coahuila

Cuatrociénegas

Mestizo

Coahuila

Cuatrociénegas

Mestizo

Coahuila

Cuatrociénegas

Mestizo

Coahuila

Cuatrociénegas

Mestizo

Guanajuato Guerrero

Xichú Eduardo Neri

Mestizo Nahuatl (continued)

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

51

Communities San José Huitziltepec Acateyahualco

Region Balsas basin

State Guerrero

Municipality Eduardo Neri

Ethnic groups Nahuatl

Mountain

Guerrero

Ahuacuotzingo

Agua Zarca

Mountain

Guerrero

Ahuacuotzingo

Alcozauca Amapilca Chilapa

Mountain Mountain Mountain

Guerrero Guerrero Guerrero

Alcozauca Alcozauca Chilapa

Copanatoyac Huamuxtitlán

Mountain Mountain

Guerrero Guerrero

Copanatoyac Huamuxtitlán

Ixcuinatoyac Olinalá San José Laguna Tecolcuautla Tehuitzingo (Tlahuitzingo) Tlapa

Mountain Mountain Mountain Mountain Mountain

Guerrero Guerrero Guerrero Guerrero Guerrero

Alcozauca Olinalá Alcozauca Ahuacuotzingo Olinalá

Náhuatl/ Mestizo Náhuatl/ Mestizo Mixtec/Mestizo Mixtec/Mestizo Náhuatl/ Mestizo Mixtec Náhuatl/ Mestizo Mixtec Náhuatl Mixtec Náhuatl Náhuatl

Mountain

Guerrero

Tlapa

Trapiche Viejo

Mountain

Guerrero

Chilapa

Xocoyolzintla San Miguel Xicalco Cañada del Agua Pino Real Real de Otzumatlán Rio de Parras Cuanajo

Mountain Southeast of Mexico City Basin Cuitzeo Basin Cuitzeo Basin Cuitzeo

Guerrero Mexico city Michoacan Michoacan Michoacan

Ahuacuotzingo Tlalpan

Náhuatl, Mixtec, Tlapanec Náhuatl/ Mestizo Náhuatl Mestizo

Indaparapeo Charo Queréndaro

Mestizo Mestizo Mestizo

Basin Cuitzeo Lake Patzcuaro region Lake Patzcuaro region Lerma-Chapala region Tierra Caliente region Tierra Caliente region Zitacuaro region Lake Patzcuaro region Monarca region

Michoacan Michoacan

Queréndaro Pátzcuaro

Mestizo Purhepechas

Michoacan

Quiroga

Mestizo

Michoacan

Sahuayo

Mestizo

Michoacan

Churumuco

Mestizo

Michoacan

La Huacana

Mestizo

Michoacan Michoacán

Zitácuaro Erongarícuaro

Mazahua Purhépecha

Michoacán

Zitácuaro

Mazahua/ Mestizo

Icuacato Barranca del Aguacate El Chocolate Ichamio Francisco Serrato Erongarícuaro Zitácuaro

(continued)

52

Communities Undameo Pichátaro Infiernillo

J. Caballero et al.

Region Morelia Purhépecha region Tierra Caliente region Tierra Caliente region Balsas basin Balsas basin Balsas basin Balsas basin Balsas basin

State Michoacán Michoacán Michoacán

Municipality Morelia Pichátaro Infiernillo

Ethnic groups Mestizo Purhépecha Mestizo

Michoacán

Infiernillo

Mestizo

Morelos Morelos Morelos Morelos Morelos

Jantetelco Cuautla Cuernavaca Tepalcingo Tepalcingo

Mestizo Mestizo Mestizo Mestizo Mestizo

Balsas basin Balsas basin Balsas basin Balsas basin Balsas basin

Morelos Morelos Morelos Morelos Morelos

Puente de Ixtla Jojutla Miacatlán Tepalcingo Huitzilac

Morelos

Huitzilac

Morelos

Huitzilac

Morelos

Tepoztlán

Morelos

Tlayacapan

Morelos

Totolapan

Nahuatl

Cuilapam de Guerrero

Highlands of the state of Morelos Highlands of the state of Morelos Highlands of the state of Morelos Highlands of the state of Morelos Highlands of the state of Morelos Central Valleys of Oaxaca

Mestizo Mestizo Mestizo Mestizo Nahuatl, Mestizo Nahuatl, Mestizo Nahuatl, Mestizo Nahuatl, Mestizo Nahuatl

Oaxaca

Cuilapam de Guerrero

Coyula

Cuicatlán valley

Oaxaca

Cuicatlán

Cuicatlán

Cuicatlán valley

Oaxaca

Cuicatlán

Dominguillo Ixcatlán Jocotipac Nodón Quiotepec

Cuicatlán valley Cuicatlán valley Cuicatlán valley Cuicatlán valley Cuicatlán valley

Oaxaca Oaxaca Oaxaca Oaxaca Oaxaca

Cuicatlán Ixcatlán Jocotipac Cuicatlán Cuicatlán

San Lorenzo Pápalo Tecomavaca

Cuicatlán valley

Oaxaca

Cuicatlán

ZapotecoMixtecoMestizo Cuicatec/ Mestizo Mestizo/ Cuicatec Mestizo Ixcatec Mixtec Mixtec Mestizo/ Cuicatec Cuicatec

Cuicatlán valley

Oaxaca

Tecomavaca

Pitirera Chalcatzingo Cuautla Cuernavaca Ejido Los Sauces El Limón de Cuauhchichinola El Zapote Jojutla Palpan de Baranda Tepalcingo Tres Marías Coajomulco Huitzilac Tepoztlán Tlayacapan Totolapan

Nahuatl/ Mazatec/ Mestizo (continued)

Ethnobotanical Knowledge in Mexico: Use, Management, and. . . Communities Santa Catalina Chinango Tequixtepec Tonaguia

53

Region Low Mixteca

State Oaxaca

Municipality Tequixtepec

Ethnic groups Mixtec

Low Mixteca Sierra de Juarez (north region) Sierra Sur

Oaxaca Oaxaca

Mixtec Mixe

Oaxaca

San Juan de Los Cúes

Tehuacán valley

Oaxaca

Tequixtepec Santo Domingo Roayaga Putla Villa de Guerrero Teotitlán

Teotitlán del Camino

Tehuacán valley

Oaxaca

Teotitlán

Chazumba Acateno

Low Mixteca Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla

Puebla Puebla

Chazumba Acateno

Nahuatl/ Mazatec/ Mestizo Nahuatl/ Mazatec/ Mestizo Mixtec Nahuatl

Puebla

Ahuacatlán

Nahuatl

Puebla

Nahuatl

Puebla

Ayotoxco de Guerrero Chignahuapan

Puebla

Cuetzalan

Nahuatl

Puebla

Nahuatl

Puebla

Francisco Z. Mena Huachinango

Puebla

Huehuetla

Nahuatl

Puebla

Hueyapan

Nahuatl

Puebla

Hueytamalco

Nahuatl

Puebla

Jonotla

Nahuatl

Puebla

Libres

Nahuatl

Puebla

Naupan

Nahuatl

Puebla

Nauzontla

Nahuatl

Puebla

Olintla

Nahuatl

Puebla

Pahuatlan

Nahuatl

El Campanario

Ahuacatlán Ayotoxco de Guerrero Chignahuapan Cuetzalan Francisco Z. Mena Huachinango Huehuetla Hueyapan Hueytamalco Jonotla Libres Naupan Nauzontla Olintla Pahuatlan

Nahuatl

Nahuatl

(continued)

54

Communities Tepecintla

J. Caballero et al.

Region Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Northern sierra of Puebla Sierra Negra Sierra Negra Sierra Negra

State Puebla

Municipality Tepecintla

Ethnic groups Nahuatl

Puebla

Teziutlan

Nahuatl

Puebla

Tlatlauquitepec

Nahuatl

Puebla

Tuzamapan de Galeana Cuetzalan

Nahuatl

Venustianao Carranza Xicotepec

Nahuatl

Nahuatl

Puebla

Xochitlán de Vicente Suarez Zacapoaxtla

Puebla

Zacatlán

Nahuatl

Puebla

Nahuatl

Puebla

Zapotitlán de Méndez Zautla

Puebla

Zongoxotla

Nahuatl

Puebla

Zoquiapan

Nahuatl

Puebla Puebla Puebla

Coyomeapan Coyomeapan Coyomeapan

Nahuatl Nahuatl Nahuatl

Sierra Negra

Puebla

Coyomeapan

Nahuatl

Sierra Negra

Puebla

Coyomeapan

Nahuatl

Sierra Negra Sierra Negra Tehuacan valley

Puebla Puebla Puebla

Coyomeapan Coyomeapan Ajalpan

Aticpac Caltepec Chilac

Sierra Negra Tehuacan valley Tehuacan valley

Puebla Puebla Puebla

Coyomeapan Caltepec Tehuacán

Chimalhuaca Coatepec

Sierra Negra Tehuacan valley

Puebla Puebla

Coyomeapan Caltepec

Nahuatl Nahuatl Nahuatl/ Mestizo Nahuatl Mestizo Nahuatl/ Mestizo Nahuatl Nahuatl/ Mestizo

Teziutlan Tlatlauquitepec Tuzamapan de Galeana Tzinacapan Venustianao Carranza Xicotepec Xochitlán de Vicente Suárez Zacapoaxtla Zacatlán Zapotitlán de Méndez Zautla Zongoxotla Zoquiapan Caxalli Matlahuacala San Gabriel Vista Hermosa San Marcos Tlatlalkilotl Santa María Coyomeapan Xochitlalpa Ahuatla Ajalpan

Puebla Puebla Puebla Puebla

Nahuatl

Nahuatl

Nahuatl

Nahuatl

(continued)

Ethnobotanical Knowledge in Mexico: Use, Management, and. . . Communities Coxcatlán

Region Tehuacan valley

State Puebla

Municipality Coxcatlán

Guadalupe Victoria Ixtacxochitla Reyes Metzontla

Tehuacan valley

Puebla

Coxcatlán

Sierra Negra Tehuacan valley

Puebla Puebla

Zoquitlan Zapotitlán

San Juan Raya San Luis Atolotitlán San Nicolás Tepoxtitlán San Rafael Santiago Tilapa

Tehuacan valley Tehuacan valley

Puebla Puebla

Zapotitlán Caltepec

Tehuacan valley

Puebla

Atexcal

Tehuacan valley Tehuacan valley

Puebla Puebla

Tilapa Tilapa

Tehuacán Yohuajca Zapotitlán Salinas

Tehuacan valley Sierra Negra Tehuacan valley

Puebla Puebla Puebla

Tehuacán Coyomeapan Zapotitlán

Zinacatepec

Tehuacan valley

Puebla

Zinacatepec

Zoquitlán Acaxochitlan

Sierra Negra Northern sierra of Puebla Yucatan Península

Puebla Puebla

Zoquitlán Acaxochitlan Felipe Carrillo Puerto Real de Catorce

Xkon Ha Wirikuta (Las Margaritas ejido) Aquismón

Huasteca

Tancuime

Huasteca

El Rosario San Isidro Buen Suceso Maxcanú Sucilá

North region South region

Quintana Roo San Luis Potosí San Luis Potosí San Luis Potosí Tlaxcala Tlaxcala

Yucatán Península Yucatán Península

Yucatán Yucatán

Altiplano region

55

Ethnic groups Nahuatl/ Mestizo Mestizo Nahuatl Popoloca/ Mestizo Mestizo Mestizos Mestizo/ Nahuatl Mestizo Nahuatl/ Mestizo Mestizo Nahuatl Mixtec/ Popoloca/ Mestizo Nahuatl/ Mestizo Nahuatl Nahuatl Maya

Aquismón

Mestizo, Wixarika Huastec

Aquismón

Huastec

Tlaxco San Pablo del Monte Maxcanú Sucilá

Mestizo Nahua Maya Maya

References Abad-Fitz I, Maldonado-Almanza B, Aguilar-Dorantes K, Gómez-Caudillo L, Casas A, Blancas J, Beltrán-Rodríguez L, Sierra-Huelsz JA, Cristians-Nizawa S, Moreno-Calles AI, GarcíaRodríguez Y, Espinosa-García F, Sánchez-Méndez L. Consequences of traditional management

56

J. Caballero et al.

in the production and quality of copal resin (Bursera bipinnata (Moc. & Sessé ex DC) Engl.) in Mexico. Forests 2020;11(9):991. Aguirre-Dugua X, González A, Eguiarte L, Casas A. Round and large: morphological and genetic consequences of artificial selection on the gourd tree Crescentia cujete by the Maya from the Yucatán Peninsula, Mexico. Ann Bot. 2012;109:1307–16. Aguirre-Dugua X, Casas A, Pérez-Negrón E. Phenotypic differentiation between wild and domesticated varieties of Crescentia cujete and culturally relevant uses of fruits as bowls in the Yucatan Peninsula, Mexico. J Ethnobiol Ethnomed. 2013;9:76. Aguirre-Dugua X, Llanderal-Mendoza J, González-Rodríguez A, Eguiarte L, Casas A. Anthropogenic dispersion of selected germplasm creates a new phylogeographic mosaic of Crescentia cujete populations in Mesoamerica. Tree Genet Genomes. 2018;14:18. Alcorn JB. Huastec Mayan ethnobotany. Austin: University of Texas Press; 1984. Alvarado-Sizzo H, Casas A, Parra F, Arreola-Navas HJ, Terrazas T, Sánchez C. Species delimitation of Stenocereus griseus (Cactaceae) species complex reveals the new species S. huastecorum PloS One 2018;13(1):1–25. Álvarez-Ríos GD, Pacheco-Torres F, Figueredo-Urbina CJ, Casas A. Management, morphological and genetic diversity of domesticated agaves in Michoacán, México. J Ethnobiol Ethnomed. 2020;16(1):1–17. Ardelean CF, Becerra-Valdivia L, Pedersen MW, Schwenninger JL, Oviatt CG, Macías-Quintero JI, Arroyo-Cabrales J, Sikora M, Ocampo-Díaz YZE, Rubio-Cisneros II, Watling JG, Medeiros VB, Oliveira PE, Barba-Pingarón J, Ortiz-Butrón A, Blancas-Vázquez J, Rivera-González I, Solís-Rosales C, Rodríguez-Ceja M, Gandy DA, Navarro-Gutierrez Z, De La Rosa-Díaz JJ, Huerta-Arellano V, Marroquín-Fernández MB, Martínez-Riojas LM, López-Jiménez A, Higham T, Willerslev E. Evidence of human occupation in Mexico around the Last Glacial Maximum. Nature. 2020;584:87–92. Arellanes Y, Casas A, Arellanes A, Vega E, Blancas J, Vallejo M, Torres I, Rangel-Landa S, Moreno-Calles AI, Solís L, Pérez-Negrón E. Influence of traditional markets on plant management in the Tehuacan Valley. J Ethnobiol Ethnomed. 2013;9(1):38. Arévalo-Marín E, Casas A, Landrum L, Ruíz-Sánchez E, Alvarado-Sizzo H, Clement CR. The taming of Psidium guajava: natural and cultural history of a tropical plant variation. Front Plant Sci. 2021;12:714763. Avendaño A, Casas A, Dávila P, Lira R. Traditional use, management and commercialization of “pochote” Ceiba aesculifolia (H. B. & K.) Btitten & Baker subsp. parvifolia (Rose) P. E. Gibbs & Semir (Bombacaceae) in the Tehuacán Valley, Central Mexico. J Arid Environ. 2006;67: 15–35. Avendaño A, Casas A, Dávila P, Lira R. In situ management and patterns of morphological variation of Ceiba aesculifolia subsp. parvifolia in the Tehuacán Valley. Econ Bot. 2009;63(2):138–51. Barrera-Marín A, Barrera-Vásquez A, López-Franco RM. Nomenclatura etnobotánica maya. Mérida: Instituto Nacional de Antropología e Historia, Centro Regional del Sureste; 1976. Benz BF. Maize in the Americas. Maize in the Americas. In: Staller J, Tykot R, Benz BF, editors. Histories of maize: multi-disciplinary approaches to the prehistory, linguistics, biogeography and evolution of maize. Boston: Elsevier Academic Press; 2006. p. 9–20. Benz BF, Santana FJ, Pineda MR, Cevallos J, Robles L, De Niz D. Characterization of Mestizo plant use in the Sierra de Manantlán, Jalisco-Colima, Mexico. J Ethnobiol. 1994;14:23–41. Berlin B, Breedlove DE, Raven PH. Principles of Tzeltal plant classification: an introduction to the botanical ethnography of a Mayan-speaking people of Highland Chiapas. New York: Academic Press; 1974. Blancas J, Casas A, Rangel-Landa S, Moreno-Calles AI, Torres I, Pérez-Negrón E, Solís L, Delgado-Lemus A, Parra F, Arellanes Y, Caballero J, Cortés L, Lira R, Dávila P. Plant management in the Tehuacán-Cuicatlán Valley, Mexico. Econ Bot. 2010;64:287–302. Blancas J, Casas A, Pérez-Salicrup D, Caballero J, Vega E. Ecological and socio- cultural factors influencing plant management in Nahuatl communities of the Tehuacan Valley, Mexico. J Ethnobiol Ethnomed. 2013;9(1):39.

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

57

Boege E. El patrimonio biocultural de los pueblos indígenas de México. Mexico City: Instituto Nacional de Antropología e Historia, Comisión Nacional para el Desarrollo de los Pueblos Indígenas; 2008. Bunge-Vivier V, Martínez-Ballesté A. Factors that influence the success of conservation programs in common property resources in Mexico. Int J Commons. 2017;11(1):102088. Bye RA. Ethnoecology of the Tarahumara of Chihuahua, Mexico. Ph.D. dissertation, Department of Biology, Harvard University, Cambridge, MA; 1976. Caballero J. Recursos Comestibles Potenciales. In: Reyna T, editor. Seminario sobre la alimentación en México. Mexico City: Instituto de Geografia, Universidad Nacional Autónoma de México; 1984. p. 114–25. Caballero J. La Dimension Culturelle de la Diversite Vegetale au Mexique. J Agric Trad Bot Appl Nouv Sér. 1994;XXXVI(2):1–12. Caballero J, Cortés L. Percepción, uso y manejo tradicional de los recursos vegetales en México. In: Rendón B, Rebollar S, Caballero J, Martínez MA, editors. Plantas Cultura y Sociedad. Mexico City: Secretaría del Medio Ambiente Recursos Naturales y Pesca, Universidad Autónoma Metropolitana; 2001. p. 79–100. Caballero J, Mapes C. Gathering and subsistence patterns among the Purhepecha of Patzcuaro, Mexico. J Ethnobiol. 1985;5(1):31–4. Caballero J, Rubluo A, Bye R, Palomino G. La Unidad de Investigación sobre Recursos Genéticos del Jardín Botánico de la UNAM. In: Memorias de la I Reunión Nacional de Jardines Botánicos. Mexico City: Asociación Nacional de Jardines Botánicos; 1985. p. 44–53. Caballero J, Casas A, Cortés L, Mapes C. Patrones en el conocimiento, uso y manejo de plantas en pueblos indígenas de México. Estud Atacameños. 1998;16:181–96. Camou-Guerrero A, Reyes-García V, Martínez-Ramos M, Casas A. Knowledge and use value of plant species in a Raramuri community: a gender perspective. Hum Ecol. 2008;36:259–72. Cano M, De la Tejera B, Casas A, García-Barrios R, Salazar L. Migración y estructura de huertos familiares en la comunidad tlahuica de San Juan Atzingo, México. Bot Sci. 2012;90(3): 287–304. Casarrubias-Hernández MG. Uso, manejo tradicional y aspectos ecológicos del maguey zacatoro (Agave angustifolia Haw.) en el municipio de Ahuacuotzingo, Guerrero, México. Bachelor thesis, Universidad Autónoma de Guerrero, Chilpancingo; 2019. Casas A, Caballero J. Traditional management and morphological variation in Leucaena esculenta (Moc. et Sessé ex A.DC.) Benth. in the Mixtec region of Guerrero, Mexico. Econ Bot. 1996;50(2):167–81. Casas A, Viveros JL, Caballero J. Etnobotánica mixteca: sociedad, cultura y recursos naturales en la Montaña de Guerrero. Mexico City: Instituto Nacional Indigenista CONACULTA; 1994. Casas A, Vázquez MC, Viveros JL, Caballero J. Plant management among the Nahua and the Mixtec of the Balsas river basin: an ethnobotanical approach to the study of plant domestication. Hum Ecol. 1996;24(4):455–78. Casas A, Caballero J, Mapes C, Zárate S. Manejo de la vegetación, domesticación de plantas y origen de la agricultura en Mesoamérica. Bot Sci. 1997;61:31–47. Casas A, Valiente-Banuet A, Viveros JL, Caballero J, Cortés L, Dávila P, Lira R, Rodríguez I. Plant resources of the Tehuacán Valley, México. Econ Bot. 2001;55(1):129–66. Casas A, Otero-Arnaiz A, Peréz-Negrón E, Valiente-Banuet A. In situ management and domestication of plants in Mesoamerica. Ann Bot. 2007;100:1101–15. Casas A, Camou A, Otero-Arnaiz A, Rangel-Landa S, Cruse-Sanders J, Solís L, Torres I, Delgado A, Moreno-Calles AI, Vallejo M, Guillén S, Blancas J, Parra F, FarfánHeredia B, Aguirre-Dugua X, Arellanes Y. Manejo tradicional de biodiversidad y ecosistemas en Mesoamérica: el Valle de Tehuacán. Invest Ambient Cienc Polít Públ. 2014;6:23–44. Casas A, Parra F, Aguirre-Dugua X, Rangel-Landa S, Blancas J, Vallejo M, Moreno-Calles AI, Guillén S, Torres I, Delgado-Lemus A, Pérez-Negrón E, Figueredo CJ, Cruse-Sanders JM, Farfán-Heredia B, Solís L, Otero-Arnaiz A, Alvarado-Sizzo H, Camou-Guerrero A. Manejo y

58

J. Caballero et al.

domesticación de plantas en Mesoamérica. Una estrategia de investigación. In: Casas A, TorresGuevara J, Parra F, editors. Domesticación en el Continente Americano. Volumen 2. Perspectivas de investigación y manejo sustentable de recursos genéticos en el Nuevo Mundo. Morelia: Universidad Nacional Autónoma de México/Universidad Nacional Agraria La Molina/CONACYT; 2017. p. 69–102. Clement CR, Casas A, Parra-Rondinel FA, Levis C, Peroni N, Hanazaki N, Cortés-Zárraga L, Rangel-Landa S, Alves RP, Ferreira MJ, Cassino MF, Coelho SD, Cruz-Soriano A, PancorboOlivera M, Blancas J, Martínez-Ballesté A, Lemes G, Lotero-Velásquez E, Bertin VM, Mazzochini GG. Disentangling domestication from food production systems in the Neotropics. Quaternay. 2021;4(1):4. Colunga-GarcíaMarín P, Zizumbo-Villarreal D, Casas A, Delgado-Lemus A, Figueredo CJ, RangelLanda S, Eguiarte L, Aguirre-Dugua X, Vargas O, Cabrera-Toledo D. Los agaves y las prácticas mesoamericanas de aprovechamiento, manejo y domesticación. In: Domesticación en el Continente Americano. Volumen 2. Perspectivas de investigación y manejo sustentable de recursos genéticos en el Nuevo Mundo, pp. 273–310, Casas A, Torres-Guevara J, Parra F. (eds). Universidad Nacional Autónoma de México/Universidad Nacional Agraria La Molina, Morelia, Michoacán, Mexico. 2017. CONANP/SEMARNAT (Comisión Nacional de Áreas Naturales Protegidas/Secretaría de Medio Ambiente y Recursos Naturales). Programa de Conservación y Manejo Reserva de la Biosfera Los Tuxtlas. México City: Comisión Nacional de Áreas Naturales Protegidas; 2006. Corona E, Casas A, Argueta-Villamar A, Alvarado-León CI. La domesticación de especies y paisajes en México. In: Bali J, editor. México. Grandeza y diversidad. Mexico City: Instituto Nacional de Antropología e Historia (INAH); 2021. p. 78–98. Cruse-Sanders JM, Friar EA, Parker KC, Huang DI, Mashayekhi S, Prince LM, Otero-Arnaiz A, Casas A. Managing diversity: domestication and gene flow in Stenocereus stellatus Riccob. in Mexico. Ecol Evol. 2013;3(5):1340–55. Cuevas E, Blancas J, Caballero J, Hinojosa-Díaz IA, Martínez-Ballesté A. Agricultural management and local knowledge: key factors for the conservation of socio-ecosystems in the face of the pollinator world crisis. Bot Sci. 2021;99(2):305–20. Darwin C. On the origin of species by means of natural selection, or the preservation of favored races in the struggle for life. London: John Murray; 1859. De Candolle A. Darwin considéré au point de vue des causes de son succès et de L’importance de ses travaux. Geneva: H. Georg; 1882. Delgado-Lemus A, Casas A, Téllez-Valdés O. Distribution, abundance and traditional management of Agave potatorum in the Tehuacán Valley, Mexico: perspectives for sustainable management of non-timber forest products. J Ethnobiol Ethnomed. 2014;10:63. De-Nova JA, González-Trujillo R, Castillo-Lara P, Fortanelli-Martínez J, Mora-Olivo A, SalinasRodríguez MM. Inventario florístico de la Reserva de la Biosfera Sierra del Abra Tanchipa, San Luis Potosí, México. Bot Sci. 2019;97(4):761–88. Eberhard DM, Simons GF, Fennig CD, editors. Ethnologue: languages of the world. 25th ed. Dallas: SIL International. Online version: http://www.ethnologue.com; 2022. Farfán B, Casas A, García-Frapolli E, Moreno-Calles AI, Castilleja A. Ethnoecology and interchange of wild and weedy plant products in the traditional Phurépecha markets. J Ethnobiol Ethnomed. 2018a;14(5):1–19. Farfán B, Casas A, Rangel-Landa S. Cultural, economic and ecological factors influencing management of wild plants and mushrooms interchanged in Purépecha markets of Mexico. J Ethnobiol Ethnomed. 2018b;14(68):1–21. Felger RS, Moser MB. People of the desert and sea: ethnobotany of the Seri Indians. Tucson: University of Arizona Press; 1985. Figueredo-Urbina CJ, Casas A, Rocha V, Colunga-GarcíaMarín P, Nassar J, González A. Genetic structure of coexisting wild, silvicultural managed and cultivated populations of Agave inaequidens in Michoacán, Mexico. AoB Plants. 2015;7:plv114.

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

59

Figueredo-Urbina CJ, Casas A, Torres-García I. Morphological and genetic divergences among the wild Agave inaequidens and A. cupreata with the domesticated species Agave hookeri. An analysis of their evolutionary relationships. PLoS One. 2017;12(11):e0187260. Figueredo-Urbina CJ, Casas A, Martínez-Díaz Y, Santos-Zea L, Gutiérrez-Uribe JA. Domestication and saponins contents in a gradient of management intensity of agaves: Agave cupreata, A. inaequidens and A. hookeri in central Mexico. Genet Resour Crop Evol. 2018;65:1133–46. Flannery K, editor. Guilá Naquitz. New York: Academic Press; 1986. Flores JS. Etnobotánica de las leguminosas en la Península de Yucatán: uso y manejo entre los mayas. Ph.D. dissertation, Facultad de Ciencias, Universidad Nacional Autónoma de México, Mexico City; 1999. González-Insuasti MS, Caballero J. Managing plant resources: how intensive can it be? Hum Ecol. 2007;35:303–14. González-Insuasti MS, Martorell C, Caballero J. Factors that influence the intensity of non-agricultural management of plant resources. Agrofor Syst. 2008;74:1–15. González-Insuasti MS, Casas A, Méndez-Ramírez I, Martorell C, Caballero J. Intra-cultural differences of the importance of plant resources and their impact on management intensification. Hum Ecol. 2011;39(2):191–202. González-Soberanis C, Casas A. Traditional management and domestication of tempesquistle, Sideroxylon palmeri (Sapotaceae) in the Tehuacán-Cuicatlán Valley, Central Mexico. J Arid Environ. 2004;59(2):245–58. Hanelt P. Mansfeld’s encyclopedia of agricultural and horticultural crops. Gatersleben: Institute of Plant Genetics and Crop Plant Research (IPK); 2001. Harlan J. Crops and man. Madison: American Society of Agronomy and Crop Science Society of America; 1975. Harris DR, editor. The Origins and spread of agriculture and pastoralism in Eurasia. London: UCL Press; 1996. Harris DR, Hillman GC, editors. Foraging and farming: the evolution of plant exploitation. Abingdon/New York: Routledge; 1989. Hawkes JG. The diversity of crop plants. Cambridge, MA: Harvard University Press; 1983. Illsley C, Torres-García I, Hernández-López JJ, Morales-Moreno P, Varela-Álvarez R, IbañezCouch I, Nava-Xinol H. Manual de manejo campesino de magueyes mezcaleros forestales. Grupo de Estudios Ambientales, Mexico City, Mexico. 2018. Instituto Nacional de Lenguas Indígenas (INALI). Catálogo de las lenguas indígenas nacionales: variantes lingüísticas de México con sus autodenominaciones y referencias geoestadísticas. Ciudad de México: Diario Oficial; 2008. Kirchhoff P. Mesoamérica: sus límites geográficos, composición étnica y caracteres culturales. Acta Am. 1943;1:92–107. Kistler L, Thakar HB, VanDerwarker AM, Domic A, Bergström A, George RJ, Harper TK, Allaby RG, Hirth K, Kennett DJ. Archaeological Central American maize genomes suggest ancient gene flow from South America. PNAS. 2020;117(52):33124–9. Larios C, Casas A, Vallejo M, Moreno-Calles AI, Blancas J. Plant management and biodiversity conservation in Náhuatl homegardens of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013;9:74. León-Jacinto A, Torres I. Polinización del Maguey Alto en el Municipio de Morelia. In: MartínezPalacios A, Morales-García J, Guillén-Rodríguez S, editors. Aspectos sobre el manejo y la conservación de agaves mezcaleros en Michoacán. Morelia: UMSNH, SAGARPA, CECTI; 2015. p. 143–54. Lima-Nascimento AM, da Silva JSB, Casas A, Marques-Lucena C, Lucena RFP. Traditional management of cactaceae: Cereus jamacaru DC as the native cactus most managed by rural communities in areas of Caatinga in Brazil. Ethnobot Res Appl. 2021;21:1–12. Lins Neto EMF, Peroni N, Casas A, Parra F, Aguirre X, Guillén S, Albuquerque UP. Brazilian and Mexican experiences in the study of incipient domestication. J Ethnobiol Ethnomed. 2014;10:33.

60

J. Caballero et al.

Lira R, Casas A, Rosas-López R, Paredes-Flores M, Rangel-Landa S, Solís L, Torres I, Dávila P. Traditional knowledge and useful plant richness in the Tehuacán Cuicatlán, México. Econ Bot. 2009;63:271–87. López-Austin A, López-Luján L. La periodización de la historia mesoamericana. Arqueol Mex. 2002;E11:14–23. Lotero-Velásquez E, Garcia-Frapolli E, Blancas J, Casas A, Martínez-Ballesté A. Eco-symbiotic complementarity and trading networks of natural resources in Nahua communities in Mountain Regions of Mexico. Hum Ecol. 2022;50:307–19. Lucena RFP, Marques-Lucena C, Meiado MV, Dias da Cruz D, Oliveira de Andrade M, Casas A. Uso, manejo y domesticación de cactáceas en Brasil. 12. In: Casas A, Torres-Guevara J, Parra F, editors. Domesticación en el continente americano Volumen 2. Perspectivas de investigación y manejo sustentable de recursos genéticos en el Nuevo Mundo. Morelia: Universidad Nacional Autónoma de México/Universidad Nacional Agraria La Molina/ CONACYT; 2017. p. 311–26. MacNeish RS. A summary of subsistence. In: Byers DS, editor. The prehistory of the Tehuacán Valley: environment and subsistence, vol. 1. Austin: University of Texas Press; 1967. p. 290–309. MacNeish RS. The origins of agriculture and settled life. Norman: Oklahoma University Press; 1992. Maffi L. Linguistic, cultural, and biological diversity. Annu Rev Anthropol. 2005;34:599–617. Mapes C, Guzmán G, Caballero J. Elements of mycological classification. J Ethnobiol. 1981;1(2): 231–7. Mapes C, Caballero J, Espitia E, Bye RA. Morphophysiological variation in some Mexican species of vegetable Amaranthus: evolutionary tendencies under domestication. Genet Resour Crop Evol. 1996;43(3):283–90. Mapes C, Basurto F, Bye R. Ethnobotany of “Quintonil”. Knowledge, use and management of edible greens Amaranthus spp. (Amaranthaceae) in the Sierra Norte de Puebla, Mexico. Econ Bot. 1997;51(3):293–306. Martínez M. Plantas útiles de la flora mexicana. Mexico City: Ediciones Botas; 1959. Martínez M. Catálogo de nombres vulgares y científicos de plantas mexicanas. Mexico City: Fondo de Cultura Económica; 1979. Martínez M. Las plantas medicinales de México. Mexico City: Ediciones Botas; 1989. Martínez-Alfaro MA, Evangelista V, Mendoza M, Morales G, Wong A. Catálogo de plantas útiles de la Sierra Norte de Puebla, México. Cuadernos 27. Mexico City: Instituto de Biología, Universidad Nacional Autónoma de México; 1995. Martínez-Ballesté A, Caballero J. Management compromises and the sustainability of palm populations in Mayan homegardens. Bot Sci. 2016;94(2):291–301. Martínez-Ballesté A, Martorell C. Effects of harvest on the sustainability and leaf productivity of populations of two palm species in Maya homegardens. PLoS One. 2015;10(3):e0120666. Martínez-Ballesté A, Martorell C, Martínez-Ramos M, Caballero J. Applying retrospective demographic models to assess sustainable use: the Maya management of xa’an palms. Ecol Soc. 2005;10(2):1–35. Martínez-Ballesté A, Martorell C, Caballero J. Cultural or ecological sustainability? The effect of cultural change on Sabal palm management among the Lowland Maya of Mexico. Ecol Soc. 2006;11(2):27. Martínez-Ballesté A, Martorell C, Caballero J. The effect of Maya traditional harvesting on the leaf production, and demographic parameters of Sabal palm in the Yucatán Peninsula, Mexico. For Ecol Manag. 2008;256(6):1320–4. Matos-Moctezuma E. Mesoamérica. In: Manzanilla L, López Luján L, editors. Historia Antigua de México: El México Antiguo, sus Áreas Culturales, los Orígenes y el Horizonte Preclásico, vol. I. México City: Instituto Nacional de Antropología e Historia; 1994. p. 49–74. Moreira PA, Aguirre-Dugua X, Maria C, Casas A, Clement CR, Vigouroux Y. Prehistoric routes of treegourd introduction and dispersal into Amazonia. Front Ecol Evol. 2017;5:150.

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

61

Nabhan GP. Native crop diversity in Aridoamerica: conservation of regional gene pools. Econ Bot. 1985;39(4):387–99. Pancorbo M, Parra F, Torres-Guevara J, Casas A. Los otros alimentos: plantas comestibles silvestres y arvenses en dos comunidades campesinas de los Andes Centrales del Perú. Etnobiología. 2020;18(1):8–36. Parra F, Casas A, Peñaloza-Ramírez JM, Cortés-Palomec AC, Rocha-Ramírez V, GonzálezRodríguez A. Evolution under domestication: ongoing artificial selection and divergence of wild and managed Stenocereus pruinosus (Cactaceae) populations in the Tehuacán Valley, Mexico. Ann Bot. 2010;106:483–96. Parra F, Blancas J, Casas A. Landscape management and domestication of Stenocereus pruinosus (Cactaceae) in the Tehuacán Valley: human guided selection and gene flow. J Ethnobiol Ethnomed. 2012;8(1):1–17. Parra F, Casas A, Rocha V, González-Rodríguez A, Arias-Montes S, Rodríguez-Correa H. Spatial distribution of genetic variation of Stenocereus pruinosus in Mexixo: Analyzing the origins of its domestication. Genetic Resources and Crop Evolution 2015;62(6):893–912. Parra-Rondinel F, Casas A, Begazo D, Paco A, Márquez E, Cruz A, Segovia J, Torres-García I, Zarazúa M, Lizárraga L, Torres-Guevara J. Natural and cultural processes influencing gene flow among wild (atoq papa), weedy (araq papa and k’ipa papa), and crop potatoes in the Andean region of southern Peru. Front Ecol Evol. 2021;9:617969. Pease JB, Haak DC, Hahn MW, Moyle LC. Phylogenomics reveals three sources of adaptive variation during a rapid radiation. PLoS Biol. 2016;14(2):e1002379. Pennington CW. The Tarahumara of Mexico: their environment and material culture. Salt Lake City: University of Utah Press; 1963. Pennington CW. The Tepehuan of Chihuahua: their material culture. Salt Lake City: University of Utah Press; 1969. Pérez-Negrón E, Casas A. Use, extraction rates and spatial availability of plant resources in the Tehuacán-Cuicatlán Valley, Mexico: the case of Santiago Quiotepec, Oaxaca. J Arid Environ. 2007;70(2):356–79. Piperno D, Pearsall D. Phytoliths in the reproductive structures of maize and teosinte: implications for the study of maize evolution. J Archaeol Sci. 1993;20:337–62. Piperno DR, Ranere AJ, Holst I, Iriarte J, Dickau R. Starch grain and phytolith evidence for early ninth millennium BP maize from the Central Balsas River Valley, Mexico. Proc Natl Acad Sci U S A. 2009;106:5019–24. Pulido MT, Caballero J. The impact of shifting agriculture on the availability of non-timber forest products: the example of Sabal yapa in the Maya lowlands of Mexico. For Ecol Manag. 2006;222:399–409. Rangel-Landa S, Casas A, Dávila P. Facilitation of Agave potatorum: an ecological approach for assisted population recovery. For. Ecol. Manage. 2015;347:57–74. Rangel-Landa S, Casas A, García-Frapolli E, Lira R. Socio-cultural and ecological factors influencing management of edible plants: The case of Ixcatlán, Mexico. J Ethnobiol Ethnomed 2017;13(1):59. Rangel-Landa S, Rivera-Lozoya E, Casas A. Uso y manejo de las palmas Brahea spp. (Arecaceae) por el pueblo ixcateco de Santa María Ixcatlán Oaxaca, México. Gaia Sci. 2014;8:62–78. Rangel-Landa S, Casas A, Rivera-Lozoya E, Torres-García I, Vallejo M. Ixcatec ethnoecology: biocultural principles of plant management in Oaxaca, Mexico. J Ethnobiol Ethnomed. 2016;12:30. Rodríguez-Arévalo I, Casas A, Campos J, Lira R. Uso, manejo y proceso de domesticación de Pachycereus hollianus (F.A.C. Weber) Buxb. (Cactaceae) en el Valle de Tehuacán-Cuicatlán, Mexico. Interciencia. 2006;31(9):677–85. Santana Michel F, CONABIO (Comisión Nacional para el Conocimiento y Uso de la Biodiversidad). Actualización de la base de datos sobre la flora de la Reserva de la Biósfera Sierra de Manantlán, Jalisco-Colima, México. https://www.gbif.org/es/dataset/bddca4a4-60b54db5-95fa-be0cce979d3a (2021). Actualized in July 2021; consulted 09/05/2022.

62

J. Caballero et al.

Smith BD. Origins of agriculture in eastern North America. Science. 1989;246(4937):1566–71. Smith BD. The initial domestication of Cucurbita pepo in the Americas 10,000 years ago. Science. 1997;276:932–4. Smith BD. Eastern North America as an independent center of plant domestication. Proc Natl Acad Sci U S A. 2006;103:12223–8. Toledo VM, Barrera-Bassols N. La memoria biocultural. Barcelona: Icaria; 2008. Toledo VM, Ordóñez, MJ. The biodiversity scenario of Mexico: A review of terresrial habitats. In Biodiversity of Mexico. Origins and distribution, pp. 757–778, Ramammoorthty TP, Bye R, Lot A, Fa J (eds.). Oxford University Press, Oxford, UK. 1993. Toledo VM, Batiz AI, Becerra R, Martínez E, Ramos CH. La selva útil: etnobotánica cuantitativa de los grupos indígenas del trópico húmedo de México. Interciencia. 1995;20:177–87. Torres I. Aspectos etnobotánicos y ecológicos de los recursos vegetales en la comunidad de San Luis Atolotitlán, municipio de Caltepec, Puebla, México. Bachelor thesis, Universidad Michoacana de San Nicolás de Hidalgo, Morelia; 2004. Torres-García I, Casas A, Delgado-Lemus A, Rangel-Landa S. Aprovechamiento, demografía y establecimiento de Agave potatorum en el Valle de Tehuacán, México: Aportes etnobiológicos y ecológicos para su manejo sustentable. Zonas Áridas. 2013;15(1):92–109. Torres-García I, Casas A, Vega E, Martínez-Ramos M, Delgado-Lemus A. Population dynamics and sustainable management of mescal agaves in central Mexico: Agave potatorum in the Tehuacán-Cuicatlán Valley. Econ Bot. 2015a;69(1):26–41. Torres-García I, Blancas J, León A, Casas A. TEK, local perceptions of risk, and diversity of management practices of Agave inaequidens in Michoacán, Mexico. J Ethnobiol Ethnomed. 2015b;11:61. Torres-García I, Rendón-Sandoval FJ, Blancas J, Casas A, Moreno-Calles AI. El género Agave en los sistemas agroforestales de México. Bot Sci. 2019;97(3):261–88. Torres-García I, León A, Vega E, Moreno-Calles AI, Casas A. Integral projection models and sustainable forest management of Agave inaequidens in western Mexico. Front Plant Sci. 2020;11:1224. Torres-Guevara J, Parra-Rondinel F, Casas A, Cruz-Soriano A. De los cultivos nativos y el cambio del clima. Hallazgos (Huánuco y Apurímac). Lima: Universidad Nacional San Antonio Abad de Cusco, Universidad Nacional Agraria La Molina, Universidad Nacional Autónoma de México, Programa Nacional de Innovación Agraria, IDMA, CADEP; 2019. Trindade MRO, Gomes-Jardim J, Casas A, Medeiros-Guerra N, Lucena RFP. Availability and use of woody plant resources in two areas of Caatinga, Northeastern Brazil. Ethnobot Res Appl. 2015;14:313–30. Ubiergo-Corvalán PA, Rodríguez-Galván G, Castro-Laportee M, Zaragoza-Martínez L, Casas A, Guevara-Hernández F. El solar maya ch’ol y sus saberes etnobotánicos en comunidades al norte de Chiapas, México. Ethnoscientia. 2019;4(1):1–19. Ubiergo-Corvalán PA, Rodríguez-Galván MG, Zaragoza-Martínez ML, Ponce Díaz P, Casas A, Mariaca-Méndez R. Agrobiodiversidad vegetal comestible en el territorio indígena Maya-Ch’ol de Chiapas, México. Trop Subtrop Agroecosyst. 2020;23(2020):46. Ubiergo-Corvalán P, Rodríguez-Galván G, Casas A, Zaragoza-Martínez L, Ponce-Díaz P. Plantas comestibles del solar Maya Ch’ol, Chiapas. Tuxtla Gutiérrez: Universidad Autónoma de Chiapas; 2021. Ulloa CU, Acevedo-Rodríguez P, Beck S, Belgrano MJ, Bernal R, Berry PE, Brako L, Celis M, Davidse G, Forzza RC. An integrated assessment of the vascular plant species of the Americas. Science. 2017;358:1614–7. Vavilov NI. Origin and geography of cultivated plants. Cambridge, UK: Cambridge University Press; 1992. Velásquez-Milla D, Casas A, Torres-Guevara J, Cruz-Soriano A. Ecological and socio-cultural factors influencing in situ conservation of crop diversity by traditional Andean households in Peru. J Ethnobiol Ethnomed. 2011;7:40.

Ethnobotanical Knowledge in Mexico: Use, Management, and. . .

63

Villaseñor JL. Checklist of the native vascular plants of Mexico. Rev Mex Biodivers. 2016;87: 559–902. Zárate S, Pérez-Nasser N, Casas A. Genetics of wild and managed populations of Leucaena esculenta subsp. esculenta (Fabaceae: Mimosoideae) in La Montaña of Guerrero, Mexico. Genet Resour Crop Evol. 2005;52:941–57. Zarrillo S, Gaikwad N, Lanaud C, Powis T, Viot C, Lesur I, Fouet O, Argout X, Guichoux E, Salin F, Solorzano RL, Bouchez O, Vignes H, Severts P, Hurtado J, Yepez A, Grivetti L, Blake M, Valdez F. The use and domestication of Theobroma cacao during the mid-Holocene in the upper Amazon. Nat Ecol Evol. 2018;2:1879–88. Zeder M. Domestication and early agriculture in the Mediterranean Basin: origins, diffusion, and impact. Proc Natl Acad Sci U S A. 2008;105(33):11597–604. Zeder M. The origins of agriculture in the Near East. Curr Anthropol. 2011;52(S4):S221–35. Zeder M. The domestication of animals. J Anthropol Res. 2012;68(2):161–90. Zeder M. Core questions in domestication research. Proc Natl Acad Sci U S A. 2015;112(11): 3191–8. Zeder MA. Domestication as a model system for the extended evolutionary synthesis. Interface Focus. 2017;7:20160133. Zizumbo D, Colunga P. Los huaves: la apropiación de los recursos naturales. Universidad Autónoma de Chapingo. Chapingo, Estado de México. Mexico. 1982.

Agroforestry Complexes in the Mountain Regions of Mexico Ana Isabel Moreno-Calles, Gerardo Herna´ndez-Cendejas, Wilfrido Lo´pez-Martínez, Alexis Daniela Rivero-Romero, Yessica Ange´lica Romero-Bautista, Karla Guzma´n-Ferna´ndez, Ana Mitzi García-Leal, Ernesto Gutie´rrez-Coatecatl, Cloe X. Pe´rez-Valladares, Ana Rojas-Rosas, Ignacio Torres-García, and Selene Rangel-Landa Abstract

Mountain regions in Mexico represent outstanding environments where agroforestry processes are particularly important due to their complexity, difficult access, steep slopes, frosts, among other social and environmental aspects. From a historical perspective, mountain inhabitants are in a constant endeavor to develop management strategies to satisfy their needs. Their knowledge on surrounding ecosystems is constructed from the interaction’s experiences though generations, evaluated through trial and error and constantly improved. This knowledge harbors a profound relevance as repository of biocultural strategies. Agroforestry systems are ongoing processes of deliberate human decisions, relating the articulation of wild, managed, and domesticated species at multiscale and multitemporal dimensions. These processes aim to derive ecological, economic, and social benefits integrating all these aspects in production, which in turn has A. I. Moreno-Calles (*) · G. Hernández-Cendejas · W. López-Martínez · A. D. Rivero-Romero · Y. A. Romero-Bautista · K. Guzmán-Fernández · A. M. García-Leal · E. Gutiérrez-Coatecatl · C. X. Pérez-Valladares · A. Rojas-Rosas Escuela Nacional de Estudios Superiores Unidad Morelia, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected]; [email protected]; [email protected]; [email protected] I. Torres-García Escuela Nacional de Estudios Superiores (ENES), Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] S. Rangel-Landa Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Consejo Nacional de Ciencia y Tecnología (CONACYT) – Escuela Nacional de Antropología e Historia (ENAH), Mexico City, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_10

65

66

A. I. Moreno-Calles et al.

important implications in terms of governance of the social units. In this chapter we show eight case studies from Mexico City, the State of Mexico, and the states of Puebla, Michoacán, Oaxaca, and Tlaxcala in temperate, subhumid, and semiarid conditions, at elevations ranging from 1300 to 2700 m, characteristic of the mountains of Central Mexico. From homegardens and collective gardens, slash and burn agroforestry, the several forms of multicrop milpas, profoundly and diversely interconnect with the wild surroundings; terraces and semi terraces in temperate, subhumid, and semiarid conditions are agroforestry systems that conform food landscapes. This chapter summarizes our views on these systems in the region.

Introduction Steep hill lands, dry areas with long seasons without precipitation and low temperatures throughout the year, high biological diversity, ancient and complex human molded environments, are some of the predominant features of the mountain regions of Mexico (Boege 2008; Rojas 1991; Toledo and Barrera-Bassols 2008). From elevations about 1000 m it is possible to find these environments in Mexico. “Monte,” as it is called in Spanish, in the conceptions of local ethnogeomorphology, has been and continues to be recognized as one of the main living spaces of original and other peasant’s groups of rural México. It is in these spaces where peri-urban and urban vulnerable poor groups of numerous Mexican settlements are established. In environments at these elevations live a high number of local and migrant populations coming from the countryside to peri-urban areas of medium and large size cities of this country. In this context, the inaccessibility and the continual struggle to control soil erosion and landslides, maintenance of soil fertility, humidity, and diminishing the effect of frosts on agricultural systems are some of the most recurrent concerns of peasants groups since pre-Hispanic times until the present (Moreno-Calles et al. 2013, 2016, 2020; Wilken 1990). Borning and growing in the mountains, people learn from previous generations and continually innovate strategies to maintain the relationship with these environments. One of these strategies is the close interaction with the orographic, edaphic, climatic, and biological diversity of these spaces (Stepp et al. 2005), and it is in such context that agroforestry complexes make sense. Agroforestry is conceptualized as the deliberate human articulation of wild and managed plants, animals, and fungi coexisting in spatial and time with crops and/or domestic animals in systems of agricultural management. Agroforestry systems integrate cultivated and/or domesticated elements in shrublands, temperate, and tropical forests with the purpose of obtaining ecological, economic, and social benefits (Moreno-Calles et al. 2014). Agroforestry worldwide is mainly practiced by farming families and is associated with landscape management practiced by original and peasant villages as well as those in urban and peri-urban contexts. Agroforestry complexes (including landscapes, systems, and agroforestry practices) integrate multiscale and multitemporal management strategies of biological and biocultural diversity that provide environmental benefits to human beings at local, regional, and global levels. With this management form, it is possible to generate

Agroforestry Complexes in the Mountain Regions of Mexico

67

economic outcomes, strengthen food security, satisfy local and global needs, and environmental benefits mitigating effects of climatic phenomena like frosts, droughts, atypical rains, and wind gusts. Also, its components provide shade and protection, shaping the habitat of other useful species, favoring, maintaining, or increasing the formation of soil and maintaining or increasing its fertility, reducing erosion, and the effect of non-beneficial arthropods. Also these systems increase the control capacity of fire, maintaining hydrological benefits and in sum, constituting an important alternative for mitigation and adaptation to climate change (Jose 2009; Montagnini et al. 2015). In addition, these forms of management are a great scenario that articulates social learning, collective creation of knowledge, worldviews and practices, social and environmental values, and governance models of the social units who practice them and those of the social actors who are interested in them. These social actors include farmers, groups of native people, government entities, civil society, and international organizations (Toledo and Barrera-Bassols 2008). The questions that guided this chapter are: (i) Are there consistent patterns between the mountain regions of Mexico and agroforestry complexes? and (ii) What processes allow to understand these patterns? For analyzing these questions we analyze eight case studies that our group conducted in communities of Mexico City, the State of Mexico, and the states of Puebla, Michoacán, Oaxaca, and Tlaxcala in temperate, subhumid, and semiarid conditions.

Mexican Mountains, Biocultural Diversity, and Agroforestry Complexes In Mexico, mountain areas are extensive and bioculturally relevant. The work of Boege (2008) analyzing the biocultural heritage of indigenous peoples of Mexico does not make a special emphasis on mountain areas, but it is possible to realize the special relationship between mountain regions, biodiversity, and cultural diversity in the biocultural heritage of Mexico. Previously, Wolf and Cirlot (1971) invited to observe the relationships between the relief, especially the slopes of the Gulf of Mexico and the Pacific to understand the distribution of the different cultures in Mesoamerica. The historian Bernardo García (2008) takes up this idea to analyze the conformation of the country’s regions. The relationship between mountain areas, biodiversity, and cultural diversity is an important topic in Mexico. For this chapter we based on the map of physical-geographic regions proposed by Cervantes Zamora et al. (1990), identifying that 47 of the 88 provinces that they propose can be considered as mountain landscapes. Mountain areas are of great importance to Mexico both for their distribution and for their relationship with the conformation of different environmental factors such as climate, vegetation, and the settlement and distribution patterns of the population of native peoples. Mountain areas are present in 28 of the 32 states of Mexico, the exceptions are the states of Tabasco, Campeche, Yucatan, and Quintana Roo. The provinces that can be considered as mountain areas are: Altos de Chiapas, Chiconquaco, Cordillera Costera del Sur, Depresión del Balsas, El Cabo, Escarpe Limítrofe del sur, Gran Meseta y Cañones Chihuahuenses, Gran meseta y Cañones Duranguenses, Gran Sierra

68

A. I. Moreno-Calles et al.

Plegada, Karst Huasteco, Lagos y Volcanes de Anáhuac, Lomeríos de la Costa del Golfo Norte, Mesetas y Cañadas del Sur, Mil Cumbres, Mixteca Alta, Neo volcánica Tarasca, Pie de la Sierra, Pliegues Saltillo Parras, El Burro, Cuatralba, Sierra de Guanajuato, Sierra de Jalisco, Sierra de la Paila, Sierra de los Tuxtlas, San Carlos, Sierra de Tamaulipas, la Giganta, Sierra La Cardona, Sierra de Baja California Norte, Sierra de la Costa de Jalisco y Colima, Sierra del Norte de Chiapas, Sierras Neo volcánicas Nayaritas, Sierras Orientales, Sierras y Llanuras Coahuilenses, Sierras y Llanuras Sonorenses, Sierras y Llanuras Tarahumaras, Sierras y Valles del Norte, Sierras y Valles de Oaxaca, Sierras y Valles Guerrerenses, Sierras y Valles Zacatecanos, Volcanes de Colima y Volcanes de la Costa Golfo Norte. The surface of these provinces is 980,852 km2, that is, 49.93% of the surface of Mexico. Mountain areas contain 58 of the 61 climate types of Mexico since a factor of climatic variability is given by the altitude gradient. Thus, climate types are distributed in part following the elevation levels. The diversity of climates is also expressed in the diversity of vegetation types existing in these mountain areas. Based on the map of INEGI (2020b) series 6 on land uses and vegetation we found 135 different vegetation types for mountain areas. There are different types of tropical forests, like deciduous, subperennial, and perennial, different types of scrublands, as well as temperate forests like those dominated by oaks, pines, fir, sabinos, and mesophylous mountain forests. In many cases, the differences in vegetation may be present in a region in relation to altitude. Mountain areas are diverse in plants and animals. Different types of agriculture are also practiced in those areas, forming agroforestry complexes that include agroforestry terraces and semi-terraces, agroforestry fields where most of the richness of native maize, beans, squash, and edible native herbs (quelites) and local fruit species are preserved. Family and collective homegardens, agroforests where coffee, cinnamon, vanilla, pineapple, and cocoa are produced for income generation through local, regional, and global markets and agrosilvopastoral systems of colonial and recent origin. The seasonal agriculture practiced in the mountain areas has an extension of 111,561 km2 (Series VI, INEGI 2020b), which represents 52.15% of the seasonal agriculture in the country. In the map of Fig. 1, we show the proposal by Stepp (2005) to explore the relationship between mountain areas and biocultural diversity according to the records of the data base “Sistemas Agroforestales de Mexico” (www.red-sam.org).

The Mountains, Their People, and History There is a clear relationship between mountain areas and cultural diversity in Mexico for most of the original groups, except for the Yucatan peninsula and Tabasco. This settlement pattern, where the presence of indigenous groups coincide with the mountain areas, has its origin in geographic, ecological and historical processes. Three major processes in the history of Mexico influenced these areas, which are the Spanish conquest and colonization in the 16th to 18th centuries, the policies on indigenous lands and communities carried out by the laws of the Reform period in the nineteenth century, and the processes of land restitution and endowment of ejidos made by the Agrarian Reform in the 20th century. Taking these processes into account, we can

Agroforestry Complexes in the Mountain Regions of Mexico

69

Fig. 1 Map of Mexico with the identified mountains and the points of the ethnoagroforestry systems registered in the database of the Network of Agroforestry Systems of Mexico (www. redsam.org)

see a coincidence between the regions of native peoples and the mountain areas, which have perhaps their explanation in what Aguirre Beltrán (1991) called refuge regions. In other words, throughout these processes, this population moved from their former territories to settle mountain territories where they could escape from the pressure of the colonial system, the dispossession of the haciendas, and other processes of deterritorialization, in order to maintain their ways of life. That is why in some cases it is not by chance that indigenous groups currently occupy the territories that were once of difficult access or distant. In these places, agroforestry systems and their benefits have a relevant meaning (Fig. 2).

Case Studies Living in the Cold Mountains of Mexico Through Agrosilviculture Agroforestry and Rituality in Ayuujk Montains The municipality of Tlahuiltoltepec is located in the “Sierra Norte” of the state of Oaxaca, a region called in the local native Ayuujk language Xaamkëjxp, which means “cold place” or “place among clouds.” Its territory comprises an altitude range between 1000 and 3400 m. The community located at the highest elevation is Santa María Yacochi (2332 m) where the fir (Abies) forest is the dominant type of

70

A. I. Moreno-Calles et al.

Fig. 2 Agroforestry complexes and indigenous peoples in Mexico. Only the Yucatan Peninsula and Tabasco are excluded from being considered mountainous areas

vegetation, in the middle elevations are the communities of Tlahuiloltepec, Mosca, and Flores (2090–2260 m) with temperate pine-oak (Pinus – Quercus) forest as dominant vegetation, and in the lowest elevation is the community of Nejapa (1800 m), where tropical deciduous forests is dominant. The inhabitants of these communities select the places in mountains that will be used for cultivation, the trees, shrubs and herbs and other elements that will be removed from the land and the area that will be used for agriculture (Fig. 3a, b). Among the usually tolerated species in Tlahuiltoltepec we found ash trees (Fraxinus udehi), oaks (Quercus spp.), and “palo de águila” (Alnus acuminata), these species are valued for their timber and for construction and firewood together with pine trees (Pinus spp.). The leaves of “palo de águila” are considered a good fertilizer, a natural dye is obtained, and is considered by people as a tree that attracts water “that tree has deep roots that make the water rise, that’s why you find a lot around here” (Díaz 2017). The latter alludes to the hydraulic lifting that the tree performs since it has deep roots can reach the deep-water levels. When the inhabitants of these localities have a spare landspace inside or in the premises of their houses, they also use it for plant culture. In both places, plot and house, the “milpa” system, consisting of maize (Zea mays), beans (Phaseolus vulgaris), and “chilacayote” (Cucurbita ficifolia) is stablished, in addition of growing fruit trees such as apple (Malus domestica), “tejocote” (Crataegus mexicana), “ocote” (Pinus lawsonii), and the shrub “chamizo” (Baccharis conferta).

Agroforestry Complexes in the Mountain Regions of Mexico

71

Fig. 3 Eight case studies are presented in Mexico City, State of Mexico (Transfiguración and Noxtepec), Puebla (Zapotitlán Salinas), Michoacán (Cuanajo), Oaxaca (Tlahuiltoltepec and Ixcatlán) and Tlaxcala (Carmen Tequexquitla) in temperate, sub-humid and semi-arid conditions in the altitudinal ranges since 1300 to 2700 masl

The relevance of species like “palo de águila” is expressed where it is spatially located and in the wealth of agroforestry practices in which it is found. For instance, Alnus acuminata is tolerated for planting maize and without no specific spatial arrangement, they can be located at the top, bottom, middle or in the bank of plots, or at the end of a maize row. Although these trees need time to be harvested, their use is limited to the time it takes for the tree to grow and the number of trees they have in their land or house is directly related to the areas they manage and own. These areas are relatively small and can stand only three to five trees. The decision to maintain few or several trees is taken based on their needs, such as using a large part for plant culture, a pen for domestic animals or the construction of a house. A close, ancient and ritual connection with the mountains, the forests, and the conception of the wild, is clear, and without its complementarity and antagonism, agriculture would not be conceived. Rituals related to agroforestry management and the connection that farmers have with nature are carried out in the mountain or the hill. When lands are cleared, people perform rituals directed to ask nature its permission to clear and to plant; mescal is offered to nature, otherwise peasants could suffer a bad augury. Living in the mountains also means that a collective organization of the agricultural activities is needed in order to reduce the effect of inaccessibility. This organization requires the necessary strength to achieve the essential transformations

72

A. I. Moreno-Calles et al.

and maintenance for its subsistence, an example of this is the planning to start sowing. The family that will sow invites other people, family or friends, they have to provide food all along the working day, “tepache” and mescal; in addition, they commit to help sowing the land of those people that helped them. Sowing starts early with breakfast, then a ritual is performed on the land to sow, the performance of this ritual will depend on each family. The way of sowing by the Ayuujk people is practiced as follows: people are placed vertically in rows, the first one starts sowing and leaves 60–80 cm between each hole where he sowed. The person below starts sowing a moment later to the first and will sow the parallel row below using the spaces left, being careful not to pass the first one, as well as the other people who are sowing. When reached the row end, the last person will return taking first place to sow, the others will follow his footsteps. This practice has been transmitted throughout generations. “Mano-vuelta” (hand and return) is practiced, it consists in the commitment of returning the favor of participation in sowing, without any economic retribution. The performance of the ritual involves the offering of a fowl, the seeds to be sown, mescal, “tepache” and the “coas” (ancient Mesoamerican instrument used for sowing). The ritual begins with a prayer, in Ayuujk language, invoking a deity that represents nature and wellness, the so-called Rey Konk, to ask for his permission to sow, for the seeds to germinate, for good augury in all the labors and good harvest. A rooster is sacrificed with a machete, its blood is scattered on the soil and maize seeds as a sign of fertility. Its head is then buried by the landowner, near a specific tree or in the center of the field, then mescal and tepache are poured to the ground three times, to feed the earth. Then mescal is distributed among those who will participate in sowing, three drops most be poured on the ground before they drink. Finally, the rooster’s feathers are settled around the place where its head was buried, then the sowing begins (Fig. 4).

Fig. 4 Agricultural ritual performed by the family of Doña Catalina on her plot, a rooster was sacrificed and its blood was used to water the seeds that will be used for planting, mezcal and tepache are offered to the land and then among the participants who will sow. (Photo by the authors)

Agroforestry Complexes in the Mountain Regions of Mexico

73

Every year in December, the Ayuujk people make a visit to a sacred mountain called “Cempoaltépetl” to show their appreciation for the wellness of the year, for the harvest, for the good health and for the work done. This visit is also made during the first days of the new year to ask for health, work or a good harvest. This journey takes 3 h, the people walks up to the highest point of elevation of the mountain (3420 m). The Ayuujk also draw upon the “xëmaapyë,” wise people that serve as fortune tellers, who help people to prepare the offerings they must present to the mountain to deal with specific illness and the commitment with the mountain. Some people don’t search fortune tellers help because they know what they have to offer, but then when people encounter with a snake or dreams with it, they know it is a sign that a ritual or offering is needed to be in harmony with the mountain and then they recur to the fortune teller to ask for his guidance.

Homegardens and Ornamentality in Cuanajo, Michoaca´n The town of Cuanajo is in northern Michoacán. In Purépecha (the original language), it means “place of frogs.” The valley of Cuanajo is located at 2320 m, surrounded by mountains that surpass 3300 m, the main vegetation types are temperate forests dominated by the associations of Pinus-Quercus, Pinus, Quercus-Pinus and Abies religiosa (Caballero 1982; Farfán-Heredia 2019). The area has undergone major transformations because of excessive logging (a process that began since the conquest) and agriculture. In recent years, the increase of intensive cropping of some crops like avocado (Persea americana), blackberry (Rubus ulmifolius), and strawberry (Fragaria vesca) in the surroundings have also contributed to deforestation. The resulting scenario is a heterogeneous matrix composed by crop areas and forest remnants. Cuanajo is under a process of urbanization and transformation. In the community, the inheritance and distribution of land to male children is a tradition, so when a son begins his own family, he is granted a space of land in which to live. This aspect has had a direct impact on the accommodation and use of space. Formerly, the existence of an agroforestry system called “ekuaros” or family gardens as a productive space for species valued as food, medicinal, forage, ornamental, incomes security, textile, tannins, timber, and aromatic uses (Alarcón-Chaires 2009). Today, within the locality, family gardens or “patios” are mostly recognized as areas with similar purposes to that of the “ekuaro.” Ekuaros are described as adjacent areas to the house room composed by different useful plant species, which are under incipient management. Within the patios of this town, it is common to find introduced fruit trees that tolerate cold and drought such as peaches (Prunus persica), lemons (Citrus x limon), oranges (Citrus x sinesis), nísperos (Eriobotrya japonica), manzana (Malus domestica), chabacano (Prunus armeniaca), pear (Pyrus communis), and membrillo (Cydonia oblonga), and also native species such as avocado (Persea americana), capulín (Prunus serotina), tejocote (Crataegus mexicana), guava (Psidium guajava), and chirimoya (Annona cherimola), other common crops are chile perón (Capsicum pubescens) and chayote (Sechium edule). Medicinal plants are also cultivated. Predominance of ornamental plants, in terms of abundance and richness, is evident. Improvised pots, like food cans and paint buckets, are abundant in the patios, these are used to cultivate

74

A. I. Moreno-Calles et al.

ornamental plants, highlighting the diversity of succulent plants. The cultivation and propagation of these plants is encouraged because they are used to trade or barter in neighboring markets. Pátzcuaro barter market stands out, a market with Purépecha tradition frequented by the inhabitants of towns nearby. The importance of ornamental plants is related to their economic and commercial value in response to demand and in coherence with an economy based on the use value and the diversification of goods produced. The presence of ornamental plants in house rooms is closely linked with the perception of femininity in Cuanajo. A well-kept house should be adorned and embellished with many flowering plants. Likewise, admiration for these spaces increases when the diversity and uniqueness of the species is greater. An ancient relationship with this ornamental diversity has been documented, we can refer to this connection as ornamentality, seen as a long-term and continuous relationship or process in a territory. This interaction favors landscapes, spaces, species and cultivars for the purpose of decoration, contemplation, beautify, influence mood, and make day to day bearable. To remember the colors, smells and shapes is a remembrance of mothers and grandmothers, because in this community, flowers are a matter for women, and it is considered a luxury and a way to provide economic incomes, food or material resources through exchange. Therefore, women are always looking to increase their collection of ornamental plants, adding those that they consider scarce or unique or, conversely, other plants that have become very popular due to their novelty. One of the most relevant strategies for this purpose is the existence of a living exchange network of ornamental plants, in which women look for new additions or to recover lost cultivars. This network is supported by women and their patios and “ekuaros” and the regional markets. The heterogeneity mountain can provide color palette or the adjacent areas, the other patios or ekuaros or the markets exchange. A particular group of plants occuring in patios that stands out for its beauty and diversity is a group of cacti called “teresitas” and “pitayitas,” species of the genera Disocactus (D. phyllanthoides and D. speciosus), Epiphyllum, Selenicereus, and hybrids between genera and species (Disocactus x hybridus; Disocactus x Epiphyllum, and Disocactus x Selenicereus) (Fig. 5). Ornamental species management differs to the management of other species present in this system. They are planted and cultivated in pots, an aspect that allows their mobility, transport and the colonization of new spaces like roofs and walls; the propagation method is mostly asexual by means of cutting and shoots; and its support is given from two main inputs, soil and water. The soil is gathered from the forest mulch of the mountains that surround the community. Ornamentality has a different meaning than ornamentalization, a process registered in other homegardens in Mexico, this concept refers to a process of diminishing the richness of uses and the relationships with the ornamental richness of plants and their unique dominance toward a gardening process (Blanckaert et al. 2004). An additional way is related to the mountain by women in cold and dry contexts and in the decrease of spaces and resources.

Agroforestry Complexes in the Mountain Regions of Mexico

75

Fig. 5 Teresitas and pitayitas in Cuanajo, Michoacan, Mexico. (a) Disocactus speciosus (b) Disocactus phyllanthoides (c) Disocactus crenatus x (d) Disocactus x Selenicereus (e) from left to right: Disocacus crenatus x, Disocactus crenatus x, Disocactus x Selenicereus, Disocactus x Epiphyllum, Disocactus x Selenicereus. (Photos by the authors)

Milpas of the Mountain as Agroforestry Complex The locality of Transfiguration is in an altitude rage of 2750 to 3200 m in the interior of the Sierra de Monte Alto, a peri-urban zone of Mexico City. Its climate is temperate sub-humid with rains in summer and an average temperature of 16  C. Its vegetation is pine-oak forest (Pinus, Abies and Quercus) and riparian forest (García 1998). Around 6000 people inhabit the community and no native language is spoken. The region was formerly occupied by Hñahñu people, who have carried out agricultural activities since ancient times. Later, these groups moved to the northeast of the state of México (Esparza 1999). Currently, in addition to agriculture, fishing, harvesting, charcoal production and the provision of services are carried out. The “milpas” are the predominant systems in the region and their arrangement is based on the integration of wild and cultivated plant and fauna diversity, these characters make them systems under agroforestry management. Its basic composition includes “criollo” maize (Zea mays) including black, white and red varieties, fava bean (Vicia faba); bean (Phaseolus vulgaris) including black, yellow, red, brown and purple pinto varieties; squash (Cucurbita pepo) and “chilacayote” (Cucurbita ficifolia). Other species of nutritional and cultural value within these systems are the native or wild potato (Solanum sp.) and the foreign potato (Oxalis tuberosa), which are normally cultivated in the highest and temperate zones of the community. In the agroforestry milpas (AFM) trough semi terrace additionally to the annual and perennial crops represented by fruit trees such as “capulín” (Prunus serotina), “pear” (Pyrus communis) and “tejocote” (Crataegus mexicana) are maintained; as well as other plants such as prickly pear (Opuntia spp.) and maguey

76

A. I. Moreno-Calles et al.

Fig. 6 In Transfiguración community, the presence of semi-terraces and agroforestry practices for soil retention was recorded, these house fruit trees such as capulín (Prunus serotina) and tejocote (Crataegus mexicana), in addition to magueyes (Agave salmiana), a species used for the extraction of sap for “pulque” production. (Photo by the authors)

(Agave salmiana). In addition, forest or wild species that are under some degree of management such as oak (Quercus sp.), pine (Pinus spp.), oyamel (Abies religiosa), madroño (Arbutus xalapensis), and aile (Alnus acuminata) (Fig. 6). This agrobiodiversity is included as part of the different agroforestry practices such as isolated trees, boundaries, vegetation patches, and islands of vegetation. Finally, it is important to mention the fauna composition of the AFM, which is represented by a total of seven species of mammals, one reptile, and 25 species of birds. Some of these species are cacomixtle (Bassariscus astutus), wild rabbit (Sylavilagus floridanus), cascabel (Crotalus sp.), jilguero (Myadestes occidentalis), zopilote (Coragyps atratus), and pavito (Myioborus pictus).

Agrosilviculturization in Urban Contexts: Community and Collective Gardens in Peri-Urban Mexico City The Valley of Mexico where Mexico City is located is at an altitude of 2428 m, with an extension of 9600 km2, surrounded by a volcanic mountain range. The climate is temperate with summer rains, the average temperature is between 18 and 24  C, the average annual precipitation ranges from 1000 and 1400 mm (Torres et al. 2000). In Mexico City and its metropolitan area, around 21 million people lived in 2017 (ONU HABITAT 2020; Torres et al. 2000). According to the 2015 intercensal survey, of the total of employed people only 0.39% worked in the agricultural sector (INEGI 2020a).

Agroforestry Complexes in the Mountain Regions of Mexico

77

Fig. 7 The oak-pine vegetation that surrounds the collective garden “Los Cedros” is found in the peri-urban context of Mexico City; this is an example of implementation of collective gardens that are affected by the urbanization of rural areas. The trees in the place allow to obtain food, materials, medicines, they serve as windbreaks. Agaves allow to stabilize the terraces that must be used in the context of the slopes of the elevations. (Photo by the authors)

The space and time for agriculture is conceived as limited by the elevation and the population density and the influence of the city in daily life. The community gardens of Mexico City are characterized by possessing a high diversity of managed species arranged in polycultures and agroforestry models which include native plants, vegetables, medicinal plants, edible and ornamental flowers, fruit trees, among others. Its origin may be related to the inclusion of agriculture and agroforestry in urban and peri-urban contexts or due to the expansion of the city into areas that were previously agricultural (Fig. 7) (Borelli et al. 2017). There are several reasons for the selection of species, among which are temporality, personal taste, commercial demand or for having a specific role in the garden. However, there are also species that were already in the plots before orchards were implemented, this especially occurs with fruit, timber or fuel trees. Yields of this type of garden can be for self-consumption, sale, or used in pedagogical activities. Another important characteristic of community gardens is that they are managed through agroecological practices. Regarding the management of weeds, this can be done through biological control, physical protections or manual weeding. In some cases, material is previously used as a “bed” for domestic animals like fowl, where it mixes with its guano and is later used in the cultivation beds, increasing soil fertility. To control pests and diseases biological management is used, like the cultivation of

78

A. I. Moreno-Calles et al.

ladybugs, physical traps, attracting trap plants, association of crops (Beilin and Hunter 2011). Fertilization processes are mainly managed by compost and vermicompost methods, where wastes from the garden are processed, as well as tree pruning. Also, specific species are planted that contribute to this end. Finally, there are strategies for efficient use of water, among which are the use of mulch, drip irrigation and directed irrigation (for example, clay pots buried and filled with water). In some cases, there is a presence of fauna that interacts with different processes in the garden, such as chickens and ducks that control pests and weeds, wild birds that spreads seeds, cats and dogs’ control urban pests and serve as guardians, as well as insects that promote pollination. As for the social part of collective gardens, these are managed by groups of people, which can be organized in different ways, for example, being under the coordination of a civil society, extensive family, community, or a government entity. On the other hand, the forms of participation are usually by volunteering or employment. Most of the land does not belong to the people who maintain the gardens, so there is certain insecurity about it. Finally, these spaces usually have objectives beyond production, such as being an educational medium, recovering public spaces or promoting environmental issues.

Subhumid and Semiarid Mountains of Mexico and Agrosilviculture Tempered-Subhumid Transition and Agroforestry Management Noxtepec is a town located in the State of Mexico at an altitude range of 1600–2500 m, a small community of about 250 mestizo settlers who manage communally their territory. It is a rural population with social backwardness (INEGI 2020). It has a temperate sub-humid climate with summer rains, average temperature of 19  C and an average rainfall of 1600 mm. It is a reservoir of climates, traditions and species (Tuirán et al. 2000). Its reddish soils are clayey, mostly of forest vocation (50%), followed by agricultural use (40%). It is found in the floristic province of the Balsas River basin, the main types of vegetation are temperate forests dominated by Pinus forest, fir forest, Quercus forest, mesophyllous forest, and tropical deciduous forest (INEGI 2020). Maize (Zea mays), squash (Cucurbita spp.), and beans (Phaseolus vulgaris) are mainly sown in these communal lands. People who still dedicate to the labor do it for self-consumption where they have no opportunity to obtain surpluses. The agricultural activities take place in three management spaces: the mountain/hill, the plot or maizefield and the patio where domesticated and non-domesticated species interact and coexist. The forest area (1900 to 2500 m) is used mainly for grazing animals, far from the village and are managed by both adults and young people who are responsible for their care. In this area agricultural activities cannot be carried out, due to the distance, making it difficult to take care for them, in addition to the fact that soils are shallow and extremely rocky, and only grow some plant species like

Agroforestry Complexes in the Mountain Regions of Mexico

79

Fig. 8 Heterogeneous matrix of the village of Noxtepec, were agroforestry practices like patios, maize fields, orchards, and native vegetation interacts in a visible continuum. (Photo by the authors)

grass, cacti, some herbs and flowering plants. Further away, where there are no more management activities there are pines (Pinus spp.), oaks (Quercus spp.), amates or saiba (Ficus spp.), and the huizaches (Acacia spp.). The milpa/plot area is located between 1800 and up to 2500 m, this zone allows the subsistence of the inhabitants where maize, sometimes squash and sometimes beans are planted. These areas may be close to homes, where only community farmers can carry out activities, these areas are in variable altitudes because it may be near the forest area or near the patio in the village. To carry out clearance in forest patches to cultivate, slash and burn is practiced. Besides that, the whole area is destined to the management of domesticated species, the presence of non-domesticated native species, such as pines, oaks, and cacti that are used to delimit the land is remarkable. Between 1600 to 1800 m “patios” are located, next to houses being their composition mainly of fruiting and ornamental plants. This area is managed by the inhabitants of the home but is mostly fostered by women. We can find exotic fruit species such as banana (Musa x paradisiaca), coffee (Coffea arabica), lemons (Citrus spp.), peaches (Prunus persica); and native like guavas (Psidium guajava), papayas (Carica papaya), avocado (Persea americana), among others; we can also find ornamental species. This area features shallow fertile soils that require constant management (Fig. 8). Although other existing agroforestry systems where forest species interact with maize at higher altitudes of the village, where slabs stone are abundant and are used to stablish terraces, in Noxtepec, also exists a system that is located at the vicinity of

80

A. I. Moreno-Calles et al.

the river and that allows the use and implementation of riparian species intermingled with exotic crops such as coffee.

Agrosylviculture of Palms: Landscape Traditional Management One of the most bioculturally important palm species in Mexico is Brahea dulcis. This species is mainly found at elevations between 800 and 1600 m, in limestone soils, in deciduous tropical forests, oak forests, and xerophytic scrubland from subhumid to semidesertic climates (Pérez-Valladares et al. 2020; Quero 1994). At its distribution range B. dulcis has a wide variety of uses, from which the weaving of diverse objects with fibers obtained from its leaves, is an activity that continues to be basic for the economy of several rural communities in Mexico (Pérez-Valladares et al. 2020), as Santa María Ixcatlán and San Juan Sosola in the region known as Mixteca Alta in the state of Oaxaca. In this region B. dulcis outstands for the high number of managements practices related to it, and for the close relationship humans have established with it since ancient times (Smith 1965). The continuous management on the geographic space, has derived in variable socio-ecosystems where B. dulcis is the dominant species, these are known as palmonares (palm stands), which are almost always actively associated with agricultural production and livestock grazing. Palmonares are also source of a diversity of non-timber forest resources, as in Santa María Ixcatlán where 104 of the 115 species recorded in these systems have at least one use, some of them considered basic for subsistence, such as palm, “quelites” (edible herbs), medicinal plants, forages that keep small livestock (Rangel-Landa et al. 2016). The process of palm stands formation take place in such a gradual and unperceivable way, that in occasions the change may go unnoticed, but with time, it gives rise to a complex agrosilvopastoral matrix (Fig. 9). This conforms a type of landscape domestication (Terrell et al. 2003), a process by which people transforms their natural spaces to derive specific resources from their surroundings (Laland et al. 2000; Odling-Smee et al. 1996). In natural places where the palm distributes, when a plot of native vegetation (for instance, in oak forest or xerophytic shrubland) is cleared for cultivation, individuals of B. dulcis are tolerated inside the plot or as part of the boundaries (Fig. 10a) (Blancas 2001; Illsley et al. 2001; Rangel-Landa et al. 2014, 2016; Vallejo et al. 2014). Agricultural practices as soil enrichment, shrubs removal, and fire dedicated to the milpa, enhance the palm; but there are also particular practices explicitly performed for the palms, such as tolerance, pruning dry leaves, and cutting of shoots. The grazing of minor livestock (sheep and goats mostly), as management practice is relevant to maintain palmonares since B. dulcis is more resistant to grazing effects compared to other species of bushes and trees (Fig. 10b); this favors its dominance. Two types of palm stands are recognized (Fig. 10c), those known as manchoneras, where constant young leaves extraction derives in small height plants with colonial individuals; and the soyacahuiteras, where less intensive leave harvest, allows the development of arborescent habit and the prevalence of sexual reproduction. There is also an important temporal dimension context of these interactions. Since this area has been occupied by human cultures over millennia. Throughout time, the

Agroforestry Complexes in the Mountain Regions of Mexico

81

Fig. 9 Mosaic complex of agricultural land use, parcels in fallow and palm groves. Brahea dulcis is managed together with other species with high cultural value such as Juniperus flaccida (used as living fence, firewood, shadow, forage, medicine and construction) and Agave salmiana subsp. tehuacanensis (used as living fence, soil control, make the traditional food barbacoa, construction). Santa María Ixcatlán, Oaxaca. (Photo by the authors)

same plots that were once natural vegetation under communitarian management, later transformed to agricultural areas by familiar unities; with time left on fallow, the palm can easily proliferate, enhancing the formation of palm stands that are managed by the communities, which in a future could be reconverted to agriculture again. In this way, management of palmonares is articulated as a peculiar geographic expression at landscape level, in which management decisions also change through the time. Since agroforestry and management are inter-scalar processes, local actions exercised continually on individuals on places, over time have the faculty of modify large areas through the promotion of plant communities. This particular way of landscape management, the geographic locations commonly designated for these activities and the historical perspective of land use, project the palmonares, as systems that still have a lot to tell. They are related to other forms of human adaptation of natural spaces that modify large areas through the promotion of plant communities. At landscape level, these spaces can be seen as a mosaic of cultural and natural land units intermingled among them. This is accomplished without the profound disruptive outcomes of other kinds of land use and land use change like monocultures. Nevertheless, palm stands shows a decrease in biological diversity in comparison with natural communities, as was documented in Santa María Ixcatlán (Rangel-Landa et al. 2014; Vallejo et al. 2014) and some topics still need further study, such as: effects on soil and hydric dynamics. These are

82

A. I. Moreno-Calles et al.

Fig. 10 Phases and components of the agrosilvopastoral complex of the palmonares. (a) Recently opened land for agriculture in Santa María Ixcatlán. Oaks and palms are tolerated inside the cultivation area; (b) Grazing goat cattle in a palmonar in Santa María Ixcatlán. (Photo by the authors)

priority issues due to commonly observed erosion processes in these systems (Fig. 10d) that directly affect agricultural production and water supply for human activities.

Agroforestry Complexes in the Mountain Regions of Mexico

83

Interactionships Between Habitants and Diversity in Semiarid Agroforestry Complexes Within the Tehuacán-Cuicatlán Biosphere Reserve, in the state of Puebla, the town of Zapotitlán Salinas, is located at an altitude range of 1380–1700 m, where soils with high salt content dominate (Dávila et al. 1993). It has a dry, semi-warm climate, with a rainy season in summer, with an average rainfall of 425 mm and an average temperature of 21.2  C (Dávila et al. 1990). In these adverse conditions, people have developed multiple activities to support their economy. This locality is inhabited by a mestizo population, descendants of the Popoloca or “Injiva” indigenous culture. The inhabitants are mainly engaged in agriculture, gathering agroforestry diversity, goat breeding, mining and elaboration of onyx handicrafts, salt extraction, and offering tourist services. According to Rzedowski (1978), the vegetation types of this area are thorn forest, grassland and xerophytic scrub. The main plant communities present are the mezquital dominated by Prosopis laevigata, the thorn scrub, the tetechera dominated by Nebuxbaumia tetetzo, the cardonal dominated by Chepalocereus coluna-trajani, the izotal of Beaucarnea gracilis, and the tropical deciduous forest (Valiente-Banuet et al. 2000). Within the forms of management of the agro-diversity of Zapotitlán Salinas, agroforestry systems of semi-terraces are in plains with an average altitude of 1546 m. These semi-terraces, also known as “cuaxustles,” are agroforestry systems where maize beans and squash are traditionally cultivated (Fig. 11). Its structure is composed by trees, columnar cacti, globose cacti and shrubs, distributed as isolated individuals, islands or strips of vegetation within the cultivation land, as well as live fences (Moreno-Calles and Casas 2010). Species such as mezquite (Prosopis laevigata), manteco (Parkinsonia praecox), sotolín (Beaucarnea gracilis), garambullo (Myrtillocactus geometrizans), tetecho (Neobuxbaumia tetetzo), pitahaya (Selenicereus undatus), xoconostle (Stenocereus stellatus), izote (Yucca periculosa), pitzometl (Agave marmorata), prickly pear cactus (Opuntia spp.), and some globose cacti (Ferocactus latispinus and Echinocactus platyacanthus) (Paredes-Flores et al. 2007). Derived from a sampling of bird richness within these agroforestry systems, 89 bird species were recorded; 59 species are resident and 30 are migratory. This represents 68% of the documented bird richness for the Zapotitlán Salinas Valley. The families Parulidae (13 species), Tyrannidae (12 species), and the families Trochilidae, Troglodytidae and Passerellidae with seven species were the most abundant. 21 species (23.6%) show some degree of endemism and four species are under some protection category in the Official Mexican Norm (CONABIO 2020); three under special protection (Parabuteo unicinctus, Falco peregrinus, Aimophila notosticta) and one threatened (Geothlypis tolmiei). The agroforestry practices present in the semi-terraces of Zapotitlán Salinas, may be providing the birds with foraging and nesting sites, as well as facilitating their movement on a local and landscape scale, due to their proximity to the cacti forest patches and hills.

84

A. I. Moreno-Calles et al.

Fig. 11 Both images illustrate that the plots of these agroforestry systems are surrounded by the forest of columnar cacti and maintain interactions with the “forest,” sharing some plant species and maintaining a great associated diversity. (Photo by the authors)

Agroforestry Complexes in the Mountain Regions of Mexico

85

Metepantles: Agroforestry Terraces and Semiterrace of a TemplateSemiarid Zone of Tlaxcala, Mexico The municipality of El Carmen Tequexquitla is located east of the state of Tlaxcala at an altitude range between 2400 to 2700 m. It has a temperate semi-dry climate with summer rains, a dominant vegetation of xerophilous scrub, ecosystem where succulent prickly and rosetophylous plants dominate; there are also coniferous scrub (Juniperus spp.), rosetophylous scrub (Yucca spp.), sandy desert (dunes), and pine forest, essentially “stone pine” (Pinus cembroides) (Muñoz et al. 2006). El Carmen and neighboring municipalities are post-Columbian settlements and that responded to a strategic geographical location, since these places are called “transit sites” and have historically represented highly important trade routes between the center and east of the country, especially toward the Veracruz port. To a lesser extent, some people still make handicrafts made of “ocoxal” or “paja” (dried Pinus leaves), tule (Typha sp.), and sotol (Nolina sp.) basketry (Enciclopedia de los Municipios y Delegaciones de México 2020). The characteristic agroforestry systems of the region correspond to agroforestry terraces, their arrangement include the incorporation of wild or forest elements, especially plants, as strategies to retain soil, collect water, and reduce the effect of droughts on the main crops. Especially in the state of Tlaxcala, these terraces are known as “metepantles” and one of the main elements that make them up is the maguey (Agave salmiana), which gives its name to the system (Fig. 12). Metepantles have an area between half and one hectare, where two or three rows of vegetation, called bordos or terraces, are distributed generally at contour lines, in

Fig. 12 The main arrangement of the El Carmen agroforestry terraces, these strategies are of great importance in maintaining agrobiodiversity since they protect many wild and domesticated plant species, in addition to serving as retainers of water and soil. (Photo by the authors)

86

A. I. Moreno-Calles et al.

which wild and cultivated plants can be found, mainly magueyes, izotes, sotol (Dasylirion spp.), cocoons, walnut trees (Juglans regia), peaches, and plums (Prunus spp.). The main crops associated with this system are maize, beans, and squash, as well as oats and wheat.

Concluding Remarks • There is an important coincidence between the distribution of the Mexican mountains and the agroforestry complexes. • This relationship is old, in those with less time it is around one hundred years at least in various places. • Agroforestry complexes exist in temperate, subhumid and semi-arid mountains. The environmental conditions of the mountains promote diversification strategies that are addressed through agroforestry complexes. • Social organization is essential in the conditions of low humidity, cold, reduced space, steep slopes, and saline soils, among others, however the surrounding forest diversity is implemented and used, as well as introduced diversity. • Agroforestry means different paths of relating with the mountains these variety of agroforestry complexes allows people to continue inhabiting them and that favor unique processes such as the domestication of cultivars, species, systems and landscapes ornamentality, edibles, rituality, and social organization that only occurs at these altitude environments. Acknowledgments DGAPA-UNAM PAPIIT IG200720 “Agricultura y Agroforestería Social y Familiar en Contextos de Cambios Locales y Globales” Project and IA207721 Manejo biocultural del territorio agavero de México: Aporte desde las ciencias agroforestales.

Appendix 1 (IPNI) Family Acanthaceae Alstroemeriaceae Amaranthaceae

Specie Justicia spicigera Schltdl. Alstroemeria sp. Amaranthus hybridus L.

Beta vulgaris subsp. cicla (L.) Schübl. & G. Martens

Common name Mucle/Muicle

Use Medicine

SAF Homegardens

Astromelia; Peruvian Lily Quelite quintonil

Ornamental

Homegardens

Edible

Chard

Edible

Agroforestry milpas and agroforestry terraces of arid zones Homegardens

(continued)

Agroforestry Complexes in the Mountain Regions of Mexico

Common name Garlic Azucena; Lily Onion

87

Family Amaryllidaceae

Specie Allium sativum L. Hippeastrum sp. Allium cepa L.

Anacardiaceae Annonaceae

Mangifera indica L. Annona cherimola Mill. Pimpinella anisum L. Eryngium carlinae F. Delaroche Foeniculum vulgare Mill. Brahea dulcis (Kunth) Mart.

Mango Cherimoya

Use Edible Ornamental Food, pest control Edible Edible

Anise

Medicine

Toad Herb

Medicine

Fennel

Medicine

Soyate; Creole palm

Agave marmorata Roezl

Maguey pitzometl

Edible, handcrafted, ceremonial and construction Beverage and medicine

Yucca periculosa Baker

Izote

Fiber

Yucca sp.

Izotes

Edible (flowers)

Agave salmiana Otto ex. Salm Dyck

Boundary and edible

Agave sp.

Maguey cenizo o manso; Ash-colored or tame maguey Maguey

Boundary

Dasylirion sp.

Sotol

Handcrafted

Beaucarnea gracilis Lem.

Sotolín

Fodder and ornamental

Aloe vera (L.) Burm. f. Heterotheca inuloides Cass.

Aloe

Medicine

Agroforestry milpas and agroforestry terraces of arid zones Agroforestry terraces of arid zones Agroforestry terraces of arid zones Homegardens

Mexican arnica

Medicine

Homegardens

Apiaceae

Arecaceae

Asparagaceae

Asphodelaceae Asteraceae

SAF Homegardens Homegardens Homegardens Homegardens Long fallow systems Homegardens Long fallow systems Homegardens Agroforestry terraces of arid zones Agroforestry terraces of arid zones Agroforestry terraces of arid zones Agroforestry terraces of arid zones Agroforestry milpas and agroforestry terraces of arid zones

(continued)

88

Family

A. I. Moreno-Calles et al.

Specie Baccharis conferta Kunth

Betulaceae

Dahlia coccinea Cav. Gazania sp. Lactuca sativa L. Artemisia ludoviciana Nutt. Senecio rowleyanus H.Jacobsen Ustilago maydis (DC) Corda Begonia sp. Begonia spp. Alnus sp.

Brassicaceae Burseraceae

Alnus acuminata Kunth Raphanus sativus L. Bursera sp.

Basidiomycota Begoniaceae

Cactaceae

Common name Chamizo

Dalia

Ornamental

SAF Agroforestry milpas and Long fallow systems Homegardens

Sleepy plant Lettuce Prodigiosa, estafiate Rosary

Ornamental Edible Medicine

Homegardens Homegardens Homegardens

Ornamental

Homegardens

Huitlacoche

Edible

Angel wing Begoña Aile

Ornamental Ornamental Boundary and windbreakers Timber and firewood Edible

Agroforestry milpas Homegardens Homegardens Agroforestry milpas Long fallow systems Homegardens Long fallow systems Agroforestry terraces of arid zones Agroforestry terraces of arid zones Homegardens

Eagle stick Radish Copal

Use Timber and firewood

Echinocactus platyacanthus Link & Otto Ferocactus latispinus (Haw.) Britton & Rose Disocactus flagelliformis (L.) Lem. Myrtillocactus geometrizans (Mart. ex Pfeiff.) Console Opuntia sp.

Biznaga

Ornamental

Biznaga with hooked spines Floricuerno

Ornamental

Garambullo

Edible

Nopales

Boundary

Hylocereus undatus (Haw.) Britton & Rose Disocactus spp. (Disocactus speciosus (Cav.) Barthlott,

Dragon fruit

Edible

Agroforestry terraces of arid zones Agroforestry milpas and agroforestry terraces of arid zones Homegardens

Teresita/ pitaya / palmita

Ornamental

Homegardens

Ornamental

(continued)

Agroforestry Complexes in the Mountain Regions of Mexico

Common name

Use

SAF

Tetecho

Edible

Xoconostle

Edible

Papaya Batatilla Sheep tail

Edible Ornamental Ornamental

Agroforestry terraces of arid zones Agroforestry terraces of arid zones Homegardens Homegardens Homegardens

Child finger

Ornamental

Homegardens

Rococo; Cloud Rabbit ear

Ornamental

Homegardens

Ornamental

Homegardens

Miss nail

Ornamental

Homegardens

Squash

Edible

Chayote

Edible

Agroforestry milpas and Long fallow systems Homegardens

Chilacayota or/and chilacayote

Edible

Juniperus deppeana Steud.

Sabino

Juniperus flaccida Schltdl.

Táscate, nebro

Boundary, windbreakers, shade and firewood Firewood

Arbutus xalapensis Kunth Euphorbia milii Des Moul.

Madroño; madrone

Family

Specie

Caricaceae Convolvulaceae Crassulaceae

Disocactus phyllanthoides (DC.) Barthlott, Disocactus x Selenicereus, Disocactus x Epiphyllum) Neobuxbaumia tetetzo (J.M. Coult.) Backeb. Stenocereus stellatus (Pfeiif.) Riccob. Carica papaya L. Ipomoea sp. Sedum morganianum E. Walther Sedum rubrotinctum R.T. Clausen Sedum dasyphyllum L. Kalanchoe tomentosa Baker Sempervivum calcareum Jord. Cucurbita pepo L.

Cucurbitaceae

Sechium edule (Jacq.) Sw. Cucurbita ficifolia Bouché.

Cupressaceae

Ericaceae Euphorbiaceae

89

Boundary and windbreakers Ornamental

Agroforestry milpas and Long fallow systems Agroforestry terraces of arid zones Agroforestry terraces of arid zones Agroforestry milpas Homegardens (continued)

90

Family

Fabaceae

A. I. Moreno-Calles et al.

Specie

Common name Corona de Cristo; Crown of christ Nochebuena

Use

SAF

Ornamental

Homegardens

Colorín, pipí

Timber and firewood

Phaseolus vulgaris L. Leucaena leucocephala (Lam.) de Wit

Frijol; Bean

Edible

Guaje

Edible

Vicia faba L.

Haba; Faba bean Huizache

Edible

Manteco; Greasy tree

Shade

Mezquite; Mesquite

Edible, fodder and shade Boundary, shade, attractor of rain,, against frost, timber and firewood Ornamental Ornamental

Long fallow systems and silvopastoril systems Agroforestry milpas Long fallow system and agroforestry terraces of arid zones Agroforestry milpas Agroforestry terraces of arid zones Agroforestry terraces of arid zones Agroforestry terraces of arid zones Agroforestry milpas and Long fallow systems

Euphorbia pulcherrima Willd. ex Klotzsch Erythrina americana Mill.

Acacia farnesiana(L.) Willd. Parkinsonia praecox (Ruiz & Pav.) Hawkins Prosopis laevigata M. C. Jhonst Fagaceae

Quercus sp.

Encino; Oak

Geraniaceae Hydrangeaceae

Geranio Hortensia

Juglandaceae

Geranium sp. Hydrangea acuminata Siebold & Zucc. Juglans regia L.

Lamiaceae

Mentha spicata L. Lavandula angustifolia Mill. Origanum majorana L. Mentha piperita L.

Nogal, nuez; Walnut Hierbabuena; Peppermint Lavanda; Lavender Mejorana; Marjoram Menta; Mint Mirto

Fodder

Edible Medicine

Homegardens Homegardens

Agroforestry milpas Homegardens

Edible and pollination Medicine

Homegardens

Edible and medicine Medicine

Homegardens

Homegardens

Homegardens (continued)

Agroforestry Complexes in the Mountain Regions of Mexico

Family

Lauraceae

Lythraceae Malvaceae Moraceae

Musaceae Myrtaceae Nyctaginaceae

Specie Salvia microphylla Kunth Rosmarinus officinalis L. Thymus acicularis Waldst. & Kit. Plectranthus coleoides Benth. Persea americana Mill. Cuphea ignea A. DC. Malva sp. Ficus carica L. Ficus petiolaris Kunth Musa  paradisiaca L. Psidium guajava L.

Oleaceae

Bougainvillea glabra Choisy Fraxinus sp.

Onagraceae

Fuchsia sp.

Orchidaceae

Epidendrum radicans Pav. ex Lindl. Oxalis tuberosa Molina

Oxalidaceae

Pinaceae

Pinus lawsonii Roezl ex Gordon Abies religiosa (Kunth) Schltdl. & Cham Pinus cembroides Zucc. Pinus sp.

Poaceae

Zea mays L.

Common name

91

Use

SAF

Romero; Rosemary Tomillo; Thyme Vaporup

Medicine

Homegardens

Medicine

Homegardens

Medicine

Homegardens

Aguacate; Avocado

Edible

Cigarro de cantinflas Malva Higo; Fig Saiba Amarilla Plátano; Banana Guayaba; Guava Bugambilia

Ornamental

Homegardens, Long fallow systems Homegardens

Ornamental Edible Shade and boundary Edible

Homegardens Homegardens Long fallow systems Homegardens

Edible Ornamental

Long fallow systems Homegardens

Fresnos

Timber and firewood Ornamental

Long fallow systems Homegardens

Ornamental

Homegardens

Edible

Agroforestry milpas

Timber and firewood Boundary, firewood and recreational Edible (seeds)

Long fallow systems Agroforestry milpas

Arete/ fucsia; Earring Espíritu Santo; Holy Spirit Papa extranjera; Foreign potato Ocote Oyamel

Pino piñonero; Pinyon Pine Pino: Pine

Timber, firewoood and boundary

Agroforestry terraces of arid zones Agroforestry milpas and Long fallow systems

Edible (continued)

92

A. I. Moreno-Calles et al.

Family

Specie

Rosaceae

Cymbopogon citratus (D.C.) Stapf Prunus serotina subsp. capuli (Cav.) McVaugh

Medicine Food, boundary

SAF Agroforestry milpas Homegardens Agroforestry milpas and Long fallow systems Homegardens ans agroforestry milpas Homegardens ans agroforestry milpas Agroforestry milpas and Long fallow systems Homegardens

Chabacano; Apricot

Edible

Prunus domestica L.

Ciruelos; Plum

Edible

Prunus persica L.

Duraznos; Peach

Food and windbreakers

Fragaria sp.

Fresa; Strawberry Manzano; Apple

Edible

Membrillo; Quince Níspero; Japanese medlar Pera; Pear

Edible

Homegardens ans agroforestry milpas Homegardens

Edible

Homegardens

Boundary, edible and windbreakers

Crataegus mexicana DC.

Tejocote; Mexican hawthorn

Boundary, winbreaks and edible

Coffea arabica L. Citrus x limon (L.) Obsbeck Citrus sinesis (L.) Osbeck Ruta graveolens L. Capsicum annuum L.

Café; Coffee Limón; Lemon Naranja; Orange Ruda Chile; Chili pepper

Edible Edible and shade Edible

Homegardens ans agroforestry milpas Homegardens ans agroforestry milpas Homegardens Homegardens

Cydonia oblonga Mill. Eriobotrya japonica (Thunb.) Lindl. Pyrus communis L.

Solanaceae

Use

Prunus armeniaca L.

Malus domestica (Suckow) Borkh.

Rubiaceae Rutaceae

Common name Maíz; maize, corn Té Limón: Lemon tea Capulines; Black cherry

Edible and shade

Medicine Food and pest control Edible

Homegardens Homegardens Homegardens Homegardens (continued)

Agroforestry Complexes in the Mountain Regions of Mexico

Family

Specie Capsicum pubescens Ruiz & Pav. Physalis philadelphica Lam

Solanum lycopersicum L. Solanum sp.

Common name Chile perón, canario

93

Use

SAF

Miltomate

Edible

Jitomate; Tomato Papa criolla; Creole potato

Edible

Agroforestry milpas and agroforestry terraces of arid zones Homegardens

Edible

Agroforestry milpas

Appendix 2 Common name Bassarisk Mountain rabbit Snake Brown-backed solitaire Black vulture Painted redstart

Specie Bassariscus astutus Sylavilagus floridanus Crotalus sp. Myadestes occidentalis Coragyps atratus Myioborus pictus

Use Food

SAF Agroforestry milpas Agroforestry milpas Agroforestry milpas Agroforestry milpas Agroforestry milpas Agroforestry milpas

References Alarcón-Chaires P. Etnoecología de los indígenas P’urhépecha. Una guía para el análisis de apropiación de la naturaleza. Consejo Nacional de Ciencia y Tecnología; Morevallado; Universidad Nacional Autónoma de México, UNAM. 2009. Beilin R, Hunter A. Co-constructing the sustainable city: how indicators help us “grow” more than just food in community gardens. Local Environ. 2011;6:523–38. Beltrán AG. Obra antropológica, IX. Regiones de refugio: El desarrollo de la comunidad y el proceso dominical en Mestizoamérica, FCE: UV: INI: GEV; 1991. Blancas J. Estudio etnobotánico del “Soyatl” o palma (Brahea dulcis (HBK) MARTIUS) en la comunidad náhuatl de Huitziltepec, Eduardo Neri, Guerrero. Thesis, Universidad Nacional Autónoma de México, Ciudad de México. 2001. Blanckaert I, Swennen R, Paredes-Flores M, Rosas-López R, Lira-Saade R. Floristic composition, plant uses and management practices in homegardens of San Rafael Coxcatlán, Valley of Tehuacán-Cuicatlán. Mex J Arid Environ. 2004;57:179–202. Boege E. El patrimonio biocultural de los pueblos indígenas de México: hacia la conservación in situ de la biodiversidad y agrodiversidad en los territorios indígenas. Instituto Nacional de Antropología e Historia: Comisión Nacional para el Desarrollo de los Pueblos Indígena; 2008. Borelli S, Conigliaro M, Quaglia S, Salbitano F. Urban and Peri-urban agroforestry as multifunctional land use. In: Dagar J, Tewari V, editors. Agroforestry. Singapore: Springer; 2017.

94

A. I. Moreno-Calles et al.

Caballero J. Notas sobre el uso de los recursos naturales entre los antiguos purépecha. Biótica. 1982;1:31–42. Cervantes-Zamora Y, Cornejo-Olguín SL, Lucero-Márquez R, Espinoza-Rodríguez JM, MirandaViquez E, Pineda-Velázquez A. Provincias Fisiográficas de México. UNAM: Atlas Nacional de México. Instituto de Geografía; 1990. CONABIO. Categorías de riesgo en México. [Internet] 2020 [cited 2020 April 07]. Available from: https://www.biodiversidad.gob.mx/especies/catRiesMexico.html Dávila P, Chiang F, Villaseñor JL. Fitogeografía del Valle de Tehuacán. Cuicatlán Bol Soc Bot México. 1990;50:135–49. Dávila P, Villaseñor JL, Medina R, Ramírez A, Salinas A, Kent J, Tenorio P. Listado florístico de México X. UNAM: Flora del Valle de Tehuacán Cuicatlán. Instituto de Biología; 1993. Díaz GI. Entrevista realizada por Wilfrido López Martínez en la localidad de Santa María Tlahuitoltepec Mixe. Oaxaca: Marzo; 2017. Enciclopedia de los Municipios y Delegaciones de México. [Internet] cited 2020 April 07]. Available from: http://www.e-local.gob.mx/wb2/ELOCAL/EMM_tlaxcala Esparza SX. Isidro Fabela. Monografía Municipal. México: Instituto Mexiquense de Cultura; 1999. Farfán-Heredia B. Etnoecología Purépecha: motivos y consecuencias del manejo de plantas silvestres y hongos intercambiados en mercados tradicionales. thesis, Universidad Nacional Autónoma de México, Ciudad de México. 2019. García RA. Análisis integrado de paisajes en el occidente de la cuenca de México. La vertiente oriental de la Sierra de las Cruces, Monte Alto y Monte Bajo. Madrid: Universidad Complutense de Madrid; 1998. García B. Las regiones de México: breviario geográfico e histórico (Vol. 5). El Colegio de México AC; 2008. Illsley C, Aguilar J, Acosta J, García J, Gómez T, Caballero N. Contribuciones al conocimiento y manejo campesino de los palmares de Brahea dulcis (HBK) Mart. En la región de Chilapa, Guerrero. In: Rendón A, Rebollar S, Caballero N, Martínez A, editors. Universidad Autónoma Metropolitana. 2001. INEGI. Censo Nacional de Población y Vivienda 2010. [Internet] cited 2020 April 07]. Available from: https://www.inegi.org.mx/programas/ccpv/2010/ INEGI. Encuesta intercensal 2015. [Internet] cited 2020a April 07]. Available from: https://www. inegi.org.mx/programas/intercensal/2015/default.html#Tabulados.%20M%C3%A9xico;% 202015 INEGI. Uso of suelo y vegetación [Internet] cited 2020b April 07]. Available from: https://en.www. inegi.org.mx/temas/mapas/usosuelo/ International Plan Names Index (IPNI). Available from: https://www.ipni.org/ Jose S. Agroforestry for ecosystem services and environmental benefits: an overview. Agrofor Syst. 2009;1:1–10. Laland KN, Odling-Smee J, Feldman MW. Niche construction, biological evolution, and cultural change. Behav Brain Sci. 2000;23:131–46. Montagnini F, Kaimowitz D, Concepción GD, Khalid AR, Braden JB, Hansen PE, Encinas RA. Sistemas agroforestales: principios y aplicaciones en los trópicos. CATIE. 2015. Moreno-Calles AI, Casas A. Agroforestry systems: restoration of semiarid zones in the Tehuacán Valley, Central Mexico. Ecol Restor. 2010;28:361–8. Moreno-Calles AI, Toledo VM, Casas A. Los sistemas agroforestales tradicionales de México. Una aproximación biocultural. Bot Sci. 2013;4:375–98. Moreno-Calles AI, Galicia LV, Casas A, Toledo-Manzur V, Vallejo M, Santos-Fita D, Camou A. La Etnoagroforestería: El estudio de los sistemas agroforestales tradicionales de México. Rev Etnobiología. 2014;3:1–16. Moreno-Calles AI, Casas A, Rivero-Romero AD, Romero-Bautista YA, Rangel-Landa S, FisherOrtíz RA, Santos-Fita D. Ethnoagroforestry: integration of biocultural diversity for food sovereignty in Mexico. J Ethnobiol Ethnomed. 2016;54:2–21.

Agroforestry Complexes in the Mountain Regions of Mexico

95

Moreno-Calles AI, Cariño M, Sosa V, Soto-Pinto L, Rosales-Adame JJ, Montañez EP, Palma M, Pérez MS, Ruenes R, López-Martínez W, coordinadores. Sistemas Agroforestales de México: Formación, Investigación y Acciones Emergentes. CONACYT: UNAM; 2020. ISBN en trámite Muñoz L, Ibañez-Bernal S, Corona M. Los mosquitos (Diptera: Culicidae) de Tlaxcala, México: Lista comentada de especies. Fol Entomol Mex. 2006;45:223–71. Odling-Smee F, Laland K, Feldman M. Niche construction. J Am Nat. 1996;147:641–8. ONU HABITAT. Superficie de CDMX crece a ritmo tres veces superior al de su población. [Internet] cited 2020 April 07]. Available from: https://www.onuhabitat.org.mx/index.php/ superficie-de-cdmx-crece-a-ritmo-tres-veces-superior-al-de-su-poblacion Paredes-Flores M, Lira R, Dávila P. Estudio Etnobotánico de Zapotitlán Salinas, Puebla. Acta Bot Mex. 2007;79:13–61. Pérez-Valladares CX, Moreno-Calles AI, Casas A, Rangel-Landa S, Blancas J, Caballero J, Velazquez A. Ecological, cultural, and geographical implications of Brahea dulcis (Kunth) Mart. Insights for sustainable management in Mexico. Sustainability. 2020;12:412. Quero HJ. Las palmas de México: presente y futuro. Bol Soc Bot Méx. 1994;55:123–7. Rangel-Landa S, Rivera-Lozoya E, Casas A. Uso y manejo de las palmas Brahea spp. (Arecaceae) por el pueblo ixcateco de Santa María Ixcatlán Oaxaca, México. Gaia Scientia. 2014;17:62–78. Rangel-Landa S, Casas A, Rivera-Lozoya E, Torres-García I, Vallejo-Ramos M. Ixcatec ethnoecology: plant management and biocultural heritage in Oaxaca, Mexico. J Ethnobiol Ethnomedicine. 2016;12:30. Rojas RT. La agricultura en tierras mexicanas desde sus orígenes hasta nuestros días (No. 080 L6 v. 71). Consejo Nacional para la Cultura y las Artes; 1991. Rzedowski J. Vegetación de México. Limusa; 1978. Smith CE. Plant fibers and civilization- cotton, a case in point. Econ Bot. 1965;19:71–85. Stepp JR, Castaneda H, Cervone S. Mountains and biocultural diversity. Mt Res Dev. 2005;3: 223–7. Terrell JE, Hart JP, Barut S, Cellinese N, Curet A, Denham T, Kusimba CM, Latinis K, Oka R, Palka J, Pohl ED, Pope KO, Williams PR, Haines H, Staller JE. Domesticated landscapes: the subsistence ecology of plant and animal domestication. J Archaeol Method Theory. 2003;10: 323–68. Toledo VM, Barrera-Bassols N. La memoria biocultural: la importancia ecológica de las sabidurías tradicionales. Icaria. 2008. Torres P, Rodríguez L, García B. Mexico City: the integration of urban agriculture to contain urban sprawl. In: Bakker M, Dubbeling S, Gündel U, Sabel-Koschella S, Zeeuw H, editors. Growing cities, growing food: urban agriculture on the policy agenda. A reader on urban agricultura; Colombo, Sri Lanka: International Water Management Institute; 2000. Tuirán R, Rodolfo A, Ávila JL, Ramos LF, Córdova M. Índice de intensidad migratoria. MéxicoEstados Unidos; 2000. Valiente-Banuet A, Casas A, Alcántara A, Dávila P, Flores-Hernández N, Del Coro Arizmendi M, Villaseñor JL, Ortega J. La vegetación del Valle de Tehuacán-Cuicatlán. Bol Soc Bot Méx. 2000;67:24–74. Vallejo M, Casas A, Blancas J, Moreno-Calles AI, Solís L, Rangel-Landa S, Dávila P, Téllez O. Agroforestry systems in the highlands of the Tehuacán Valley, Mexico: indigenous cultures and biodiversity conservation. Agrofor Syst. 2014;88:125–40. Wilken GC. Good farmers: traditional agricultural resource management in Mexico and Central America. University of California Press; 1990. Wolf ER, Cirlot JE. Los campesinos. Barcelona: Labor; 1971.

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty Alejandro Casas, Berenice Farfa´n-Heredia, Andre´s Camou-Guerrero, Ignacio Torres-García, Jose´ Juan Blancas Vázquez, and Selene Rangel-Landa

Abstract

Archaeological studies have revealed that humans have inhabited the territory that today is Mexico for more than 20,000 years, while ethnobotanical research has identified that these peoples have made use of nearly 8,000 plant species, according to current information. More than 2,000 species are edible plants, that were and are included in their diet together with animals, mushrooms and a diverse microbiota responsible of fermentation of plant and animal substrates. Most edible species were and are obtained through gathering, hundreds of species have been managed in forests and anthropogenic areas to ensure their availability A. Casas (*) Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] B. Farfán-Heredia Universidad Intercultural Indígena de Michoacán, Pichátaro, Michoacán, Mexico A. Camou-Guerrero · I. Torres-García Escuela Nacional de Estudios Superiores (ENES), Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected]; [email protected] J. J. Blancas Vázquez Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico e-mail: [email protected] S. Rangel-Landa Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Consejo Nacional de Ciencia y Tecnología (CONACYT) – Escuela Nacional de Antropología e Historia (ENAH), Mexico City, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_3

97

98

A. Casas et al.

and a small fraction (nearly 250 species) became fully domesticated crops, managed in production systems that started practicing more than 10,000 years ago. Gathering and food production have coexisted for thousands of years, and crops have progressively displaced wild and weedy plants as main components of diet, but these plants are still important components of traditional food, especially in rural areas. This chapter gathers information from different regions and human cultures of Mexico that our research group has studied. It is directed to evaluate how important are wild and weedy products in human diet at present, how they are used and managed, and what role they play in food sovereignty in rural contexts. Information from 19 communities and regional markets among the Mixtec, Nahua, Mazahua, Rarámuri, Cuicatec, Ixcatec, Popoloca, and mestizo people was analyzed in La Montaña of Guerrero, the Sierra Tarahumara, the Monarch Butterfly Biosphere Reserve, and the Tehuacán-Cuicatlán Biosphere Reserve. In each community we carried out inventories of plants and other organisms used as food. The amounts of wild and weedy organisms consumed throughout the year were evaluated and compared with the food provided by the principal components of diet obtained from local production systems and/or interchanged in markets. We found that the proportion of wild and weedy plants that forms part of the annual diet may vary from 6.4% in mestizo communities of the Tehuacan Valley to nearly 21.9% in the Rarámuri community of Cuiteco in the Sierra Tarahumara, northern Mexico. The weight of wild and weedy plants in diet is not necessarily related to the diversity of products that are available; for example, the diversity of edible plants is higher in the Tehuacan Valley than in the Sierra Tarahumara. It is seemingly more related to human culture since, in general, the indigenous communities and households studied consume higher amounts of wild and weedy food than the mestizo communities and households. In some cases, it is also related to the level of deficit in the production of staple products through agriculture, which was recorded in all households sampled. Wild and weedy food importantly complement the edible biomass that the production systems are unable to cover; in addition, in all cases studied the highest consumption of wild and weedy products occurred during the period of greater scarcity of products from agriculture. The complementarity of products obtained through gathering is therefore clear in all cases analyzed. But, as mentioned, the cultural aspects were also relevant. Consuming wild and weedy products is associated to the indigenous people diet. In some cases, it is a motive of proud and is an element of cultural identity, in some others it is motive of discrimination. It is imperative to recognize that wild and weedy products complement the diet in biomass and nutriments, thus contributing to food security. But in addition, these edible products are part of a culinary culture and therefore important pieces of food sovereignty. In some areas, these products are being displaced by industrialized food of poor nutrimental quality but that, for some sectors of the population, have higher prestige. Therefore, promoting the cultural value of native food may be a way to combat the consumption of harmful industrialized food, maintaining culinary culture and enhancing food sovereignty.

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

99

Introduction The high biological and cultural diversity characterizing Mexico (Toledo and Ordóñez 1993; Boege 2008, 2009; Toledo 2001a, b; Toledo et al. 2010; Vidal and Brusca 2020) has determined the construction of landscapes exceptionally diverse. Throughout history, their components have been incorporated to human cultures, providing direct benefits as resources to satisfy multiple needs (Casas et al. 2016; Clement et al. 2021). Humans started to occupy the territory that today is Mexico more than 20,000 years ago (Ardelean et al. 2020) and, since the earliest occupations, peoples have developed different types of interactions with thousands of species of plants (nearly 8,000 species have been documented until now by ethnobotanists), animals, mushrooms and microorganisms occurring in the local ecosystems. Most organisms have been directly collected from the surrounding environments where humans have lived, but a fraction of these species have been under some type of management and an even smaller fraction of the managed species became domesticated in some degree, and with them humans started to practice production systems (MacNeish 1967, 1992; Flannery 1986; Smith 2005; ZizumboVillarreal and Colunga-GarcíaMarín 2010, 2017; Clement et al. 2021). Ethnobotanical studies have revealed that medicinal plants are generally the most diverse group of organisms that humans that have lived in Mexico have interacted with, followed by edible plants (more than 2,000 species; Caballero et al. 1998; Casas et al. 2016; Clement et al. 2021). Plant management have very much focused on medicinal and edible plant species (Caballero and Cortés 2001), but domestication has by far focused on plants providing food. Nearly 88% of the 251 native crops of Mexico are edible plants (Clement et al. 2021). This general pattern suggests that attention of health problems has based more on diversity of options and incipient management, whereas satisfaction of food needs has based more on both diversity and management practices directed to ensure their availability and domestication directed to increase and improve the edible products obtained from them. Recent inventories of edible plants occurring in Mexico have identified 2,168 edible plant species (Mapes and Basurto 2016), nearly 1,800 of them are native to Mexico and neighboring Mesoamerican areas (Clement et al. 2021), and the rest were introduced to its territory at different times (prehistory, pre-Columbian times, post-European conquest and modern times; Corona et al. 2021). We have identified 1,665 edible species that are obtained through gathering in the wild and weedy environments (Clement et al. 2021). It draws the attention the fact that, after probably more than 10,000 years of practicing production systems such as horticulture and agriculture, the rural peoples of Mexico continue carrying out gathering of food, and this activity is still very important to their subsistence pattern (Viveros et al. 1993, Casas et al. 1994). Our more recent records have identified 630 edible plant species under some management form, and we have identified that these species are nearly 80% of all native plant species receiving management practices, some of them (at least 364 species) with signs of incipient domestication. And we have recorded 251 native species that are domesticated crops (Clement et al. 2021), nearly 88% of them (220 species) being edible plants.

100

A. Casas et al.

Before such panorama, some important research questions that we explore in this chapter are how important are wild, weedy and domesticated plants in rural people’s diet? How gathering, management and domestication are related to such cultural importance? Rural people of Mexico are nearly 21.4% of the total population of the country (INEGI 2021). It is currently composed by thousands of communities of mestizo people, many of them with strong indigenous influence because their relatives recognized themselves as indigenous people, or because they coexist and interact with neighbors recognized themselves as indigenous people, or/and because at regional and national levels numerous aspects of the Mexican identity are rooted in indigenous cultures. An important portion of rural people of Mexico recognize themselves as indigenous (according to INEGI 2021, and INPI 2015, nearly 21.5% of all Mexican people define themselves as indigenous, and more than half of them live in rural contexts), and another portion is Afro-Mexican, among other cultural groups with their own identity. In general, rural peoples include in their diet numerous wild and weedy plant species, the number of species and the proportion of the diet that is composed by these elements vary extraordinarily among regions, human cultures and households, but there is not a clear evaluation about it and what influences on it. We explore in this chapter how variable can these aspects be and the possible factors that contribute to explain such patterns. Gathering consists, in general terms, in collecting products existing in populations of organisms occurring in forests or anthropized areas independently of human intention (Casas et al. 1996, 1997, 2017). In plants, gathering is recognized to be carried out on products of non-crop species. But this activity is commonly more complex than one can imagine. People commonly construct rules and agreements to planning the spatial area, the season, and amounts of products that can be gathered; sometimes gathering involves especial tools and techniques; and some products involve certain human selection since the quality of products may be variable (Casas et al. 1996, 2017; Blancas et al. 2010). Most edible products are gathered in wild areas, but many of them also occur in secondary forests, in patches of wild vegetation in agroforestry systems, and nearly 300 species may be weedy edible plants associated to crop fields. Nearly 730 species of edible plants (approximately 40% of edible plants recorded until now in Mexico) are under different types of management. More than 440 species are tolerated or maintained in forest areas that are cleared for different purposes, or weedy plants that are let standing during the weeding of crop fields because these will eventually be used, mainly as food (Casas et al. 1996, 2017; Blancas et al. 2010, 2013). About 120 species receive different practices of protection to decrease the pressure of herbivores (some common practices are to cover the base of the plants with spiny branches, which are effective against vertebrates herbivores; sometimes people put white caustic lime around or on the stems, which protect the desirable plants against ants, snails and caterpillars); other practices are directed to diminish pressure from competition with other plants, by directly removing competitors using their hands, cutting tools or fire; other practices are directed to benefit the desirable plants through irrigation systems, removing plants to procure them light capture or, conversely, adding covers to protect plants against excess of light or frost (Casas

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

101

et al. 1996, 1997, 2017; Blancas et al. 2010, 2013; Torres et al. 2015). Nearly 400 species have been described to be fomented or enhanced by the deliberate propagation of seeds, vegetative propagules or transplanting; this is a practice resembling those commonly considered as cultivation. It is pertinent to mention that all the practices described above may involve human selection (Casas et al. 1996, 1997, 2017; Blancas et al. 2010, 2013). Like in some practices of gathering, people favor or benefit through these practices those plants with phenotypes with favorable attributes; therefore, they selectively tolerate, protect or enhance those phenotypes with better attributes (Casas et al. 1996, 1997, 2017; Blancas et al. 2010, 2013). These selective practices have been studied in different species and have demonstrated to have consequences as part of domestication processes (Casas and Caballero 1996; Casas et al. 1999, 2007; Cruz and Casas 2002; Arellano and Casas 2003; Otero-Arnaiz et al. 2005; Blancas et al. 2009; Parra et al. 2010, 2012; Aguirre-Dugua et al. 2012; Abad-Fitz et al. 2020). We have recorded 251 native plant species that are cultivated with advanced signs of domestication (Clement et al. 2021). These are properly the native crop species of Mexico and the neighboring Mesoamerican area. In total, approximately 40% of all edible plants recorded in Mexico are managed, and around 88% of crops native to Mexico are edible plants.

General Research Methods Our research group has conducted ethnobotanical studies focusing on management and domestication of edible plants and their relation to food patterns. Food is a primary need and edible resources have been part of the culture throughout human history. Therefore, understanding management and domestication of edible resources (plants, animals, fungi and microbiota of ferments) is necessarily based on the study of food patterns and culinary culture. We have conducted studies of diet and food resources and culture among different cultural groups and regions of Mexico. In these studies, we have centered our attention in (1) inventorying the species used as food, their forms of use, preparation and management; (2) we have evaluated their importance in diet through semi-structured interviews, periodic surveys to reconstruct food patterns from the diet, based on the “diet of the day before” approach, as well as participatory workshops to explore communitarian consensus about the main basic components of diet and those components considered as complementary food; (3) we invested efforts to evaluate (weighting) the amounts of components of the diet in kilograms; (4) we have evaluated the level of food sufficiency by contrasting information of consumption with information about the amounts of food produced by households and those obtained through gathering. The latter approach involved analyses about the local availability and abundance of edible wild and weedy products through vegetation sampling, phenological studies and interviews; and (5) we have conducted studies of inter-communitarian interchange and commercialization of products that allow obtaining what is locally absent or scarce.

102

A. Casas et al.

Fig. 1 Regions and human cultures where our research group has conducted ethnobotanical studies. In green, the regions sampled for analyzing rural food pattern in this chapter

Our studies have comprised different regions and cultural groups (Fig. 1), but in this chapter we summarize the main findings in contrasting biocultural areas, mainly among the Rarámuri of the Sierra Tarahumara in Chihuahua, northern Mexico; also, among the Mixtec, Nahua and Mestizo communities of La Montaña de Guerrero region, the Mazahua people of the Monarch Butterfly Biosphere Reserve, and the Ixcatec, Mixtec, Cuicatec, Popolocan, Nahua, and Mestizo people of the TehuacánCuicatlán Biosphere Reserve. Based on a sample of case studies, in this chapter we explore general patterns of traditional rural diet, looking for providing elements to answer the questions mentioned above.

La Montan˜a de Guerrero Region This region is located at the northeast of the state of Guerrero, limiting with the state of Oaxaca (Fig. 1). La Montaña region is characterized by a complex mountainous landscape, with elevations from 1,100 to nearly 3,000 m and an extraordinary mosaic of ecosystems, including different types of thorn-scrub and tropical deciduous forests in the lowlands up to elevations around 1,400 m; in addition, different types of oak and pine oak forests, some relics of cloud forest in the highlands, and a

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

103

high diversity of riparian vegetation strongly varying with elevations and soil types (Casas et al. 1994). The region is inhabited by Mixtec, Nahua, Tlapanec, and Mestizo people and is recognized to be one of the regions of Mexico with the highest indexes of poverty (Casas et al. 1994; Morales-Hernández 2015; Hernández-Corchado 2019). In this region we conducted ethnobotanical studies inventorying the flora the regional cultures interact with. In this chapter we center our attention in four communities of the municipality of Alcozauca, where we carried out in depth studies of food patterns. Our studies in this region comprised the communities of Alcozauca, Amapilca, San José Lagunas, and Ixcuinatoyac with Mixtec and mestizo households (20, 10, 12 and 14 households, respectively). In those communities we found 162 edible plant species; 102 species are wild and weedy plants obtained from gathering, 20 are native species of the area introduced into cultivation, and the rest are crops native to Mexico and introduced from other areas of the world. We classified the types of food provided by these plants as (1) greens or vegetables with edible leaves and young stems, flower buds and flowers and in some cases (Leucaena esculenta, L. macrophylla, L. leucocephalla, Gonolobus spp. Matelea spp., among others) their immature fruits; (2) fruits, which are plant species whose mature fruits are consumed fresh; (3) roots, bulbs and tubers, whose subterranean parts are edible; (4) seeds and grains, providing edible seeds, nuts and dry fruits (outstandingly mesquite pods and oak acorns); (5) condiments, which provide different parts consumed in small amounts, mainly used as flavouring; and (6) beverages, including fermented beverages (mainly those produced with agave sap or pulque, those produced with maize or chicha, and several species whose leaves and seeds are used as infusions or substitutes of coffee accompanying food). Table 1 helps to show a general panorama of how these food types included wild, weedy and domesticated species found in the communities studied, which illustrates the type of information documented in other communities and regions commented ahead in this chapter. The readers may notice that, in all cases, the sum referred to does not check with the total numbers mentioned on the text. This is because several species are at the same time wild, weedy and domesticated and their different parts may be used as different food types. The main sources of wild edible plants are the riparian vegetation (26 species), the tropical deciduous forest (24 species), oak forest (20 species), pine-oak forest (8 species) (Fig. 2). Crop fields, mainly milpa system and homegardens are sources of 30 weedy edible species, mainly the greens called quilite in Nahuatl and yiwa in

Table 1 Number of wild, weedy, and crop plant species used for different food types in the diet of people of the La Montaña de Guerrero in the state of Guerrero, central Mexico

Use form Greens and vegetables Fruit Roots, bulbs, tubers Seeds and grains Condiments Beverages

Wild 30 23 9 17 3 4

Weedy 27 2 1 4 4 1

Domesticated 27 40 3 10 5 2

104

A. Casas et al.

Fig. 2 General aspects of the environments that mostly provide wild edible species (occurring in the forest types), weedy edible plants, mainly quilites and tomatoes, and cultivated plants maintained in agricultural fields. Some species occurring in the forest are carried into the agricultural areas. (Photos: José Blancas)

Mixtec and several species of tomatoes of the genera Solanum and Physalis, and nearly 80 species of cultivated plants (domesticated and not domesticated).

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

105

The interviews and periodic surveys allowed identifying the general pattern of the diet, composed by a group of frequent food that we considered “basic components of diet,” and a group of food seasonally consumed or complementing the basic components of the diet, which we call here “complementary components of diet” (Table 2). Again, the example of the La Montaña de Guerrero region helps to illustrate the type of information documented in the other regions and cultures shown in this chapter. Wild and weedy food is consumed seasonally and for this reason it is considered complementary. Some species produce edible parts throughout the year, but most of them during specific seasons. However, during the seasons when edible plants are available, they become main components of the diet, especially during the periods when the main basic components of diet are scarce (Fig. 3). This period may be two to 3 months a year. A balance of the amounts of products consumed throughout the year allows estimating that the complementary products constitute on average 12% of the biomass consumed (considering the total of basic and complementary components of diet together) by the general population studied. It is nearly 14.5% for the diet of indigenous people and about 7.1% for the mestizo people. The latter sector consumes more industrialized products, outstandingly sugar and sodas, pasta and other industrialized fast food products. Some mestizo people refer to wild and weedy food, Table 2 General pattern of the daily/weekly diet of adult indigenous and mestizo people of the La Montaña de Guerrero region, in the state of Guerrero, central Mexico. Digits in parentheses are numbers of pieces consumed by children Meal Breakfast

Indigenous people diet (n ¼36) Maize tortillas: 2–4 (1–2) Maize dough atole 3 cups/week Sauce: 1/2 bowl/day Water

Lunch

Maize tortillas 4 (1–2) Beans: 1 dish Eggs 1–2 pieces per week Meat: 150 g/month Sauce: ½ bowl/day Water

Dinner

Maize tortillas: 1–3 (1–2) Sauce: ½ bowl/day Water

Mestizo people diet (n ¼ 20) Maize tortillas: 2–3 (1–2) Beans: 1 dish Eggs: 3–4 pieces/week Milk 1–2 cups/week Coffee: 3–5 cups/week Chocolate: 1 cup/week Bread 3 pieces/week Sauce: 1/2 bowl/day Meat 150 g /week White or brown sugar Water Maize tortillas: 3 (1–2) Beans: 1 dish Eggs: 3 pieces/week Sauce Meat: 300–450 g/week Rice: 2–3 dishes/week Pasta soup: 2–4 dishes/week Water Bread: 2(1) pieces/day Coffee: 7 cups/week Chocolate: 1–2 cups/week

106

A. Casas et al.

Fig. 3 A general panorama of the temporal availability of basic and complementary components of diet in the La Montaña of Guerrero region. The thickness of the bars represent amounts of products available, estimated based on the interviews and measurements of products harvested throughout the year. The red square indicates the periods of scarcity of the components of the basic diet coinciding with the periods when wild and weedy food increase their availability

especially the quilite or yiwa, as “indian food” in a derogatory way (Viveros et al. 1993; Casas et al. 1994), an attitude identified in La Montaña Region two decades later, as well as in other regions of Mexico where our research group has conducted ethnobotanical research. The balance of the amounts of agricultural products produced is about 60% of the food annually required; in other words, it is in deficit in all households studied in this and other regions. Such unbalance indicates that wild and weedy products in part complement the amount of food that is not produced by households and such complementarity is particularly important during the periods when the basic components of the diet are scarcer; the latter is the period when the agricultural cycle starts at the beginning of the rainy season. In this period the products of the previous agricultural cycle are in the lowest level of availability in the households’ stores, but this is the season when wild and weedy greens, fruits and seeds are more abundant (Fig. 3). The complementarity is also relevant in terms of nutriments since agricultural products are particularly rich in carbohydrates and proteins, while the wild and weedy components provide vitamins, fiber, oils, proteins, and carbohydrates (see details in Viveros et al. 1993; Casas et al. 1994).

The Mazahua of the Monarch Butterfly Biosphere Reserve The Monarch Butterfly Biosphere Reserve is located in the highlands of the states of Mexico and Michoacán. This area forms part of the Trans-Mexican Volcanic Belt, mainly covered by temperate forest of oak, pine-oak, pine and Abies

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

107

(Farfán-Heredia et al. 2007). It is a protected area that called the attention of conservationists of the world because it is an important point of the migration of the monarch butterfly during winter. This area has been inhabited by Otomi, Matlatzinca, P’urhépecha, and, majorly, by Mazahua people. We have conducted ethnobotanical studies in Mazahua communities, part of them can be visualized in this book (FarfánHeredia and Casas). As part of that ethnobotanical project, we documented the local people’s diet. An in depth study was conducted in the community of Francisco Serrato (Farfán Heredia 2001). There, we documented that local people live based on agricultural practices. They mainly cultivate maize, wheat and beans, but they produce 60.60%, 60.78% and 36.84% of the amounts they require to consume annually of these foods, respectively. The deficit of staple food is compensated through wild and weedy plants (51 species) and, more importantly than in any other community studied by our group, mushrooms (31 species), which are mostly available between June and early September, when products of agriculture are not available yet (Fig. 4). Therefore, the pattern of complementarity is similar to that described for the Mixtec people of La Montaña de Guerrero region. In this community we studied the food patterns of 20 households, all of them Mazahua, and we estimated that on average wild and weedy plants and mushrooms provide 14.2% of the annual edible food.

The Rara´muri of Cuiteco As described in the chapter by Camou and collaborators in this book, the Rarámuri live in the Sierra Tarahumara, which is the northern portion of the Western Sierra Madre. People of Cuiteco are mainly agriculturalists, practicing cultivation of maize, beans, wheat, and potatoes. They conceded the exploitation of timber from their forests to private companies, and they obtained very low incomes; therefore, they decided to stop the concession. Their lives depend mainly on agricultural products they produce (Fig. 5), but this is insufficient to satisfy their needs throughout the year (Camou-Guerrero et al. 2008). They complement their diet based on 27 non-crop edible plants (wild and weedy plants), which represents 48% of the edible species used by local people. Our studies of diet and evaluations of amounts of products included in it, indicate that on average 21.9% of the annual biomass consumed by the Rarámuri people is obtained from wild and weedy plants. From all villages studied by our research team, this is the highest value recorded (Camou-Guerrero et al. 2008). And again, although wild and weedy plants used as food are available throughout the year, their abundance is higher during the first months of the rainy season, when they become the basic component of the diet.

Peoples of the Tehuaca´n-Cuicatla´n Biosphere Reserve The Tehuacán-Cuicatlán Valley is an iconic biocultural region of Mexico, partly because during the 1960s a team of archaeologists leaded by Richard MacNeish was able to reconstruct the most complete records of the prehistory of Mesoamerica

108

A. Casas et al.

Fig. 4 Basic and complementary components of the Mazahua people diet. (a) cultivation of maize and beans in milpa system, (b) cultivation of wheat, (c) maize harvest, (d) wheat threshing, (e, f) bean shelling, (g) preparation of maize tortillas and tamales, the staple food. (h–k) wild food products quelites (Amaranthus hybridus), zarzamora (Rubus liebmannii), different species of fungi (Ramaria spp., Hypomyces lactifluorum, Amanita caesarea sensu lato) and capulines (Prunus serotina). (Photo: Berenice Farfán-Heredia)

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

109

Fig. 5 Food preparation and the diet of the Rarámuri people in Cuiteco, Chihuahua northern Mexico. (a) Typical metallic stove of the Sierra Tarahumara used for cooking tortillas prepared with native varieties of maize, (b) Harvesting the “heart” of nopal cladodes (Opuntia spp.) by removing their spines, (c) Preparing maize tamales for the “fiesta,” (d) Drying fruit (peaches and apples) for consuming them during periods of scarcity, (e) The Rarámuri milpa with corn and squash, (f) Peach harvest from the local homegardens. (Photos: Andrés Camou)

(MacNeish 1967). In addition, since biologists working in the region have documented one of the richest biodiversity of Mexico (Dávila et al. 2002; Casas et al. 2017), while ethnobiologists have documented the highest ethnobotanical knowledge of Mexico at regional level (Casas et al. 2016, 2017). The earliest occupations of the region by humans are approximately 12,000 to 11,000 years old (MacNeish 1967) when, according to archaeological records, human subsistence based on hunting of large animals or megafauna (Flannery 1967; Callen 1967; MacNeish 1967). Extinction of several species of megafauna led eventually humans to increase consumption of small animals and edible plants in their diet, during the phase called El Riego. The Tehuacán Valley has been recognized as one of the areas of Mexico, and the Americas, with the earliest remains of domestication of plants and practice of food production, particularly clear during the phase named Coxcatlán. The archaeological records indicate that early forms of food production started to provide significant resources to human diet some 7,000 years ago, and that this activity gained progressively a principal role in human life, displacing the consumption of wild animals and plants (Fig. 6). Presence of domestic animals in the region was recorded in relatively late strata. Dogs and turkeys appeared in the area approximately 3,000 years ago, when people lived in

110

A. Casas et al.

Fig. 6 Changes in the availability of components of human diet throughout time in the Tehuacan Valley, central Mexico. The general scheme, originally elaborated by the team of archaeologists leaded by Richard S. MacNeish (1967), based on the amounts (and proportions represented by the areas of each polygon) of remains of plants and animals at different strata in the floor of caves inhabited by humans in the region. (Modified from MacNeish 1967)

agriculturalist villages. Since then, domestic animals contributed to human life in relatively small amounts, even smaller than wild animals. Figure 6 illustrates the proportions of different components of human diet throughout time, according to MacNeish (1967). It is a general panorama deduced by archaeologists based on the abundance of remains in the floors of the caves they studied, as well as the study of human coprolites (MacNeish 1967; Callen 1967). But it represents a reference point about what happened from prehistory to until nearly 500 years ago, and how this panorama drastically changed during the Colonial period and, especially, during the last century. According to that scheme, about 9,000 years before the present, in the Tehuacan Valley the diet of humans was more than 60% constituted by wild animal products and nearly 30% edible wild plants. Then, during the Coxcatlan phase, wild animals and plants had similar weight in human diet, while crops contributed nearly 10% of the diet. But some 6,000 years later the cultivated plants constituted nearly 75% of diet and wild animals and plants the rest, with a small amount of domestic animal products (Fig. 6, MacNeish 1967). As in other regions, our study in the Tehuacan Valley included surveys, interviews, measurements of amounts of food components throughout the year, evaluations of yields of agricultural products, especially staple food, and vegetation sampling exploring the spatial abundance of the plant resources (useful plants in general and edible plants in particular; in one case study we included estimations of abundance of some species of edible fauna; see Solis and Casas 2019). All of them

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

111

have been strategies to reconstruct the current general diet of people in a sample of rural communities. We analyzed the level of sufficiency of households to acquire their food, the spectrum of food that forms part of forests and agrobiodiversity local people interact with, the balance of consumption of local products and industrialized components in diet. We therefore characterized food patterns among households in rural villages of the region, investigated limitations and problems to access food, the local resources available for designing food security programs, and outlined general strategies that are possible to achieve regional food security and sovereignty based on local resources and production systems. We conducted our regional studies in 13 rural communities (Coxcatlán, San Rafael, Zapotitlán, San Juan Raya, Santa María Ixcatlán, San Lorenzo Pápalo, Quiotepec, Metzontla, San Luis Atolotitlán, San Pedro Nodón, Aticpac, Ahuatla, and Coyomeapan) distributed throughout the region (Fig. 7). These communities were sampled in order to have a representation of different human cultures (Náhuatl, Mixtec, Ixcatec, Cuicatec, Popolocan, and mestizo people) and

Fig. 7 The Tehuacán-Cuicatlán Valley region, indicating the location of the communities included in the analysis of food patterns and the role of wild, weedy, and cultivated plant resources

112

A. Casas et al.

environments (territories generally including tropical dry or humid forest, thornscrub forest, cloud forest, and temperate forests with oaks and pines). We conducted semi-structured interviews, surveys and participatory observation during people’s daily life in relation to local natural resources (mainly plants and animals), production systems (mainly agriculture, raising of domestic animals, extraction of forest products, and manufacturing of handcrafts), and food patterns. We characterized the food patterns following similar methods as referred to above, evaluating amounts of food consumed in typical preparation of the main stoves, amounts of products obtained from agriculture, and amounts of edible products obtained from agroforestry systems and forests. We also inventoried all species of plants and animals consumed as food, recording amounts and frequency of consumption throughout the year or during seasons of availability. Details of the ethnobotanical studies and subsistence patterns can be followed in Casas et al. (2001, 2008), Rangel-Landa and Lemus-Fernández (2002), Echeverría (2003), Torres (2004), Solís (2006), Pérez Negrón and Casas (2007), Paredes-Flores et al. (2007), Blancas et al. (2010, 2013), Arellanes et al. (2013), Olvera (2016), Rangel-Landa et al. (2016, 2017). Our general ethnobotanical studies have comprised 21 communities (we focus now in 13 of them) and 6 regional markets, and these studies have contributed to identify the richest inventory of plant species used and managed by peoples at regional level in Mexico. At present this inventory includes more than 2,000 species, approximately 85% of which are native from the local forests. On average, each community uses 275  43 spp. From the general inventory, we have identified 424 species of plants used as food (Fig. 8); on average each community reported 92  12.7 edible species, and, also on average, we have estimated that the annual consumption of wild and weedy plants is 9.3  3.9%, compared with the basic components of the diet (Table 3; Fig. 9). Consumption of wild animal protein is still important. Although difficult to evaluate, some ethnozoological studies in the area have identified 38 species of insects, 21 of mammals, 20 of birds and 5 of reptiles (Fig. 10). Among the general patterns that can be visualized from Table 3 are: (1) most of the diet is based on plants either crops, wild and weedy species and, in some communities, especially those of the highlands, mushrooms; (2) the main components of diet are staple crops (maize, beans and, in communities of the highlands, wheat) and fruit produced by households, an commonly rice, obtained in the markets; (3) animal protein is the following important component of diet, mainly provided by eggs, meat and milk, mostly obtained in the market; however, meat consumed in the communitarian parties is derived from animals raised in the village; (4) wild meat is important, as revealed by few studies carried out in the region; documenting its role is difficult and there is scarce information; but the information available indicates that wild meat and a high variety of insects are important in the local diet; (5) industrialized food, mainly soda and pasta is relatively highly consumed, it is more consumed than wild and weedy plants; it is unhealthy and should be attended by public policy programs.

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

113

Fig. 8 General panorama of the variation of wild and weedy edible plant species in the TehuacánCuicatlán Valley. Our inventory has recorded 424 plant species in the region, included herbaceous,

114

A. Casas et al.

Discussion Biocultural Diversity and Diversity of Food Patterns The cases analyzed in this chapter are a small sample of the extraordinary diversity of ecological and cultural contexts of Mexico. It is far to be representative and, therefore, still limited to arrive to conclusions. A comprehensive understanding of food patterns, the cultural, ecological, and economic factors influencing on them deserve more extensive research and systematization of the information available. Such analysis will be helpful to evaluate the past and current value of food non-crop resources, and their potential to be included in programs for improving healthy food, recovering ancient culinary culture and establishing the bases for its innovation. However, the cases studied show general trends that can help to identify important patterns: (1) the high value that still have numerous wild and weedy plants in the diet of rural people, which include millions of Mexican people; (2) the extraordinary diversity of species, edible products and nutriments that can provide to diet; (3) the high diversity of forms of preparation that can be documented and that conform an ancient culinary culture with great possibilities of innovation; (4) the general context of cultural discrimination associated to the consumption of these resources; (5) the general trend toward losing so important food that is being rapidly replaced by harmful industrialized foods, which are expensive, poor in nutrients and are cause of critical public health problems such as obesity, diabetes, cardio-vascular diseases, among others; (6) the need to promote ecological diagnoses of the conservation status of wild and weedy plants before the scenarios of deforestation and use of herbicides in agricultural systems; (7) the need to enhance ethnobotanical studies from transdisciplinary perspectives directed to promote a multisectoral collaboration for constructing a vast systematized program of culinary culture; (8) the importance of carrying out research to evaluate the nutritional value and toxicological aspects of these hundreds of resources; (9) the generation of spaces in massive communication media and scholar programs directed to promote their healthy value and the multiple ways to include in diet of people of both rural and urban contexts; and (10) the importance of studying current and potential circuits of interchange that are and may be able to provide products among communities, regions and rural and urban

ä Fig. 8 (continued) shrubby, and arboreal plants, producing edible leaves, flower buds, fruits, roots and tubers, seeds and grains, some of them with commercial value in the regional markets. (a) Floral bottoms of the matzitzi (Dasylirion serratifolium), commercialized in the regional markets of mountain region of Tehuacan Valley, Mexico; (b) Recollection and commercialization of individual complete of tequilitl (Peperomia peltilimba), a cultural relevant edible plant in Aticpac, Coyomeapan, Mexico; (c) Floral bottoms of the mexcalli cacaya (Agave obscura) recollected and sold in the regional markets of Tehuacan Valley. (Photos: José Blancas)

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

115

Table 3 Average components of diet in communities of the Tehuacán-Cuicatlán Valley

Food product Maize Beans Rice Bread Fruit crops Eggs Meat Milk Wild and weedy plants Wild meat Industrialized food

Consumption per household per week (kg) 17.10  11.90 kg 4.30  2.00 1.41  0.98 5.60  1.98 12.50  4.80 3.50  1.25 3.85  1.80 12.20  7.20 5.44  1.60

Selfsufficiency (%) 47–59.73% 65–100% 0.00 0–15.00 10–100 10–20 10–15 0–8 100

6.4–11

? 14.50  4.80

100 0.00

2–5.3 12–15.3

% of diet

40–48

18–23

contexts. All these facts and linked actions could be the basis to seriously contribute to maintain such important biocultural heritage. In all actions enunciated above, ethnobotany should play a central role. As scientific discipline, ethnobotany has the opportunity to include taxonomic, ecological, phytochemical, and evolutionary approaches and/or linking to specialists in these areas to continue investigating and systematizing information on the general inventory of food resources (wild, weedy, and crops), their distribution, abundance, biotic interactions and the current state of their availability, their nutritional value, their state of management and domestication. Similarly, ethnobotany may include approaches and/or links with specialists in studies of governance, nutrition, anthropology of nutrition, food science, gastronomy, among other fields, to strengthen its capacities to analyze diverse topics related to the nutritional aspects of traditional food, the quality of the rural and urban diet and the ways to improve them based on local products, the steps necessary to rescue and innovate the culinary culture, the better ways of social organization to harvest and manage food products to guaranty equitability in the access to benefits, among other important issues. From the general diagnosis of this chapter, several questions and hypotheses arise. Some of them are related to the variation of diet and the amount of wild and weedy components in it, and the causes of the variation. We found that variation in the proportion of wild and weedy plants may range from 6.4% of the annual diet in mestizo communities of the Tehuacan Valley (even lower among some mestizo households of Alcozauca, Guerrero) to nearly 21.9% in the Rarámuri community studied in the Sierra Tarahumara, in northern Mexico (even higher among some households of the Rarámuri and the Mazahua communities studied). These ciphers are lower than what the archaeologists have estimated could have been the diet of indigenous people in the Tehuacán valley just before the arrival of the Spaniards when, according to estimations by MacNeish (1967), wild animals and plants could

116

A. Casas et al.

Fig. 9 Different dishes of the regional diet including wild and weedy edible plants and animals in the Tehuacán Valley. Their complementarity in terms of edible biomass, seasonal availability, and

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

117

constitute around 25% of people’s diet in that region. The level of importance of wild and weedy plants in diet is not necessarily related to the diversity of products that are available in the surrounding ecosystems. For example, the diversity of edible plants is much higher in the Tehuacan Valley than in the Sierra Tarahumara or in the Monarch Butterfly Biosphere Reserve, while the level of consumption of wild and weedy plants, on average, is higher in the latter two regions. It is partly related to human culture since, in general, the indigenous communities or households studied consume higher amounts of wild and weedy food than the mestizo communities or households. In such context, the pejorative expression among some mestizo households that those plants are “indian food” reflects that certainly, people that recognize themselves as indigenous consume more these plants than other people that deny being indigenous. For some households, it is a motive of proud and is an element of cultural identity, for some others it is motive of discrimination, and they hide information about their consumption. We recorded this information in the interviews conducted, confirming this general pattern, probably because people interviewed recognized or not through their answer their cultural condition. But, anyway, the consideration of this food as “Indian” reveals that it is a sign of the cultural value of these components of diet. It is in addition necessary considering that changes in land use impact the availability of wild and weedy products; when forests are transformed into crop lands the weedy food may increase its abundance while the wild one decreases. However, it is in this context necessary considering that the use of herbicides in crop lands may affect the availability of weedy food. In addition, it is important to mention that since the 1990s, in Mexico the collective tenure of land (the Ejido and Comunidad) have suffered changes, since reforms to the Mexican laws allowed privatizing the collective areas. Such changes have established restrictions to access resources that in the past were available to all households in a community. This process has become dramatic in some regions, where the organized crime has taken the previous collective areas and the restriction to access forest products may be violent. And certainly, it is important to mention the negative effect that harmful food industry has had increasing its net of commercialization and the great support that find in the massive media and the active marketing promoting their products. Seasonal migration, which is high in rural Mexico, contributes to root new values of “prestigious” food acquired in other regions.

ä Fig. 9 (continued) nutrimental composition are part of the explanation of why after more than 10,000 years of practicing agriculture in the area gathering an hunting continue being important activities to satisfy the alimentary needs. (a) Xochiquilitl (Phaseolus coccineus), a weedy plant recollected in the milpa; (b) Gathering of male inflorescences of tepejilote (Chamaedorea tepejilote) a wild plant of rainforest; (c) Cultivation of tlanilpaquilitl (Piper auritum) in homegardens of Sierra Negra, Mexico. (Photo: José Blancas)

118

A. Casas et al.

Fig. 10 Edible animals. Regional studies by Zarazua-Carbajal et al. (2020) have identified 38 species of insects, 21 of mammals, 20 of birds and 5 of reptiles that form part of the diet of people of the Tehuacán-Cuicatlán Valley. (a) Roasted larvae Polybia sp. prepared in San Luis Atolotitlán, Puebla, Mexico (Photo: Selene Rangel-Landa); (b) Roasted larvae Polybia sp. prepared in Aticpac, Coyomeapan, Puebla, Mexico (Photo: Mariana Zarazúa); (c) Larvae Arsenura polyodonta (Poxokuili) prepared in sauce (Photo: Mariana Zarazúa); (d, e) Thassus gigas (Cocopache) an edible hemiptere (Photos: Ignacio Torres-García); (f) Larvae Arsenura armida (Cuetlas) (Photo: Ignacio TorresGarcía), (g, h) Larvae Paradirphia fumosa (Cuchamá) (Photos: Ignacio Torres-García); (i) Sciurus sp. (Moto or ardilla; squirrel) (Photo: Mariana Zarazúa); (j) Nasua narica (Pesojtli or tejón; badger) (Photo: Mariana Zarazúa); and (k) Roasted squirrel (Sciurus sp.) (Photo: Ignacio Torres-García)

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

119

But, in counterpart, the consumption of wild and weedy plants is also a question of food security. It is related to the level of the deficit in the production of staple products through agriculture, which was recorded in all households sampled. Wild and weedy food importantly complement the edible biomass that the production systems are unable to cover. We in general registered the information for a year; but we have recorded in several studies (Blancas et al. 2013; Rangel-Landa et al. 2016, 2017) that people are aware about the interannual variation in the availability of some edible species. During dry years crop production is severely affected, and although some wild and weedy edible plants decrease their availability as well, they offer options to face years of food scarcity. In addition, in all cases studied the highest consumption of wild and weedy products in a year occurred during the period of more scarcity of products from agriculture. The complementarity of products obtained through gathering is therefore clear in all cases analyzed in this chapter. It is imperative to recognize that wild and weedy products complement the diet in biomass and nutriments, thus contributing to food security. But in addition, these edible products are part of a culinary culture and therefore important pieces of food sovereignty. In some areas these products are being displaced by industrialized food of poor nutrimental quality but that some sectors of the population consider having higher prestige. Therefore, promoting the cultural value of native food may be a way to combat the consumption of harmful industrialized food, maintaining culinary culture and enhancing food sovereignty. Documenting, constructing, sharing and promoting a culinary culture is an important part of a strategy for maintaining and innovating gastronomy, attending issues of food security and, overall, food sovereignty. All these are dynamic issues, cultural and ecological contexts are continually changing, as well as gastronomy. The traditional food is the base for continuing constructing local gastronomies and may play a significant role in the purpose of gaining autonomy and recovering healthy food to face the harmful food. A transdisciplinary approach of ethnobotany with other disciplines, and scholars interacting with the local people and other actors (NGOs, governmental agencies, social organizations, among others) may construct vigorous programs for constructing, promoting, and rooting food sovereignty. Our research team has developed experiences in alliance with other sectors that inspire the enormous horizon of possibilities to carry out programs linking culinarian culture with food sovereignty at communitarian, regional, and national levels. First, in La Montaña de Guerrero we produced an illustrated catalogue of edible plants (Cabrera et al. 1998), based on ethnobotanical and ecological research and direct evaluations and a compilation of nutritional information for some plant species. Together with this material, distributed in schools of the region, we collaborated with the radio station XEZV La Voz de la Montaña, which transmits in Nahuatl, Tlapanec, Mixtec, and Spanish. We collaborated providing basic information on a group of edible plants, their nutritional properties and some recipes. The radio speakers were women, who made dramatizations based on the information we provided, transmitting the programs in the different languages referred to. They invited to the public to share their own recipes and the program had a great response. This experience was carried out in the 1990s, but at present the radio, television and other media may

Fig. 11 Aspect of the recipe book elaborated by Olvera-Espinosa (2016) based on the local ways of preparation of quilite in the Sierra Negra, neighboring the Tehuacán Valley. This book is a useful instrument for promoting the local food patterns and contribute to maintain both plant resources and food sovereignty

120 A. Casas et al.

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

121

Fig. 12 Aspect of the didactic guide for the identification of fungi, elaborated by Farfán-Heredia (2010), based on the traditional ecological knowledge of edible, medicinal, and toxic fungi. It was developed with the purpose of diffusing the traditional ecological knowledge, use forms, management, and promote the consumption of wild mushrooms from temperate forests of central Mexico

122

A. Casas et al.

Fig. 13 Workshops for home processing of wild, weed and cultivated plants as a strategy for productive diversification and to contribute to construct processed toward food sovereignty. A to F participatory workshops to produce fruit liqueurs, preserves and packaging of fruits, vegetables and mushrooms in indigenous and peasant communities; G capulín fruit liqueur (Prunus serotina); H Chilacayote candies (Cucurbita ficifolia); I to L different types of preserves, sweets, and canned fruits and mushrooms. (Photos: Berenice Farfán-Heredia)

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

123

expand options of communication. Schools, social organizations, NGOs and governmental agencies working in environmental and nutrition issues are crucial allies. A second experience was carried out in the Sierra Negra, a region neighboring the Tehuacan Valley, where our team worked in producing a recipes book of quilite, illustrated, written in Spanish and Nahuatl (Fig. 10). This material has been very much appreciated by local authorities, schools and families of the region. It is an instrument for working with groups of different sectors and to provide material to communicators using different media. Another interesting experience was carried out in Michoacan, where a guide of edible, medicinal and toxic mushrooms (FarfánHeredia 2010) has helped in diffusing information to recognize them and recommending forms of management the forest where these mushrooms grow (Fig. 11). In addition, several workshops have been organized in communities and with people of the markets of Michoacán, showing different forms of preparing canned food with different presentations (Fig. 12). These workshops have explored different possibilities to innovate food preparation for their commercialization and interchange among communities and among rural and urban areas. These experiences are initiatives that emerged from local people and our research group, but that have important potential to actions. Ethnobotany is a research field connected with different social and natural scientific disciplines and is making important theoretical contributions. But also, ethnobotany has an extensive horizon to project proposals and link sector for attending different social and environmental problems. It is an important area for constructing the sustainability science, with high potential to contribute to attend different social-environmental issues, among them food sovereignty (Fig. 13). Acknowledgments The authors thank the Consejo Nacional de Ciencia y Tecnología (CONACYT research project A1-S-14306), Mexico, the Dirección General de Asuntos del Personal Académico (DGAPA, UNAM, research project IN206520, IN224023), and the Comisión Nacional para el Conocimiento y Uso de la Biodiversidad (CONABIO/GEF/FAO, research project RG023, GEF project ID 9380 770) for financial support.

References Abad-Fitz I, Maldonado-Almanza B, Aguilar-Dorantes K, Gómez-Caudillo L, Casas A, Blancas J, Beltrán-Rodríguez L, Sierra-Huelsz JA, Cristians-Nizawa S, Moreno-Calles AI, GarcíaRodríguez Y, Espinosa-García F, Sánchez-Méndez L. Consequences of traditional management in the production and quality of copal resin (Bursera bipinnata (Moc. & Sessé ex DC.) Engl.) in Mexico. Forests. 2020;11(9):991. Aguirre-Dugua X, González A, Eguiarte L, Casas A. Large and round: morphological and genetic consequences of artificial selection on the gourd tree Crescentia cujete by the Maya from the Yucatán Peninsula, Mexico. Ann Bot. 2012;109(7):1307–16. Ardelean CF, Becerra-Valdivia L, Pedersen MW, Schwenninger JL, Oviatt CG, Macías-Quintero JI, Arroyo-Cabrales J, Sikora M, Ocampo-Díaz YZE, Rubio-Cisneros II, Watling JG, Medeiros VB, Oliveira PE, Barba-Pingarón J, Ortiz-Butrón A, Blancas-Vázquez J, Rivera-González I, Solís-Rosales C, Rodríguez-Ceja M, Gandy DA, Navarro-Gutierrez Z, De La Rosa-Díaz JJ, Huerta-Arellano V, Marroquín-Fernández MB, Martínez-Riojas LM, López-Jiménez A,

124

A. Casas et al.

Higham T, Willerslev E. Evidence of human occupation in Mexico around the Last Glacial Maximum. Nature. 2020;584:87–92. Arellanes Y, Casas A, Arellanes-Meixueiro A, Vega E, Blancas J, Vallejo M, Torres I, Solís L, Pérez-Negrón E. Influence of traditional markets and interchange on plant management in the Tehuacán Valley. J Ethnobiol Ethnomed. 2013;9:38. Arellano E, Casas A. Morphological variation and domestication of Escontria chiotilla (Cactaceae) under silvicultural management in the Tehuacán Valley, Central Mexico. Genet Resour Crop Evol. 2003;50(4):439–53. Blancas J, Casas A, Lira R, Caballero J. Traditional management and morphological patterns of Myrtillocactus schenckii (Cactaceae) in the Tehuacán Valley, Central Mexico. Econ Bot. 2009;63(4):375–87. Blancas J, Casas A, Rangel-Landa S, Torres I, Pérez-Negrón E, Solís L, Moreno AI, Delgado A, Parra F, Arellanes Y, Caballero J, Cortés L, Lira R, Dávila P. Plant management in the TehuacánCuicatlán Valley. Econ Bot. 2010;64(4):287–302. Blancas J, Casas A, Pérez-Salicrup D, Caballero J, Vega E. Ecological and sociocultural factors influencing plant management in Náhuatl communities of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013;9:39. Boege E. El patrimonio biocultural de los pueblos indígenas de México: hacia la conservación in situ de la biodiversidad y agrodiversidad en los territorios indígenas. Mexico City: Instituto Nacional de Antropología e Historia, Comisión Nacional para el Desarrollo de Pueblos Indígenas; 2008. Boege E. El reto de la conservación de la biodiversidad en los territorios de los pueblos indígenas. In: Capital natural de México (Vol. II): Estado de conservación y tendencias de cambio. Mexico City: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad (CONABIO); 2009. Caballero J, Cortés L. Percepción, uso y manejo tradicional de los recursos vegtales en Mexico. In: Rendón-Aguilar B, Rebollar-Domínguez S, Caballero J, Martínez-Alfaro MA, editors. Plantas, cultura y sociedad. Estudios sobre la relación entre seres humanos y plantas en los albores del siglo XXI. México City: Universidad Autónoma Metropolitana, Secretaría del Medio Ambiente, Recursos Naturales y Pesca; 2001. p. 79–100. Caballero J, Casas A, Cortés L, Mapes C. Patrones en el conocimiento, uso y manejo de plantas en pueblos indígenas de México. Estudios Atacameños. 1998;16:181–96. Cabrera J, Casas A, Rojas MC, Viveros JL. Alimentos en la Naturaleza. Algunas plantas comestibles, silvestres, arvenses y ruderales. Mexico City: Secretaría del Medio Ambiente Recursos Naturales y Pesca; 1998. Callen EO. Analysis of the Tehuacan coprolites. In: The prheistory of the Tehuacan Valley, Volume one, environment and subsistence pp. 261–289, Byers DS (ed.), University of Texas Press, Austin. 1967. Camou-Guerrero A, Reyes-García V, Martínez-Ramos M, Casas A. Knowledge and use value of plant species in a Raramuri community: a gender perspective. Hum Ecol. 2008;36:259–72. Casas A, Caballero J. Traditional management and morphological variation in Leucaena esculenta (Moc. et Sessé ex A.DC.) Benth. in the Mixtec region of Guerrero, Mexico. Econ Bot. 1996;50(2):167–81. Casas A, Viveros JL, Caballero J. Etnobotánica mixteca: sociedad, cultura y recursos naturales en la Montaña de Guerrero. México City: Instituto Nacional Indigenista-Consejo Nacional para la Cultura y las Artes (CONACULTA) Mexico City, Mexico; 1994. Casas A, Vázquez MC, Viveros JL, Caballero J. Plant management among the Nahua and the Mixtec of the Balsas River basin: an ethnobotanical approach to the study of plant domestication. Hum Ecol. 1996;24(4):455–78. Casas A, Caballero J, Mapes C, Zárate S. Manejo de la vegetación, domesticación de plantas y origen de la agricultura en Mesoamérica. Bol Soc Bot México. 1997;61:31–47. Casas A, Caballero J, Valiente-Banuet A, Soriano JA, Dávila P. Morphological variation and the process of domestication of Stenocereus stellatus (Cactaceae) in Central Mexico. Am J Bot. 1999;86(4):522–33.

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

125

Casas A, Valiente-Banuet A, Viveros JL, Dávila P, Lira R, Caballero J, Cortés L, Rodríguez I. Plant resources of the Tehuacán Valley, Mexico. Econ Bot. 2001;55(1):129–66. Casas A, Otero-Arnaiz A, Pérez-Negrón E, Valiente-Banuet A. In situ management and domestication of plants in Mesoamerica. Ann Bot. 2007;100(5):1101–15. Casas A, Rangel-Landa S, Torres-García I, Pérez-Negrón E, Solís L, Parra F, Delgado A, Blancas J, Farfán B, Moreno AI. In situ management and conservation of plant resources in the TehuacánCuicatlán Valley, Mexico: an ethnobotanical and ecological approach. In: Albuquerque UP, Alves-Ramos M, editors. Current topics in ethnobotany. Kerala: Research Signpost; 2008. p. 1–25. Casas A, Torres-Guevara J, Parra F. Domesticación en el continente americano. Volumen 1. Manejo de biodiversidad y evolución dirigida por las culturas del Nuevo Mundo. Universidad Nacional Autónoma de México, Universidad Nacional Agraria La Molina. Lima, Perú. 2016. Casas A, Torres-Guevara J, Parra F. Domesticación en el continente americano. Volumen 2. Perspectivas de investigación y manejo sustentable de recursos genéticos en el Nuevo Mundo. Universidad Nacional Autónoma de México, Universidad Nacional Agraria La Molina. Morelia, Michoacán, México. 2017. Clement CR, Casas A, Parra-Rondinel FA, Levis C, Peroni N, Hanazaki N, Cortés-Zárraga L, Rangel-Landa S, Alves RP, Ferreira MJ, Cassino MF, Coelho SD, Cruz-Soriano A, PancorboOlivera M, Blancas J, Martínez-Ballesté A, Lemes G, Lotero-Velásquez E, Bertin VM, Mazzochini GG. Disentangling domestication from food production systems in the Neotropics. Quaternay. 2021;4(1):4. Corona E, Casas A, Argueta-Villamar A, Alvarado-León CI. La domesticación de especies y paisajes en México. In: Bali J, editor. México. Grandeza y diversidad. Mexico City: Instituto Nacional de Antropología e Historia (INAH); 2021. p. 78–98. Cruz M, Casas A. Reproductive biology and morphological variation of Polaskia chende (Cactaceae) under domestication in Central Mexico. J Arid Environ. 2002;51:561–76. Dávila P, Arizmendi MC, Valiente-Banuet A, Villaseñor JL, Casas A, Lira R. Biological diversity in the Tehuacán-Cuicatlán Valley, Mexico. Biodivers Conserv. 2002;11:421–42. Echeverría Y. Ecología de los recursos vegetales en una comunidad Mixteca del Valle de Tehuacán. Bachelor thesis, Universidad Michoacana de San Nicolás de Hidalgo. Morelia; 2003. Farfán Heredia B. Ecología de los recursos vegetales en una comunidad Mazahua de la reserva de la Mariposa Monarca, Michoacán. Bachelor thesis, Universidad Michoacana de San Nicolás de Hidalgo. Morelia; 2001. Farfán-Heredia B. Guía didáctica para la identificación de hongos silvestres del oriente de Michoacán. Michoacán: Editorial Universidad Intercultural Indígena de Michoacán/ELIFAR Grupo especializado en proyectos ambientales S.C; 2010. Farfán-Heredia B, Casas A, Ibarra G, Pérez-Negrón E. Mazahua ethnobotany and people subsistence in the Monarch Butterfly Biosphere Reserve, Mexico. Econ Bot. 2007;61(2):173–91. Flannery KV. The vertebrate fauna and hunting patterns. In: The prhistory of the Tehuacan Valley. Volume one, environment and subsistence, pp. 132–177, Byers DS (ed.). University of Texas Press, Austin. 1967. Flannery KV. Guilá Naquitz. New York: Academic; 1986. Hernández-Corchado R. Proletarianization and dispossession of Mixtec workers: roots of indigenous migration to New York. Íconos Rev Ciencias Soc. 2019;63:19–36. INEGI Instituto Nacional de Estadística y Geografía. Censos y Conteos de Población y Vivienda. Series históricas. Consulted in https://www.inegi.org.mx/datos/?ps¼Programas. 30 Mar 2021. Instituto Nacional de Pueblos Indígenas de México (INPI). Indicadores Socioeconómicos de los Pueblos Indígenas de México. https://www.gob.mx/inpi/articulos/indicadores-socio economicos-de-los-pueblos-indigenas-de-mexico-2015-116128. Mexico City; 2015. MacNeish RS. A summary of the subsistence. In: Byers D, editor. The prehistory of the Tehuacan Valley. Environment and subsistence. Austin: University of Texas Press; 1967. p. 290–310. MacNeish RS. The origins of agriculture and settled life. New York: Oklahoma University Press; 1992.

126

A. Casas et al.

Mapes C, Basurto F. Biodiversity and edible plants of Mexico. In: Lira R, Casas A, Blancas J, editors. Ethobotany of Mexico. New York: Springer; 2016. p. 83–132. Morales-Hernández R. Análisis regional de la marginación en el estado de Guerrero, México. Papeles de Población. 2015;84:251–74. Olvera Espinosa A. ¡Que viva el quelite! ¡Mayolliti in kilitl! Recetario de quelites de Coyomeapan, Puebla. Bachelor thesis. Environmental Sciences Programme, UNAM, Morelia; 2016. Otero-Arnaiz A, Casas A, Hamrick JL, Cruse J. Genetic variation and evolution of Polaskia chichipe (Cactaceae) under domestication in the Tehuacán Valley, Central Mexico analyzed by microsatellite polymorphism. Mol Ecol. 2005;14(6):1603–11. Paredes-Flores M, Lira Saade R, Dávila-Aranda PD. Estudio etnobotánico de Zapotitlán Salinas, Puebla. Acta Bot Mex. 2007;79:13–61. Parra F, Casas A, Peñaloza-Ramírez JM, Cortés-Palomec AC, Rocha-Ramírez V, GonzálezRodríguez A. Evolution under domestication: ongoing artificial selection and divergence of wild and managed Stenocereus pruinosus (Cactaceae) populations in the Tehuacán Valley, Mexico. Ann Bot. 2010;106:483–96. Parra F, Blancas J, Casas A. Landscape management and domestication of Stenocereus pruinosus (Cactaceae) in the Tehuacán Valley: human guided selection and gene flow. J Ethnobiol Ethnomed. 2012;8:32. Pérez-Negrón E, Casas A. Use, extraction rates and spatial availability of plant resources in the Tehuacán-Cuicatlán Valley, Mexico: the case of Quiotepec, Oaxaca. J Arid Environ. 2007;70(2): 356–79. Rangel-Landa S, Lemus-Fernández R. Aspectos etnobotánicos y ecológicos de los recursos vegetales entre los Ixcatecos de Santa María Ixcatlán, Oaxaca. Bachelor thesis, Universidad Michoacana de San Nicolás de Hidalgo. Morelia; 2002. Rangel-Landa S, Casas A, Rivera-Lozoya E, Torres I, Vallejo M. Ixcatec ethnoecology: biocultural principles of plant management in Oaxaca, Mexico. J Ethnobiol Ethnomed. 2016;12:30. Rangel-Landa S, Casas A, García-Frapolli E, Lira R. Socio- cultural and ecological factors influencing plants management: comparing edible, medicinal and ceremonial plants in Santa María Ixcatlán, Oaxaca, Mexico. J Ethnobiol Ethnomed. 2017;13:59. Smith B. Reassessing Coxcatlan Cave and the early history of domesticated plants in Mesoamerica. Proc Natl Acad Sci. 2005;102(27):9438–45. Solís L. Etnoecología cuicateca en San Lorenzo Pápalo, Oaxaca. Master thesis. Postgraduate in Biological Sciences, UNAM, Morelia; 2006. Solís L, Casas A. Cuicatec ethnozoology: traditional knowledge, use and management of fauna by people of San Lorenzo Pápalo, Oaxaca, Mexico. J Ethnobiol Ethnomed. 2019;15:58. Toledo VM. Biocultural diversity and local power in Mexico. In: Maffi L, editor. On biocultural diversity: linking language, knowledge and the environment. Washington, DC: Smithsonian Institution Press; 2001a. p. 472–88. Toledo VM. Indigenous peoples and biodiversity. In: Levin SA, editor. Encyclopedia of biodiversity. San Diego: Academic; 2001b. p. 451–63. Toledo VM, Ordóñez MJ. The biodiversity of Mexico: a review of terrestrial habitats. In: Ramamoorthy TP, Bye R, Lot A, Fa J, editors. Biological diversity of Mexico. Origins and distribution. Oxford: Oxford University Press; 1993. p. 757–78. Toledo VM, Boege E, Barrera-Bassols N. The biocultural heritage of Mexico: an overview. Landscape. 2010;2(6):7–13. Torres I. Aspectos Etnobotánicos y Ecológicos de los Recursos Vegetales en la comunidad de San Luis Atolotitlán, Municipio de Caltepec, Puebla, México. Bachelor thesis, Universidad Michoacana de San Nicolás de Hidalgo. Morelia; 2004. Torres I, Blancas J, León A, Casas A. TEK, local perceptions of risk, and diversity of management practices of Agave inaequidens in Michoacán, Mexico. J Ethnobiol Ethnomed. 2015;11:61. Vidal O, Brusca RC. Mexico’s biocultural diversity in peril. Rev Biol Trop. 2020;68(2):669–91.

Wild, Weedy and Domesticated Plants for Food Security and Sovereignty

127

Viveros JL, Casas A, Caballero J. Las plantas comestibles y la alimentación entre los mixtecos guerrerenses. In: Leff E, Carabias J, editors. La dimensión cultural del desarrollo sustentable. Mexico City: Centro de Investigaciones Interdisciplinarias en Humanidades, U.N.A.M-Porrúa Editores Mexico City, Mexico.; 1993. Zarazúa-Carbajal M, Chávez-Gutiérrez M, Romero-Bautista Y, Rangel-Landa S, Moreno-Calles AI, Alvarado-Ramos F, Blancas-Vázquez J, del Val de Gortari E, Arizmendi MC, Casas A. Use and management of wild fauna by peoples of the Tehuacán-Cuicatlán Valley and the surrounding areas, Mexico. J Ethnobiol Ethnomed. 2020;16(4):1–23. Zizumbo-Villarreal D, Colunga-GarcíaMarín P. Origin of agriculture and plant domestication in West Mesoamerica. Genet Resour Crop Evol. 2010;57:813–25. Zizumbo-Villarreal D, Colunga-GarcíaMarín P. La milpa del occidente de Mesoamérica: profundidad histórica, dinámica evolutiva y rutas de dispersión a Suramérica. Rev Geogr Agrícola. 2017;58:33–46.

Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers, Conduits, and Generators of Plant-People Interactions and Relationships Robert Bye and Edelmira Linares

Abstract

Mountains provide a backdrop to study ethnobotanical processes. Their varied topography, orientation, and gradients allow the development of an array of floristic and vegetational mosaics. The resulting variability necessitates adjustments by inhabitants through a mixed strategy of resource exploitation. In turn, they appropriate and modify surrounding plant resources. In this chapter we analyze the interactions and relationships between the Rarámuri people of the Sierra Tarahumara in northern Mexico. We centered our attention in the diversity of strategies in different vegetation types, which have been studied by our team for nearly 50 years working in the region.

Introduction Because many ethnic groups continue to live in the mountains of Mexico, the Mexican Sierra Madres (SM) favor many ethnobotanical processes.1 Within the SM system (Occidental, Oriental, Neovolcanic, Sur, and Chiapas) where temperate 1

Ethnobotany used here in the context of the study of the interactions and relationships between people and plants in the various contexts (e.g., culture, geography, society, economy, ecology, and evolution, among others) and across different scales of time and space. Interactions involve processes that have reciprocal impacts on plants and people; relationships result in modifications of one of the components’ genetic or cultural information systems. Emic and etic perspectives may differ. Ethnobotanical studies should assure that all participants be respected, and they share benefits from the research that are fair and equative. R. Bye (*) Laboratorio de Etnobotánica, Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico E. Linares Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_5

129

130

R. Bye and E. Linares

forests dominate, 47 native ethnic groups are guardians for 31.3% of the territory (Boege 2008). Of the 49 primary vegetation types in Mexico found in the Indigenous territories, that of the Tarahumara or Rarámuri people rank 2nd (after the Maya people’s territories) in the extension of land area (2,647,372 of 28,033,092 ha). The Tarahumara people are principal inhabitants of primary vegetation for: one subhumid tropical zone (tropical deciduous forest) and 4 subhumid temperate forests zones (pine, pine-oak, oak, and oak-pine forests). Of the SM system, the Sierra Madre Occidental (SMOc) has the greatest latitudinal range extending 1,200 km from the north in Sonora and Chihuahua to the south in Jalisco. Three ecosystems (Madrean, Xerophytic Madrean, and Tropical) converge in the cordillera presenting 15 primary vegetation types. Located in the 560-km-long northern portion of the SMOc, the Sierra Tarahumara (ST) occupies the western part of the state of Chihuahua where 4 Indigenous peoples live: Tarahumara (Rarámuri), Tepehuan, Warihio, and Pima Alta. Subsequent to the Conquest and later Mexican Independence, European and mestizo peoples have settled in the region, especially during the twentieth century. Prior to the colonization, other Indigenous peoples inhabited the region but became extinct, among them the Chinipas, Cocoyome, Conchos, Eudeve, Guazapares, Joyas, Opata, Salineros, Tobosos, and Tubares. Based upon the 2010 data collected by INEGI (2010) in 15 municipos of the Sierra Tarahumara, 278,516 individuals self-identified as being Indigenous. Of these, 94,988 spoke one of the four native languages of the language family Yuto-nahua in the ST: Guarijío (Warihó [waɾiˈho]; 0.9%), Pima (Oob no’ok [ʔo:b noʔok]; 0.4%), Tarahumara (Ralámuli [raˈlamuli]; 89.8%), and Tepehuano del norte (Odami [ʔˈodami]; 8.8%). The Sierra Tarahumara is relatively recent in geomorphic formation (Reyes-Gómez and Núñez-López 2014; Reyes-Cortés and Reyes-Cortés 2014). It initiated with the uplifting along the southwestern portion of the North American Plate. This large, elevated plateau consists of middle to late Tertiary volcanic rocks (Paleogene-Neogene Periods, 54 to 28 Mya) that overlay Mesozoic and older rocks (Cretaceous-Paleogene Periods, 64 to 54 Mya). East of the Continental Divide (CD) are gently sloping foothills and basinand-range features with only internal drainage. West of the CD is densely dissected by barrancas exhibiting the greatest pluvial gradient in Mexico as a result of accelerated erosion from the higher altitude (Cerro Mohinora with elevation of 3,300 m) down to the Pacific coast plains in Sonora and Sinaloa (at elevations 0–300 m) (Schmidt 1973). The SMOc serves as an N-S corridor as well as an obstacle to E-W transit. Along the N-S axis, Nearctic biota have penetrated southward at the higher altitudes while Neotropical biota moved northward at the lower levels, especially on the western slopes. Humans continued to migrate along this axis as evidenced in Paquimé, a city occupied between 1,130 and 1,350 CE by the Mogollon culture with regional trade with Mesoamerica (Whalen and Minnis 2001). As recent as the late nineteenth century, nomadic Apache bands wandered throughout the region until they were subdued by the US government (Orozco 2000). The impenetrability of the SMOc allowed limited E-W movement. A few restricted passes (p. ej., Janos) permitted the exchange of coastal and inner basin goods between inhabitants on either side of the divide (Wilcox et al. 2008). After his shipwreck, Álvar Núñez Cabeza de Vaca returned to the New Spain from Florida on the Gulf of Mexico via Sinaloa on the Pacific slope and became the first European to

Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers,. . .

131

traverse this geographic barrier between 1527 and 1536 CE (Núñez Cabeza de Vaca 1984). Later in that century, other Europeans penetrated the western barrancas and eastern foothills in search of mineral wealth and religious conversion of Indigenous souls. Because of the difficulty the Europeans faced in dominating this barrier, the various Indigenous groups sought to maintain their independence by retreating into the SMOc and distancing themselves from the foreigners. Such strategy afforded them certain autonomy (except near isolated mining and missionary centers) until the mid-1900s when roads and railroad began to cut across mountain chain. With the opportunity to channel N-S migration and the altitudinal gradient afforded by the geomorphic forces, the SMOc is clothed with layers of biological diversity. The flora of the ST of ca. 2,200 vascular plant species is composed of 20% boreal species and 50% tropical species (Bye 2004). Because of the isolation in the barrancas and peaks, the remaining 30% represent endemic species. The plant species are layered in distinct vegetation types that transect from East to West: Oak woodland in the eastern foothills, piñon-juniper woodlands, pine-oak forests, coniferous forests at the highest elevations, descending down the steep western barrancas with tropical oak woodland, and seasonally dry tropical forest and spreading out into the coastal thorn forest.

Rara´muri: Guardians of the Sierra Tarahumara Given that the Rarámuri is the dominant ethnic group in the Sierra Tarahumara, focus is placed on a summary of information recorded among the Tarahumara. Rarámuri is the language spoken by the Tarahumara. Many of the origin myths of the Tarahumara focus on their ancestors who are products of the Great Father (onorúame) and the Devil (diablo) and emerged from previously destroyed worlds onto the mountains where rocks grew and softened into moist earth (Lumholtz 1902). The young Tarahumara couple danced yúmari, planted maize, beans, and potatoes, and slept; upon awakening from their dreams, the Tarahumara population increased. Different elements of the myths reflect the syncretism of Indigenous and Christian cosmovision. Common elements include the following: earth, flora, and fauna as beings; sustenance based upon maize and other cultivated plants; and importance of water and fertility of Earth. Offerings (wirómera) through feasts are important to maintain Great Father’s health and that of the Rarámuri. Plants are manifestations of beings that are respected and are often an entity of ceremonial attention. Offerings may be made during ceremonies to plicate cosmological beings manifested as plants, so they do not provoke harm on people, crops, and domestic animals. Agricultural fields and livestock are periodically blessed with specific herbal mixtures, often containing chuchupate (Ligusticum porteri J. M. Coult. & Rose). Altars and sacred places are sanctified with aromatic smoke produced from burning earth-copal (morewaka) from Bursera resin transformed by insects. Protected pineoak groves with medicinal plants may have one of the various types of peyote transplanted to provide “fuerza” to the herbal remedies. Traditional Tarahumara social organization is based upon the nuclear family who lives in isolated ranchos scattered across the mountain landscape. On the one hand, the

132

R. Bye and E. Linares

arable land is fragmented in small plots. Often those near the principal residence are insufficient to produce all their dietary needs; hence the dispersed plots allow them to produce more food. This mosaic spreads centrifugally the fields across the mountainscape. With the normal growing season attenuated or canceled by extreme weather conditions or other environmental factors (which are usually localized rather than widespread in mountains), the risk of damage is spread across the environmental gradient and, consequently, reduces complete crop loss due to hail, frost, floods, droughts, or diseases and pests. The necessity to attend distant fields for long periods of time in different places results in a settlement pattern with residential mobility (Ruiz-Funes 1995). The cultivation of detached fields is reinforced among the Tarahumara by their bilateral inheritance pattern and the propensity for individuals to marry outside their immediate community. On the other hand, local social networks draw centripetally outlying households together periodically to attend to their community responsibilities such as feasts, reciprocal labor obligations, political actions, and government programs, among others. Many social gatherings revolve around mutually beneficial activities that terminate with the consumption of the esteemed fermented beverage, tesgüino; hence Kennedy (1978) labels the tesgüino complex as the central theme for social organization. Over the centuries, the Tarahumara have resisted religious and governmental efforts to reduce their dispersed rancheria communities into settlements. Today, church centers and municipal and ejido government centers draw Tarahumara for transitory activities such as feasts and political events; nonetheless, they return to their isolated houses to carry out their livelihoods. The nuclear family Tarahumara rancho is integrated into a plastic network composed of other scattered ranchos and coalesced by feasts and traditional authorities. The various feasts and communal work that move from site to site are shared among the members reinforcing social ties. In some cases, they center on territories once associated with colonial churches. After the Mexican Revolution, ejidos were formed; they reflected more the geography of resource exploitation by mines, forest logging, and agricultural production than that of the autonomous Indigenous groups. Hence the current municipal and ejidal partitioning does not mirror the Rarámuri perception of territory and its resources.

Aspects of Interactions and Relationships Between the Rara´muri and Their Plant-world Environment The ecological gradient of the Sierra Tarahumara is compatible with the transhumance character of the Tarahumara (Fig. 1). Some Tarahumara migrate seasonally between separate ranchos in order to take advantage of the pleasant summers in the Sierra and the warmer winters in the barranca (Fig. 2). Others may live permanently in their rancho at a particular spot along the ecological gradient but travel for short periods to other zones so as to gather resources not available locally or from dispersed milpas (to spread out geographically the risks of environmental disasters). Moving across this variable vegetation tapestry, they have access to different floristic

Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers,. . .

133

Fig. 1 The ecological gradient allows a mosaic of agricultural fields in the oak woodland between the pine-oak forest of the Sierra and the subtropical forests of the barranca. (Photo by the authors)

elements that afford a diversity for household consumption. Some Tarahumara dedicate all or part of their life to moving and trading these plant products with people in other zones (Fig. 3). Examples of the interregional trade of useful plants between the Sierra and barranca inhabitants are summarized in Table 1. Of such importance is this interregional trade that a specific native term, norawa, applies to this ancient social relationship. For the past 120 years, 32 species have been documented; 14 from the Sierra and 18 from the barranca. Under the influence of religious, mining, and lumbering settlements in the Sierra and barranca over the centuries, some Tarahumara are dedicated to regional commerce of plants products valued by non-Rarámuri humans (chabochis). With the penetration of the railroad and roads since the twentieth century, 47 species of medicinal plants from the Sierra Tarahumara have entered the urban markets of Ciudad Juarez, Chihuahua City, Delicias, Camargo, and Parral in Chihuahua as well as other cities in Durango, Sonora, and Sinaloa (Bye 1986). The well-being of the Tarahumara involves the use of plants that affect the spiritual and physiological health of the individual, the family, and their communities. The Tarahumara share the Earth with other beings including chabochis, plants, and animals. Today, we all coincide on the fourth plane of a seven-tiered cosmos. Humans are the product of Great Father and the Devil; Rarámuri are migrating to God and chabochis are migrating to the Devil. There are variants of the origin legend of the Tarahumara; some aspects appear to be syncretic with Christian concepts. A pair of Tarahumara children (or three couples) with maize emerged onto the rocky

134

R. Bye and E. Linares

Fig. 2 A Tarahumara transporting dried maize stalks grown in his milpa in the Sierra to his winter residence in the barranca. They will be used to feed his goats as pasturage during the dry season. (Photo by the authors)

fourth plane that was drying from the flood. After planting the kernels, they fell into a dreamy sleep; upon waking, abundant maize harvest permitted the growth of the Tarahumara population. All beings of the creation have souls which need attention and nurturing. In humans, the souls are located in various organs of the body as well as the articulations (Merrill 1988). The souls can travel when one is sleeping or when one is sick. The safe return of one’s souls assures the health of the person. Foods are key to Tarahumara life not only for their corporal existence but also for sustaining the balance and movement of the Earth. Feasts for blessings, petitions, and thankfulness are centered around communal foods (e.g., tónari), music, and dancing to which all attendees are expected to contribute and partake. In the Sierra Tarahumara with a diverse flora along the 2,000-m ecological gradient, a wide variety of food plants are exploited. For instance, the Tarahumara of the rancherias of Bacusínare in the Chínapas barranca, consume 153 species of plants throughout the year (Bye 1982; Mares-Trías and Burgess 1999). The majority of the food plants are seasonal and gathered in the forests (71%), while the remainder are managed by cultivation in the milpa, as spontaneous in the milpa, and by cultivation in gardens (10%, 5%, and 19%, respectively; note that some species are present in more than one forest type). The 147 edible plants of the forests are distributed along the

Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers,. . .

135

Fig. 3 A Tarahumara trader rests along the route of his journey in the Sierra. He is returning to the barranca with baskets (warí) made of Nolina sp. (gurú), brooms of Muhlenbergia porteri (pichira), and dippers of Arbutus sp. (madroño) derived from the Sierra. (Photo by the authors)

gradient from the high pine-oak forest, the tropical oak woodland, and the seasonally dry tropical forest (41%, 28%, and 41%, respectively; note that some species are present in more than one forest type). Since the early reports of the eighteenth century to the present, about 120 quelites (guiribaka) derived from edible, tender leaves and shoots have been important in the Tarahumara diet (Bye 1981). For the families with scarce food supplies at the end of the dry season in April, the appearance of emerging seedlings and sprouting buds is considered “regalo del cielo” (González-Mendoza 2016). From the Sierra to the barranca, 120 types of quelites are consumed raw or cooked (Linares et al. 2016). Tender shoots of wild shrubs (Jacobinia candicans Benth. & Hook. f.) and trees (Eysenhardtia byei Cruz Durán & M. Sousa) add diversity to the dishes in the barranca. Unfolding stinging leaves of nettles (various species of Urtica and Tragia) jutting from the rock crevices in the Sierra are carefully plucked and seared to provide tasty meals. Green sprouts of tender leaves protruding from the barren earth (Asclepias euphorbiifolia Engelm. ex A. Gray, various species of Arracacia and Tauschia) are excised from their crowns on top of thick, deeply rooted perennials of the llanos. Germinating plantlets of many annual weeds of the milpa (various species of Amaranthus, Chenopodium, Bidens, Cosmos, and many more) are carefully picked from among the emerging seedlings of maize, bean, and squash. The tender shoots, juvenile leaves, and flowers of beans and squash are gathered from the expanding plants cultivated in the milpas. As mentioned below, some native and

136

R. Bye and E. Linares

Table 1 Plant products exchanged between the Tarahumara of the Sierra (S) and Barranca (B) reflect the importance of these biocultural resources to the Rarámuri who live along the steep environmental gradient in the mountains of the Sierra Tarahumara, Chihuahua, as documented in 1890 (Lumholtz 1902), 1910 (Edward Palmer reported in Bye 2004), 1930 (Bennett and Zingg 1935), and 2019 (personal observations by Robert Bye, 1971–2019) Scientific name

Common name

S

B

1890

1910

1930

2019

Arbutus spp.

madroño

Dysphania graveolens

epazote de zorrillo

Cydonia oblonga

membrillo

+

Dichondra argentea

oreja de ratón

+

Haplopappus spinulosus subsp. australis

siwátcoli

Ligusticum porteri

chuchupate

+

Malus pumila

manzana

+

Manfreda guttata

amole

Muhlenbergia porteri

pichira

Pediomelum pentaphyllum

contrayerba

+

Pinus cembroides

piñón

+

Prunus persica

durazno

+

Psacalium decompositum

matarique

+

+

Quercus emoryi

bellota

+

+

Tagetes lucida

yerbanís

+

+

Agave spp.

mescal

+

+

Buddliea cordata

tepozán

+

+

Bursera grandifolia

palo mulato

+

+

Caesalpinia platyloba

palo colorado

Capsicum annuum var. glabriusculum

chiltepin

+

+

Citrus spp.

naranja

Coursetia glandulosa

arí

Erythrina flabelliformis

chilicote

Haematoxylon brasiletto

palo brasil

Hintonia latiflora

copalquín

Indigofera suffruticosa

añil

+

Jatropha malacophylla

sangregrado

+

+ +

+

+

+ +

+

+

+

+

+

+ +

+

+ +

+

+

+

+

+

Litsea glaucescens

laurél

Nolina sp.

gurú

+

Nolina matapensis

ruyá

+

Selaginella lepidophylla, S. pilifera

flor de peña

+

Yucca madrensis

flor de yucca

+

+

+

+

+

+

+

+

Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers,. . .

137

exotic quelites have been incorporated into the management practices so as to extend their production and availability. The principal sustenance of the Tarahumara is based upon cultivating the traditional Mesoamerican milpa (Bye and Linares 2018). The broken terrain provides a variety of options, though limited in size and scattered in distribution. Floodwater farming in the flood plains of barranca bottoms, alluvial flats, seep fields where springs exit on the slopes, and the upland valley bottoms may have been the earliest agricultural practices (Pennington 1963). Vegetation clearing along with nutrient and soil accumulation in the swidden agricultural system continues today. Soil and water retention practices with terraces slopes and arroyos with check dams permit temporal patches for cultivation. Irrigation is limited to shelves perched along permanent streams and rivers. The mesas and llanos opened up to agriculture with the arrival of the Spanish missionaries who introduced the plow and domestic animals. Because of limited archeological studies in the area, the antiquity of maize agriculture is tentative. Charred maize and ceramics suggest that maize was present no later than the late Middle Archaic Period (between 3,500 BC – 1,500 BC, and potentially much earlier) (Hard et al. 2006). The earliest charred maize sample in the ST dates to 2,280 ybp. Different ingredients/foods are produced on an annual cycle that provide nutriments for humans as offerings for the feasts. The basic cultivated field of maize, beans, and squash and associated plant and animal are produced in mawechi and ikwechi (place of swidden agriculture) (Figs. 4 and 5) and wasá (land of work,

Fig. 4 A Rarámuri tending his mawechi in the upper dry tropical dry forest; note that plants of Yucca madrensis are tolerated and the stumps of Eysenhardtia byei are sprouting. (Photo by the authors)

138

R. Bye and E. Linares

Fig. 5 Mawechi field in the dry tropical forest of the barranca. (Photo by the authors)

plowed land) (Fig. 6). Gardens or small plots of cultivated plants (both native and exotic cultivated plants) are maintained in solid rich niches near houses, terraced arroyos, and pockets in rocks (Fig. 7). The milpa of the Tarahumara Alta is planted at the end of the dry season (after the snow and frost of the winter are plowed into the soil). The harvest occurs at the end of the rainy season. This cultivation period extends from May to October; nonetheless, it can be shortened to 3 months due to late spring hailstorms and early fall frosts. Throughout this cycle, various products are gathered starting with the first germinating weeds considered as quelites and ending with mature ears of maize, dried bean pods and plump, and ripe squashes. In between, different stages of development and parts of cultivated and synanthropogenic plants are gathered. For instance, Cucurbita provides fresh food from the young stem tips, followed by the male flowers, then the immature fruits, the mature fruits, and finally the mature seeds. In the subtropical barrancas, milpa cultivation also occurs during the dry winter season. Because the steep slopes, the fields are usually on terraces slopes and pockets of the soil along the flats of the arroyos and rivers. Water is the principal limiting factor (in contrast to the cold temperatures in the high mountains); hence many fields are near watercourses from which simple canals draw water to irrigate the fields. Similarities and contrasts exist in the milpa agriculture along the gradients of the mountains. Although a single species (Zea mays L.), maize (sunú) is the keystone species of the Rarámuri foodways; it is differentiated among the 12 races between the Tarahumara Alta and Tarahumara Baja (Bye 2010, 2019). Those of the high sierras have shorter production cycle, with “Apachito” race being the most

Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers,. . .

139

Fig. 6 This wasá agriculture field in an upland valley bottom of the Sierra was constructed by selective deforestation and controlled erosion so that the nutrients and soil particles were captured by the terraces constructed at the base of the slope. Note the seasonal dwelling where Rarámuri farmer resides during the growing and harvest season. (Photo by the authors)

Fig. 7 An alluvial flat in the Sierra that has been terraced for cultivation by expansion and stabilization with a 2-m-high rock wall. (Photo by the authors)

140

R. Bye and E. Linares

precocious to assure earlier harvest before the killing frost. These “Cristalino” races are derived from maize that migrated northward along the corridor and diversified under preferential selection for culinary properties. The maizes of the Tarahumara Baja are related to floury tropical races that migrated northward along the Pacific coastal plains. Three domesticated beans are present. At the higher zones, tekómari (Phaseolus coccineous L.) is valued because of the large seeds and earlier harvest dates. The midlevel fields are dominated by the common muní (P. vulgaris L.) which in the past had over 15 local varieties (most of which have been replaced by ojo de cabra types). In the lower barranca, P. acutifolius A. Gray was common but has been replaced by the introduced yorímuni (Vigna spp.). A similar gradient pattern appears with squash. At high altitudes, Cucurbita ficifolia Bouché resists the cool climate; at middle altitudes, C. pepo L. dominates; and in the lower tropical barrancas, C. argyrosperma Huber is commonly cultivated. Tesgüino (batari, suiki) is the fermented beverage that is essential for Tarahumara life throughout the mountains. Various carbohydrate substrates are transformed by microbes through batch fermentation under strictly controlled cultural and environmental conditions. It is the food medium which dominates the traditional Tarahumara feasts at all social levels. Depending upon their availability and preference, various substrates can be fermented such as maize stalks, various grains, and variety of fruits. The most common form of tesgüino is derived from germinated maize kernels that are sequentially fermented by various bacteria and yeasts (Litzinger 1983; Lappe and Ulloa 1989). Many additives can be added to the fermenting mash to modify the fermenting process in order to ensure a quality beverage (Pennington 1963); some provide supplementary substrates, additional fermenting and stimulatory agents, or substances that may control/depress undesirable microorganisms. The 30+ types of microorganisms that provide the stepwise conversion of the sugars to alcohol and nutriment enhancement are usually found in the pores of the inner walls of the clay fermenting pots. A critical ingredient in the final phase of the batch fermentation process is the basiawi, heads of the brome grass which harbor inoculants of beer yeast (Saccharomyces cerevisiae Meyen ex E. C. Hansen). The brome grass encouraged along the margins of the milpas is considered the best source of basiawi. With the declining availability of ceramic tesgüino pots as well as encouraged stands of Bromus spp., some Tarahumara now use bread yeast purchased in chabochi stores. In other cases, there are differentiations among taxa. For instance, cultivated root crops in the Sierra include introduced and native potatoes (Solanum tuberosum L. and S. fendleri A. Gray ex Torr., respectively) while the barranca have sweet potatoes (Ipomoea batatas L.). Intrageneric differentiation can be seen with edible anthropogenic quelites in the milpas. In the sierras, Amaranthus powellii S. Wat. is a preferred edible green while in the barrancas A. palmeri S. Wats. is a popular edible green. Also, the Tarahumara have a cultural response to the climatic limitation of milpa agriculture, especially in the higher mountains. There, the short growing season can be further curtailed by extreme weather phenomena at the terminals of the agricultural cycle. One process is the dehydration the milpa products at the opportune

Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers,. . .

141

Fig. 8 A Tarahumara milpa in the early phase of its cultivation cycle showing maize, tekómari (both the red- and white-flowered varieties), and various quelites. (Photo by the authors)

moment in order to extend the shelf life of the foods and to assure staple reserves in anticipation of harvest shortages. The products are referred to as “pasados.” The initial agricultural cycle produces are great diversity of spontaneous edible plantlets in the milpa known as quelites (Fig. 8). These are eagerly consumed fresh during a six-week period. Before they become unpalatable, they are gathered and specially processed as “quelite pasado” so as to be dried in such a manner that they are flexible and retain their color, texture, and flavor. During the flowering period of the squash, the male flowers are processed as “flor pasada” while the immature fruits are dried in form of disks. The mature squashes may be stored intact over the winter season for later consumption. In order to extend the storage for more years, they are peeled, spirally cut, and dried as bichikori; not only does this extend the shelf life of the squash fruit, but it also becomes palatable and sweeter. If an early frost threatens the freezing and loss of underdeveloped maize ears, the Tarahumara prepare chacales. The immature maize ears with kernels in the milk stage are blanched in hot water and dried in the sun; the dried, shriveled kernels can be stored on the cob or degrained to be stored in containers. Later, they can be rehydrated and prepared in various dishes. They are especially appreciated as a ceremonial food during Holy Week at the end of the dry period when food shortages are more common. Although the domesticated milpa plants originated in central and southern Mexico, the complementary diversification process continues in the mountains of Chihuahua. In the case of maize, the Sierra Tarahumara is one of four major centers of diversification due to the selection of various races of the cristalino types; this form is preferred

142

R. Bye and E. Linares

from popping, a characteristic appropriate for making pinole, the main form of maize consumption by the Rarámuri (Kato-Yamakake et al. 2009; Bye 2010). Due to the short growing season, various races of maize have been selected to take advantage of the residual ground moisture (some of which consists of snow and frost that was plowed under during the winter) deep in the soil. In contrast to most Mexican maize races, the “Apachito,” “Chihuahua cristalino,” and “Azul” races possess two features that allow the seedlings to grow below the dry soil surface. On the one hand, the monocotyledon elongates more than 20 cm so that the roots of deeply planted seeds can become successfully established while the leaves emerge above the soil surface to photosynthesize. Also, they have an additional rare character of being hydrotropic so that the root apex actively seeks greater moisture concentration. In the case of tekómari (ayocote of central Mexico), the Rarámuri believe that it is strengthened by planting periodically this large bean near the forest. As a result, the productivity of cultivated tekómari is increased by permitting the movement of leaves to maintain a warmer microclimate around the plants as well as permitting the initial flowers (normally sterile) to produce fruits. These features are common in the wild Phaseolus coccineus of the mountain forests and are transferred from the wild plants to those in the cultivated milpa by hummingbirds and bumblebees. Domestication processes are also applied to exotic species. A popular native quelite mustard, rochíwari (Lepidium virginicum L.) that was traditionally planted in fertile soils in the milpa has been displaced by an exotic crucifer. Brassica rapa L. (coles, mekuásare), is a European mustard probably introduced into the area as a contaminant of the Old Word grains such as wheat and later incorporated into the Mesoamerican milpa. It is highly prized as a quelite and was quickly merged into the ethnotaxonomic category of “mekuásare” along with its biological relatives, Thelypodiopsis wootoni (Robinson) Rollins of the Sierra and Th. byei Rollins of the barrancas. This summer annual readily flowers during the long-day season; when planted in the fall under shorter days (Fig. 9) as the Tarahumara have been doing since the eighteenth century, it produces edible enlarged hypocotyl and leafy rosette. This process highlights the Rarámuri perception of phenotypic plasticity under different management regimes and expression of desirable characteristics that can be selected (Bye 1979). Homogeneous experimental cultivation has demonstrated that continuously cultivated mustard populations for generations (in contrast to spontaneously growing weeds of the same region) have suppressed the flowering in favor of the basal leaves by several weeks (McAlvay 2018). Exotic plants contribute to the Tarahumara diet and are diversified along the mountain gradient. In the mountains, the Jesuit and Franciscan missionaries introduced various cultivars of apples, quinces that are associated with houses as well as former habitations. In the barrancas, bananas, citrus, occupy similar sites. Small plots in the transition zone between the Sierra and barranca are often planted with winter wheat. During the last century, private and governmental agencies have introduced commercial cultivars such as different pome fruits (e.g., “Golden Delicious” apples) in the temperate zone and mangos in the tropical zone. Plants are important in the maintenance of Rarámuri health. They are recognized as beings with specific powers. Many have been shown to have bioactive

Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers,. . .

143

Fig. 9 A Tarahumara mother with her child tending the cultivated plot of mekuásare (Brassica rapa) in October. (Photo by the authors)

constituents that exhibit curative properties. The popularity of these herbal remedies has not only promoted interregional trade among Rarámuri settlements along the mountain floristic gradient but also exported selected medicinal plants to state, national, and international markets. Curative rituals are essential in both preventative and curative Rarámuri ethnomedicine (Bye 1985). People, fields, and animals are ceremonially blessed with decoctions of chuchupate (Ligusticum porteri). The central altar of the high sierras supports offerings of brazilwood from the barrancas. Tarahumara shaman employ psychotropic plants to search for and retrieve lost souls of the sick people. Híkuri (peyotes of various species of Cactaceae including Lophophora, Ariocarpus, and Echinocereus) are gathered during pilgrimages to the Chihuahua Desert or in the Sierra. Roots of bakánowa (Scirpus, Coryphanta) are procured and returned after three years to waterfalls along the canyon slopes; during the loan period, they assist the shaman in rescuing sicken people’s souls that are trapped in the watery underworld. In order to not upset the diablo that can cause illness if mistreated, offerings (even if trivial) are made on the side of a ceremonial altar to plicate him. Other illnesses can be caused by spiritual beings as well as by accidents, unhealthy eating practices, unsanitary conditions, and other agents. Nearly 300 plant species are employed in the Sierra Tarahumara for remedial purposes and distributed along the mountain gradient. The majority are found in the Sierra. Of the medicinal flora, 7% are endemic to the region. Ongoing laboratory studies have demonstrated that many have anti-inflammatory, antiseptic,

144

R. Bye and E. Linares

gastroprotective, antinociceptive, and antidiabetic properties. The efficacy of these herbal remedies not only encouraged the interregional trade between the inhabitants of the Sierra and barranca but also encouraged their export to markets beyond the mountains. Chuchupate (Ligusticum porteri), which is restricted to the Chihuahuan pine-oak forests with demonstrated gastroprotective properties, is a locally popular medicine to relieve gastrointestinal ailments (Juárez-Reyes et al. 2014; VelázquezMoyado et al. 2015). Copalquin (Hintonia latiflora Schltdl. & Cham.), which grows in the tropical barrancas where its natural distribution in tropical Mexico reaches its northern limit, has been a traditional treatment for malaria and now been accepted in treating diabetes and other contemporary metabolic ailments; the bark contains bioactive principals with gastroprotective and hypoglycemic properties (Cristians et al. 2013, 2014). These two medicinal plants from the Sierra Tarahumara are among more than 50 species from the mountains that are readily available in the herbal markets of Chihuahua and adjacent states in Mexico and the USA. In fact, they are main active ingredients in a commercial preparation manufactured in Chihuahua City and widely available nationally in natural product stores. However, caution must be exercised in the employment of Indigenous herbal remedies. Chukaka (Packera candidissima (Greene) W. A. Weber & Á. Löve) is an example of where departure from Tarahumara practice can be dangerous (Fragoso-Serrano et al. 2012). This endemic perennial herb of the ST carpets clearings in the pine-oak forest and is readily available. Traditionally, it was drunk sparingly as a tea to alleviate gastric ulcers and applied topically to heal wounds. In recent years, it has gained popularity among chabochis who constantly drink the infusion (as “agua de uso”) to treat diabetes. Because of the presence of hepatoxic pyridazine alkaloids, sustained consumption of the tea can provoke live illness and death. The increase in the rate of death associated with cirrhosis in Chihuahua may be linked to the inappropriate consumption of Tarahumara remedial herb. The mountain gradient offers a diverse array of plants that benefit the Rarámuri culture. Music played on native mouth harp (chapareke, fabricated from the quiote of the sacred Agave wocomahi Gentry) as well as the colonial violin (fabricated from various woods, glues, and resin from the different forests) is essential to the Tarahumara feasts. Hard fibers (gurú and sereke) for basketry are derived from Nolina sp. in the Sierra and Dasylirion sereke Bogler of the barranca, respectively. The degree of resistance to pulling each leaf determines if it is appropriate to be harvested. The various species of trees provide firewood with different qualities for different tasks. Various species of oak (Quercus spp.) provide durable heat for warming the houses and cooking, although only the ash of the tropical species can be used in the nixtamalization process of maize. Tascate (Juniperus spp.) provides a fire with fast heat bursts for popping maize kernels. Pines (Pinus spp.) (especially those struck by lightning) provide resin-rich ocote (chopeke) used for torches to light the trails at night, to start fires, and to concoct and remedy for soothing the throat. The wood’s properties are well recognized for fabricating various structural elements used in building construction, fencing, ladders, and implements. Traditionally, these trees are managed using pruning, pollarding, and coppicing techniques to assure to production of quality prime material and sustainable growth.

Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers,. . .

145

Benefits of Ethnobotanical Research in the Sierra Tarahumara Ethnobotanical studies have an additional obligation to revert the results of the studies to the community who have participated. They are important contributions to biocultural heritage and ecological conservation (Bye and Linares 2018). The conservation of biodiversity through generating patchiness and regenerating plant cover employing traditional management practices is seen as the key to sustainability in the Sierra Tarahumara (LaRochelle and Berkes 2003). Tarahumara women and children are key actors in the conservation of bioculturally important plants in the Sierra (Camou-Guerrero et al. 2008; Wyndham 2010). Ethnobotanical studies have contributed complementary information to bilingual (Rarámuri-Español) publications on food and maize (Mares-Trías and Burgess 1999). Recently, products of the ethnobotanical studies in biodiversity conservation have responded to requests from the Tarahumara authorities to produce bilingual videos that document traditional food ways such as pinole, calabaza, chacales, quelites, chiles, and tesgüino (Appendix 1).

Conclusion The Sierra Tarahumara is the northern section of the Mexico´s longest mountain chain, Sierra Madre Occidental. Over time, the flora and people responded to patterns of flow and blockage that this corridor presents as well as to the endemics that have arisen within the dissected ranges. The Rarámuri people are the dominant Indigenous culture of the region. Their traditional management practices and knowledge of the diverse flora in the mosaic of tropical and temperate vegetation developed such that they have been applied to exotic species as well. About one quarter of the flora of more than 2,000 plant species in the temperate Sierra and tropical barranca has cultural importance. The intimate interactions and intensified relationships of the Tarahumara people with these biocultural resources contribute to the Rarámuri’s resilience when confronting extreme environmental challenges and acculturation pressures.

Appendix 1 Videos related to conservation of agrobiodiversity in the Sierra Tarahumara (produced by CONABIO – https://www.biodiversidad.gob.mx/videos/videosConabio. html) El pinole y el esquiate en la Sierra Tarahumara https://www.youtube.com/watch?v¼j9BumBiSPZ0 El pinole y el esquiate en la Sierra Tarahumara (Rarámuri) https://www.youtube.com/watch?v¼84TqrYajflk Los chacales, anticipando la escasez de maíz en la Sierra Tarahumara https://www.youtube.com/watch?v¼hYSiFBNZz2I

146

R. Bye and E. Linares

Los chacales en la Sierra Tarahumara (Rarámuri) https://www.youtube.com/watch?v¼-QFPXnwWJEw La calabaza y su aprovechamiento en la Sierra Tarahumara https://www.youtube.com/watch?v¼E3cZ3pwSNUM La calabaza en la Sierra Tarahumara (Rarámuri) https://www.youtube.com/watch?v¼zOGuholVvCo El tesgüino, bebida ancestral rarámuri. https://www.youtube.com/watch?v¼UtjvUJX6RsY

References Bennett WC, Zingg RM. The Tarahumara, an Indian tribe of northern Mexico. Chicago: University of Chicago Press; 1935. Boege E. El patrimonio biocultural de los pueblos indígenas de México. México: Instituto Nacional de Antropología e Historia: Comisión Nacional para el Desarrollo de los Pueblos Indígenas; 2008. Bye R. Incipient domestication of mustards in northwest México. Kiva. 1979;44(2-3):237–56. Bye R. Quelites – ethnoecology of edible greens – past, present and future. J Ethnobiol. 1981;1(1): 109–23. Bye R. Comentarios sobre las plantas que comen los tarahumaras del oeste. In: Mares-Trías A. Ralamuli Nu’tugala Go’ame (Comida de los Tarahumaras). Chihuahua, Chih.: D. Burgess; 1982. p. 491–495. Bye R. Medicinal plants of the Tarahumara Indians of Chihuahua, México. In: Tyson RA, Elerick DV, editors. Two Mummies from Chihuahua: a multidisciplinary study. San Diego: San Diego Museum; 1985. p. 77–104. Bye R. Medicinal plants of the Sierra Madre: comparative study of Tarahumara and Mexican market plants. Econ Bot. 1986;40(1):103–24. Bye R. Plantas vasculares de la Sierra Tarahumara listado florístico, etnobotánico e histórico. Universidad Nacional Autónoma de México, Instituto de Biología. Informe final SNIBCONABIO proyecto No. X006. Bases de datos SNIB-CONABIO proyecto No. X006. México D. F. 2004. http://www.conabio.gob.mx/institucion/proyectos/resultados/X006_Caratula_Base %20de%20datos.pdf Bye R. El maíz y su diversidad en la Sierra Tarahumara. In: Burgess D, Mares A, editors. Sunute we’ká e’karúgame newalime ju / con el maíz se pueden hacer muchas cosas. Chihuahua: Published by the authors; 2010. p. 105–11. Bye R, Linares E. La milpa Tarahumara, Patrimonio cultural. In: López-Morales FJ, QuirozMoreno EG, coord. El Patrimonio Cultural Inmaterial: Usos Sustentables del Patrimonio. México, CDMX: Instituto Nacional de Antropología e História; 2018. p. 36–45. Bye, R. Chihuahua y su importancia en la diversificación de maíz. In: Secretaría de Cultura del Estado de Chihuahua. Maíces de vida: Visiones del maíz nativo de Chihuahua. Chihuahua: Gobierno del Estado; 2019. p. 17–25. Camou-Guerrero A, Reyes-García V, Martínez-Ramos M, Casas A. Knowledge and use value of plant species in a Rarámuri community: a gender perspective for conservation. Hum Ecol. 2008;36(2):259–72. Cristians S, Bye R, Navarrete A, Mata R. Gastroprotective effect of Hintonia latiflora and H. standleyana aqueous extracts and compounds. J Ethnopharmacol. 2013;145(2):530–5. https://doi.org/10.1016/j.jep.2012.11.025. Cristians S, Mata R, Bye R. Phenological and geographical influence in the antidiabetic compounds concentration of Hintonia latiflora (Sessé et Mociño ex DC.) Bullock (Rubiaceae) leaves. J Ethnopharmacol. 2014;152:308–13.

Ethnobotany in the Sierra Tarahumara, Mexico: Mountains As Barriers,. . .

147

Fragoso-Serrano M, Figueroa-González G, Castro-Carranza E, Hernández-Solis F, Linares E, Bye R, Pereda-Miranda R. Profiling of alkaloids and eremophilanes in miracle tea (Packera candidissima and P. bellidifolia) products. J Nat Prod. 2012;75(5):890–5. González-Mendoza O. Crecer en la gran tarahumara. In: Jáquez-Rosas MR, coord. Patrimonio Gastronómico de Chihuahua. Chihuahua, Chih.: Instituto Chihuahuense de la Cultura. 2016. p. 123–152. Hard RJ, Roney JR. A massive terraced Village Complex in Chihuahua, Mexico, 3000 years before present. Science. 1998;279:1661–4. Hard RJ, MacWilliams AC, Roney JR, Adams KR, Merrill WL. Early agriculture in Chihuahua, Mexico. In: Staller JE, Tykot RH, Benz BF, editors. Histories of Maize: multidisciplinary approaches to the prehistory, linguistics, biogeography, domestication, and evolution of maize. New York: Elsevier; 2006. p. 471–85. INEGI. Censo de Población y Vivienda. 2010. www.inegi.org.mx/programas/ccpv/2010/default. html#Tabulados Juárez-Reyes K, Angeles-López GE, Rivero-Cruz I, Bye R, Mata R. Antinociceptive activity of Ligusticum porteri preparations and compounds. Pharm Biol. 2014;52(1):14–20. Kato-Yamakake TA, Mapes-Sánchez C, Mera-Ovando LM, Serratos-Hernández JA, Bye-Boettler R. Origen y Diversificación del Maíz – una revisión analítica. México, DF: Instituto de Biología, Universidad Nacional Autónoma de México; 2009. Kennedy JG. Tarahumara of the Sierra Madre – Beer, ecology and social organization. Arlington Heights: AHM Publishing Corporation; 1978. LaRochelle S, Berkes F. Traditional ecological knowledge and practice for edible wild plants: biodiversity use by the Raramuri in the Sierra Tarahumara, Mexico. Int J Sust Dev World Ecol. 2003;10:361–75. Lappe P, Ulloa M. Estudios étnicos, microbioanos y químicos del tesgüino Tarahumara. México, DF: Universidad Nacional Autónoma de México; 1989. Linares E, Bye R, Mera LM. Alimentos de la milpa rarámuri y su entorno. In: Jáquez-Rosas MR, coord. Patrimonio Gastronómico de Chihuahua. Chihuahua, Chih.: Instituto Chihuahuense de la Cultura; 2016. p. 37–50. Litzinger WJ. The ethnobiology of alcoholic beverage production by the Lacandon, Tarahumara and other aboriginal Mesoamerican peoples. Ph. D. Dissertation. Boulder/Ann Arbor: University of Colorado/University Microfilms; 1983. Lumholtz C. Unknown Mexico: a record of five years exploration among the tribes of the Western Sierra Madre; in the tierra caliente of Tepec and Jalisco; and among the Tarascos of Michoacan, vol. 1. New York: Scribners and Sons; 1902. Mares-Trías A, Burgess D. Comida de los Tarahumaras – Ralámuli Nu’tugala Go’ame. México, DF: CONACULTA – Culturas Populares; 1999. McAlvay AC. Domestication, invasion, and ethnobotany of Brassica rapa. Ph.D. Dissertation. Madison: University of Wisconsin; 2018. Merrill WL. Rarámuri Souls: knowledge and social process in Northern Mexico. Washington, DC: Smithsonian Institution Scholarly Press; 1988. Núñez Cabeza de Vaca A. Naufragio. México, DF: Editorial Origen, S.A.; 1984. Orozco V. El conflicto entre Apaches, Rarámuris y mestizos en Chihuahua durante el siglo XIX. In: Hers MA, Mirafuentes JL, Soto MD, Vallebueno M, coord. Nomadas y Sedentarios en el Norte de México, México, DF: Universidad Nacional Autónoma de México; 2000. p. 683–692. Pennington CW. The Tarahumar of Mexico: their environment and material culture. Salt Lake City: University of Utah Press; 1963. Reyes-Cortés IA, Reyes-Cortés M. Geología. In: La Biodiversidad en Chihuahua – Estudio de Estado. México, DF: CONABIO; 2014. p. 23–5. Reyes-Gómez VM, Núñez-López D. Geografía y fisiografía. In: La Biodiversidad en Chihuahua – Estudio de Estado. México, DF: CONABIO; 2014. p. 23–5. Ruiz-Funes R. Caminando por las cuevas de gigantes: patrón de asentamiento y distribución espacial en la Sierra Tarahumara. Un estudio etnoarqueológico. Tesis Licenciatura Arqueología, ENAH; 1995.

148

R. Bye and E. Linares

Schmidt RH. A geographic survey of Chihuahua. El Paso: University of Texas at El Paso. [Southwestern studies; Monography number 37]; 1973. Velázquez-Moyado JA, Martínez-González A, Linares E, Bye R, Mata R, Navarrete A. Gastroprotective effect of diligustilide isolated from roots of Ligusticum porteri Coulter & Rose (Apiaceae) on ethanol-induced lesions in rats. J Ethnopharmacol. 2015;174(4):403–9. Whalen ME, Minnis PE. Casas Grandes and its Hinterland: prehistoric regional organization in Northwest Mexico. Tucson: University of Arizona Press; 2001. Wilcox DR, Wiegand PC, Scott J, Howard JB. Ancient cultural interplay of the American Southwest in the Mexican Northwest. J Southwest. 2008;50(2):103–206. Wyndham F. Environments of learning: Rarámuri children’s plant knowledge and experience of schooling, family, and landscapes in the Sierra Tarahumara, Mexico, Hum Ecol 2010;38(1):87-99.

Rara´muri Ethnobotany: Peasant Subsistence and Biodiversity Conservation at Local Scale Andre´s Camou-Guerrero, Juan Vega, María Teresa Guerrero-Olivares, and Alejandro Casas

Abstract

This study documents and analyzes from ethnoecological approaches the Rarámuri forms of use, management, and conservation of plant resources and forests. It was part of a participatory work that started from an initiative of the Rarámuri community of Cuiteco, Chihuahua, Northern Mexico, in connection with the local NGO CONTEC A.C. We analyzed the general role plants have had in people’s subsistence and the species with higher potential for improving their lives while conserving both plant species and ecosystems. The research started with meetings in the General Assembly of the community, and then a series of workshops directed to: (1) recognize and map the main environmental units within the territory of the community; (2) methods for sampling vegetation to obtain information about distribution and abundance of plant species; (3) methods for collecting plants, preparing herbarium specimens, and documenting the Rarámuri knowledge on plant uses and ecology, and (4) methods for evaluating amounts of plants used. After a period of training, people organized groups

Juan Vega: deceased. A. Camou-Guerrero (*) Escuela Nacional de Estudios Superiores (ENES), Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] J. Vega · M. T. Guerrero-Olivares Consultoría Técnica Comunitaria A.C., Chihuahua, Chihuahua, México A. Casas Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_6

149

150

A. Camou-Guerrero et al.

of volunteers, including women and men who participated in different activities together with the researchers. We firstly analyzed people’s spatial view of their territory, the environmental units they recognize, their role as sources of specific plant resources, and conditions for sustainable forms of using forests. Local people identified the following environmental units: (1) the Sierra or repárabo, (2) the slops or gallena, (3) the lajero or í'pichí, (4) the crop lands (agrachí), including the milpa or mawechi and homegardens (kumerachi), (5) the rivers or gomichi, and (6) the plains or valleys (eepó). Our study reports a general floristic inventory of the ejido, which is composed by 356 plant species, 226 (63%) of which are used in some way by local people. A total of 14 general use categories were identified: medicine (116 species, 31.1% of all plant species used), fodder (89 species, 23.9%), food (56 species, 15.0%), domestic goods (35 species, 9.4%), firewood (31 species, 8.3%), building materials (21 species, 5.6%), ornamental (10 species, 2.7%), ritual (5 species, 1.3%), and tannins (3 species), among others. For the main-use categories, we documented 12 subcategories and 42 specific uses of plants. The most extended vegetation type is pine forest, which also provides the highest number of plant resources. However, other ecosystems are sources of specific resources. The study reveals the high importance of plants in the multiple use of resources of the Rarámuri subsistence, outstandingly, the high volume of species used as firewood, food, medicine, and fodder. The highest economic potential was identified among some medicinal plants like Cosmos pringlei (babiza), Zornia reticulate (hierba de la víbora), Ligusticum porteri (chuchupate), and Psacalium decompositum (matarike). These resources have demand in national and international markets and may be profitable. However, their sustainable use require ecological studies for designing the bases of sustainable harvesting, as well as the construction of local institutions to guarantee equity in work and benefit sharing.

Introduction The idea that traditional rural communities of Mexico practice a pattern of multiple use of natural resources and ecosystems was first documented and proposed by Toledo et al. (1976). Later, such pattern has been analyzed by several scholars like Bye (1976), Zizumbo-Villarreal and Colunga-Garcia Marín (1982), Alcorn (1984), Caballero and Mapes (1985), Toledo (1990), and Casas et al. (1994, 2001), among others, in different cultural and environmental contexts. Currently, it is widely recognized that the multiple use strategy allows rural communities to maximize and optimize the use of different environments, obtaining from them a high diversity of products, frequently associated to specific local ecosystems. It is also recognized that, in the case of biotic resources, their multiple use integrates management practices at different spatial and temporal scales, as well as different levels of organization. At the level of individuals or populations, several management techniques have been described, for instance, gathering, tolerance, enhancement, and

Rara´muri Ethnobotany: Peasant Subsistence and Biodiversity Conservation. . .

151

cultivation of plant species of high cultural value (Bye 1993; Casas et al. 1996; Casas et al. 1997). At the levels of biotic communities and landscapes, several techniques have been identified to be practiced for managing vegetation, which may drive ecological processes like ecological succession (Berkes and Davidson-Hunt 2006), or those looking for maintaining high levels of diversity of species or genetic diversity of a given species (Casas et al. 1997), as well as interactions among the species composing the managed communities (Casas et al. 2016). Ethnoecology is nowadays a conceptual framework useful to analyze the traditional knowledge and its relationship with the multiple use of resources and ecosystems. According to Barrera-Bassols and Toledo (2005), ethnoecology is an interdisciplinary approach to understanding the perception, intervention, and transformation of nature by human societies. These processes may be analyzed based on a complex system of appropriation of nature by humans, which include the kosmos (the system of beliefs and worldview that allows analyzing the conception of the world by people of a human group), the corpus (the cognitive and knowledge system of that human group), and the praxis (the use and management practices and techniques of elements and systems of nature). A basic supposition in this general framework is that Indigenous rural societies conform socio–ecological systems highly resilient (Alcorn and Toledo 1998; Toledo et al. 2003), compared with other communities, either rural or urban, strongly assembled to the global capitalist world. In the Indigenous societies, the social structures, knowledge, and technological practices converge in a model of integral management of ecosystems, which is continually adjusted to changing conditions of culture and environment and based on the also continuous development of experiences and knowledge. This management form corresponds to what Berkes et al. (2000) have called adaptive management, which involves an iterative process of reviewing knowledge and interventions to improve management of resources and systems toward sustainability. The relevance of the cognitive and practical systems in traditional societies has motivated numerous studies documenting the patterns of using natural resources in Indigenous/campesino communities (see Berkes et al. 1998). Such studies have been part of a strategy to generate innovative management techniques, recognizing that the traditional knowledge is crucial for reaching goals of conservation and sustainability linked to use of natural resources (Berkes and Folke 1994; Hetch and Posey 1989). Documenting the complex kosmos-corpus-praxis of Indigenous cultures may undoubtedly be a key work to find tested ways of sustainable forms of resource management, but, as some authors suggest, such possibility may and should be reinforced by linking the traditional knowledge systems with the system of scientific and technological elements provided by sciences like ecology, agronomy, and sustainability science, among others (Pérez-Negrón and Casas 2007). In this context, our study focuses on documenting and analyzing ethnoecological aspects of the Rarámuri culture, particularly in relation to the use, management, and conservation of plant resources. We conducted this study as part of a participatory work that started from an initiative of the Rarámuri community of Cuiteco in connection with CONTEC A.C. (a local NGO). In the early twenty-first century, people from Cuiteco decided

152

A. Camou-Guerrero et al.

to stop exploiting timber from their forests, an activity that generated negative ecological and socio-economic impacts, and agreed to look for possible economic alternatives with nontimber forest products. For the first time, the ejido was under the control of Indigenous leaders, and the Indigenous people of the local community assumed the decision to safeguard and take care of plant resources (both timber and nontimber) and to stop logging. Therefore, the local Rarámuri people and CONTEC asked collaboration of our research team at the Institute of Research on Ecosystems and Sustainability (IIES for its acronym in Spanish), UNAM, to conduct ethnoecological studies focused on developing the basis of proposals of sustainable forest management. The main purpose of our study was to document knowledge and practices traditionally carried out by the Rarámuri of Cuiteco with plants. We analyzed the general role plants have had in people subsistence and the potential plants may have for the purposes of local people to improve their lives while conserving their plant resources and ecosystems. We focused our attention on people’s spatial view of their territory, the environmental units they recognize in the landscape, and their role as sources of specific plant resources as a basis for planning activities of sustainable use of their forests.

The Sierra Tarahumara The Western Sierra Madre of Mexico is 1,250 km long, running from northwest to southeast through the states of Sonora, Chihuahua, Durango, Nayarit, and Jalisco. In the state of Chihuahua, the Sierra Madre is called Sierra Tarahumara (ST), because the most numerous Indigenous peoples inhabiting the area are the Tarahumara, called to themselves Rarámuri. The ST covers approximately 20% of the territory of the state of Chihuahua (Fig. 1). It is a region of high biological importance, recognized as one of the 152 priority terrestrial regions of Mexico for conserving biological diversity, because of its biogeographic importance and the high variation of ecosystems occurring there (Arriaga et al. 2000). In addition, it is one of the 625 Indigenous territories of Mexico (Boege 2006), which constitutes the main reservoir of culture and knowledge of the Rarámuri, Odami, O’óba,

Fig. 1 Localization of the Ejido Cuiteco, municipality of Urique in the Sierra Tarahumara region of the state of Chihuahua, Mexico

Rara´muri Ethnobotany: Peasant Subsistence and Biodiversity Conservation. . .

153

and Warijios peoples. From this perspective, the ST is an area of great biocultural importance. Despite the important environmental and cultural heritage that the region harbors, the Sierra Tarahumara, as other Indigenous territories of Mexico, has been fragmented and lives a complex reality. The ST has been an economic enclave for the industrial extraction of raw matters, mainly mining and timber, which are activities of high environmental impact (Guerrero et al. 2000). But contrastingly, the Indigenous communities of the area have been included within the catalogue of the 250 microregions of priority attention in Mexico because of their poverty (SEDESOL 2001). The circumstances referred to above suggest that the governmental plans for economic and social development implemented in the region not necessarily have contributed to the well-being of Indigenous people nor established the bases for a social development based on equity (see Sariego 2002). For the contrary, the main social trend that can be observed in the area is the increase of poverty of local people and a progressively higher environmental deterioration, exacerbated by clandestine logging (which is expressed in massive felling of trees) and the presence of organized crime. The arising of Indigenous and campesino movements in defense of forests of the ST (Guerrero et al. 2000) reflects, as Martínez-Alier (2009) said, that the inequitable access to the benefits derived from natural resources and ecosystem services increases conflicts related to the management and administration of the resources on a basis of high social, economic, and ecological cost. Efforts to describe and understand the complex relationships between society and nature among the Rarámuri started in 1902, with the early ethnographic studies by Carl Lumholtz (1986), and some decades later by the ethnobiological studies by Pennington (1963) and Bye (1976), and more recently those by DeBano et al. (1995) and Camou-Guerrero et al. (2008). The ethnobotanical studies by Felger and Wilson (1995) documented that, from the richness of the flora of the Sierra Madre, nearly 1000 wild plant species are used by the Indigenous groups that live in the region. Bye (1981, 1995) estimated that only the Rarámuri traditional medicine makes use of nearly 300 species of plants, whereas nearly 120 species are wild edible plants. The pattern of multiple use of natural resources can be seen not only in the diversity of species and the different forms of the species that are used by the Rarámuri (Bye 1976, 1986, 1995; Pennington 1963), but also in the variety of forms of ecosystems management, for instance, the cultivation system called wika which is carried out on pronounced slopes, as well as the different strategies for harvesting wild plants (Bye 1981; LaRoche and Berkes 2003), and in the complex forms of differential use of environments occurring in the landscapes of their territories. The Rarámuri people recognize areas with different capacities and requirements for practicing agriculture, livestock raising, forest uses, fruticulture, hunting, fishing, and gathering specific forest products. In a similar context, the management system has been described based on the use of fire, through controlled firing of forests to create propitious spaces for agriculture and subsequent processes of ecological succession that favor the availability of specific resources (Berkes and Davidson-Hunt 2006; Davidson-Hunt 2003).

154

A. Camou-Guerrero et al.

The Ejido Cuiteco Cuiteco is an ejido (a form of collective land tenure in Mexico) mostly composed by Indigenous households. It belongs to the municipality of Urique, in Chihuahua, Mexico, within the ST (Fig. 1). Its territory is 8,561 ha extent and has elevations from 1,700 a 2,575 m, with a climate temperate subhumid with the rainy season in the summer (García 1981). The predominant vegetation type is pine-oak forest, but other ecosystems occur in the area. The main economic activities are agriculture and cultivation of perennial plants in homegardens; they are also seasonally employed as agricultural jornaleros in the states of Sinaloa, Chihuahua, and Sonora. Other common activities are small-scale livestock (mainly goats and cattle) and jobs like brickwork, carpentry, and handcrafting. People live in 90 households composing the ejido: 75 of them are Rarámuri and 15 Mestizo. The total number of inhabitants is 535, 455 of them being Rarámuri and 80 Mestizo. People of Cuiteco have a traditional Indigenous organization for making decisions, based on traditional principles closely related with religious festivities, ceremonies carried out throughout the year. The main responsibility in the organization structure falls in the Indigenous Governor and the Fiesteros (the feast people), through whom the traditional rites and the nawésare (council) are maintained for ensuring the best coexistence among households and people of the community. In addition, the ejido land tenure regime has its own regulations established in the Ley Agraria (Agrarian Law), which defines the charges, functions, and directions. The ejido authorities are integrated by the Comisariado Ejidal and the Consejo de Vigilancia, but the maximum instance of decisions is the General Assembly formed by all men and women that are ejidatarios, all of them having the right of voice and vote in the decision-making processes.

Research Methods This study started with the interest of people from Cuiteco to inventory their nontimber forest products and to design strategies for their sustainable use. They expressed to the local NGO CONTEC their decision to stop deforestation and timber exploitation, substituting such process by other forms of using the forest while conserving and recovering the damage caused by the external companies. This study is the first step for accomplishing that purpose. It was directed to diagnosing the plant resources they have, where these resources occur, and how much is available for planning viable sustainable forms for using their forests. The research initiated with several meetings in the General Assembly of the community, where they expressed their interest to have our collaboration as academic group, and we expressed our disposition to collaborate with them, based solely on an academic interest (nonprofitable). We also expressed that their own knowledge on their territory and plant resources would be the basis for designing any further plan, enriched with the pertinent ecological academic knowledge. Then, we together with the local authorities organized a series of workshops to (1) recognize and

Rara´muri Ethnobotany: Peasant Subsistence and Biodiversity Conservation. . .

155

map the main environmental units within the territory of the community; (2) impart methods for sampling vegetation which would allow having information about distribution and abundance of plant species; (3) provide methods for collecting plants and preparing voucher specimens, as well as a collective strategy for documenting what people know about each species, including their Rarámuri and/or Spanish names, attributes, properties and uses, and general ecological information about their distribution, abundance, animals interacting with them, and forms of reproduction, among other topics; (4) contribute methods for evaluating amounts of plants used as food, medicine, fuel, and fodder, among other purposes; and (5) organize and plan a schedule of work and meetings for reviewing advances and have reflections on them (Fig. 2). After a period of training, people organized groups of volunteers, including women and men, to participate in the different activities of the project, forming work teams together with the researchers.

Plant Species Used in Cuiteco We found in Cuiteco a general floristic inventory composed by 356 plant species, 226 (63%) of which are used in some way. The details of this inventory can be consulted in Camou-Guerrero et al. (2008). A total of 14 general categories were

Fig. 2 Aspects of the participatory collaboration between local people and researchers: (a) General Assembly of Cuiteco; (b) workshop for constructing the territory of the community and its environmental units; (c) workshop for ethnobotanical collecting of plants and ecological sampling; and (d) workshop of reflections about the information documented in the diagnosis on local nontimber forest products. (Photos: Andrés Camou-Guerrero)

156

A. Camou-Guerrero et al.

identified: medicine (116 species, 31.1% of all useful plants); fodder (89 species, 23.9%); food (56 species, 15.0%); domestic goods (35 species, 9.4%); firewood (31 species, 8.3%); building materials (21 species, 5.6%); ornamental (10 species, 2.7); ritual (5 species, 1.3%); tannins (3 species, 0.8%); colorants (2 species, 0.5%); poisons (2 species, 0.5%); glue (1 species, 0.3%); resins (1 species, 0.3%); and fibers (1 species, 0.3%). In turn, for the main-use categories we documented 12 subcategories and 42 specific uses of plants (Camou-Guerrero et al. 2008).

Environmental Units The Rarámuri people of Cuiteco recognized six environmental units in their territory, which are defined based on elevation, topography, soil type, management, vegetation types, and particular plant associations. The units are the following: 1) the Sierra or repárabo, which are areas of highlands in the hills surrounding the town; 2) the slopes of hills or gallena; 3) the lajero or i’pichí, which are stony areas; 4) the crop lands (agrachí) including the milpa or mawechi and homegardens (kumerachi); 5) the rivers or gomichi; and 6) the plains or valleys (eepó) (Fig. 3a). According to people, the most conspicuous units are steep slopes, the lajero, and the Sierra, in the higher elevations of the territory dominated by pine forest. Less conspicuous are the croplands and plains.

Fig. 3 Participatory map of the territory of the Ejido Cuiteco, elaborated by local people, showing: (a) environmental units; (b) vegetation types; and (c) types of land use

Rara´muri Ethnobotany: Peasant Subsistence and Biodiversity Conservation. . .

157

Within these environmental units, people recognized at least eight types of plant cover (Fig. 3b). One is the pine forest (PF), locally called ojkorare, which has the most extensive coverage in the territory (3,314 ha, 38.7% of the total surface) (Table 1), dominated by Pinus arizonica, Pinus chihuahuana, Pinus ayacahuite, Quercus sideroxyla, and Juniperus deppeana. All tree species (16 spp.) of this vegetation type are used by people of Cuiteco (Table 2). We in addition identified 61 species of herbaceous plants, 36.1% used in one or more ways by people of Cuiteco (Table 2). The useful species recorded in this vegetation type allow covering needs for eleven use categories, most of them (22 species) for medicinal purposes, followed by species used to manufacturing utensils (17), as fuel (17), fodder (16), food (15), and construction (12) (Table 3). People also recognize the oak forest (OF), locally called rojarare which covers 2,625 ha (nearly 30.7% of the total surface of the territory of Cuiteco) (Table 1). In this forest, we recorded the highest richness of trees and shrubs, with 17 species (Table 2), dominant being Quercus sideroxyla, Quercus scytophylla, Quercus arizonica, Quercus mcvaughii, Pinus arizonica, Pinus ayacahuite, and Pinus chihuahuana. All the tree species recorded were reported being used for different purposes. In addition, we recorded 53 species of herbaceous plants, 39.6% of them being useful (Table 2). The useful species recorded allow satisfying needs of ten use categories, mostly medicinal (26 species) elaboration of utensils (15 species), fuel (15 species), fodder (15 species), food (13 species), and construction (12 species) (Table 3).

Table 1 Local Rarámuri knowledge on vegetation types occurring in the ejido Cuiteco: (l) lajero, (p) plain, (s) slope, (r) river, and (si) sierra Vegetation type Pine forest Oak forest Crop field/fallow Juniperus forest Riparian vegetation

Rarámuri name Ojkorare Rojárare I’chiuare Aurirare –

Environmental unit si, s p, l, s p, s, si s, l, si r

Surface (ha) 3,314 2,625 480 349 30

Table 2 Richness and proportion of useful plant species per vegetation type. Oak Forest (OF); Pine Forest (PF); Riparian Vegetation (RV); Juniperus Forest or tascatal (JF); milpa (M); and fallow agricultural fields (F) Vegetation type OF PF RV JF M F

Trees and shrubs (species richness) 17 16 9 6 – –

Useful species (%) 100 100 100 100 – –

Herbaceous (species richness) 53 61 19 51 43 48

Useful species (%) 21 (39.6) 22 (36.1) 9 (47.4) 23 (45.1) 19 (44.2) 22 (45.8)

M

F

JF

RV

PF

VT OF

Medicinal 26 (28.9) 22 (24.4) 5 (5.6) 17 (18.9) 11 (12.2) 9 (10.0)

Food 13 (23.2) 15 (26.8) 7 (12.5) 7 (12.5) 5 (8.9) 9 (16.1)

Fodder 15 (19.5) 16 (20.8) 9 (11.7) 16 (20.8) 13 (16.9) 8 (10.4) –



Construction 12 (36.4) 12 (36.4) 4 (12.1) 5 (15.2)





Utensils 15 (32.6) 17 (37.0) 7 (15.2) 7 (15.2)

Ornament 2 (16.7) 3 (25.0) 2 (16.7) 3 (25.0) 1 (8.3) 1 (8.3) –



Pigments – – 1 (33.3) 1 (33.3) 1 (33.3) – 1 (20.0)

Ritual 1 (20.0) 1 (20.0) 2 (40.0) – –





Fuel 15 (34.1) 17 (38.6) 6 (13.6) 6 (13.6)





Tannins 2 (40.0) 3 (60.0) – – – –





Resins 2 (50.0) 1 (25.0) – – 1 (25.0)

28

30

63

43

108

Total 103

Table 3 Number of useful species per vegetation type (VT): Oak Forest (OF), Pine Forest (PF), Riparian Vegetation (RV), Juniperus Forest (JF), Fallow Agricultural Fields (F), and Milpa (M). The percentage is shown in parentheses

158 A. Camou-Guerrero et al.

Rara´muri Ethnobotany: Peasant Subsistence and Biodiversity Conservation. . .

159

The “tascatal” or Juniperus Forest (JF) is called aurirare in Rarámuri and covers 394 ha (4.6% of the surface of Cuiteco), dominated by Juniperus deppeana (Table 1). We recorded 6 species of trees and shrubs, all of them used in some way (Table 2). Richness of herbaceous species is high (51 species), 45.1% of them being useful (Table 2), covering needs of eight use categories, mainly medicinal (17 species) and fodder (16 species) (Table 3). The riparian vegetation (RV) has lower cover than the forests mentioned above (nearly 30 ha, 0.3% of the Surface of Cuiteco’s territory) (Table 1), where we identified the Cupressus forest (wa’arare), the Abies forest (materare), and the Alnus forest (ropgarare). The RV harbors nine species of trees and shrubs, Cupressus arizonica, Abies durangensis, and Alnus acuminata being the dominant species. All perennial and 47.4% of herbaceous species are useful (Table 2) to satisfy needs of nine use categories, mainly food (7 species), elaboration of utensils (7 species), fuel (6 species), and medicinal (5 species) (Table 3). Finally, crop field areas (agrachí) and fallows (i’chiuare) represent nearly 5.6% of the surface of the territory (480 ha) (Table 1). We recorded 43 species of herbaceous plants in the milpas; nearly 44.2% of them were recorded as useful, while in fallows we recorded 48 species, 45.8% of them useful (Table 2). Useful plant species recorded in milpas allow satisfying needs of five general use categories mainly medicinal (9 species), food (9 species), and fodder (8 species). Useful species from fallows satisfy needs of four categories, mainly fodder (13 species), medicinal (11 species), and food.

Rara´muri Economy and Plant Resources The conceptual framework of the local Rarámuri economy, shown in Fig. 4, considers household as the basic unit, in which the activities are organized through a labor division considering age and gender (Camou-Guerrero et al. 2008; Vatant 1990). The main subsystems of activities in which the subsistence is based on are agricultural, livestock, homegardens (kumerachi), forest (obtaining products for domestic and commercial use), the temporary salaried jobs within the community, and migration (CONTEC 2008). The activities related to the productive subsystems indicated in Fig. 4 are carried out in the immediate social environment (according to family relationships and cultural traditions) and the extended social environment (the organizational ejido structure). The main authority in the immediate social environment is the traditional Siríame (the Indigenous Governor), while the governing structure of the ejido is regulated by the Ley Agraria, which does not take cultural traditions into account, but is based on modern conceptions of natural resource management and social coexistence. Decisions related to natural resource management are made at the ejido level, as well as those related to the commercial use of timber resources (CONTEC 2008).

160

A. Camou-Guerrero et al.

Fig. 4 Conceptual framework of the Rarámuri economy. (Adapted from CONTEC (2008))

Timber Commercial Extraction In general, the forests of the Sierra Tarahumara region are exploited not only by the ejidos, but also by external actors such as forestry engineers, sawmill owners, and carriers who obtain the concession and agreements with communities to do it. The benefits of forest exploitation for the communities are given through a system of profit sharing. According to Camou-Guerrero (2008), in the early twenty-first century, the profit sharing in Cuiteco was on average $1,545.00 Mexican pesos (141.87 US dollars) per ejidatario per year, an income nearly ten times lower compared with the official minimum salary for the region (nearly $4.3 dollars per day, approximately $1,575 dollars per year in the municipality of Urique). This situation is an overexploitation, which may be even worst in other communities where 106.8 dollars per ejidatario per year has been reported (Guerrero et al. 2002).

Rara´muri Ethnobotany: Peasant Subsistence and Biodiversity Conservation. . .

161

In other words, the profit sharing from timber exploitation was not profitable neither in Cuiteco nor in any other community of the region. Another form of access to benefits of forest use is through employment of the local inhabitants, which generates additional incomes. The main activities are felling of trees, cleaning (removing branches), log skidding, and charging. Transportation and other activities are carried out by other external persons. However, according to interviews with local people, the employment involved very few local people. According to Guerrero et al. (2002), in a typical ejido of the ST, the logging activities only generate jobs for 10% of the community members. According to Vatant (1990), forest management directed to industrial purposes has included salaried work that enhanced the monetarization of the ejidos’ economy. But in Cuiteco, the incomes were extremely low, and people therefore decided stopping this type of forest (and people) exploitation. Paradoxically, such unprofitable activities help to explain the permanence and relevance of traditional Rarámuri activities in local people subsistence, since the diversified production allows households higher self-sufficiency than specialized activities. A similar situation was documented by Ortiz-Avila and Masera (2008) among rural people in the state of Michoacán. According to these authors, households practicing diversified agrosilvopastoral systems had more food self-sufficiency than households whose subsistence depended exclusively on forest plantations. Although it is true that a tension is observed between subsistence activities and commercial forest management, it is necessary to reflect on the results presented here with current forest management and conservation policies in the state of Chihuahua that integrate local perspectives.

Firewood Among the species used as firewood, the most relevant for their cultural importance are Quercus crassifolia (u’turi), Quercus viminea (achíchuri), Quercus arizonica (mapaka), or Juniperus deppeana (auarí). Among the main subsistence activities carried out by the community of Cuiteco, obtaining of fuel is one of those activities that demand a significant labor hand; 91 jornales1 per household are required per year (nearly half, 49.7%, the time invested in subsistence activities). It also represents the highest volume of plant biomass used from forest products (nearly 5.0 ton per household per year), with a return rate2 of 0.34 (Table 4). However, and even when fuelwood is the main fuel used by the Rarámuri households of Cuiteco (CONTEC 2006), its consumption is lower than in other rural regions of Mexico. For instance, Farfán et al. (2007) reported that in a community of the Monarch Butterfly Biosphere Reserve, the Mazahua households use on average 9.4 tons of fuelwood per year.

1

The return rate is the profit obtained from production ($/kg) divided by production costs (total costs plus opportunity cost). 2 An 8-h work day.

162

A. Camou-Guerrero et al.

Table 4 Labor force (LF), Consumption and Annual Returning Rate (ARR), for the main systems providing plant resources in the Ejido Cuiteco Use category Fuel Maize cultivation Edible Construction Medicinal Total

LF (jornal/household/ year) 91.0 73.0

% 49.7 39.9

Consumption (kg/household/year) 5,000.5 515.5

% 45.4 4.7

ARR 0.34 0.15

9.0 7.0 3.0 183.0

4.9 3.8 1.6 100

299.3 5,171.9 17.9 11,005.1

2.7 47.0 0.2 100

1.5 – 2.4 –

Table 5 Percentage of households’ production that are edible plant species Crop in fields Maize Beans Wheat Potato Oat Peas Lentils Total

kg/year 515.5

Homegardens Apples

kg/year 168.2

102.4 44.2 65.9 36.0 8.5 4.9 777.4 (56.9%)

Peaches Apricots Plums Quince Fig Walnuts

80.7 9.7 11.0 5.8 0.1 13.1 288.6 (21.1%)

Noncrops Quelites (greens) Flowers/roots Wild fruit Condiments – – – –

kg/year 251.9 7.1 37.1 3.5 – – – 299.6 (21.9%)

Agriculture In the period of the study (agricultural cycle 2004–2005), seven crops were registered: corn, beans, peas, potatoes, wheat, lentils, and oat. As shown in Table 5, maize is the most important crop due to its production volume (369.2 kg on average per family), and it should be noted that in all crops there was a production deficit, and therefore the households must buy part of their annual food to cover the basic requirements of diet. Corn production demands 73 jornales per family per year (for 1.3 ha on average of corn cultivation), which represents 39.9% of the annual time invested per family in the productive activities analyzed but has the lowest return rate (0.15) (Table 4). Despite its low profitability, the relevance of agriculture lies in its role as a pillar of the Rarámuri culture and a factor of social cohesion (Pennington 1963; Vatant 1990).

Noncrop Edible Plants In a sample of families from Cuiteco (20% of the total), 27 species were documented as the main noncultivated resources used as food. Of these, 37% are used as quelites,

Rara´muri Ethnobotany: Peasant Subsistence and Biodiversity Conservation. . .

163

Table 6 Crop productivity (2004–2005) Cultivo Maize Bean Wheat Potato Oat Pea Lentil

Average production (Kg) 369.2 78.5 42.9 42.0 34.1 6.5 2.2

Max. production (Kg) 1000.0 600.0 400.0 400.0 425.0 60.0 20.0

Min. production (Kg) 70.0 10.0 25.0 25.0 100.0 5.0 7.0

s.d. 277.7 121.2 100.6 92.9 101.9 13.7 5.6

Notes: s.d. standard deviation

30% as fruits, and 26% as condiments, and flowers and / or roots of 7% species are consumed. Noncrop edible plant species of great cultural importance gathered for direct consumption or commercialization include Opuntia sp. (wirá), Brassica campestris (Quelite mostaza), Prionoscidium madrense (sarabiki), and orégano de la sierra,3 among others. We estimated that these species have a return rate of 1.5, which is higher than those of fuelwood and maize. Considering all edible plant species, we estimated that people invest nearly 4.9% of the time invested in all subsistence activities (Table 4). Comparing with maize production, extraction of fuelwood and medicinal plants, the volume of noncrop edible plants represents 5.1% of the total volume used. However, compared with the total volume of cultivated species, it represents 21.9% (Table 6). The consumption of noncrop plant species in the ejido is high. We recorded that nearly 3.9 tons of Brassica campestris and 0.5 tons of Chenopodium berlandieri are consumed per year in Cuiteco, which is close to amounts estimated by Farfán et al. (2007) among the Mazahua of the Monarch Butterfly Biosphere Reserve (4.3 and 0.7 tons/year, respectively).

Medicinal Ethnobotanical work in the sample of households studied allowed identifying 57 plant species with medicinal use. Diarrhea is the disease for which the largest number of medicinal species are used (20) followed by cough (19), flu and colds (18 species) (Fig. 5). The diseases with the fewest number of medicinal species were malaria (2) and measles (1). Among the species of medicinal use of greater cultural importance are Cosmos pringlei (babiza), Zornia reticulate (Hierba de la víbora), Ligusticum porteri (chuchupate), and Psacalium decompositum (matarike).

3

Three main genera have been recorded in ST: Hedeoma, Lippia, and Monarda. The species that has been documented in the Alta Tarahunara area is M. austromontana (some think it is a variety or subspecies of M. citriodora). In the Baja Tarahumara, there are several species of Hedeoma, and the species that is commercialized in the municipality of Urique seems to be a new species (Bye, R., Pers. Comm.).

164

A. Camou-Guerrero et al.

Fig. 5 Number of medicinal plant species used for each disease (n ¼ 173). Repeating species with more than one use are included

According to Bye (1986), nearly 30 plant species used as medicine by the Rarámuri are used and commercialized in the urban centers of the state of Chihuahua and have high therapeutic and commercial potential (Bye 1995). We identified two species of medicinal plants that are commercialized at local level in Cuiteco, which have a high return rate (2.4), the lowest investment of time for their gathering (1.6% of the total), and the lowest volume of use, 0.2% (Table 4). These are plant resources of promising value, to develop potential alternative production projects, and are L. porteri, which may be commercialized in the city of Chihuahua with a price of $170.0 per kg of dry root (ITESO 2004), and P. decompositum with a price of $100.0 per kg of dry root (ITESO 2004). It is important to indicate that medicinal plants continue having a relevant role in maintaining the health of the Rarámuri families. Studies carried out in different communities identified that although 99% of the population go to clinics and health centers, 75–90% of them maintain traditional practices to attend some illnesses (CONTEC 2004a, b). Although collecting and commercialization of these species are already carried out, planning for their alternative of nontimber forest product requires detailed evaluations in relation to their distribution and abundance, evaluating the damage of collecting to plants’ survival, an evaluation of the demographic impact to identify the optimum harvest, as well as alternative forms of management and propagation. Other resources with potential

Rara´muri Ethnobotany: Peasant Subsistence and Biodiversity Conservation. . .

165

value should be evaluated similarly; the rhythms and forms of extraction to attend traditional needs are different to those implemented for commercialization. Ecological information like that referred to above, together with traditional techniques of extraction and propagation, are undoubtedly the basis to design sustainable forms of use of these and other nontimber forest products. For the moment, the current diagnosis helps to identify some of the potential resources, and new studies are needed to establish the new forms of sustainable use of local forests.

Conclusion Historically, the Rarámuri economy has been affected by social, political, and technological processes, as well as by climatic and environmental factors, dismantling traditional subsistence activities. Despite these factors, activities such as agriculture and harvesting wild plant products have been preserved in time. The diversity of plant resources that is currently used by the Rarámuri families of Cuiteco offers a wide spectrum of options that maintain and promote the satisfaction of the community needs. However, given the possible loss of biocultural diversity, it is of great importance to know in detail the resources that currently or potentially can be used. The inhabitants of Cuiteco practice a multiple use of resources that is expressed in a high diversity of useful plant species. Of the 356 registered plant species, 63% present some type of use grouped into 14 general categories. Cuiteco represents approximately 0.16% of the total area of the Sierra Tarahumara but includes 18.7% of the floristic diversity reported for the Sierra Madre and 22.6% (226 species) of the useful flora. Of the 300 species of medicinal plants and 120 food plants used by the Rarámuri culture, Cuiteco includes 40.7% and 55.0%, respectively, which highlights the importance of Indigenous cultures in general – and the Rarámuri culture in particular, in the maintenance of biological diversity and useful flora, in relatively small portions of territory. Among the broad spectrum of plant resources documented, some have high local cultural importance, and some others like matarike and chuchupate have high economic potential, but their use require detailed studies about conditions for sustainable harvest. This first approach offers a general panorama about the potential according to the purposes of local people, the distribution and abundance of plant resources. We put in practice and visualized together with local people the great value of the joint work among local people and scholars, of both traditional and academic botanical and ecological knowledge, to construct strategies of sustainable use of resources and ecosystems, as well as the importance of constructing local institution to regulate new decisions and programs for using the forest. This is the first step of a way in which the local people decide the direction and distance. Acknowledgments The authors thank financial support from the Consejo Nacional de Ciencia y Tecnología (CONACYT AS-1-14306), the Dirección General de Asuntos del Personal Académico (DGAPA-PAPIIT, research project IN206520 and IN224023), UNAM, and the Comisión Nacional para el Conocimiento y Uso de la Biodiversidad (CONABIO/GEF/FAO project ID 9380 770, research project RG023), Mexico.

166

A. Camou-Guerrero et al.

References Alcorn JB. Huastec Mayan ethnobotany. Austin: University of Texas Press; 1984. Alcorn J, Toledo VM. Resilient resource management in Mexico’s forest ecosystems: the contribution of property rights. In: Berkes F, Folke C, Colding J, editors. Linking social and ecological systems. Management practices and social mechanisms for building resilience. Cambridge, UK: Cambridge University Press; 1998. p. 216–49. Arriaga L, Espinoza JM, Aguilar C, Martínez E, Gómez L, Loa E. Regiones Terrestres Prioritarias de México. Mexico: Comisión Nacional para el Conocimiento y uso de la Biodiversidad (CONABIO); 2000. Barrera-Bassols N, Toledo VM. Ethnoecology of the Yucatec Maya: symbolism, knowledge and management of natural resources. J Lat Am Geogr. 2005;4(1):10–41. Berkes F, Davidson-Hunt I. Biodiversity, traditional management systems, and cultural landscapes: examples from the Boreal Forest of Canada. ISSJ 187. UNESCO; 2006. Berkes F, Folke C. Linking social and ecological systems for resilience and sustainability. Paper presented at the work shop on property rights and the performance on natural resources system. Stockholm; 1994. Berkes F, Folke C, Colding J, editors. Linking social and ecological systems. management practices and social mechanisms for building resilience. Reino Unido: Cambridge University Press; 1998. Berkes F, Colding J, Folke C. Rediscovery of traditional ecological knowledge as adaptative management. Ecol Appl. 2000;10(5):1251–62. Boege E. Territorios y diversidad biológica: la agrobiodiversidad de los pueblos indígenas de México. Pp. 237–298. In Concheiro, B., and F. López (Coords.). Biodiversidad y conocimiento tradicional en la sociedad rural. Entre el bien común y la propiedad privada. Centro de Estudios para el Desarrollo Rural Sustentable y la Soberanía Alimentaria. Cámara de Diputados, LIX Legislatura, Universidad Autónoma Metropolitana. Mexico; 2006. Bye RA. Ethnoecology of the Tarahumar of Chihuahua, Mexico. Tesis Doctoral. Cambridge, MA: Universidad de Harvard, E.U; 1976. Bye RA. Quelites – ethnoecology of edible greens: past, present, and future. J Ethnobiol. 1981;1(1): 109–23. Bye RA. Medicinal plants of the Sierra Madre: comparative study of Tarahumara and Mexican market plants. Econ Bot. 1986;40(1):103–24. Bye RA. The role of humans in the diversification of plants in Mexico. In: Rammamoorthy TP, Bye RA, Lot A, Fa J, editors. Biological diversity of Mexico. New York: Oxford University Press; 1993. p. 707–31. Bye RA. Prominence of the Sierra Madre occidental in the biological diversity of Mexico. In: DeBano LF, Ffolliott PF, Ortega-Rubio A, et al., coord. Biodiversity and management of the Madrean Archipelago: the Sky Islands of Southwestern United States and Northwestern Mexico. Technical report: RM GTR-264. Fort Collins: Department of agriculture, USDA Forest Services, Rocky Mountain Forest Range Experimental Station, University of Arizona; 1995, pp. 19–27. Caballero J, Mapes C. Gathering and subsistence patterns among the Purhepecha. J Ethnobiol. 1985;5(1):135–46. Camou-Guerrero A. Los Recursos Vegetales en una Comunidad Rarámuri: Aspectos Culturales, Económicos y Ecológicos. Tesis de Doctorado. Posgrado en Ciencias Biológicas, Instituto de Investigaciones en Ecosistemas y Sustentabilidad. UNAM; 2008. Camou-Guerrero A, Reyes-García V, Martínez-Rámos M, Casas A. Knowledge and use value of plant species in a Rarámuri community: a gender perspective for conservation. Hum Ecol. 2008;36:259–72. Casas A, Viveros JL, Caballero J. Etnobotánica Mixteca: sociedad, cultura y recursos naturales en la Montaña de Guerrero. Mexico: INI/CONACULTA; 1994. Casas A, Vázquez MC, Viveros JL, Caballero J. Plant management among the Nahua and the Mixtec of the Balsas river Basin: an ethnobotanical approach to the study of plant domestication. Hum Ecol. 1996;24(4):455–78.

Rara´muri Ethnobotany: Peasant Subsistence and Biodiversity Conservation. . .

167

Casas A, Caballero J, Mapes C, Zárate S. Manejo de la vegetación, domesticación de plantas y origen de la agricultura en Mesoamérica. Bol Soc Bot México. 1997;61:31–47. Casas A, Valiente Banuet A, Viveros JL, Caballero J, Cortés L, Dávila P, Lira R, Rodríguez EI. Plant resources of the Tehuacán – Cuicatlán Valley, Mexico. Econ Bot. 2001;55(1):129–66. Casas A, Lira R, Torres I, Delgado-Lemus A, Moreno-Calles AI, Rangel-Landa S, Blancas J, Solís L, Pérez-Negrón E, Vallejo M, Parra F, Farfán-Heredia B, Arellanes Y. Ethnobotany for sustainable ecosystem management: A regional perspective in the Tehuacán Valley. In: Lira R, Casas A, Blancas J, editors. Ethnobotany of Mexico. Interactions of peoples and plants in Mesoamerica. Utrecht/Holanda: Springer; 2016. p. 179–206. Chapter 8. Consultoría Técnica Comunitaria (CNTEC A.C.). Ordenamiento ecológico comunitario del ejido El Consuelo, municipio de Carichí. Chihuahua: SEMARNAT – CONTEC A.C; 2004a. Consultoría Técnica Comunitaria (CNTEC A.C.). Ordenamiento Ecológico Comunitario del ejido de Magullachi, municipio de Carichí. Chihuahua: SEMARNAT - CONTEC A.C; 2004b. Consultoría Técnica Comunitaria (CNTEC A.C.). La leña: el recurso olvidado. Una experiencia de participación social y cambio tecnológico en dos microregiones de la Sierra Tarahumara. Colección: Cosechando Palabras, No. 1. Chihuahua: CONTEC A.C. – Fundación Sierra Club; 2006. 28 p. Consultoría Técnica Comunitaria (CNTEC A.C.). Fortalecimiento de los procesos de apropiación tecnológica comunitaria para el manejo de los recursos naturales de la Sierra Tarahumara. Chihuahua, México. Chihuahua: MISERIOR, CONTEC; 2008. Davidson-Hunt I. Journeys, plants and dreams: adaptative learning and social-ecological resilience. Doctoral thesis. Winnipeg: Natural Resources Institute, University of Manitoba; 2003. DeBano LF, Ffolliott PF, Ortega-Rubio A, coord. Biodiversity and management of the Madrean Archipiélago: the Sky Islands of Southwestern United States and Northwestern Mexico. Technical report, RM-GTR-264. Fort Collins: Agriculture Department, USDA Forest Services, Rocky Mountain Forest Range Experimental Station, University of Arizona; 1995. Farfán B, Casas A, Ibarra G. Plant resources in the Monarch Butterfly biosphere reserve, Mexico: Mazahua ethnobotany and peasant subsistence. Econ Bot. 2007;61:2173–91. Felger SR, Wilson M. Northern Sierra Madre Occidental and its Apachian Outliers: a Neglected Center of Biodiversity. In: DeBano LF, Ffolliott PF, Ortega-Rubio A, et al., coord. Biodiversity and management of the Madrean Archipelago: the Sky Islands of Southwestern United States and Northwestern Mexico. Technical report: RM GTR-264. Fort Collins: Department of agriculture, USDA Forest Services, Rocky Mountain Forest Range Experimental Station, University of Arizona; 1995, pp. 36–59. García E. Modificaciones al Sistema de Clasificación Climática de Koppen. Mexico: Instituto de Geografía, Universidad Nacional Autónoma de México; 1981. Guerrero M, Reed C, Vegter B. Informe: La industria forestal y los recursos naturales en la Sierra Madre de Chihuahua: impactos sociales, económicos y ecológicos. Comisión de Solidaridad y Defensa de los Derechos Humanos, A.C. (COSIDDHAC), Chihuahua, México. Austin: Texas Center for Policy Studies; 2000. Guerrero MT, Villa F, Nelly M, Reed C, Vegter B. La industria forestal del estado de Chihuahua: efectos económicos, ecológicos y sociales del TLCAN. Environmental effects of free trade. Paper presented at the North American Symposium on Analysis of the Linkages between Trade and the Environment (October 2000). Comisión para la Cooperación Ambiental (CCA). Montreal, Canadá; 2002. Hetch SB, Posey DA. Preliminary results of soil management techniques of the Kayap Indians. Adv Econ Bot. 1989;7:174–88. Instituto Tecnológico y de Estudios Superiores de Monterrey (ITESO). Investigación de mercado: plantas medicinales. Chihuahua: ITESO-CONTEC; 2004. LaRoche S, Berkes F. Traditional ecological knowledge and practice for edible wild plants: biodiversity use by the Rarámuri in the Sierra Tarahumara, Mexico. Int J Sust Dev World. 2003;10:361–75. Lumholts C. El México desconocido. Mexico: Instituto Nacional indigenista; 1986. p. 516.

168

A. Camou-Guerrero et al.

Martínez-Alier J. El ecologismo de los pobres: conflictos ambientales y lenguajes de valoración. Barcelona: Editorial Icaria; 2009. Ortiz-Avila T, Masera O. Subsidios y estrategias de producción campesina: el caso de Casas Blancas, México. Revista Iberoamericana de Econmía Ecológica. 2008;7:61–80. Pennington CW. The Tarahumar of Mexico. Their environment and material culture. Salt Lake City: University of Utah Press; 1963. Pérez-Negrón E, Casas A. Use, extraction rates and spatial availability of plant resources in the Tehuacán–Cuicatlan Valley, Mexico: the case of Santiago Quiotepec, Oaxaca. J Arid Environ. 2007;70:356–79. Sariego JL. El indigenismo en la tarahumara. Identidad, comunidad, relaciones interétnicas y desarrollo en la Sierra de Chihuahua. Mexico: Instituto Nacional Indigenista (INI)–Instituto Nacional de Antropología e Historia (INAH); 2002. SEDESOL. Las 250 micro-regiones de México. 2001. http://www.sedesol.gob.mx Toledo VM. La perspectiva etnoecológica, cinco reflexiones acerca de las “ciencias campesinas” sobre la naturaleza con especial referencia a México. Ciencias Número especial. 1990;4:22–9. Toledo VM, Argueta A, Rojas P, Mapes C, Caballero J. Uso múltiple del ecosistema: estrategia del desarrollo. Ciencia y Desarrollo. 1976;2(11):33–9. Toledo VM, Ortíz-Espejel B, Cortés L, Moguel P, Ordóñez MJ. The multiple use of tropical forest by indigenous peoples in Mexico: a case of adaptive management. Conserv Ecol. 2003;7(3):9. http://www.consecol.org/vol7/iss3/art9 Vatant F. La explotación Forestal y la Producción Doméstica Tarahumara. Serie Antropología Social. Mexico: Instituto Nacional de Antropología e Historia (INAH); 1990. Zizumbo-Villarreal D, Colunga-Garcia Marín P. Los huave. La apropiación de los recursos naturales. Departamento de Sociología Rural. Mexico: Universidad Autónoma de Chapingo; 1982.

Wild Food and Traditional Knowledge of the Kumiai from Baja California Carolina Gutie´rrez-Sa´nchez and Claudia Leyva-Aguilera

Abstract

This work aims to documenting use and importance of natural resources identified as edible by the Kumiai, native people from the northwestern region of Baja California. For this study a qualitative investigation was carried out in both historical and current contexts, to identify edible plants with cultural value and symbolism for the native cuisines. In Baja California, native groups have accumulated a great ecological knowledge by moving as nomads through different ecosystems. This life pattern allowed them to take advantage and survive based on the natural resources those ecosystems provide. For native people it has been important the use of natural corridors containing high richness of medicinal and edible plants. Although the tradition and knowledge are no longer practiced as in the past, people make efforts to revitalize them through individual and collective actions. In the context of modernity, the valuation of the edible elements found in nature with which the native groups of the region subsisted remained outdated. Through fieldwork, this study looked for documenting the diversity of edible components and demonstrating the differential and complementary composition of traditional food among different ecosystems. The design of this research is qualitative, following ethnographic methods, including tools such as interviews and participant observation. For documenting information, we especially worked with elders and traditional authorities within the Kumiai community, considering that they are those people maintaining the most representative ancestral knowledge. As result of our investigation, we identified the influence of the Kumiai C. Gutiérrez-Sánchez Universidad Autónoma de Baja California, Ensenada, Mexico e-mail: [email protected] C. Leyva-Aguilera (*) Department of Sciences, Environmental Education Specialty, Universidad Autónoma de Baja California, Ensenada, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_7

169

170

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

knowledge of biodiversity and varieties of native resources and recognized their cuisine and traditions. This work contributes information that adds elements to the scarce data available about food and biodiversity recognized, used and managed by the native people of Baja California and aspires to contribute to recognize their food heritage. This study is mainly directed to show local initiatives that have been implement for the valuation of natural resources and their relationship with culture and complementary information about the role of edible plants and their uses.

Introduction: Wild Food as Cultural Heritage Food is one of the most important worries of humanity in the world today, and plants are important components in relation to this concern; plants form part of food and involve cultural and spiritual practices related to it. In many cultures, the traditional knowledge about the use of plans is transmitted from generation to generation mainly though oral means. Indigenous and local societies are facing changes that threaten the preservation of their traditional knowledge, and maintaining and recovering traditional food systems can contribute to this direction and bring people closer to a conscious and harmonious connection with the natural environment and health (Kuhnlein et al. 2009). Indigenous groups use plants in multiple ways during their daily life (Nabhan 2004). There are thousands of edible wild plants recognized by different cultures around the world, but paradoxically, less than 10 species supply most of the calories consumed in the world, just three of them, rice, maize, and wheat, provide 60% of the world’s food energy intake (FAO 2009). Commonly, ethnobotanical surveys of relatively small samples of respondents can surprisingly yield high numbers of species used by the different cultures around the world (Bharucha and Pretty 2010). Among indigenous communities, use of wild species is part of the food culture and economy (Anderson 2005). In their study, Kuhnlein and collaborators (2009) registered seven communities in Canada knowing around 120 species per community, rising to 194 in those communities formally designated as agriculturalists. In addition, researchers have also recorded that wild species have diverse uses. For instance, 80% of 65 wild food plants consumed in Nepal have multiple uses (Shrestha and Dhillon 2006). Numerous rural and indigenous territories have been transformed into urban areas, devaluing the rural landscapes. The way of life of many rural groups has been disrupted by urban contexts, causing changes in diet and health of populations and ecosystems (Ericksen 2008). In this way, the traditional food practices of rural groups fulfill the important role of making visible the wealth and heritage found in local ecosystems. In rural areas, different social groups, particularly those most vulnerable because of their poverty, take biodiversity from natural environments as a main source of

Wild Food and Traditional Knowledge of the Kumiai from Baja California

171

good to satisfy food security and generate income (Pretty and Pervez 2015). The biodiversity of a wild environment contributes to food security, through thousands of edible wild species that continue being part of the diets of about one billion people (Aberoumand 2009). Some studies indicate that recognizing the importance of the natural environment requires the development of an intrinsic relationship between species conservation policies and food security (Bharucha and Pretty 2010). For many cultures, the exact meaning of the term “wild” is a problematic concept. According to Pretty and Pervez (2015), this term is commonly used to refer to ecosystems and contexts where people have not intervened. However, in the case of the native groups of this study, as in many other native groups around the world, wild commonly include areas, places and species where and which are intervened and managed, including much of the ecosystems of their place of origin. In Baja California, like in most arid and semiarid zones of the world, water is a main limiting factor, its scarcity determining poor agricultural development at different scales, while food demand increases due to population growth (Cortes 1994). However, in arid areas, a broad spectrum of plant species forming part of the natural vegetation can transform a high percentage of their energy into food energy concentrated in their seeds, which makes them potentially rich in nutrients; therefore, it is pertinent documenting such resources, encouraging their proper use and conservation, and this effort may involve numerous plant species of these areas (Lucero 1995). At present, in Baja California there are indigenous human groups related to communities located in the states of California, Arizona and Utah, in southern United States (Fig. 1). Native peoples from Baja California are related to the local ecosystems and landscapes, which play an important role in people’s food systems. In turn, interactions between people and environments contribute to the maintenance of species and genetic diversity of the original products they have used for thousands of years (Williams 2004; Fig. 2). The native groups have become adapted to the diverse environments of the region through millenary processes of cultural development. These human cultures show their creativity based on meticulous observations of nature and applied to construct techniques to survive in the difficult conditions that the local environments present to them (Wilken-Robertson 2012), such as droughts, heat waves, food scarcity, and cold weather. Most native people’s food was wild and its abundance markedly associated to some seasons, so that survival demanded a constant and permanent search of resources, which was progressively available at greater distances. In times of drought, the native people became practically nomads, migrating incessantly in search of food (Aschman 1954). The importance of the ecological knowledge of the local environment by native human communities of Baja California, is reflected on the still persistent activity of seasonal harvesting of fruits, flowers, roots, seeds and leaves as mentioned in the upcoming paragraphs (Wilken-Robertson 2018).

172

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

Fig. 1 Map of the Yuman language family. This map explore historical linguistic approaches to explore the cultural relationships between language varieties and prehistoric migrations. (This map was taken from the book of Kumeyaay Ethnobotany by Michael Wilken R. (2018). Map elaboration in collaboration with Gerardo Chavez)

Wild Food and Traditional Knowledge of the Kumiai from Baja California

173

Fig. 2 Rush basket made it by a native woman from the community of San José de la Zorra. (Photo by Samantha Caldera Delgado, 2016)

From the High Sierra to the Coast: The Wild Food Corridor The interactions with nature are for the Kumiai a representative aspect of their culture. Walking through different ecosystems, has constituted a long-term practice that has constructed their botanical and ecological knowledge. Oaks, pines and other trees, shrubs and herbs, are located in different ecosystems but these are united by natural corridors where people practice gathering of their products in different seasons. All these components of vegetation were considered fundamental elements for life, uses and customs of these groups (Anderson 2005). Landscapes are constructions and represent a certain culture, since these are based on the subjectivity and imagination that project the value of natural resources (Schama 1995). In the Peninsula of Baja California, the chroniclers of the eighteenth century describe that the way of life of the native people was nomadic, mainly characterized by their mobility and the seasonal consumption of wild resources, which allowed them to subsist in the rugged and limited environment (Piñón 2000). Throughout the nineteenth century, the peninsular native people changed their lifestyle and management practices of natural resources by adopting cultural and economic exchange with other social actors (mainly missionaries and ranchers) (Wilken-Robertson 2018). The collection of food played an important role in the way of subsistence of these communities; this activity was intense during the seasons of abundance, to survive during the periods of food scarcity (Shipek 1982). Groups of native people settled for short periods of both seasons in different areas, from the highest point of the peninsula in the Sierra de San Pedro Mártir, which ranges elevations higher than 3,000 m. a. s. l., where they collected pine nuts, to the coastal plateaus oriented toward the Pacific Ocean, where they had the opportunity to collect leaves and seeds of several plant species, in addition to fishing (Garduño 2010).

174

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

In the Kumiai territory, there are elements of the chaparral that were used, like the barrel cactus (Ferocactus spp.), jojoba (Simmondsia chinensis), canutillo (Ephedra californica), sweet mesquite and cat’s claw, among others. In this ecosystem, the oaks occur only in the valleys that accumulate enough rain moisture, as well as in damps and springs. Oaks, according to the Kumiai worldview, are important as source of food, since the acorns are prepared and consumed in different ways. The oak forest is considered their roof, shade and food. The association of the Kumiai with this forest is as important as that established with corn by the Mesoamerican cultures (Tapia and Grijalva 2012). In this way, the oaks constituted symbolic representation of identity and life of the Kumiai, providing to them food (acorns), water (oaks are effective rain catcher), shelter, human culture (important part of their diet) and energy (fuelwood) (Small and Ramirez 2015). The Kumiai culture has a knowledge of medicinal plants that is relevant, usually older people possess it, but they do not devote themselves to healing. This knowledge is the product of hundreds of years of talks between grandparents, parents, children, and grandchildren. Children learn the names of the plants, their location, some curious facts, but not their use, knowledge about use is reserved for some persons (Gutiérrez and Von Glascoe 2019).

Discovering Landscape and Culture Through Wild Foods: Experiences with Plants and Animals According to Harris (1989), the most carnivore cuisines are related to low density of population and lack of need for croplands whereas the most herbivore cuisines are associated with dense populations whose habitat cannot support breeding cattle (Harris 1989). Based on this pattern, it has been demonstrated that the great differences between the cuisines of the world can help to observe the ecological weaknesses and opportunities that can be found according to the characteristics of the zone (Álvarez 2002). In Baja California, the indigenous communities have gathered rich botanical knowledge by the simple means of moving as nomads through different ecosystems to make use and survive of their natural resources. In the Kumiai communities of Ensenada, one can still perceive the aroma of these traditional plants; the elderly remembers its use whether for edible or medicinal use some of them still practicing the gathering by seasons. Currently in the Kumiai communities, you can find variants of the eddible flora that this culture traditionally used to use (Wilken 2012), like in their traditional festivities and special occasions as weddings or birthdays. The weather and soil conditions are necessary for the reproduction of different species that are part of the traditional Kumiai cuisine. Since in recent years the climatic conditions have been changing surprisingly, storms tend to delay or anticipate to what people normally expect according to their knowledge about seasons. In addition, the frequent fires, and permits for collecting (that the indigenous people have not been able to pay or are not easy to get.) have influenced decreasing of traditional practices (Morales 2010), particularly the harvesting of edible plants.

Wild Food and Traditional Knowledge of the Kumiai from Baja California

175

One of the informants with empirical knowledge of her community’s ecosystems, Mrs. Rogelia, mentions that: “The frog herb (Eryngium carlinae) is a plant that gets out the excess of fat and cleans your brain, heart, liver, and kidney. Plants have what we need, that’s why we have to share the knowledge we have about them La hierba del sapo es una hierba que saca del cerebro, del corazón, del hígado, del riñón, el exceso de la grasa y te desintoxica. Las plantas son las que tienen lo que necesitamos. Es por ello que se debe transmitir esos usos (Rogelia, personal communication, 18 of july of 2016).”

Sharing meals with people to understand their practices is essential to preside over dialogue and a closer understanding of customs. In this part, the experience of putting into practice field knowledge narrated, bringing this narration of the experience and knowledge of the inhabitants. Bendimez (1987) mentions that before the arrival of the Spaniards, the original groups of the Baja California region were seasonal nomadic groups that moved along the peninsula between the coast and the mountains. Looking for better life conditions in that cycle, taking advantage of species such as acorn oaks, pitahaya (Stenocereus thurberi), biznaga (Ferocactus spp.), nopal (Opuntia basilaris), chollas (Cylindropuntia cholla), jojoba (Simmondsia chinensis), pine nut (Pinus quadrifolia), among others (Barco 1988). Among the species of the regional flora, the more peculiar found within the ecosystems are fruits and seeds of the pitahayas, nopales cladodes which are consumed cooked or roasted. The roasted stem and scape of the inflorescence of agave are consumed; they also obtain from this preparation a very sweet liquid that is consumed as honey. Honey was an ingredient that they used in many of their preparations; currently the Kumiai continue to collect wild honey, this is one of the pre-Hispanic activities still preserved. Seeds such as acorns are processed into flour, grinding them into natural metates1 that found and manufactured with stones (Aschman 1949). The native people also used the wood to carve and make a kind of ladles that employed to move the large stews or simply as a tool to take some food. Their meals were and are spiced with vegetable powders and grinding products. These powders are put in small containers that the natives from the peninsula call “coras,” which were and are made of herbs and branches, mainly of reed and willow (Aschmann 1952). The Kumiai complemented their diet with animals occurring in local ecosystems like deer, rabbits, black tailed jackrabbit, squirrels, mice, seabirds, and

1

The metate, or grinding stone, is a tool used to process (grind or crush) food by different cultures since prehispanic times. In the case of the ancient cultural groups of Baja California, the metate was any long rock that used to grind grains, seeds or dried meat.

176

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

migrant animals such as seagulls, ducks, and others, as well as vipers, snakes, lizards, and various reptiles. Similarly, as with plants, the Kumiai had into account meat consumption during the more appropriate time, considering the life cycle and breeding seasons of the animals. During the right time, pigs, chickens, wild rabbits, deer and chacuaca (quail) can be the more important food (Aschmann 1952). Another attractive resource of the Kumiai culture is the rattlesnake, which was symbolized in their worldview. To eat it as food, the head was cut so that it is not toxic; they dried it and milled it, and the flavor of this mass is commonly compared in taste with that of shrimp or fish. Mr. Silva, from the community of San José de la Zorra mentions what the ancient people ate: “Uh well deer meat, wheat tortilla, mustard leaves, maguey flowers, and rabbit and quail meat. For flavoring water, we used the manzanita fruits (Arctostaphylos pungens) “that was very good”. All that was prepared by my grandmother and besides that pennyroyal found on the hill; pennyroyal is really good and nutritious. Uh pues lo que era la carne de venado, tortilla de trigo, la mostaza, la flor de quiote, la carne de conejo, la codorniz. Para el agua fresca pues la manzanita, “ese sí que estaba bien buena”. Todo eso lo preparaba mi abuelita y pues de ahí el poleo que había en el cerro; el poleo es muy bueno y nutritivo. (R. Silva and R. Carrillo, personal communication, 3 of October, 2016)”

Utensils Mortars and metates (a stone milling instrument fabricated with a plain stone at the basis and a long cylindrical stone manipulated with both hands for pushing and grounding different types of matters) were among the tools that people used during the food processing season. People used to remove the shells from acorns as it has a very hard layer, then they grounded usually 5 to 6 acorns depending on the mortar pit; later, in the metate the acorn flour was milled and prepared to make the acorn food. The metate and mortar were also used to soften dried meat and to process some medicinal plants. To clean the acorns, baskets called sawiles (Kumiai language) made of rush or pine were used. After removing the first layer of acorns it was necessary to remove a second layer of shell that gives a bitter taste to the seed and after washing it Cloak came off. Large baskets of willow used to put food. This material together with the eucalyptus leaf is a natural insect repellent, which in that way helped to maintain food well-preserved. Clay pots used to cook and maintain the temperature of certain foods and beverages (Fig. 2). They use the agave leaves (penca) as a source of water, a piece of the leaf kept it in their mouths and they were sucking it, so they could last several days without needing to drink water, sometimes they add flavor through herbs they find around, especially chía (Salvia columbariae) (Grijalva and Tapia 2012).

Wild Food and Traditional Knowledge of the Kumiai from Baja California

177

From the Sierra: Pin˜o´n (Pine Forest) Seeds, green leaves, wild flowers and wild fruits are basic components of the Kumiai traditional diet, part of these seeds ingredients are transformed into pastes which are rich in oleic beneficial for health, they were and are important in human nutrition. In addition, these products have advantages in a nomadic life in arid ecosystems, since these products are easily preserved foods and therefore were used for long periods of travel. In the past, these wild foods were only consumed directly or grounded or roasted. This helped the body take advantage of all the nutrients. In such a way, the Kumiai said that people lived longer because of the constant and traditional consumption of these products. Among the foods in the category of fruits that the Kumiai have in their diet the pinion with shell and the acorns are among the most important (WilkenRobertson 2018). These foods belong to the group of nuts. Due to their high amount of calories, these nuts are recommended for long trips or hunting activities (Fig. 3). The pine nut is a main food mostly consumed by those groups of people settled in the mountains. This fruit found in the pines (Pinus quadrifolia) occur in the Sierra de Juárez, in La Huerta, as well as in the Sierra Blanca in San Antonio Necua and San Pedro Mártir further south. These pines are 5–15 m height, with a rounded crown, leafy and short trunk (Rebman and Roberts 2012). Pine cones are the carriers of this fruit that is approximately 5 cm long, between its scale’s dark brown seeds and hard surface produced, which once removed its shell offers the seed that has a golden film, which confers a certain bitterness when consumed.

Fig. 3 Pines from the region, before extract the seeds. (Photo by Carolina Gutiérrez, 2019)

178

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

Among the beverages is also the mead to cool off, which was prepared with water and honey. Another beverage and or meal is the pine nut pinole that they started to prepare when people started to raise and milk animals; they then started to add milk to its preparation. The basic preparation was to grind wheat, making with it masita2 and then add the milk and a little honey to sweeten the mixture. Wheat was also used to make a coffee substitute, where wheat was roasted, grounded, and a kind of infusion prepared (Fig. 5). The Kumiai are a people that stands out for being great gatherers. Kumiais from the Sierra area still practice the gathering of pine nuts as a family ritual, as a way to maintain traditions alive and as part of the local economy.

Jojoba In the region, several other plant species are representative of the traditional Kumiai cuisine and provide healing benefits, that improve local economy. In this case, in the chaparral environment, we find the jojoba (Simmondsia chinensis), which is one of the appreciated plants by Kumiai native people from community of “La Huerta,” since its presence is more frequent in those directions. There is a registration that in the past, Kumiai people used the seed oil to keep their hair and skin healthy. Currently, jojoba is known for the skincare products manufactured, such as shampoo, soaps and hair oils. Traditionally it was collected during the months of July and August. There is a perception of accessibility to different wild food species, according to the Kumiai community location, even though they are located in different areas they share food ways and traditions. In a conversation with a member of San Antonio Necua, Kumiai community, comment: It varies a lot because it has not rained, the weather has changed a lot. There are years when it ripens early and others when it ripens late, you have to be on watch for example at La Huerta (another Kumeyaay community), they have a lot of jojoba and lots of pine nuts that we don’t have, they are different seeds, acorns thrive better here, in fact we use to exchange seeds between people when they started to establish in places as sedentary. When people started to establish the communities, they had to adapt to wheat, maize, is to say to anything that could be sowed and harvested that Europeans and the missionaries brought, as well as fruit trees, lemons, peaches and apples. There were many fruit trees beneath the old establishments at Sierra Blanca one of the oldest (Personal comunitation, San Antonio Necua, 2015). Varía mucho porque no ha llovido, el clima ha cambiado bastante, hay veces que madura antes o madura después, tienes que estar pendiente, -por ejemplo, en La Huerta (otra comunidad Kumiai) tienen mucha jojoba y también tiene mucho piñón que nosotros no tenemos-. Son semillas diferentes, aquí se da más la bellota. De hecho, se practicaba el trueque de semillas entre las personas cuando se empezaron a establecer en los lugares como sedentarios. Cuando se empezaron a establecer las comunidades tuvieron que adaptarse al trigo, el maíz, es decir, a lo que se podía sembrar y cosechar que habían traído los europeos y In Mexican food culture, masita is the diminutive of dough, preparation made from kneading flour with a liquid agent.

2

Wild Food and Traditional Knowledge of the Kumiai from Baja California

179

los misioneros; así como los árboles frutales, limones, duraznos, manzanos, etc. Se pudieron encontrar muchos árboles frutales debajo del asentamiento que antes había en la sierrita blanca que es de los más antiguos (Comunicación personal, San Antonio Necua, 2015).

From the Valleys: Bellota (Acorns) Even at present, the oak acorns constitute a wealth for many people. For the Kumiai people of Baja California, these plant parts are emblematic elements of their diet. The acorns are currently used as food and fodder for different types of livestock as well as in different cities in Europe where acorns still have a culinary use. The ancient Kumiai alike had their unique preparations, but acorn mush (atole) is the most common. The high importance of acorns in diet is based on the evidence that the acorns are characteristic resources of the riparian ecosystem and can be harvested as a food resource (Wilken-Robertson 2018). The acorns are very bitter in the mouth, which allows assuming that contain tannins and acids, which makes them also long-lasting food. It is a fruit rich in carbohydrates, proteins and oils (Lucero 1995), some older people say that they only needed to consume the acorn mush pasta once a day to be satisfied (Fig. 4). Oaks are some of the most abundant trees in the region, the fruiting of oaks begins with a green color and grows darker and tannin as the summer passes. The tree begins to bear fruit when they are between 8 and 10 years old, the acorns ripen in the

Fig. 4 Riparian ecosystem, olds oaks in San José de la Zorra. (Photo by Adrian Macias, 2019)

180

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

Fig. 5 Gather acorns for preparation of “atole de bellota.” (Photo by Carolina Gutiérrez, 2016)

fall during between October and December, sometimes delaying until January (Gutiérrez and Von Glascoe 2019). According to the taste of sweet or bitter acorns, the Kumiai distinguish two types of oaks, one of them is Quercus agrifolia, which has elongated leaves and produces bitter acorns, with a distribution in the riparian areas of the Mediterranean-type vegetation and Quercus subspecies rotundifolia with a rounder leaf and sweeter acorn. The use of acorns requires much less work, few or no fertilizers and water to obtain a good harvest (Sieso et al. 2002). In relation to the transmission of knowledge about medicinal and culinary plants, Mrs. Beatriz of the community of San José de la Zorra says that they learned from native plants by the fact of using them (empirically) in case of any discomfort. In turn, she comments that in 1918 a man arrived in the community with influenza that cured with the cow’s weed (Rumex sp.) and that is why the medicinal property of this herb was known. It is essential to see in a holistic way the connection that exists with

Wild Food and Traditional Knowledge of the Kumiai from Baja California

181

natural resources and food, but especially to take actions that guarantee the revaluation of wild species. In the ecosystems occurring in Ensenada, especially those around the native people communities, it is possible to find areas with high potential as botanical alternatives including medicinal and culinary wild species representing important ecosystem services for local landscapes. To prepare the acorn mush, the seasonality (october, november, december) must be considered, is part of an individual preparation, as some people said, the preparation steps must be carefully followed. If the acorn is closed, it should be put on a stone, the skin it has is opened and the yellow seed can be seen, they are very bitter if tasted. The seed produced in November ripe in December, they fall alone, and people collect them. The acorns collected are left exposed to the sun for long time to get dry and after they are dry they are muddled with a stone; following this way it is easier to open them. My mother use to do it [grinding acorns] in a metate she would grind it like that, is a long stone that has a hole to muddle against another flat stone with that she would turn the acorns to flour, she would rinse the flour two to three times. For that, you have to set a clean cloth. She would make a sieve with some branches; she would pour the powder (acorn flour) and would rinse and rinse. She would taste it and taste it again until she got rid of the bitterness, then it was time to make the mush. To be honest it is a lot of work and people fear not make it right, so they do not try. However, I have a lot of experience at it. The acorn mush has to be done only by one person, for example if I am making it, I have to be the one to pour the water and everything, if someone else tries to help me the water in the mush wont drain it will pool, and that delays the process and dilutes the flavor. That is why my mom told us to stay away when she was making atole, so we would hide to see. (B. Carrillo, personal conversation, July 21, 2016) Mi ama lo hacía en metate lo molía así, era una piedra que tienen un hoyo y larga y eso lo machucaba quedaba como una harina, y ya eso lo lavaba como unas dos o tres veces. Para eso tienes que poner una tela limpiecita como de manta. Hacía con ramas y la manta como un colador y pues le echaba el polvo y ese lo estaba lavando y lavando. Lo probaba y lo probaba y ya de tantas veces de haberlo lavado que se fue todo lo amargoso, se hace el atole. La verdad que es mucho trabajo y pues la gente se asusta de que no le va a salir y pues no lo hace, pero pues yo ya tengo práctica. El atole de bellota lo tienen que hacer una persona nomas por ejemplo si yo lo hago, tengo que ser yo quien le eche el agua y todo, porque si va a otra persona a ayudarme el agua no se va se queda estancado y pues ya tardas y no te sabe. Por eso mi mama nos decía “no se arrimen”, cuando hacia el atole de bellota, luego se enojaba y decía “ya vieron de seguro ya no se va el agua. Nos escondíamos y espiábamos para ver y pues nos regañaba por eso. (B. Carrillo, conversación personal, 21 de julio del 2016)

In the past, the persons responsible for preparing the mush were old women and only they could know how was done, in fact, they hid among the herbs to prepare it because at that time there were not many houses. Once prepared, the mush was accompanied with dried deer or rabbit meat. Concerning the accompaniments of the acorn that recalls what the ancients did, Tabita S., a native from the Kumiai community of San Antonio Necua, said: “In the past, there was a custom of preparing barbecue with animals heads, which should be skinned and seasoned with garlic, oregano and bay leaf. To then wrap it again with its own skin and leave

182

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

it to cook on a pit, we did not use pepper, it was introduced over time, what we did was to use a lot was onion and oregano. In those days we made deer machaca, beef or rabbit and accompanied nopales, meats would be salted and hung to air dry a day before, so the preparation was easier the next day” (Tabita S., personal communication, august 12, 2015). “Antes también se acostumbraba mucho la barbacoa de cabeza y se preparaba con la misma cabeza del animal la cual se le retiraba la piel y se le agregada ajo, orégano y laurel para después volverla a envolver en sus mismas pieles y se cocía dentro de un pozo. No se usaba el chile para agregar a los platillos se fue introduciendo conforme el tiempo, lo que si se utilizaba mucho era la cebolla y el orégano. En aquel entonces la machaca que se preparaba era de venado, res o de conejo y se acompañaba con nopales, las carnes se dejaban orear un día antes con sal para que su preparación fura más fácil al día siguiente” (Tabita Sandoval, comunicación personal, 12 de agosto del 2015). The acorn mush should be prepared with the right pasty consistency, it has a bitter taste. Many people do not like that taste since it could be unusual and tasteless, however, when combined with a freshly made flour tortilla and a salt hint, just like that, the taste becomes quite pleasant. This image shows a representation of the place in San José de la Zorra where the women of three generations used to gather and grind acorns. This place is surrounded by holm oaks, it is located on the slopes of hills and has a large stone with several indentations like the one shown in Fig. 4. Maria’s grandmother said that “when a stranger is watching the process the cloth clogs and will not let all the properties of the acorn trough. It is to say that the acorns juice is repressed and does not want to be seen,” “cuando una persona ajena está viendo el proceso la malla se tapa y no deja que pase todo lo que la bellota aporta, es decir se reprimía el jugo de la bellota, no quiere ser visto” (María Ester, personal communication, May 12, 2016). The procedure for making the old acorn mush is not simple; firstly, it is necessary to open the acorn and the seed, leave it very clean, then grind it very well, wash it very well and then cook it. Acorn cooking is a very long time process and only one person has to do it. It is said to be similar as when somebody wants to pick up an egg white, some have not been sponged. For example, it mentions that: “if someone is doing the acorn atole and someone else helps it doesn’t come out right,” “si alguien está haciendo el atole de bellota y alguien más le ayuda no sale bien.” As for the traditional cuisine of the ancient people of the community of San José de la Zorra, Beatriz C. is one of the traditional Cook in the community, she has accumulated the memories and knowledge of the cultural heritage for over half a century. It is of great importance to recognize the work of traditional Cooks; they give the chair of the transformation of the ingredients of their localities into excellent dishes, an important part of the traditions, memories and experiences of their communities over time (Contreras and García 2004).

Wild Food and Traditional Knowledge of the Kumiai from Baja California

183

Wild Leaves Within the community, some products are seasonal, that is, they are not available the year round; therefore, it is more important that older people transmit their knowledge of what they are and how they are obtained. The common food years ago was quiote flower (Agave shawii), jalpap (Dichelostemma capitatum), islaya (Prunus ilicifolia), and sage (Salvia sp.). The Kumiai take advantage of their passage and tested them to survive; such is the case of the acorn (Quercus spp.), pitahaya (Stenocereus thurberi), biznaga (Ferocactus spp.), cactus, chollas (Cylindropuntia sp.), jojoba (Simmondsia chinensis), and pine seeds among others (Barco 1988). They also used manzanita (Arctostaphylos sp.) for fresh waters, which is used until now. In the past, was more frequent to use medicinal herbs for consumption as food or remedies, such as the epazote (Dysphania ambrosioides), the olive tree (Olea europaea), the manso herb (Anemopsis californica), the valerian (Eriogonum fasciculatum), the rosemary (Artemisia californica), the gordolobo (Gnaphalium spp.), and the sage (Salvia sp.). The latter plant species also served to flavor some preparations, practicing some rituals and controlling the flu symptoms (Fig. 6). According to the cosmogony of the Kumiai, sage cleanses and purifies space and soul. The native people use white sage (Salvia apiana) as incense for some festivities related to their cultural heritage. The natives smoked with the white Sage that gave off to purify both their physical and emotional bodies, as places and objects. The abundance of this plant within the plain and the slopes of the hills in San José de la Zorra is very noticeable recently there is a concern from the Kumiai people in Baja California and California about white sage because there is a high demand from the global market above this plant. This is causing a high impact in the deforestation and availability of the plant in traditional land and local territory. Also, most of the medicinal properties of white sage are in its leaves. People use to prepare an infusion with the leaves where the essential oils of the plant are concentrated (Pijoan and Venegas 2018). It is also used and popularized in shamanic rituals around the world. Efforts are carried out by some people to transmit to the community information about this demand. The romerillo (Artemisia california) is a native plant species of great importance for medicinal purposes in the Kumiai community and mainly found within the aromatic coastal scrubland. Its aroma occurs mostly on wet days, in the morning and in the breezy afternoons typical of the coast of the NW of Baja California. Its dense ramifications of small green leaves characterize it. Like many of the other native plants already mentioned, the romerillo is a plant resistant to low water soils (Fig. 8). Yerba santa (Eriodictyum sp.) has a peculiar appearance and aroma. Its leaves are big, long, and sticky. By touching this shrub, it is possible to perceive intense odors, which reminds a peculiar sweetness combined with different species. The native people use it mainly as a remedy for respiratory conditions such as coughs, colds and that is why they say even asthma. It is also used to relieve menstrual cramps or any muscle pain in the body. Similarly, different chefs in the region test its flavor exploring this sheet. To the full knowledge that has been heard about the use that

184 Fig. 6 White sage, edible and medicinal plant. (Photo by Carolina Gutierrez Sanchez, 2017)

Fig. 7 Hierba Santa (Eriodictyon lanatum). Edible and medicinal plant. (Photo by Carolina Gutiérrez, 2016)

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

Wild Food and Traditional Knowledge of the Kumiai from Baja California

185

Fig. 8 Lentisco or Laurel sumac (Malosma laurina). Medicinal plant. (Photo by Laura Elena Jimenez, 2016)

the Kumiai give to it, one or the other, it is interesting the wonderful organoleptic and medicinal properties this plant has (Fig. 7). The Kumiai recognized the value and importance of their environment, so the description of several of the bushes, shrubs, and seedlings is of great relevance to them since the entire ecosystem is part of their lifestyle and cuisine. For this reason, Valerian (Erigonum fasciculatum) is a medicinal plant easily found in the region and surroundings of the community of San José de la Zorra, its obtaining is only for medicinal use. Its main use is related to those who suffer from sleep disorders. The part of the plant used is the root, it is said that gives a powerful relax to calm the body. Don Rito S. tells about it that: “when one wants to sleep well, you just need to drink the root of this plant, you have to let it dry and grind it, this way you make an infusion”; “que cuando uno quiere dormir bien, solo es necesario tomar la raíz de esta planta, dejarla secar un rato y después molerla, de esta manera, se calienta agua y se pone la raíz molida en una infusión”(Don Rito S., Personal Communication, May 12, 2015). The laurel sumac (Malosma laurina) is a shrub that grows more than 2 m height, this plant can be found almost everywhere in the Kumiai communities and surrounding ecosystems (Wilken-Robertson 2018). This shrub has medicinal and antiseptic uses, some elders say that it has properties that can induce abortion and can be used to control irregularities in menstruation (Fig. 8). Within the native plants, we can find several leaves of the coastal flora that are part of the menu of the inhabitants. Next, we will mention some of them. The mustard greens (Brassica rapa) are one of the typical dishes that all Kumiai communities share, thanks to the fact that this weedy plant is adaptable to all types of natural and transformed ecosystems. It is commonly found in the clears of vegetation, along the roads, inside crop fields and between meadows (Fig. 9a).

186

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

Fig. 9 (a) Nonnative edible wild plant. Mustard plant in the Baja California landscape. (Photo by Carolina Gutierrez Sanchez, 2017); (b) Preparation of leaves by the elders of the native community of San José de la Zorra. (Photo by Carolina Gutierrez Sanchez, 2017)

María Ester, native from San José de la Zorra, tells us about the selection and preparation of this plant species: “From the mustard the large leaf was eaten, which is similar to that of wild spinach and it was previously cooked and combined with the meat in stew or as with the wheat that was prepared with lard or beef” “What it used to be eaten that is still consumed is mustard. To prepare it, the water is heated, and the mustard thrown and there it boils, and the water is thrown away and then they are added again to remove the bitterness. There are people who like bitterness and then they only oractice the first wash”; “De la mostaza se comía la hoja grande que es parecida a la de la espinaca silvestre y esta se cocía previamente y se combinaba con la carne en guiso o como con el trigo que se preparaba con manteca de cerdo o res” “De lo que antes se comía que aún se sigue consumiendo es la mostaza. Para prepararlo se calienta el agua y se avienta la mostaza y ahí hierve y le tiran el agua y después le vuelven a echar más para quitarle lo amargo. Hay personas que les gusta lo amargoso y pues solo con el primer lavado” (María Ester, personal communication, May 16, 2016; Fig. 9b). One of the most curious plants found in this investigation was the Kumiai lenguage (Dichelostemma capitatum). Beatriz C. says that: “These are some potatoes that when it rains come out on the hills, children like it very much. I threw them in the comal, they are then cooked, and they taste good. For everything, we look for one. To get the jalp’ap scabs, they look like good onions; you can just remove the peels and they taste so sweet. There are also like scallions on the hill but those are different and used for cooking”; “son unas papitas que cuando llueve salen en los cerros, a los niños les gusta mucho. Los tiraba en el comal y se cocían y saben bien buenos. Para todo buscamos uno. Para conseguir el jalp’ap escarbas son parecen como cebollitas bien buenas, solo las pelas y que buen sabor tienen como dulzón. También hay como cebollines en el cerro y ese es diferente se usa para cocinar.” Doña Virginia (native of the community of San José de la Zorra) says that: “cooked jalp’ap are very good, and that when the elders eat jalp’ap they became very lazy”; que el Jalp’ap, cocido es muy bueno. Que cuando los mayores comían

Wild Food and Traditional Knowledge of the Kumiai from Baja California

187

Jalp’ap se hacían muy flojos. Eso es lo que le contaba su mama, she says: “Yo creo que siempre se la pasaban en el cerro buscando el Jalp’ap, y pues no hacían nada en la casa” (Fig. 10). María (a woman native of the community of San José de la Zorra) used to take her children to collect the flowers of quiote, in the same way they took from the earth el jalp’ap: “Which is a kind of small potato that grew on the ground gave a different leaf that was identified in sight. Any edible plant cooked with egg or simply cooked in water. Those days as now, everything that was at hand was used”; “que es una especie de papita pequeña que crecía en la tierra daba una hojita diferente que a la vista se identificaba. Cualquier planta comestible se cocinaba con huevito o simplemente cociditas en agua. En aquellos tiempos como ahorita, se aprovechaba todo lo que se tenía a la mano.” Chronicle of the jalp’ap gathering (2016): Angelica, who is 7 years old, the little granddaughter of Mrs. Beatriz, accompanied us to the adventure of discovering that little delight. In the first instance, we

Fig. 10 Traditional food Jalp’ap. (Photo by Diana Galindo, 2016)

188

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

Fig. 11 Gathering Jalp’ap after a rainy day. (Photo by Carolina Gutierrez Sanchez, 2017)

found dry land and a soil full of small twigs, which gave no trace of the jalp’ap. A few meters above the hill, we set out to find the small white flower that is on the ground, most likely in the shade of a large bush. The more we climbed; we could see several of these little flowers. At that moment, we decided to dig. Angelica looked very attentive to the way in which the jalp’ap carved, since we had to be careful when digging around the flower, otherwise we could cut it and part of its stem, and that way we would lose the bulb, which is the jalp’ap (which is buried). We continue digging for 5 min very carefully and finally we find a peculiar, brown-colored ball, with a layer similar to that of ginger, it must surely be familiar to this type of bulb. By putting together a few, we wanted to try them. The taste is too peculiar, many people say that the taste is similar to that of the jicama3; however, my perception of this peculiar bulb is that it has sweet notes. For a moment I had a memory of trying a quince mixed with jicama and potato, it had an almost neutral aroma the mouth feel was very fresh. Angelica’s grandmother was the one who taught her to find and eat this delicious bulb, but at the same time astringent, notes also of nuts. For some children who know them and who were given the task of looking for them, even in these times is a real delicacy (Fig. 11).

Agaves and Cacti Different authors mention that the native people of The Peninsula of Baja California had as main element in their diet vegetables, 57% of their food was vegetables and an important component (28%) are agaves of any of the 8 species and one variety recognized in Baja California (Tapia 2002).

3

Edible tuber of Central America, Pachyrrhyzus erosus.

Wild Food and Traditional Knowledge of the Kumiai from Baja California

189

Fig. 12 Yucca costera (Hesperoyucca whipplei). (Photo by Carolina Gutierrez Sanchez, 2017)

Other edible plants found in the communities and their surroundings are Yucca and Agave, which stand out since pre-Hispanic times as a valuable and versatile plants, since all their parts can be used (Fig. 12). Yucca has a woody stem that is usually short, which is also known as head or pineapple. Before dying, agaves produce a quiote or long fleshy stem that reaches up to 10 meters in which their flowers grow (Rebman and Roberts 2012). The leaves of some species of agaves are used as food. Leaves of some species are fodder and some of them particularly useful for their fibers. Dry fibers of some agaves are very resistant and were used in the past to make clothing, work tools, and utensils for personal arrangement. From some of the regional agaves the Kumiai obtain an ingredient of their traditional cuisine, the quiote or scape of the inflorescence in which the flowers are produced. It is produced in the spring, from February to April. The quiote are cooked in the bush in bonfires as a custom of great-grandparents, according to informants from the community of San José de la Zorra. To prepare it there is a ritual; the quiote must first be obtained, in a trip to the mountain, and after cutting it, it must be cooked. Rosaura C. de S. tells us in relation to preparation of flower buds that: “This is like the maguey flower, which is also bitter, many change the water and that’s it. Then let it cool, drain and then chop and ready to pan or just like that. Mustard and the flower of quiote are foods of very strong flavors. You eat a lot and then full. That food has a lot of energy”. Two kinds of flower are found in the flower of the maguey; there is the purple one and the white one, both must be cooked, and the water thrown out. When the water boils, the flower is thrown and then you throw it away since the water is purple, it is to remove the color. (Rosaura C., personal communication, October 3, 2016)”. “Esta es como la flor de quiote, que también es amargozona, muchos le cambian el agua y ya listo. Después se deja enfriar, se escurre y después se pica y listo al sartén o así nomás. La mostaza y la flor de quite son alimentos de sabores muy fuertes. Comes mucho y lueguito de llena. Tiene mucha energía esa comida” De la flor de quiote se encuentran dos clases; esta la moradita y la blanquita, ambas hay que cocerlas y tirarles el agua. Cuando hierve el agua se echa la flor y después la tiras ya que el agua queda morada, es como para quitarle el colorcito. (Rosaura Carrillo, comunicación personal, 3 de octubre del 2016)

Another element of native food knowledge is the use of prickly pear of several cactus species. The cacti (Cactaceae family) are part of the landscape of the area

190

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

Fig. 13 Preparation of nopales in a common meal. (Photo by Carolina Gutierrez Sanchez, 2017)

(Fig. 13). Prickly pears and leaves of agaves are used. The way they differ from one variety to another is by the shape and color of the fruits. However, as time goes by, problems related to the proliferation of extracting endemic species arise more frequently, and what was observed in the research regarding native plants is the lack of assessment and regulation for the collection of native species of each community, in this case the Kumiai of the state of Baja California (Leyva and Espejel 2013).

Uses in Contemporary Food The knowledge of plant species by the traditional chefs or cooks in the Kumiai community of San José de la Zorra, Ensenada, either for medicinal use or consumption as food, is very important and coarse. The nutritional and cultural value of those plant species and their recognition by the local language is necessary to maintain and recover it (Christie 2002). The Kumiai communities have that fusion of history and traditions; they simply complement each other along with dances, song, nature, and food. The visits, participatory observation (Kawulich 2006), and application of some interviews (Sandoval 1996) helped in the reflection on the work in the community; the stories of the elderly fulfill a fundamental role in the historical construction of the traditional cuisine of the Kumeyaay (Dieste 2006). In this way, when entering the communities, a link established with the key actors within each community will help to continue the fieldwork to enrich the information that is valuable for documenting biodiversity (Albuquerque et al. 2019) of the region and a food heritage that is alive. Leticia A., from the community of San Antonio Necua, comments: “The islaya, this fruit is collected when we go for a walk arriving almost the small mountain range. Islaya is a bittersweet fruit that children and adults like very much, when they

Wild Food and Traditional Knowledge of the Kumiai from Baja California

191

just start to appear they have a color between yellow and green, you find them in the bushes, once ripe it takes a reddish color”;“la islaya, esta fruta se conseguía cuando se salía a caminar llegando casi a la sierrita. La islaya es una fruta agridulce que le gusta mucho a los niños y a los grandes también, cuando apenas empiezan a darse están entre amarillas y verdes, las encuentras en los arbustos, una vez madura toma un color rojizo” (Leticia A., personal communication, May 16, 2016). Currently, livestock and agricultural practices are still carried out within some of the Kumiai communities, mainly in the rancherías of these communities, which involves local people in the “cowboy culture.” An atmosphere of coexistence with the other neighborhood places surrounding these communities created, since communication between Mexicans and members of the native communities has direct contact. The most accessible labor market for these communities remains in the land. Some of the techniques mentioned in previous sections are still used in preparations that accompany the daily life cooking of native communities according to the wild plants production seasons. The consumption of wildflowers and leaves of some coastal thicket, still present in some festivities and special occasion where traditional cuisine is present (Gutiérrez and Von Glascoe 2019). For example, “ejotillo del monte” (Peritoma arborea), which is a native plant from the area commonly used by the Kumiai groups (Fig. 14). To cook them, Rigo Aldama part of the Kumiai group and a cowboy from la Huerta, explains: “you have to boil water and put the collected flowers inside; then drain the water, until the flower is not bitter.” Many of the current and traditional ingredients are still being produced today. Proof of this is narrated by the traditional cooks of the Kumiai communities and in the yearnings of flavors of the old people. However, food preparation techniques are still present, for example, preservation methods such as drying and cooking with stoves. Currently, preferences regarding food and preparations have changed considerably according to the modernization and worldview of their kitchens. For example, the preferences of the local people went more to the substantial of food and Fig. 14 Season of “ejotillo del monte.” (Photo by Carolina Gutierrez Sanchez, 2017)

192

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

nowadays the perception of the inhabitants of the communities attends to this reason. Rogelia native of the Kumiai group has the following perception: “Rabbits have almost no meat and that’s why they don’t crave it; like the chacuacas (quail), when they kill them, they don’t get any meat. Many people fry them for the same reason. “Right now we see these animals and then they seem cute, also why kill them if we already have other things to eat”;“los conejos casi no tienen carne y por eso no se le antojan; al igual que las chacuacas (codorniz), cuando las matan no les sale nada de carne. Mucha gente los fríe por lo mismo. “Ahorita vez a esos animalitos y pues te dan ternura, además para que los matan si ya tenemos otras cosas que comer” (Rogelia, personal communication, Julio18, 2016). Dehydration is one of the oldest ways to process food. The Kumiai conserved food in this way, they recognized they could store their products for long periods, and these were not ruined. This method of food preservation still practiced within the Kumiai communities, well; it is a way to extend the shelf life of seasonal fruits. “The ancient Kumiai preserved food and fruits through dehydration or pickles and these were placed in the sun on long willow tables, and did not lose their flavors”; “Los antiguos Kumiai conservaban los alimentos y frutos mediante la deshidratación o salmueras y estos eran colocados al sol en largas mesas de sauce, y no perdían sus sabores” (Tabita Sandoval, personal communication, Agust12, 2015).

A Landscape of Opportunities and Diversity to Be Discovered Landscapes include species that are important for ensuring dietary diversity and cultural heritage. The Kumiai culture looks for transmitting some elements through songs inspired by the times that have already passed. Both the grandparents that remain and the parents try to transmit the traditions and language to their relatives, but not everyone wants to learn, as recognized by people in the community. The process of transmitting information begins by identifying regional natural resources and landscape biodiversity, assessing the food culture and reinforcing the use of local techniques for food preparation (Espinosa 2002). Although, future actions taken to create and support programs for traditional cooks are needed, which would be able to promote their incorporation into the circuits of production and sale of a more diverse and inclusive regional food system (Meléndez and Cañez 2009). We examine that the consumption of wild foods that are introduced or non native are considered as an ecological service, but even with this escenario, wild food receives little attention due to lack of data. The integral development of a regional cuisine requires the production and sustainable distribution of local foods produced or harvested by indigenous communities, along with participation of the larger community to achieve sustained economic viability. As mentioned earlier, the role of wild food is important and

Wild Food and Traditional Knowledge of the Kumiai from Baja California

193

requires the development of an intrinsic relationship between species conservation and food security policies (Benor et al. 2010). It has been shown that the great differences between the cuisines of the world can trace limitations in the local resources, but also opportunities of ecological and cultural management (García and Bermúdez 2014). In the case of the Kumiai cuisine, it was found through fieldwork and interviews that, despite noticing preferences for meat consumption, the consumption of plants is recurrent, various seeds, stems, and cacti, contribute to make their diet diverse and nutritious. With this study, we began to understand and document the customs and traditions that revolve around the food of the native group Kumiai from Ensenada, Baja California. However, other contribution has resulted in the knowing of potential of wild food species. For a significant portion of households in native communities, wild food species form a low-cost food source, which contributes to their food access according to the seasons. Wild food use has been vital to the nutrition of the native people for centuries. The threats derived from changes in land use and overexploitation reduce the availability of some food resources, contributing to the loss of traditional knowledge associated with their consumption (Borelli et al. 2020). This kind of collaborative endeavor aims to promote sustainable education about the traditional cuisine of the Kumiai. The main objective is to make known the traditions of the indigenous peoples of the region through its cuisine and its ingredients. In the region, various efforts have been made for documenting and integrating edible native plants into a contemporary gastronomy (Huerta 2012). Such is the case of food events (First contest of native traditional cuisine, 2017) and publications, like “Plantas comestibles de Baja California” (Pijoan and Venegas 2018) and “Ethnobotany of the Kumeyaay of Baja California” (Wilken 2018) that recently revolve around food ethnobotanical knowledge and cultural heritage. Along with the Instituto de Culturas Nativas de Baja California (CUNA) on the frame of “Nativa” festival 2017–2018, we organized the first contest of Native Cuisines of Baja California, which was important for sharing food heritage that is alive. This contest allowed appreciating the creativity of the conservation techniques applied to the hunting and gathering of wild food. In addition, in the format applied to this event, the description of the dish aims to revive the native language associated with food ways and preparations (Fig. 15). In this sense, revalue traditional food heritage in the political and development agenda associated with the origins and wild foods, can promote a culinary and medicinal importance integrated to the local context. But also, a legislation is needed to regulate the use and management of some species that have a cultural and a ecological value to the local ecosystem. We identified that there is knowledge that can be shared to provide a deep connection with the local landscape. Also, strategies can be designed to integrate native ingredients provided by local ecosystems in order to promote native flora, and revalue the native people knowledge about uses and conservancy.

194

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

Fig. 15 Traditional dish, with cactus, native plants and wild rabbit, by Rito Silva. (Photo by Laura Elena Jimenez, 2017)

References Aberoumand A. Nutrional evolution of Edible Portulaca Oleracia as plant food. Food Anal Methods. 2009;2:204–7. Albuquerque UP, Nascimento LB, Soldati T, Feitosa IS, Campos JL, Hurrell JA. Ten important questions/issues for ethnobotanical research. Acta Bot Bras [Internet]. 2019;33:376–85. Álvarez M. La cocina como patrimonio (in) tangible. Temas de Patrimonio Cultural. Secretaria de Cultura, Buenos. 2002:8–11. ISBN N 987-1037-03-1. Anderson KM. Tending the wild. Native American knowledge and the management of California’s natural resources. University of California Press; 2005. Aschman H. A metate maker of Baja California. Am Anthropol, New Series, Part 1. 1949;51(4): 682–6. Aschman H. The ecology, demography and fate of the central desert of Baja California. Ph.D. thesis. Berkeley: University of California; 1954. Aschmann H. A primitive food preparation technique in Baja California. Southwest J Anthropol. 1952;8(1):36–9.

Wild Food and Traditional Knowledge of the Kumiai from Baja California

195

Barco M. Historia Natural y crónica de la antigua California: Edición y estudio preliminar, notas y apéndice de León Portilla. México: Universidad Nacional Autónoma de México; 1988. Bendimez PJ. Antecedentes históricos de los indígenas de Baja California. México: Universidad Autónoma de Baja California, Instituto de Investigaciones Sociales. Estud Front. 1987; núm. 14: 11–46. Benor S, Blather FR, Demissew S, Hammer K. Collection and ethnobotanical investigation of Corchurus species in Ethiopia: potential leafy vegetables for dry regions. Genet Resour Crop Evol. 2010;57:293–306. Bharucha Z, Pretty S. The role and value of wild foods in agricultural systems. Philos Trans R Soc B: Biol Sci. 2010;365(1554):2913–26. https://doi.org/10.1098/rstb.2010.0123 Borelli T, Hunter D, Powell B, Ulian T, Mattana E, Termote C, Pawera L, Beltrame D, Penafiel D, Tan A, Taylor M, Engels J. Born to eat wild: an integrated conservation approach to secure wild food plants for food security and nutrition. Plants. 2020;9(10):1299. https://doi.org/10.3390/ plants9101299 Christie ME. Naturaleza y sociedad desde la perspectiva de la cocina tradicional Mexicana: Género, adaptación y resistencia. J Lat Am Geogr. 2002;1(1):17–42. http://www.jstor.org/stable/ 25765026 Contreras HJ, Gracia AM. Alimentación y cultura: perspectivas Antropológicas. 1ª, edición. Barcelona, España: 2004.; 33 pp. Cortes REA. Análisis del conocimiento tradicional de la flora medicinal de la comunidad indígena de Santa Catarina, B.C., México. México: Universidad Autónoma de Baja California, unidad Sauzal, Ensenada; 1994. Dieste CP. Cocinas regionales: principio y fundamento etnográfico, ponencia para el VII congreso internacional de sociología rural. Quito; 2006. Ericksen PJ. Conceptualizing food systems for global environmental change research. Glob Environ Change. 2008;18(1):234–45. Espinosa DG. Metiendo la cuchara. Mujeres rurales y crisis alimentaria. Estrategias organizativas y de reproducción para el desarrollo local. Universidad Autónoma de Nayarit: Juan Pablos Editor; 2002. p. 161–92. FAO (Food and Agriculture Organization of the United Nations). The estate of food insecurity in the world YEAR: economic crises-impacts and lessons learned. Rome: Food and Agricultural Organization of the United Nations; 2009. García ME, Bermúdez G. Alimentos Sustentables a la carta. De la tierra a la mesa. Tlalpan, Ciudad de México: Comisión Nacional para el Conocimiento de la Biodiversidad (CONABIO); 2014. Garduño E. Los grupos yumanos de Baja California: ¿indios de paz o indios de guerra? Una aproximación desde la teoría de la resistencia pasiva. Estud Front. 2010;11(22):185–205. Grijalva A, Tapia LA. El imaginario colectivo Kumeyaay y sus recursos naturales. 2012. Gutiérrez C, Von Glascoe CA. “The safeguarding and diffusion of native traditional foods of Baja California, Mexico”, Falk Eivind Park Seong-Yong, Traditional food: sharing experiences from the field, United Nations (UNESCO); 2019. p. 207–20. Harris M. Bueno para comer: enigmas de la alimentación y cultura. Madrid: Alianza editorial, S.A.; 1989. Huerta RJ. Representaciones de la gastronomía de Baja California. México: Colegio de la Frontera Norte, Tijuana, México; 2012. Kawulich B. La observación participante como método de recolección de datos. Forum Qual Soc Res. 2006;6(2):2–05. Kuhnlein H, Erasmus B, Spilgelski D. Indigenous peoples food systems: the many dimensions of culture, diversity and environmental for nutrition and health. Rome: Food and Agriculture Organization of the United Nations; 2009. Leyva AJC, Espejel CMI. El Valle de Guadalupe. Conjugando tiempos. Mexicali: Universidad Autónoma de Baja California; 2013. Lucero JEJ. Determinar la calidad nutricional de la flora utilizada como alimento por el grupo étnico Kolew de Arroyo de León B.C. Tesis de grado. México: Universidad Autónoma de Baja California, unidad Sauzal, Ensenada; 1995.

196

C. Gutie´rrez-Sa´nchez and C. Leyva-Aguilera

Meléndez TJM, Cañez GM. La cocina tradicional regional como un elemento de identidad y desarrollo local: el caso de San Pedro El Saucito, Sonora, México. Hermosillo, Sonora: 17 (SPE.). Estudios sociales; 2009. p. 181–204. Morales MP. Cultura y territorialidad: aportes etnológicos para la gestión ambiental comunitaria. México: Instituto de Cultura de Baja California; 2010. Nabhan GP. Por qué a algunos les gusta el picante. Alimentos, genes y diversidad cultural. Ciudad de México: Fondo de Cultura Económica; 2004. Pijoan P, Venegas I. Plantas Nativas Comestibles de Baja California. Colección cuadernos de mercado de la cocina de Baja California. Tijuana: Culinary Art School; 2018. Piñón I. Recetario indígena de Baja California. Consejo Nacional para la Cultura y las Artes. México: Edit. Cocina indígena y popular; 2000. 135 pp. Pretty J, Pervez BZ. Wild foods. Food, politics and society. Oxford Press; 2015. p. 476–504. Rebman JP, Roberts NC. Baja California plant field guide. San Diego: San Diego Natural History Museum/Sunbelt Publications; 2012. Sandoval C. Investigación cualitativa. Bogotá: Instituto Colombiano para el Fomentos de la Educación Superior (ICFES); 1996. Schama S. Landscape and memory. New York: Vintage books; 1995. Shipek F. Kumeyaay socio-political structure. J Calif Gt Basin Anthropol. 1982;4(2):296–303. Shrestha PM, Dhillon SS. Diversity and traditional knowledge concerning wild food species in a locally managed forest in Nepal. Agrofor Syst. 2006;66(1):55–63. Sieso JP, Gómez EG, de Illescas A. Bellotas, el alimento de la Edad de Oro. Arqueoweb, 4. 2002. Small D, Ramirez R. “Ethnobotany project”, Contemporary Uses of Native Plants Southern California and Northern Baja Indians. San Diego; 2015. Tapia LA. Baja California: descripción del medio natural. Material didáctico para el curso de derecho ecológico. México: Facultad de derecho. Universidad Autónoma de Baja California, UABC; 2002. Tapia LA, Grijalva A. El imaginario colectivo Kumeyaay y sus recursos naturales. Estud Front. 2012;13(25):131–56. Wilken MA. An ethnobotany of Baja California’s Kumeyaay Indians. San Diego State: Dissertation thesis. Masters of Arts in Anthropology. San Diego State University; 2012. Wilken RM. Kumeyaay ethnobotany: shared heritage of the Californias. San Diego: Sunbelt Publications; 2018. Williams A. Primeros pobladores de Baja California. Introducción a la antropología de la península. Mexicali: Instituto Nacional de Antropología e Historia (INAH); 2004.

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre Occidental in Mexico Martha Cedano-Maldonado, Luis Villasen˜or-Ibarra, and Mara Ximena Haro-Luna

Abstract

The Wixaritari, Wixarika, or Huichol is a contemporary group of Indigenous people, one of the most traditionally preserved of Mexico and probably the Americas. It is a society whose people have lived geographically isolated since the Spanish conquest and have maintained their ethnic identity and ancestral social, political, educational, and religious structure, for over hundreds of years, through oral communication and the practice of its traditions and customs. Its language derived from the Uto-Aztec-Sonorense linguistic family, and there is archaeological evidence proving their permanence in the territory during the last 1800 years. This study documents the traditional knowledge and worldview that the Wixaritari have on useful plants they interact with. An analysis is included integrating botanical, geographical, ethnological, anthropological, linguistic, sociological, and historical works on the subject. We identified 596 plant species used by the Wixaritari, 190 of them medicinal, 131 edible, 114 ceremonial, 63 timber, and 72 with other use like fibers, fodder, soap, firewood, glue, and poison. From each category, mythological and cosmogonic aspects of the most outstanding species are described, as well as their use and plant parts used. The results suggest the need to record the Wixaritari plant traditional heritage, before it gets lost by acculturation, death, migration, and ecological deterioration of their territories. In addition, we systematized an inventory of the useful species reported, mainly addressing the edible plants and emphasize the need of conducting ethnobotanical investigations in areas with scarce or null information, for instance, the communities of San Sebastián Teponohuaxtlán, Tuxpan de Bolaños, and Guadalupe Ocotán. We also emphasize the need of generating M. Cedano-Maldonado (*) · L. Villaseñor-Ibarra · M. X. Haro-Luna Instituto de Botánica, Departamento Botánica y Zoología, Centro Universitario de Ciencias Biológicas y Agropecuarias (CUCBA), Universidad de Guadalajara, Zapopan, Jalisco, Mexico e-mail: [email protected]; [email protected]; [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_4

197

198

M. Cedano-Maldonado et al.

more applied and holistic ethnobiological research, including a greater feedback from and cooperation with the Wixaritari communities.

Introduction: The Wixaritari or Huichol Culture The Wixaritari, Wixarika, or Huichol are contemporary Indigenous people, among the most traditional of Mexico and the Americas (Bauml 1994). It is a relatively close group, which still lives geographically isolated and that jealously defends its culture and territory, from the times of the Spanish conquest to the present, promoting resistance to acculturation and independent and unusual development, especially from other cultures of the Sierra Madre Occidental of Mexico (Neurath 2003). This people have maintained their ethnic identity and their ancestral social, political, educational, and religious structure, for over hundreds of years, which they have successfully replicated, through oral communication practicing their traditions and customs (Iturrioz-Leza et al. 1995) (Fig. 1). Archaeological studies evidence their

Fig. 1 Wixaritari wearing the party outfit, which is distinguished by its striking and captivating crossstitched blanket outfit, in which they capture through symbols the mythological stories of their ancestors and their gods. Accessories such as necklaces, bracelets, and earrings are made with beads. Generally, men wear more ornate clothing than women. The allegories vary in each of the regions. (a) Nuiyama (Julia González) and (b) ‘Etsiekame (Juan López). (Photo by: Luis Villaseñor Ibarra)

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

199

permanence in the territory during the last 1800 years (Bauml 1994). Its language is grouped into the Uto-Aztec-Sonorense linguistic family, from which it derived approximately 3,000 years ago, along with that of the Pima and Pápago from Arizona, the Tarahumara, Tepehuan, and Cora people. The Wixaritari people feel united in their traditions with the Tepehuan and the Cora, calling them older brothers and younger brothers, respectively. However, the Wixaritari preserve their customs more completely and purely because they were not submitted by the Spaniards, nor concentrated in towns (Bauml 1994; Neurath 2003). Currently, the Wixaritari society is organized political and economically into five communities (Villaseñor-Ibarra et al. 2017a), which derived from four pre-Hispanic tribal groups, the Cureatsarixi, Tateiquitari, Tuapuritari, and Huautuari people (Negrín 1997). They were distributed in dispersed settlements but recognized territories. In 1723, during the colonial times, their territories were recognized as the community of San Sebastián Teponohuaxtlán (Waitia) and its annex Tuxpan de Bolaños (Tutsipa), Santa Catarina Cuexcomatitlán (Tuapurie), and San Andrés Cohamiata (Tateikie) and its annex Guadalupe de Ocotán (Xatsitsarie), all of them settled in mountainous areas and canyons of the southern portion of the Sierra Madre Occidental (Barrera-Rodríguez 2004). Current population is formed by more than 52,000 people (INEGI 2010; INALI 2008), and their traditional territory covers portions of the states of Jalisco, Nayarit, and small areas of Durango and Zacatecas (Fig. 2). These communities maintain their autonomy, with its own civil and

Fig. 2 Distribution map of the Huichol territory; in green appears the Wixarika territory in Mexico; on the right, the five most important traditional Huichol communities. (Taken from Villaseñor-Ibarra 1999)

200

M. Cedano-Maldonado et al.

religious authorities and their own social structure. All of them are intimately linked retaining stable key cultural traits (Villaseñor-Ibarra et al. 2017b). The Wixaritari have particular perceptions about the world, which are described through myths and stories about their people (Kindl 2013). The myths associated with useful or significant plants in their culture may be associated with good or bad mythical characters. Their main cultural symbols are the corn-peyote-deer trilogy (Guzmán-Mejía and Anaya-Corona 2007). Their idiosyncrasy is based on three aspects: 1) religious festivities, which allow to fulfill “el costumbre” (the custom), 2) the tradition, which governs the behaviors and encloses a particular worldview constituted by a magical-religious knowledge, and 3) “el costumbre,” which are all the basic norms of behavior, ceremonies, and justice (Barrera-Rodríguez 2004; Torres 2000).

The Wixaritari Territory The Wixaritari or Huichol region covers an area of approximately 3921 km2, included in 16 municipalities and four states of Mexico. In the state of Jalisco, the Wixaritari territory overlaps with the jurisdictions of Bolaños, Colotlán, Chimaltitán, Huejucar, Huejuquilla El Alto, Mezquitic, San Martín de Bolaños, Santa María de los Angeles, Totatiche, and Villa Guerrero. In Nayarit, the territory is part of Del Nayar, Santa María del Oro, and La Yesca. In Zacatecas, it belongs to Monte Escobedo and Valparaíso, and in Durango it is part of El Mezquital (Vázquez-García et al. 2004). The Sierra Madre del Sur, the Sierra Madre Occidental, and the Trans-Mexican Volcanic Belt converge in the Wixaritari territory, which makes up a geomorphology with high mountains, deep canyons, steep plateaus, extensive mountains, semiarid areas, and a great diversity of ecosystems, with elevations ranging from 400 to 2800 m (Fig. 3). Its relief is formed by two macrobasins, one of them the Atengo river basin and two tributaries confluencing with the Huaynamota river, the Camotlán, and Huajimic that become the Chapalagana river, and the other is the Bolaños river basin (Barrera-Rodríguez 2004). Temperatures are greater than 30  C in summer and less than 5  C in winter, with maximum rainfall of 600 to 1000 mm and in some zones up to 1200 mm. The natural wealth of the area presents a mosaic of landscapes mainly consisting of pine and oak forest communities in mountainous areas. There are also tropical deciduous forest on the slopes of the mountains and deep valleys, another forest type along the canyons, wetlands at the elevations of the sierra, crassicaule scrubland in arid and semiarid areas, and zacatal (natural or induced grasslands) in flat areas of the plateaus and highlands. In addition, in some small areas there are aquatic vegetation, chaparral, tall xerophytic vegetation of the Chihuahuan desert, thicket of Dodonaea viscosa, and spiny forest (Luquín et al. 2004). This chapter aims to make a compendium on the knowledge, beliefs, traditions, utilities, and assessments that the Wixaritari ethnic group has on plants, to recognize the relationships, the position they occupy in their culture, and the cultural identity

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

201

Fig. 3 View of the Sierra Madre Occidental from the Tateikie community (San Andrés Coamiata, Mezquitic, and Jalisco). (Photo by: Luis Villaseñor Ibarra)

granted to them. We in addition aimed to establish the current state of knowledge and research needed in relation to the Wixaritari or Huichol ethnobotany.

Methodology We reviewed and contrasted botanical, geographical, ecological, ethnological, anthropological, ethnographic, linguistic, sociological, and historical studies on the Wixaritari ethnobotanical knowledge. Methodologies include qualitative analysis of the data collected through ethnographic observations, field notebooks, database consultations, interviews, and botanical explorations. The quantitative analyses are related to the diversity and floristic richness of the region. However, it is necessary to emphasize that some floristic and ethnobotanical works report incomplete and inaccurate information on the useful flora of the Wixaritari region. Generally, botanical investigations record the scientific names of the species, but lack basic ethnobiological information like common names and uses. Contrastingly, ethnobotanical studies provide more information on cultural aspects but suffer from inexact taxonomic identification, and only the medicinal plants have been detailed. This situation makes it difficult for this work to generate a general list of the species and precise data on the useful plants hitherto recognized. The documents analyzed are presented in the literature section. The section “Edible Plants” shows original information obtained through interviews, the corroboration of plants in the field and collection of samples as supporting material for their identification. The identity of species not collected but whose Wixaritari names were recorded were corroborated with local people, using field guides. Subsequently, the information was contrasted with the floristic listings generated for the region.

202

M. Cedano-Maldonado et al.

Relations Between the Wixaritari and the Plants In the Wixaritari worldview, the myth about the origin of plants is included in the legend of the origin of the universe, in which Watákame, or “the first sower,” when arriving on land, dropped the jícara (Fig. 4), which contained all beings living that would populate the world (Neurath 2005). Plants, fungi, animals, and the elements of nature continue due to the disembodiment of their ancestors, and as the members of their bodies were watered, they created the fruits of the earth that, since then, keep the Wixaritari or Huichol people. To preserve the spirits of nature, people have to perform celebrations, offerings, physical deprivations, and various rites in order to restore strength to the gods and prevent them from getting angry because of the lack of thanks and devotional attention. When people die, their lives will be judged for the fulfillment of customs and their behavior. If the person who died showed an exemplary existence, they will become divine spirits, living together with the gods in Huiricuta (Wirikuta), forming part of the sun’s rays. But if the person failed, their descendants would appear in the form of insects, coyotes, or owls (Negrín 1997). One of its deities is the goddess of vegetation Takutsi Nakawé, a female monster that reigns in the rainy season; she is the maker of the world, governing the forces of darkness of the world, before the igneous and solar powers (Villegas 2016). From the western view, it can be confirmed that the existence of the Wixaritari has always depended on nature. They have a traditional pattern of life that adapts to the rugged landscape of the mountains, which consists in the organization of “ranchos,” or groups of houses with a small population, which are freely dispersed throughout their territory (Fig. 5). This settlement pattern has prevented the depletion of water and existing resources, as well as the simultaneous use of different ecological floors (Neurath 2003). In addition, they exhibit a great knowledge about the environment and the biological wealth present in their territory. They know which, how, and when natural resources are available and how to take advantage of different species. Likewise, their forms of production depend on the natural conditions of the Fig. 4 Jicara is a bowl made with a half of the dried fruit of Lagenaria siceraria or Cucurbita argyrosperma. (Photo by: Mara Ximena Haro-Luna)

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

203

Fig. 5 Huichol ranch in the municipality of Mezquitic, Jalisco. (Photo by: Luis Villaseñor Ibarra)

environment, their beliefs, customs, and cultural traditions and only cover the subsistence needs of the population. For this reason, the harvest is destined for self-consumption, to carry out traditional festivals and as fodder for livestock or raising domestic animals (Barrera-Rodríguez 2004). However, each community safeguards its own knowledge about native flora and fauna of the region. Generally, elders of the family are the ones who have most information about elements of their environment, including an extensive classification of use types of natural resources, and they also have more techniques for environmental management and preservation. The traditional environmental knowledge is transmitted among family members and the community member (particularly important is that to the new generations), through talks, songs, and ceremonies that can last for a whole night and even more. During the broadcast, events from their origin as Wixaritari people are narrated. In the mythology, they consider that the gods themselves taught the Wixaritari or Huichol to live. Gods were the first to do everything: cultivating, hunting, and building, among other activities; therefore, now they replicate those Gods’ teachings. They also consider the natural phenomena and elements like rain, wind, plants, animals, and others as governed by deities (BarreraRodríguez 2004). The floristic studies carried out in the Wixaritari region register a biological wealth of about 2081 plant species and 139 infraspecific taxa (Vázquez-García et al. 2004). The ethnobotanical studies have reported the use of nearly 600 species, the Wixaritari useful flora, which includes the use of wild plants, archaic

204

M. Cedano-Maldonado et al.

domesticates, and some introduced plants (Bauml 1994). These species come from wild habitats, mainly pine and oak forests, oak, tropical deciduous forest, grassland, crop fields, family gardens, orchards, and other agricultural fields (Nieves et al. 2004a). The valuation of the species by categories is variable; Bauml (1994) organized a checklist of useful species of the region in 12 use categories, while Nieves et al. (2004b) identified 18 categories. For practical reasons, in this chapter we regrouped plants into six use categories, as shown in Table 1. This instrument was built with the information from different works. The medicinal use with 190 species being the most numerous and valuable category, followed by the edible with 131 species, the ceremonial plants including 114 species, those used as timber that are 63 species, and those that we include in the category of “others” comprising 72 plant species with little known uses with low-use value. However, the information is still uncertain and generates a bias in the final results, due to several inconsistencies perceived in most publications, which do not include lists of species and lack species identification; in some studies, many plants were registered only with Huichol names. In some studies, there is no direct report of uses of each species, which may have more than one use and be counted as different species. These are some of the most common problems found in the studies reviewed. The following text describes the categories established in Table 1 based on information on the most relevant species. Table 1 Categorization of the plants used by the Wixarika ethnic group. Reference: (Bauml 1994; Bauml 2004; Bye et al. 2005; Casillas 1990; Cedano-Maldonado et al. 1998; Cedano-Maldonado et al. 2017; Higareda-Rangel et al. 2004; Nieves-Hernández 2002; Nieves et al. 2004b; and Villaseñor-Ibarra et al. 2017a) 1, 4, 18, 19, 23, 32, 34, 35, 37, 38 Categories included Artisanal Ceremonials

Edibles Timber

Medicinals Others

Total

Subcategories included Utensils, household items, and toys Bow and arrows Arrows Ornamentals Pigments or paints Sacred/religious Meals and condiments Drinks Siege Construction/sacred Tools Ropes or fibers Fodder Soap Firewood Glue Poison

No. of plants 22 4 2 40 17 55 112 19 21 21 21 190 11 21 2 21 6 11 596

Total 26

114 131

63 190

72 596

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

205

Craft Plants For the manufacture of handicrafts, people use 26 plants species (Nieves-Hernández 2002). Cucurbita moschata is used to make bowls or jícaras and is the only taxon that is cultivated; the rest of the species are collected from wild environments. Such is the case of Bothriochloa laguroides and Muhlenbergia aff. grandis used for making brooms (Bauml 1994; Nieves-Hernández 2002). With the stems of Otatea acuminata (hacu ucáari), local people weave baskets, flasks, or rod beds, with the leaves of the sotol (Dasylirion spp.); they weave petates, rugs, or the traditional benches called u’pali, which have a reed support and a strong wooden hoop crafted with mesquite or pine wood, bonded with rubber obtained from Bletia sp. (kwe’tsaka) (Lumholtz 1900) or Bletia macristhmochila (cuesucua ucáari) (Bauml 1994). Also, the Wixaritari use cotton-like fibers of Ceiba pentandra (kapo’ri) and ixtle (mai’ra sikuliaiya) to make textiles for common or ritual use (Lumholtz 1900). However, most of these manufacturing practices are obsolete, and it is rare to observe them today.

Fig. 6 Mask made from Quercus sp. (a) and stamen, in which some elements of the drum dance are represented: corn plants, a three-legged drum, a bule (fruit of Lagenaria siceraria), and eagles. (Photo by: Luis Villaseñor Ibarra)

206

M. Cedano-Maldonado et al.

At present, the crafts frequently of most elaboration are masks, pictures, or figures of yarn or chaquira, which are made with wood of Quercus spp., Pinus spp., Arctostaphylos pungens, or Prosopis juliflora, and these can be of ritual or commercial use (Fig. 6). Generally, those that are sold lack meaning or symbolism and are only repetitions of figures that they like best, to be bought by tourists (Neurath 2005). In the Wixaritari mythology, the sotol stands out and is associated with a mythical evil being, and they report that in ancient times there was an evil giant woman who devoured people. One night, a brave man tired of mistreatment by the lady took a honeycomb and, in the company of other men, went to where that woman slept, they opened her head and put the honeycomb in it. The next morning, the giant woman could not torment people anymore, because her head hurt her too much, and bees came out of her ears. Because of her despair, she lost her mind and people of the town took the opportunity to throw her through a ravine. When she fell, her head detached from the body, and in all those places where her head bounced, sotol plants sprouted because they are equal to her hair [according to the story of Mrs. Josefina T.].

Ceremonial, Sacred, or Religious Use Plants This group of plants consists of 114 species representing the third most important cultural use for the Wixaritari ethnic group, but qualitatively it may be the most relevant (Table 1). It calls to attention that the ceremonial use stands out from other aspects being more useful for survival, and it can be an attribute that distinguishes the Wixaritari from other cultural groups in Mexico. However, by analyzing the mystical and ritual context of the Wixaritari, one can understand the reason of its significance. In the Wixaritari cosmogony, a great number of Gods are venerated. Those who gave them their origin, those who govern all natural elements, and those who taught them to live and do things as they do, to mention a few. To all of them, they should be adored and honored in a special way, so they have a ceremonial and mythical calendar that is divided into two periods: that of the dry season (winter and spring), with parties held at the head of the community. The Wixaritari have established dates that are governed by the Catholic calendar, such as the New Year or Change of Wands, Carnival or Pachitas, and Easter, which are celebrations offered to male deities, such as the sun, fire, deer, and closing with the peyote party. They start in the rainy season (summer and autumn) with parties of the type neixa or mitote (includes one or several nights of shamanic singing and circular dance), which are held in ceremonial centers or in the shrines. These celebrations have movable dates and are dedicated to the Mother Earth honoring to female divinities, such as the goddess of fertility, fruits, plants, and water, to mention a few (Neurath 2003) (Fig. 7). All parties are run by the “mara’akate” or shamans; they sing-dialogue with the gods to convince them to do favors to the people present, but they are also responsible for preserving and keeping alive the traditions, rites, and religious

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

207

Fig. 7 Community meeting in the central plaza of the callihuey or ceremonial center of Tateikie, to start celebration of the “Fiesta del Tambor.” (Photo by: Luis Villaseñor Ibarra)

holidays (Neurath 2003; Villaseñor-Ibarra et al. 2017a). Each celebration is aimed at giving back, thanking, and pleasing the gods for the effort they made to generate all the good that happened in their lives during that time, such as abundant harvest or hunting, health, or well-being (Barrera-Rodríguez 2004; Guzmán-Mejía and AnayaCorona 2007). Only the neixa or mitote type parties are held in the “tukipa” or “callihuey,” communal ceremonial centers of pre-Hispanic tradition dedicated to the worship of the ancestors of the community that have deified (Fig. 8). These sites are composed of several buildings grouped around a circular dance square. The main temple called “tuki” is distinguished by presenting larger dimensions and a circular or oval shape; the other shrines, or “xirikite,” are parental, smaller, and have a rectangular structure. The Tupika are also social centers and initiation schools, since during the holidays they are staged and rites are performed about the cosmic truths of what happens in nature. The experience that the original settlers of the ceremonial center had during their mythical journey is repeated near the Pacific Ocean toward the Wirikuta desert, in San Luis Potosí (Neurath 2003). Throughout the Wixaritari territory, there are 20 ceremonial centers, where several gods of the community live and where the concept they have of space and time is represented. The interior represents the dark place of origin, the underworld and the sea. The square where the dance takes place corresponds to the Wirikuta desert and the Tamatsi Parietsika shrine “Our Big Brother, the one who walks at dawn”: the deer-peyote god. Each oratory or xirikite is the abode of a Wixaritari god, and there, the sacred jícaras are protected with objects that represent it, which are protected by the “xukuri 0 ikame” or “jicareros,” who adopt the deity as their own corresponding jícara, for 5 years (the duration of the position); at that time, they live

208

M. Cedano-Maldonado et al.

Fig. 8 Wixarika ceremonial center of Rancho de En Medio, Villa Guerrero, Jalisco, where “xirikite” or small temples can be seen. (Photo by: Mara Ximena Haro-Luna)

there and become the ancestor or god who has to personify during the holidays (Neurath 2003). Each celebration or rite includes the participation of a group of plants that provide magical and protective powers. Some represent the gods themselves, others serve to purify the spaces and the spirit, and with others, people make the offerings. In addition, there are special rituals to build the temples or niericas and others to prepare the dyes that allow them to decorate their faces for some festivities. Among the most important plant species, which is considered divine, Lophophora williamsii, also called “hikuri” “hikuli” or “peyote,” is one of the central elements of culture and religion (Anderson 1980; La Barre 1975) (Fig. 9). It has psychoactive alkaloids; the best known is mescaline, which induces and alters the state of consciousness accompanied by colorful visions (Anderson 1980). In the Wixaritari cosmogony, three great mythological cycles are described, the deer included in the first one, which deals with the events that occurred to their ancestors from the sea to the place of the dawn, from the creation of the sun and the first deer hunting, in which the peyote is given voluntarily to the hunters and becomes the deer (Neurath 2003) (Fig. 10). This episode is venerated at the Hikuli Neixa festivity “the dance of peyote,” which is celebrated at the end of the dry season. During this celebration, the jicareros perform a complex dance that expresses how the deer-peyote is hunted, collected, blessed, and ritually shared. But first, they must obtain and collect it.

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

209

Fig. 9 Lophophora williamsii, “Hikuri” or “peyote,” considered a sacred plant and one of the most important gods of the Huichol culture and religion. (Photo by: Mara Ximena Haro-Luna)

For this ritual, the jicareros of several temples undertake and direct a pilgrimage toward the east, to the Wirikuta or Hiricuta desert in San Luis Potosí, following the routes drawn from immemorial time. During that trip, the jicareros are known as peyoteros (hikuritamete); they wear a special dress, which includes a hat adorned with white turkey feathers, which resemble the flowers of the peyote, and to purify themselves, they practice fasting, confession, and abstention of dream. When the hikuri is ingested, nierika is obtained “the gift of seeing.” According to their mythology, the ancestors were the first to try hikuli and have a visionary experience, after which they became gods. By reliving all this experience, the hikuritamete can become mara’akate or “shamans,” initiates, traditional doctors, or singers. In the middle of the desert, the hikuritamete also dream of the Nia’ariwame, the snake goddess of the eastern rain, whom they take back to their communities when they return from the pilgrimage, along with a large number of peyote plants to be used in various ceremonies, as medicine or to perform several types of cooperative labor that require their strength (Bauml 2004). The Huichol consider peyote to be the heart of the gods, because they first tasted it to find out its taste, to determine the amount to eat and the duration of its effects. So, when they consume it, they reveal some of their attributes and also reveal their hearts and thoughts. The celebration of the hikuri also corresponds to the first of the three main festivals that are related to the three critical moments of the corn crop cycle, which concerns the preparation of the coamil (Neurath 2003).

210

M. Cedano-Maldonado et al.

Fig. 10 Wooden hikuri in which chaquira and stamen represent symbols of the first mythological cycle of the Huichola cosmogony, of which the creation of the sun and the first deer hunt stand out, which is voluntarily given to hunters and becomes in peyote. (Crafts made by: Fidel de la Cruz †. Photo by: Luis Villaseñor Ibarra)

The corn (Zea mays) has the same cultural value as hikuri and is considered “Tatei Niwetsika,” the goddess, described in the second mythological cycle of the Huichol cosmogony (Fig. 11), which talks about the flood, the creation of corn, and the origin of human beings. During the narration, it is described how the first cultivator “Watakame” looks for the corn goddess “Tatei Niwetsika,” the way he is saved from the flood and how his dog becomes the first woman (Neurath 2003). Tatei Niwetsika is symbolized by bundle ears selected for their perfection and stored as seeds. There are two main parties in his honor and mark two of the three critical moments of the corn crop cycle: That of Namawita Neixa marks the beginning of the season and is done with the time of planting. Tatei Neixa’s “Our Mother’s dance” marks the end of the season, the ritual presentation of the first fruits (corn and tender gourds), and children under 5 years of age are celebrated. These parties are accompanied by tepu, the drum, and the champaign of the shaman is more complex and long than at the Peyote Festival, since it guides them through an imaginary trip to the Wirikuta desert and, at a certain moment, the children recognize themselves with the first fruits and are presented as such to the gods. Then, a symbolic separation is established between human beings and their food. Thus, it is also an initiatory rite to the life of pilgrims (Neurath 2003). Maize is a species originated and domesticated by Indigenous peoples of Mexico for more than 5,000 years (Benz 2001). Among the Wixarika culture, its cultivation is the most important and ancient religious practice, and most of its harvest is

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

211

Fig. 11 Nierika or stamen box representing the “multicolour” or “pinto” Tsayule corn and representing the center of the cosmos, made by Xikitakame (Mariano López de la Cruz). (Photo by: Luis Villaseñor Ibarra)

considered sacred (Bauml 1994). Mainly, it is intended for the self-consumption of those who cultivate it and to prepare food for traditional communal festivals. Only a small portion is used for the breeding of domesticated species that they own (Barrera-Rodríguez 2004). The cultivation of up to eight agricultural varieties has been recorded, but five of them are considered sacred, because they are related to the five directions of the cosmos. Maize varieties are named according to their color: Yuawime “blue,” which corresponds to the south; Tusame “white,” to the north; Tatlawime “purple,” the west; Taxicaawime “yellow” (Fig. 12), to the east; and Tsayule “multicolored” or “pinto,” to the center. In the Xirikite shrines, these ceremonies are headed, but only in those that have descendants of any of the irikate or “people-arrow” and that are protected in the jícaras. At present, the cultivation is carried out by the traditional and ancient method known as rubbing, graving, and burning (Neurath 2003). The third plant most mentioned by the Wixaritari is the “Kiéri” “Quieri” or “Planta del Dios,” and that may correspond to Solandra brevicalyx, S. grandiflora, or S. guttata. These species exhibit phytochemical, ethnobotanical, and ornamental interest (Bernardello and Hunziker 1987). In Mexico, their use is reported in several Indigenous cultures, among them, the Wixaritari. However, there are discrepancies about the species managed. Although the majority of the authors who speak on the subject point out that the species used is Solandra brevicalix (Bauml 2004), the taxonomists of the group mention S. grandiflora (Bernardello and Hunziker 1987),

212

M. Cedano-Maldonado et al.

Fig. 12 Wixarika girl helping to shed yellow corn, one of the five sacred varieties and which represents the west of the cosmos. (Photo by: Mara Ximena Haro-Luna)

but the herbarium specimens collected in the Huichol region were identified as S. guttata (Nieves et al. 2004b), while other studies suggested that these could be Datura or Brugmansia species because they have similar hallucinogenic effects (Bauml 2004). According to Bernardello and Hunziker (1987), there is archaeological evidence that supports the theory that its use as hallucinogenic plant is prior to peyote. They also point out that all species contain psychoactive compounds to a variable degree. Although S. grandiflora is the one that produces the most, having psychoactive alkaloid compounds such as atropine, noratropin, hiosciamine, and tropin, present throughout structures plant in different proportions. In the Wixaritari mythology, Kieri was an evil sorcerer, deceiving since he was a baby, because he danced before the people and told them that it was good, and they ate their hallucinogenic leaves, sap, and fruits. But he was bad; he drove them crazy and made them believe they could fly, but it was not true. This happened until Kauyumarie, the hero of the divine culture and Venado Person, the ally of the hikuri, who kills him with the fifth arrow, arrived. When he was dying, he vomited and transformed into a plant that flew to live among the boulders of the Sierra Madre like the Wind Tree (Furst 2007). Therefore, Kiéri, in fact, did not die but is the incarnation of the God of the wind, and witchcraft “kiéri tewiari” is attributed with benign and, at the same time, perverse powers (Neurath 2003; Negrín 1997) (Fig. 13). It causes narcotic and hallucinogenic effects; however, special powers that cannot be

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

213

Fig. 13 Flowers of Solandra sp.; the plant is called Kieri, “Plant of the God,” or “God of the wind and witchcraft”; it causes narcotic and hallucinogenic effects and is used to treat matters of witchcraft. (a) Nierika showing the shape of the Kieri, made by Muwieri (Alejandro García). (Photo by: Luis Villaseñor Ibarra)

obtained through peyote are required to deal with matters of witchcraft and counter terming. Wixaritari shamans consider it one of the most powerful and dangerous plants they know. Sorcerers use it to make spells, using substances or pollen to introduce it into the victim’s body, causing them dizziness and disorientation. They mention that there are very few who come to the kieri, because there are many dangers and sacrifices demanded by this powerful deity and that while the whims are fulfilled, wealth can be expected, the gift of music and art (Alcocer and Neurath

214

M. Cedano-Maldonado et al.

2004), triumph in love, or help to deer hunting. Even a piece of the plant can be wrapped as an amulet for luck; it can also reveal the cause and cure a disease. Their favors are invoked by placing offerings at the base of the plants or tying them to the branches. Appropriate songs are sung, food and drink are left, and candles are lit before pilgrims can leave freely (Bauml 2004). Within the group of plants that serve to purify, Nicotiana rustica (wiainuri or ya) or sacred tobacco is particularly relevant. It has hallucinogenic properties and is associated with the God of Fire, and it is said that the Elder Brother Tamatsi Kauyumari created it in a seedbed, its seeds spread in the ground, because he wanted to create something to smoke. The first plants that appeared were not good, because the smoke generated by its leaves was too light. Over time, Kauyumari managed to create a good tobacco, which was tested by Tatewarí “El Abuelo Fuego” and Taweviékame “Creator of the Sun”; they called it “ya” and considered it sacred. The Wixaritari mention that there are female and male plants, the latter being stronger. This sacred tobacco is only cultivated, transported in sacred recipients called bules (fruits of Lagenaria siceraria), and smoked by shamans in purification rituals in which they must sing the whole night (Negrín 2006). Its leaves are smoked in dried corn leaves, although some authors point out combinations with Tagetes lucida (tímutsari) or Santa Maria or pericón, similar to the Aztec narcotic called “yahutli” (Bauml 2004) or with peyote, as a curative used to reduce their tension (Moreno-Coutiño and Coutiño-Bello 2012) or as a formula to achieve more lucid visions; they smoke the first, and they chew the second drinking alcohol. The sorcerers use it smearing their arrows in the juice of tobacco so that they take diseases to their enemies. But the shamans carry the tobacco leaves on their sacred journey to purify the places where they will chew the peyote (Barba-Ahuatzin 2004). Other species of the genus Bursera spp. or glasses are used for this purpose. The drops of resin (ukua) they produce are placed in a three-legged smoker (putsi); the smell generated by the smoke when burned is considered food for the god Cauyumarie (Negrín 1997). For the elaboration of sacred objects, the species of Lagenaria sicerarea or the bule (Fig. 14) is used. An entire fruit is split into two; jícaras are obtained, each conceived as a matrix where life is born; the one below corresponds to Mother Yurianaca, the fertile land. The other corresponds to Mother Huerica Huimari, the celestial vault, and the Huichols live in the middle of the reconstituted bule. But if the bules are split at the base, then they become jícaras that receive blood and water, essential objects of Huichol symbolism. Also, with Cucurbita argyrosperma (xucuri) or pumpkin, jícaras or rucuri are made, the interior is lined with wasp wax, and the exterior is decorated with elements that represent all mother goddesses. They include seeds, shells, paper flowers, glass beads, stamen or wild cotton pom poms, and prayers. Also, in these jícaras the offered foods are added. With them, the family union and the fruits of the earth are represented. Also, common objects such as food dishes and water vessels can be manufactured (Negrín 1997). The species Phragmites australis (haka), or reed, is considered one of the oldest plants on earth and created by the Grandmother of the World (Santos-García and Carrillo 2012). When the world was dark and full of water, the Cauyumarie bandage

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

215

Fig. 14 Dried fruit of Lagenaria; it is used to make utensils or sacred objects, such as jicaras naming Bule. (Photo by: Luis Villaseñor Ibarra)

was born inside a reed, which forms the shaft of each sacred arrow. With it, the bodies of the votive arrows are manufactured. When a woman asks the gods to grant them a male child, they deposit a votive arrow in a sacred place, usually in a water eye. But if a female is desired, then she must offer a jícara. The species of the genus Otatea spp. are plants related to the Goddess of Growth because she carries her cane of otate. It is used to make different nierikas or objects that allow seeing the invisible. Some are in the form of a ring, in which a net is woven, and votive arrows are hung. Others, called niericate or flat discs with the center uncovered made with reed and wet otate, were colored yarn interwoven. With these same materials, the religious figure of the namma was also constructed; it has yarn frets interwoven with the otate and the reed; they are rectangular or square in shape, and feathers were fitted on both sides, which serve as a protective shield. These are the most elaborated and colorful votive offerings and represent a front shield that shows the face of a god; usually symbolic and mythological figures are represented (Negrín 1997). Generally, the plants chosen as offerings have notorious qualities in their flowers. They represent the gifts of life-giving gods, their words, thoughts, blessings, and prayers (Negrín 1997) (Fig. 15). Some of them stand out for their fragrance such as Polianthes platypylla (teaxuxuuri uquis) or wild tuberose and Macrosiphonia hypoleuca, and some for the beauty of its flowers to put in the jícaras of the peyoteros, flowers of Yucca spp. (xunuuri) and Tillandsia caput-medusae. So that lightning does not fall in the house, in the cornfield, or in a sacred place, Bessera elegans

216

M. Cedano-Maldonado et al.

Fig. 15 The Huichols give flowers to their gods, because they represent their words, thoughts, blessings, and prayers. Nierika made by Xikitakame (Mariano López de la Cruz). (Photo by: Luis Villaseñor Ibarra)

(xauricáa ucáari) is offered to honor the goddess Nüarivame and Milla biflora (siicütame ucáari) for El fuego de Tatei Núarivame. Nieves et al. (2004b) suggest that many of these species are perhaps selected for their striking colors and the large sizes of their flowers, since 32 of the species they report meet these particularities; some of them belong to the Asteraceae (7), Fabaceae (6), Liliaceae (6), Scrophulariaceae (5), Lamiaceae (4), and Orchidaceae (4). Among the latter, Laelia speciosa (cuaiyuri) is mentioned to decorate temples at the feast of corn, altars, and hats; of Spiranthes aurantiaca (sisicalaque tataa), the flowers are stirred with sacred tobacco and loaded into the sacred bules to go to Wirikuta, and then they are burned. Another species that attracts attention is Tagetes erecta, (puvaari) or cempoalxochitl, which asserts that its cultivation is archaic and that it was domesticated in the Huichol region, with magical-religious purposes, for some rituals and ceremonies. The interest of their domestication had to do with their inflorescences, as they are one of the offerings to decorate the temples in the ceremony of the end of the harvest, exclusively, in the celebration called “Tatei Neixa” or “Yuima cuari.” In addition to ceremonial use, its leaves are prepared into a medicinal infusion to cure headache (Bauml 2004). Finally, in some religious ceremonies people make use of sacred yellow paintings, to decorate their faces with symbolic designs or simply paint it. In the pilgrimage to Wirikuta to search for peyote, people use the species Berberis trifoliata var. glauca (úxa) with which an intense yellow dye is made (Bauml 1994; Bauml 2004; Negrín 1997). In the festivity of the Pachita or of the Flags, celebrated on Friday of the Holy Week, the dried flower of Cochlospermum vitifolium (ramoakari) is used, which is ground with a small amount of orange peel, to give it a pleasant fragrance. Subsequently, the resulting powder is sprinkled on the face of men,

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

217

Fig. 16 Flower of the Cochlospermum vitifolium tree called “Ramoakari,” the main or only ingredient of the yellow dye used in the ceremony called “Fiesta de las Banderas” or “Fiesta de la Pachita”: (a) Grinding of the dried petals, (b) dye obtained, (c) dye bath, and (d) appearance of the dye on the face. (Photo by: Luis Villaseñor Ibarra)

women, and children who fasted the night before the celebration, as well as the flags used there (Cedano-Maldonado et al. 1998) (Fig. 16).

Edible Plants Although information on edible plants is scarce, the Wixaritari consume more than 131 plant species as food (Bauml 1994; Nieves-Hernández 2002). Most of the taxa are wild, very few of them are introduced species, and only a few, such as corn (hico), beans (mume) (Fig. 17), squash ( jutzi), chili (haacuucuri), and chayote and avocado (yeuca-te), are cultivated, coming from an ancient domestication. Before using cultivated chili varieties, they consumed a wild species they call puruhi (Diguet 2005). Due to the diversity of ecosystems that exist in the region and the rugged orography in which the different Wixaritari communities are distributed, there is no homogeneous knowledge in this regard. As for the species considered edible, they make use of the plants that they obtain more easily growing in areas close to their community (Barrera-Rodríguez 2004). What all the communities share is the classification of edible plants as food, including vegetables associated with salty flavors and that can be consumed mostly raw, fruits associated with sweet flavors, condiments used in different foods and drinks, roots or sweet potatoes prepared in a salty or sweet way, nuts consumed as snacks and as an ingredient to prepare mole or pipian, and finally, nutraceutical foods, that is, those that have nutritional and therapeutic action (Bauml 2004, field data by the authors).

218

M. Cedano-Maldonado et al.

Fig. 17 Wixarika girl helping to clean the mume or bean (Phaseolus sp.) They grow on their land. (Photo by: Mara Ximena Haro-Luna)

As part of the Wixaritari mythology, there is a legend that mentions the participation of edible plants to save the earth from a new flood. The story begins with Tacutsi Nacahué “the Great Grandmother,” the mother of all ancestor deities and the matrix of all female spirits on earth and in the water, and which triggered the great flood. To calm her anger and not flood the world again, she demanded from Cauyumarie “Elder Brother Venadito del Sol” that she will be offered a blind and crippled boy, a peccary, an iguana, a black puppy, and a colored wooden canoe (kwaixruari) loaded with offerings, such as pumpkin nipples and plants, seeds, and corn. When the requested gifts were delivered, she consented to disappear into the ocean, and at that time the Spirit of the Waters of the South was transformed into rock (Negrín 1997). People from the different Wixaritari communities consider the plants from which they consume raw leaves and stems (e.g., Convolvulus sp., Medicago sativa, Oxalis hernandenzii, Zornia reticulata, and Manihot rhomboidea), as well as other greens, which they call quelites, such as Amaranthus hybridus and Portulaca oleracea.

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

219

These are used to accompany savory dishes such as pumpkin (Fig. 18), meat, beans, or mushrooms. Also, within the vegetables they include the tender shoots of trees like Enterolobium cyclocarpum and Leucaena macrophylla that are consumed in broths (Bauml 2004; field data by the authors). The traditional diet is complemented with young pods or immature seeds of Asclepias spp., Amoreuxia palmatifida (taraki y zurati) (Fig. 19), Bauhinia pringlei and Ceiba sp., and with different fruits of wild trees like Leucaena leucocephala, Pithecellobium dulce, Arctostaphylos pungens, Prosopis laevigata, Psidium guajava, Casimiroa sp. Spondias purpurea, Persea americana, and cactus as Steneocereus queretanoensis (Maára), Heliocereus speciosus (hapani), and Opuntia spp. From the group of introduced plants of the old world, the Wixaritari consume Mangifera indica, Musa paradisiaca, Prunus persica, and Cucumis sativus, to mention a few (Bauml 2004). Most of the species are present at different seasons of the year, and there is no processing to eat them. In addition, it is common the consumption of seeds of Amaranthus spp. (quiaoja), pine nuts of different species of Pinus, and nuts of Juglans regia and Juglans major var. glabrata as a snack, throughout the day or during long walks (Bauml 2004) [field data by the authors]. It is common to consume different species of Opuntia spp. (nakari) either the pencas (cladodes) or the cactus hearts, which are the inner part of the most mature

Fig. 18 Mrs. Robertina eating cooked pumpkin, fruit of the Cucurbita argyrosperma species. The plants are cultivated by the Huichols and are one of the most appetizing foods tasted by them. (Photo by: Mara Ximena Haro-Luna)

220

M. Cedano-Maldonado et al.

Fig. 19 Amoreuxia palmatifida in its natural habitat, called by the Huichols “Taraki”; the plant is consumed raw (its immature fruits), and a sauce with chili is made with the seeds. Also, they eat the raw root, flavored, or cooked in broth as a vegetable. When corn flour is scarce, it is mixed with the root to make it yield. (a) Wixarika boy collecting the plant in the surroundings of the Tateikie community. (Photo by: Luis Villaseñor Ibarra)

parts of the plant (Fig. 20). These are considered a strong dish, and the Wixaritari prepare them in various ways ranging from roasted, cooked, in broths, in moles, or in sauces, or combined with other elements such as beans or other vegetables. As for the agaves, they give a different name to each part of the agave they eat as sweets: the leaves (xaapa) and the inflorescence named quiote (hari) that cook on the coals or in clay ovens, the flowers (siiberi) from which the stamens and pistil are consumed (Bauml 2004) [field data by the authors]. In the category of sweet potatoes or roots, the plants are of traditional Mexican consumption such as Pachyrhizus erosus (xata-tea) or jicama, Solanum cardiophyllum (teho) or wild potato, Ipomoea murucoides (tikarixa), Dioscorea remotiflora or sweet potato, and some species of Dalembertia and Tigridia, to mention a few. Of all, only xata-tea is consumed raw; the rest are cooked, and because of its neutral flavor is used as an accompaniment with meat, nopales, and quelites, as well as for the preparation of candys, adding maguey or bee honey (Bauml 2004) [field data by the authors]. Plants in the category of condiments include the most used Dysphania ambrosioides (hapaxuki) or epazote, as well as Capsicum annuum (kukuri) or chili pepper and two introduced species, Menta piperita locally called toronjil, and Origanum vulgare or oregano. All these species are used in a wide range of salty and sweet recipes. Gathering of condiments, such as wild peppers, nuts, and some fruits such as Solanum spp. (tásiu hïxi), is carried out at different times of the year; they are dehydrated in the sun and then stored to have available in the kitchen (Bauml 2004, field data by the authors; Fig. 21). Some of the species of edible plants are considered as nutraceutical foods, so in addition to being appreciated for their nutritional value as food, they are used as remedies for different conditions such as xeuroruwame for fertility and kukiya huaye to relieve cough. Also, it is common mainly for children to eat the nectar and stamens

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

221

Fig. 20 Opuntia ficus-indica called “Nakari” or “prickly pear”; its modified stems are consumed as food; they are also used to cure constipation, dysentery, diabetes, or to deflate. (Photo by: Mara Ximena Haro-Luna)

of a large variety of flowers such as Pseudobombax palmeri (rabe) or Magnolia pugana (awkwe) (Bauml 2004; field data by the authors).

Timber Plants According to the data collected, 63 species are reported; some of them are used for constructing fences, others to make tools or to build houses and temples. Many of these plants have dual purposes, one utilitarian and the other sacred, as it is the case of the objects that are made to be offered. As part of the mythology, it is said that the trees, especially pines, symbolize the measure that the ancestors took to determine the appropriate elevation to place the clouds, so that they did not fall too strong or too faint and their jakiérite or their children do not drown in the rain (Negrín 1997). One of the most exploited species is Haematoxilum brasiletto (uca ucáari,  itsa) or stick of Brazil, which because of its red color is selected to build the “command rods” or objects from which Tayau’s power “Father Sun” emanates. These objects are carried for a year by the Tatuwani, the “traditional governor,” and the members of the

222

M. Cedano-Maldonado et al.

Fig. 21 Fruits of a species of tomatillo (Solanum sp.) put to dry in the sun: (a) when dehydrated, strings or necklaces are made that are hung on the beams of the house to be used in the kitchen. Author of the photograph: Mara Ximena Haro-Luna; (b) Solanum lycopersicum is one of the species they grow in their orchards. (Photo by: Luis Villaseñor Ibarra)

traditional government: first, second governor, judge, sheriff, captain, commissioner and Topiles, or messengers who are known as “Ttstkate or Ttsikate?” or “wand bearers.” During the New Year celebrations, there is a change of rulers, and the “change of wands” ceremony is held (Neurath 2003). In the construction of

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

223

Fig. 22 “Tuki” or main temple of the Tateikie ceremonial center (San Andres Cohamiata, Mezquitic, Jalisco). (a) Pinus durangensis one of the species that grows in the Sierra Huichola; due to the height of its stems, it is used in the construction of the Tuki. (Photo by: Luis Villaseñor Ibarra).

traditional and sacred instruments such as the three-legged drum (tepo) and the violin, the creator must locate among his dreams the tree whose wood will have the magical power to vibrate the spirits of the ancestors (Negrín 1997). The wood of Juniperus durangensis and Pinus teocote is used to build the main temples or “Tuki,” which is located at the west of the courtyard of each ceremonial center (Bauml 1994). They have a circular or oval structure, semisunk, with a high roof of grass (Muhlenbergia sp.), supported by two wooden posts of the trees referred to above, representing “the cosmic trees (haurite),” that support the sky. The size varies in each ceremonial center; generally, the average is 10 m in height and diameter (Neurath 2003) (Fig. 22). Stems of Arundo donax, Heteropogon contortus (hamürixá ucáari), and Otatea acuminata (hacu ucáari) are used as beams to build the roofs of houses and temples (Bauml 1994) (Fig. 23). A use that has already disappeared is the manufacture of nierikas with pieces of walnut, which are left uncovered in the center; stamen of colors was also interwoven and beaded with beeswax threads of beads of chaquira. Currently, it is replaced by fiberboard boards, which are decorated with Campeche wax and stamen; these are no longer sacred objects and are sold as handicrafts (Neurath 2005) (Fig. 24).

Medicinal Plants The medicinal or herbalist Wixarika is the area most approached by scholars, and it includes the use of about 190 species for the cure of about 95 different diseases. Huichols classify plants into two groups: those of cold nature and those of hot nature. The latter are the most used (Casillas 1990). They also usually grouped these plants based on the places where they grow or are collected, those of the kieta or “ranch,” those of huyetetsita or “on the side of the path,” and those of the ritsie or “forest or

224

M. Cedano-Maldonado et al.

Fig. 23 Trolley where they guard the harvested corn, the walls are built with otate (Otate sp.). (Photo by: Mara Ximena Haro-Luna)

mountain” (Bye et al. 2005). Several of these species are conferred various therapeutic uses (Nieves et al. 2004b). Currently, the Wixarrika health system has traditional and modern medicine. For them, the first is the most successful and efficient, because it is available to everyone, resolves problems immediately, and addresses the diseases sent by the gods. The system of modern medicine is an alternative to address the natural ills that come from the mestizos, and their medicines are incompetent to heal them (Casillas 1990). The legend of how diseases arose with the Wixaritari narrates that when the world originated, the Wixaritari gods Cacaiyari were gathered, the first ear was born to be able to eat it, they put it to the fire, and from the embers a smoke came out, which was actually iricayari whooping cough and infected them. When he first appeared in the sacred place Teupa, other diseases arose in the center and the four cardinal points. First, the cuitayari dysentery came out in the west, then, in the east, the Tápacuíniya pneumonia, then, in the south, the Tawaiya plague, then, in the north, the rubella isipúriquiyá, and finally, in the center, smallpox, etsá, and the measles. Then Seriecame-Marracuarrí was responsible for spreading all other diseases (Casillas

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

225

Fig. 24 Current Nierika built with material made of plywood, Campeche wax, and yarn, which are sold as handicrafts. Formerly, they were made of walnut wood and had a hole in the center, because through it they communicated with the gods and the universe was seen. When the nierika is a sacred object, it is not for sale. (Photo by: Luis Villaseñor Ibarra)

1990). Like any Indigenous group, the Wixaritari or Huichols have an appreciation about the health-disease process. Most of his discomforts are attributed to the participation of supernatural forces and consider that the source of health and long life is related to following the rayerreiya or “el costumbre,” which includes participating in ceremonies and parties, doing offerings, traveling to sacred places, cultivating their corn, and being a good hunter and excellent pilgrim, to name a few (Casillas 1990; Guzmán-Mejía and Anaya-Corona 2007). They think their gods will punish those who do not comply, throwing arrows that carry diseases. Other illnesses may be caused by black magic and sent through a sorcerer, because only they can capture and steal the Kiúri or “the soul,” which is the one that maintains health (Casillas 1990). To achieve the recovery of any disease, even when treated with modern medicine, a mara’akame or shaman is used. For only they can enter the natural world and the supernatural and implore the gods for forgiveness and restoring health to the sick. The treatment begins with some questions to find the cause of the condition. Healing begins by using the muviere, a feathered and sacred arrow; with it, on the abdomen, the shaman makes a movement to the four cardinal points and the center of the patient, also sucks to remove the disease, and spits it out to pass it to an object or thing. Subsequently, to complete the recovery, medicinal, magical, or power plants are used (Kindl 2013), and actions must be taken by the patient and his family to please the offended deities (Camberos 2004; Casillas 1990).

226

M. Cedano-Maldonado et al.

For the treatment of supernatural or nonexplainable diseases by science, such as the loss of the soul, to gather couples, against the lack of inspiration in the realization of crafts, to protect against spells and accidents, magical protection against children’s diseases, and to do the rituals of “clean,” the use of fourteen plant species is mentioned. Among the most important: Solandra guttata (kieri, quieri nanáari) for healing in rituals and to obtain abundance, Prosopis laevigata (meequi ucáari) to cure the evil eye, and Casimiroa sp. (sapú ucáari) so that the souls of the dead do not pass inside the house can be mentioned (Bauml 1994). For the cure of natural diseases, the most used species are the following: Prosopis laevigata (meequí ucáari) which helps eliminate stomachache, Enterolobium cyclocarpum (uvee tatuata ucáari), Crocanthemum glomeratum (acuitsi-huayeya), and Guazuma ulmifolia (Aaye ucáari) to stop diarrhea (Fig. 25), Brosimum alicastrum (hauri ucáari, hairite, jairi) to deflate, Psidium guajava (guayavuaxi ucáari) for stomachache, empacho for cough, flu, and fever, and Bejaria aestuans (piriiqui ucáari) to attack cough, whooping cough, asthma, and flu. Euphorbia furcillata and E. biformi (veriya uayeeya nunuveme) are used when menstruation is not abundant, while E. strigosa (veria uayeeyari) is used to fatten undernourished women, boys, or girls. E. hyssopifolia (taueaca uquisi ucáari) and E. succedanea (taueaca ucáari) are used when blood does not come out, to treat infections, and to

Fig. 25 Guazuma ulmifolia tree, “Aaye ucáari” or “guazima”; its fruits (a) are boiled in water, and the preparation obtained is taken to stop diarrhea; they are also sucked raw to quench thirst. (Photo by: Gregorio Nieves Hernández)

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

227

Fig. 26 Natural habitat of Tagetes lucida, “Tímutsari” or “Santa María”; its flowers (a) are prepared in infusion, which is ingested to calm the pain of stomach, heart or chest, and head or lower fever. (Photo by: Luis Villaseñor Ibarra)

close cuts, and E. sphaeronzi (icuexia uayeey) to whet appetite. Lophophora williamsii (hikuri, hikuli) is used to avoid tiredness, counteract scorpion bite, soothe rheumatism, and fight against mumps; Haemataxylum brasiletto (uca ucáari,- itsa) helps against dysentery, diarrhea, cough, bloody urine, chest or heart pain, tiredness, and headache. Tagetes lucida (tímutsari) (Fig. 26) and T. erecta (púuvaari, púvaari sinúuxi) are used for stomachache, heart or chest pain, headache, and fever; T. micrantha (tuumúusári uquisi) for stomachache, diarrhea, and cramps; and Pithecellobium dulce (múxúuri ucáari, mutúri-te) for stomachache, diarrhea, parasitosis, and burns (Fig. 27). Opuntia ficus-indica (nakari) cladodes are used for constipation, dysentery, and diabetes, and it is anti-inflammatory. Arctostaphylus pungens (upapaari) is used for cough, bad urine, and torzones (twist and contractures), Pinus oocarpa (hucuu) to cut nosebleed, cough, and muscle pain, and Salvia spp. (neuturica uayeeyári uquisi) is a cold grass, used for diarrhea, empacho, corn disease, acidity, parasitosis, headache, and stomachache. Prunus serotina var. capuli (tuvii uricasanaca uquisi) is used for cough, flu, and chest or heart pain, to name a few. Most of the medicinal plants used are wild, which indicates the importance of forests as sources of these resources, and the importance of protection and biodiversity conservation linked to culture in the region (Bauml 1994; Casillas 1990; Higareda-Rangel et al. 2004; Nieves-Hernández 2002; Bye et al. 2005).

Other Uses In this category, 72 species are included, 21 are used as fodder, 21 as firewood (Fig. 28), 11 as fibers (Agave spp. and Dasylirium sp.), 11 as poison (among them Agave vilmoriana), 6 as glue (outstandingly Bletia macristhmochila), and 2 as soap (Agave vilmoriniana, Manfreda rubescens). The root of Manfreda rubescens (haariuqui uquisi) is boiled and macerated in a metate or with a wooden hammer; it is used in laundry. The leaves of Agave vilmariniana (vaave ucáari) are cut and macerated raw and used as body soap and

228

M. Cedano-Maldonado et al.

Fig. 27 Fruits of Pithecellobium dulce “múxúuri ucáari uquisi” or “guamúchil,” which are consumed to cure stomach pain, stop diarrhea, and expel parasites. (Photo by: Luis Villaseñor Ibarra)

for washing hair (Bauml 1994). There are other plants for personal care, which have not been reported so far, by personal comments to one of the authors, noted the use of various lianas, as a hair treatment to stimulate hair growth. In the case of Bletia macristhmochila (cuesucuaucáari), the bulbs are torn with a knife and scrambled with black zacate ash so that it sticks hard, such as equipment wood (uveri), guitar and violin, or various offerings (Bauml 1994).

Perspectives In recent times, roads and new accesses are opened in the Wixaritari territory, and some mestizo people have penetrated their spaces and influenced their customs and thoughts of children and young people. Several of these teenagers have migrated to study or work in other regions of Mexico, causing local uprooting and tremendous cultural erosion. Although, it is not the case of the elders and adults, some go out to sell their handicrafts, but they continually return to their lands and continue transmitting their knowledge, customs, practices, and myths to those who remain in the community. However, these changes make the registration of their knowledge a priority, before they are lost due to acculturation, death, migration, and the ecological deterioration of their territories. Based on the work reviewed for the study area, it is necessary to integrate an inventory of the useful flora of the Wixaritari, with the information generated on the subject. Among the obstacles to be overcome, there is a lack of herbarium materials, the taxonomical identification of many of the species, the linguistic record of Wixaritari, or Huichol language names and the detail of their uses. Of the most

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

229

Fig. 28 Firewood collection is a task carried out by the Wixaritari children; they take advantage of around 21 species. (Photo by: Luis Villaseñor Ibarra)

important utilitarian plant groups, it is necessary to address “edible plants,” since reports on them are insipient. Also, exploration and work on the communities of San Sebastián Teponohuaxtlán, Tuxpan de Bolaños, and Guadalupe Ocotán are needed. In general, it is necessary to generate more applied and holistic ethnobiological research, addressing aspects such as values of importance for the cultural group, its worldview, its conceptions, emotions, and behaviors. As well as delving into traditional knowledge and practice to recognize how culture has been integrated, such as the use and sustainable management of resources, the conservation of their areas, the successes and mistakes of their practices, and the ecological consequences of them, to name a few. There should be greater feedback and cooperation from scholars to the communities, participatory action research should be generated, taking into account their problems and needs, to identify and solve these problems, revaluation should be involved of sociocultural and territorial identities, and work

230

M. Cedano-Maldonado et al.

should be done on the revaluation of sociocultural and territorial identity, as well as on the recognition of biocultural heritage, among others.

References Alcocer P, Neurath J. La eficacia de la magia en los ritos coras y huicholes. Arqueol Mex. 2004;69: 48–53. Anderson EF. Peyote the divine Cactus. Arizona: The University of Arizona Press; 1980. Barba-Ahuatzin B. Antropología del tabaco. Ciencia. 2004;55(4):6–16. https://www.revistaciencia. amc.edu.mx/images/revista/55_4/antropologia.tabaco.pdf Barrera-Rodríguez RO. Medios natural y ambiental del territorio huichol (norte de Jalisco, México). Guadalajara: CUCSH, Universidad de Guadalajara; 2004. Bauml JA. Ethnobotany of the huichola people of Mexico. Doctoral Thesis. Claremont: Faculty of the Claremont Graduate School; 1994. Bauml JA. Overview of huichol ethnobotany. In: Vazquez-García JA, coordinator. Flora del norte de Jalisco y etnobotánica huichola. Guadalajara, México: CUCBA-CUCSH, Universidad de Guadalajara; 2004. p. 86–92. Benz BF. Archaeological evidence of teosinte domestication from Guilá Naquitz, Oaxaca. PNAS. 2001;98(4):2104–6. Bernardello LM, Hunziker AT. A synoptical revisión of Solandra (Solanaceae). Nord J Bot. 1987;7(6):639–52. https://doi.org/10.1111/j.1756-1051.1987.tb02032.x. ISSN 1756-105. Bye R, Aedo-Gajardo A, Faba-Zuleta P. Listado florístico y etnobotánico de las plantas medicinales de los huicholes del río Chapalagana, Jalisco y del Nayar, Nayarit. México, Distrito Federal: Instituto de Biología. Informe final SNIB-CONABIO proyecto No. AE001. Universidad Nacional Autónoma de México; 2005. Camberos S. El chamanismo y las raíces de la terapéutica en la medicina tradicional de los Huicholes. In: Vazquez-García JA, coordinator. Flora del norte de Jalisco y etnobotánica huichola. Guadalajara, México: CUCBA-CUCSH, Universidad de Guadalajara; 2004. p. 109–110. Casillas RA. Nosología mítica de un pueblo. Medicina tradicional huichola. Guadalajara: Universidad de Guadalajara; 1990. 160 p. Cedano-Maldonado M, Villaseñor-Ibarra L, Gamboa-Ruiz A. Usos tradicionales de la familia Cochlospermaceae por los wixarikas (huicholes) de San Andrés Cohamiata, Mezquitic, Jalisco, México. Boletín IBUG. 1998;5(1–3):343–51. Cedano-Maldonado M, Villaseñor-Ibarra L, Vargas-Ponce O. Plantas útiles. En: La biodiversidad en Jalisco. Estudio de Estado, vol. I. Guadalajara: CONABIO; 2017. p. 177–87. Diguet L. El idioma huichol. Contribución al estudio de las lenguas mexicanas (1911). In: Diguet L, editor. Por tierras occidentales; entre sierras y barrancas. Ciudad de México: Centro de estudios mexicanos y Centroamericanos; 2005. p. 161–93. Furst PT. Visiones de un chamán huichol. Museo de Arqueología y Antropología de la Universidad de Pennsylvania; 2007. Guzmán-Mejía R, Anaya-Corona M. Cultura de maíz-peyote-venado: sustentabilidad del pueblo wixarika. Guadalajara, México: CUCSH, Universidad de Guadalajara; 2007. Higareda-Rangel Y, Nieves-Hernández G, Luquín-Sánchez H. Plantas medicinales de Santa Catarina Cuexcomatitlán (Tuapurie) Sierra Wixárika, Jalisco. In: Vazquez-García JA, coordinator. Flora del norte de Jalisco y etnobotánica huichola. Guadalajara: CUCBA-CUCSH, Universidad de Guadalajara; 2004. p. 102–108. INALI Instituto Nacional de Lenguas Indígenas. 2015. con base en los datos de la Encuesta Intercensal, INEGI 2015 y el Cátalogo de las Lenguas Indígenas Nacionales, INALI, 2008. https://site.inali.gob.mx/Micrositios/estadistica_basica/estadisticas2015/pdf/agrupaciones/ huichol.pdf.

Wixaritari or Huichol Ethnobotany of the Southern Sierra Madre. . .

231

INEGI. Instituto Nacional de Estadística, Geografía e Informática. II Conteo de Población y Vivienda. 2010. In: https://www.inegi.org.mx/app/areasgeograficas/?ag¼14. Last consultation: Access 3 Oct 2019. Iturrioz-Leza JL, Carrillo de la Cruz J, López de la Cruz J, Ortiz-Medina J, Leal-Carretero S, Aguilar-López N, Pacheco-Salvador G, Gómez-López P, Mikiri F, Ortiz-López A, GarcíaMuñoz P, Caudillo-Félix GA. Reflexiones sobre la identidad étnica. Guadalajara: Universidad de Guadalajara; 1995. Kindl O. Eficacia ritual y efectos sensibles. Revista de El Colegio de San Luis. 2013;5:206–27. La Barre W. The Peyote Cult. Hamden: Archon Books; 1975. Lumholtz C. Symbolism of the Huichol Indians. Memoirs of the American Museum of Natural History. Vol III, Anthropology II. 1900. Luquín SH, Nieves HG, Ayala RJM. Vegetación del norte de Jalisco. In: Vazquez-García JA, coordinator. Flora del norte de Jalisco y etnobotánica huichola. Guadalajara: CUCBA-CUCSH, Universidad de Guadalajara; 2004. p. 86–92. Moreno-Coutiño A, Coutiño-Bello B. Nicotiana tabacum L., usos y percepciones. Etnobiología. 2012;10(2):29–39. Negrín JE. arte contemporáneo de los huicholes. Guadalajara: Universidad de Guadalajara, INAHSEP; 1997. Negrín J. ¿Qué atrae a los indígenas huicholes hacia el Océano Pacífico? El océano es el principio de la vida: Relatos huicholes sobre Tatéi Haramara, Nuestra Madre el Mar. 2006. http://www. huicholesyplaguicidas.org/documentos/que_atrae_a_los_huicholes_al_mar.pdf Neurath J. Huichole. Comisión Nacional para el Desarrollo de los Pueblos Indígenas, PNUD. 2003. Access to pdf 14 Jan 2019. Neurath J. Máscaras enmascaradas. Indígenas, mestizos y dioses indígenas mestizos. Relaciones. 2005;101:23–50. Nieves HG, Vázquez-G M, Saldívar ESM. Rasgos físicos de la región norte de Jalisco. In: VázquezGarcía JA, coordinador. Flora del norte de Jalisco y etnobotánica huichola. Guadalajara: CUCBA-CUCSH, Universidad de Guadalajara; 2004a. p. 22–23. Nieves HG, Vázquez-García JA, Cházaro MJ, Vázquez-G. M. Uso tradicional de la flora de la región huichola. In: Vázquez-García JA, coordinator. Flora del norte de Jalisco y etnobotánica huichola. Guadalajara: CUCBA-CUCSH, Universidad de Guadalajara; 2004b. p. 93–101. Nieves-Hernández G. Flora vascular del norte de Jalisco y su uso tradicional por la etnia huichola, Jalisco, México. Master’s Thesis. Zapopan, Jalisco México: Universidad de Guadalajara; 2002. Santos-García S, Carrillo CT. Teukarita: designación de nombres de personas entre los wixaritari y sus significados. In: Conti C, Guerrero L, Santos S, editors. Aproximaciones a la documentación lingüística del huichol. Jaén: Gráficas la Paz; 2012. Torres JJ. El hostigamiento a “El costumbre” huichol. Los procesos de hibridación social. México: El colegio de Michoacán, A.C. Universidad de Guadalajara; 2000. Vázquez-García JA, Nieves HG, Cházaro MJ, Vargas-Rodríguez YL, Flores-Macías A, Luquín SH. Listado preliminar de plantas vasculares del norte de Jalisco y zonas adyacentes. In: Vázquez-García JA, coordinator. Flora del norte de Jalisco y etnobotánica huichola. Guadalajara: CUCBA-CUCSH, Universidad de Guadalajara: 2004. p. 116–168. Villaseñor-Ibarra L. Etnomicología de la etnia Wixarika (Huichol), Jalisco, México. Master’s Thesis. Zapopan, Jalisco México: Universidad de Guadalajara; 1999. Villaseñor-Ibarra L, Cedano Maldonado M, Guzmán-Dávalos L, Pacheco Salvador G (‘Iritemai). Yekwaate metá wixaritari Tateikietari. Hongos y Wixaritari de Tateikie. Guadalajara: Universidad de Guadalajara; 2017a. Villaseñor-Ibarra L, Cedano-Maldonado M, Vargas-Ponce O. Aprovechamiento y manejo de las plantas, hongos y animales silvestres por los huicholes y nahuas. In: La biodiversidad en Jalisco. Estudio de Estado Vol. I. Guadalajara: CONABIO; 2017b. p. 189–196. Villegas L. Dioses, mitos, templos, símbolos: El universo religioso de los huicholes. Americanía Revista de Estudios Latinoamericanos Nueva Época (Sevilla). 2016;3:4–48.

Traditional Management and Diversity of Opuntia: General Panorama in Mexico and a Case Study in the Meridional Central Plateau Amaranta Paz-Navarro, Ce´sar I. Ojeda-Linares, Gonzalo D. A´lvarez-Ríos, Mariana Vallejo, and Alejandro Casas

Abstract

This chapter shows a general panorama of the biocultural importance of the Opuntia genus in Mexico, and a case study directed to illustrate more specifically such relevance. The case study was conducted in one of the regions of Mexico where the highest diversity of species and ethnovarieties of Opuntia occur, and where this variation has been strongly rooted in the regional cultures. The Opuntia genus comprises more than 200 species, numerous intraspecific taxa, and even more ethnovarieties of cactus pears. It is native to the Americas, with a wide distribution in arid and semiarid lands through subpolar, temperate, subtropical, and tropical regions, from Canada to Patagonia. At least 84 species and numerous intraspecific taxa have been recorded to occur in Mexico; nearly 50 of them are used by different human cultures, and about 15 are cultivated, showing signs of domestication. Nowadays, some Mexican prickly pears like O. ficusindica, O. streptacantha, O. robusta, O. cochenillifera, and O. auberi are cultivated throughout the world due to their economic and cultural values as food, ornaments, fodder, health-promoting benefits, as main host plant of cochineal, and for multiple other uses and applications. Some species were introduced to the A. Paz-Navarro · C. I. Ojeda-Linares · G. D. Álvarez-Ríos Laboratorio Manejo y Evolución de Recursos Genéticos, Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected]; [email protected] M. Vallejo Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico e-mail: [email protected] A. Casas (*) Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_21

233

234

A. Paz-Navarro et al.

Old World after the European colonization of the Americas and currently are invasive in areas of Australia, Asia, Africa, and Europe, causing severe ecological and economic problems. Since pre-Columbian times, the cactus pear species played agroecological roles in different regions of the Americas. In Mexico, these plants have been used since prehistory by native peoples and currently are important sources of livelihood, outstandingly human food. Fruits and young tender stems of practically all species, and flower buds of some of them, are edible. Their fruits are consumed fresh, and these and other edible parts are prepared in multiple ways in the Mexican cuisine, contributing to diet-nutritious components like amino acids, vitamins, proteins, minerals, dietary fibers, betalains, and phenolic compounds. The young tender cladodes are very much appreciated as vegetables in Mexican cuisine with dozens of recipes developed through history. The mature cladodes are also used to feed livestock; when dried, these are used as firewood. Their easy vegetative propagation and fast growth make Opuntia plants favorable for their cultivation, their use in borders of terraces against soil erosion, as live fences, and as main crops in intensive and extensive plantations. Due to their crassulacean acid metabolism (CAM) and water use efficiency, Opuntia spp. provide ecological benefits for recovering and improving degraded lands, landscapes rehabilitation, biodiversity preservation, and prevention of desertification. The case study reported in this chapter was conducted in the Meridional Central Plateau of Mexico, a region historically inhabited by different Chichimec Indigenous peoples, which developed a remarkable cultural history of interactions with prickly pears since pre-Columbian times. Then, after the European conquest, peoples of the region maintained traditional use and developed innovations associated with new socio-ecological contexts, strongly influenced by livestock raising and modern agriculture intensification. The study area is a remaining of the cultural region called “Tunal Grande” or “Gran Tunal,” because of the abundance of forests dominated by Opuntia spp. The name “Tunal” derives from the word tuna, of Caribbean origin but adopted in Spanish to refer to fruits of these plants. The case study documented the local classification of varieties and species of Opuntia, their uses, and management forms in the community of Laguna de Guadalupe in the state of Guanajuato. There, we identified 30 traditional varieties of 10 Opuntia species managed in two main environmental units: the monte (patches of thorn-scrub and secondary forests) and home gardens. In the monte, people let standing phenotypes of species and varieties that are more valued because of their attributes; in addition, they carry out practices that promote the abundance of these favorable species and varieties. In home gardens, people cultivate the most appreciated species and varieties and practice more intensively human selection that guides processes of domestication. We documented the different uses of species and varieties of Opuntia, their economic importance for local people, and their historical and current role as components of the biocultural diversity. We finally discuss the main cultural and economic factors influencing contemporary changes in the local landscape and human culture, and the socio-ecological perspectives for conserving the important biocultural heritage related to Opuntia.

Traditional Management and Diversity of Opuntia: General. . .

235

Introduction The Opuntia genus of the Cactaceae family comprises more than 200 species, a number that greatly varies according to taxonomic criteria; some authors report up 350 species (Bravo-Hollis 1978; Guzmán et al. 2003; Chávez-Moreno et al. 2009; Scheinvar et al. 2011; Illoldi-Rangel et al. 2012). Opuntia plants are called cactus pears or prickly pears in English, and nopales in Spanish, a term derived from the Nahuatl word nopalli, which refers to the cladodes (the entire plant in Nahuatl is called nopalkuahuitl) and that was adopted by the Spaniards during the Colonial time in Mexico. Fruits are called nochtli in Nahuatl, but tuna, the Taino Caribbean term, was the name adopted and diffused by the Spanish conquerors through the Americas. Actually, in some countries (e.g., Peru, Colombia, Ecuador, and Argentina) the entire plants are commonly called tuna because of the Spanish adoption of this term. However, because of the wide distribution and human cultural importance of these plants, practically all languages in Mexico have a general term to group all species of Opuntia the peoples interact with (Table 1). Although the Nahua and other peoples differentiate the name of the plant, the cladodes, and fruits, several Indigenous languages of Mexico name the entire plant based on the fruit name. Although few Indigenous systems of classification of Opuntia have been studied, those available indicate that the most meaningful traits involved in traditional classification are the color and form of cladodes, plant size, spine color, and outstandingly, several attributes of fruits (size, peel and pulp color, flavor, and consistency, among others) (Casas and Barbera 2002; Casas et al. 1999). The cactus pears have their origins in the Americas (Mazri 2018). The genus is naturally distributed from south of the arctic circle in subpolar areas of Canada and their analogous areas of Patagonia, in Chile and Argentina, through temperate, subtropical, and tropical areas of both hemispheres, including the Caribbean and Galapagos Islands (Bravo-Hollis 1978; Scheinvar 1999; IlloldiRangel et al. 2012). Cactus pears have anatomical, morphological, and physiological adaptations to contrasting environmental conditions, characteristically those with harsh waterdeficit stressing areas (Nefzaoui et al. 2014; Prat and Franck 2017; Kumar et al. 2018). Also, cactus pears display wide ranges of plasticity that allow them to cope with environments showing markedly contrasting seasons of the year (ReyesAgüero et al. 2006). Opuntia plants most commonly display spines, which are modified leaves (Glimn-Lacy and Kaufman 2006), but some species or varieties may be spineless, generally associated with human selection. They commonly have fleshy stems, called pads or cladodes, with spines and glochids arranged in areoles (Glimn-Lacy and Kaufman 2006). Leaves are small or absent, the cladodes being the main parts involved in photosynthesis. The stems contain an outer surface cuticle that is thick and waxy, to prevent the water loss in drought conditions (Glimn-Lacy and Kaufman 2006). Flowers are large, hermaphrodite, composed by several sepals and petals appearing as tepals, and a single pistil with a lobed stigma at the apex (GlimnLacy and Kaufman 2006; Arba et al. 2017). The color of the flowers is highly

236

A. Paz-Navarro et al.

Table 1 Names of Opuntia plants in the main Indigenous languages of Mexico. Each main language may have several variants. Ethnologue (Eberhard et al. 2021) identified nearly 290 languages for Mexico. The different terms annotated are some of the regional names given in a main language Language Aguacateco (Awakateko, Kayol, QaꞋyool) Amuzgo (Tzañcue) Chatino (Cha’cña) Chichimeco-jonaz (Ézar) Chinanteco (Tsa jujmí) Chocho (Runixa ngiigua) Ch’ol (Winik) Chontal/Tabasco (Yokot) Chontal/Oaxaca (Slijuala sihanuk) Chuj Cochimí (Laymón, Mti’pá) Cora (Nayeeri) Cucapá (Es péi) Cuicateco (Dibaku, Dbaku) Guarijío (Makurawe) Huave (Ikööds) Huichol (Wixárika) Ixcateco (Xwja) Ixil (Ixil) Jacalteco (Abxubal) Kakchiquel (K’akchikel, Kaqchikel) Kanjobal (K’anjobal, Q’anjob’al) Kekchí (K’ekchí) Kikapú (Kikapooa) Kiliwa (Ko’lew) Kumiai (kumeyaay, Ti’pai, Tipai-ipai, Kamia) Lacandón (Hach t’an) Mame (Qyool) Matlatzinca (Botuná) Maya Mayo (Yoreme) Mazahua (Jñatio) Mazateco (Ha shuta enima) Mixe (Ayüük) Mixteco (Ñuu sávi) Mochó or motozintleco (Qatok) Nahuatl (Mexicatl) Otomí (Hñähñü) Paipai (Akwa’ala) Pame (Xigüe, Xi’iùy)

Name of Opuntia plant k’ána ndua bya, yaa, yaa jwle’, yaa tii úp’ó, ují, em’á cha’loo kánda chish-pech, wut petek’, petek’ No record labone, lhi’as gabone baj’til a naká, takera a’ yind’itu, ndudu, ndiitu jilú, ilacúrugi nüek, nak’, nüik, nɨk’ nakari ñunda kaana páak’am noxti’ a:xí:lh, a’ši:ɬ noxti’ meskupuakaa ‘aa, a’a a’a k’oj tx’ixj xöt’ö pak’am, páak’am, tsutsuy naabo, nabo taaka, navo kijñi tu rë nanda, nänta núum täät tsä’äm, täät vi’ncha, vi’ndia, bi’ndé a’xí nopalkuauitl, nopalli dokähä, ixcähä, xathä, kähä a’a nm’u (continued)

Traditional Management and Diversity of Opuntia: General. . .

237

Table 1 (continued) Language Pápago (Tohono o’odam) Pima (Tohono o’odham) Popoloca (Ngigua) Popoluca (Tuncápxe) P’urhépecha Quiché (Q’iché, K’iché’) Seri (Comcaac) Tacuate Tarahumara (Rarámuri) Teenek (Téenek) Tepehua (Hamasipini) Tepehuano (O’dami) Tlahuica (Pjiejakjo, Pijekak’joo) Tlapaneco (Me’phaa, Mè’phàà xkuà ixì ridiì) Tojolabal Totonaco (Tachihuiin) Triqui (Guii xihanjhan) Tzeltal (Winik ate) Tzotzil (Batsil winik) Yaqui (Yoeme) Zapoteco (Binizáa) Zoque (O’de püt)

Name of Opuntia plant na’abo nava, i’ibi, na’up tùchjèkíxí, kándá, tùchjèkàndà, túchi kánda pejtak, to’a pare machiti, nochiti heel, heel imám bindia’ erá, napó pak’ak’, pots’ots’ laklaba’anti náboi, náboi jibhiadï (fruit) mbakijñi byáa’ pejpa ajiit, axit, axilh dinï, tíno pejtak ch’u naboo, naabo, naabujti bia, biaa, biaagueta, yága bidxí nakpat

variable among Opuntia species; it could be yellow, cream, yellow-red, or orange through pink or red. It has been discussed that the color variation might be associated with pollinator preferences (Heuzé and Tran 2017). Fruits are ovoidspherical fleshy berries that may exhibit different colors at maturity, most commonly red, but also orange, yellow, or green peel, and even more variety of pulp colors from different shades of red, purple, violet, green, yellow, to orange. The peel exhibits areoles with spines and glochids (Beccaro et al. 2015; Mazri 2018). The taxonomic delimitation of Opuntia species has been complicated, due to the frequent interspecific hybridizations, polyploidy, human selection in favor of specific traits of fruits, cladodes, plant size, and because of their conflicting phylogenetic positions (Wang et al. 1996; Griffith 2001; Valadez-Moctezuma et al. 2015; Martínez-González et al. 2019). The taxonomic complexity of cactus pears might highlight the complex genetic background of the species of this genus (Samah et al. 2016). Efforts to discriminate cactus pear species by using both morphological traits and molecular markers are currently conducted by several research groups (Caruso et al. 2010). Nowadays, the cactus pears play important socioeconomic, agronomic, and ecological roles in the agendas of several countries around the world, since a number of species are used for human consumption, medicine, forage, and programs against soil erosion and desertification, ecological restoration, among other purposes. In addition, several biotechnological applications have been recently developed and

238

A. Paz-Navarro et al.

performed with these plants as cosmetic, pharmaceutical products and biofuels (Nefzaoui and Ben Salem 2002; Nefzaoui et al. 2014; Valadez-Moctezuma et al. 2015; Mazri 2018). But also, some Opuntia species represent severe environmental problems in extent regions of the world where they were introduced and now are aggressive invasive plants (Monteiro et al. 2005; Shackleton et al. 2017; Tesfay and Kreyling 2021). This chapter aims to show a general panorama of the biocultural importance of the genus Opuntia in Mexico, and ethnobotanical information from a case study in one of the regions of this country where the highest diversity of species and ethnovarieties occur, and where this variation has been strongly rooted in the regional cultures, the Meridional Central Plateau. Specifically, we studied the case of Laguna de Guadalupe, Guanajuato, with the purpose to: (1) document the local species and varieties of Opuntia recognized by people, their attributes, cultural value, and traditional classification, (2) the uses and management practices carried out on the different species and varieties, those involving human selection, the main targets of selection, and practices used to carry out such selection, and (3) identify the main environmental units where Opuntia plants occur, and how these are managed.

Methods Literature Review The general panorama of Opuntia as Mexican biocultural heritage was reviewed based on literature on taxonomic, biogeographic, ecological, and ethnobotanical issues of the genus among cultural and ecological regions of the territory of the whole country. We generally identified the main different approaches to catalog species of the genus, general information about their use and management, and signs of domestication reported among species of Opuntia. Nomenclature of prickly pears among the Indigenous peoples of Mexico involved a careful review of historical, ethnohistorical, linguistic sources, the dictionaries available, as well as direct recording in the field and conversations and consultations with ethnobotanists working in different regions of Mexico. The case study was conducted in an area previously studied, and therefore we collected information on cultural, ecological, and ethnobotanical information of the area and on the role of Opuntia in people’s life.

Site of the Case Study The general perspective of Opuntia in Mexico comprised most of the territory of the country, while the case study was conducted in the community of Laguna de Guadalupe, which belongs to the municipality of San Felipe, at the state of Guanajuato (Fig. 1). It is part of the Meridional Central Plateau of Mexico (MCPM hereinafter) (Reyes-Agüero et al. 2005a), which is characterized by a semiarid climate, dominated by crassicaule scrub or xerophilous scrub vegetation (Rzedowski 1978) with mesquite (Prosopis laevigata), huizaches (Vachellia schaffneri), and

Traditional Management and Diversity of Opuntia: General. . .

239

Fig. 1 Location of the study area of Laguna de Guadalupe, Guanajuato, in the Meridional Central Plateau of Mexico (MCPM)

several species of prickly pears like Opuntia robusta, O. streptacantha, and O. leucotricha, among others, being the most abundant plant components (BravoHollis 1978; Reyes-Agüero et al. 2005a).

240

A. Paz-Navarro et al.

Laguna de Guadalupe (LG hereinafter) remains as one of the few localities where artisanal products with Opuntia are still fabricated. In the community live approximately 3667 people comprising nearly 1000 families, with land mostly on the ejido collective property regime. Ejido is one of most common types of rural property in Mexico; its main characteristic is the communal type of governmental structure, in which decisions and rules are made by an Ejido Assembly in which participate all the recognized local members of the ejido, or ejidatarios. The Ejido comprises 5463.7 ha used as plots for agriculture and 5675.5 ha of forest shrublands, grassland, and secondary vegetation areas, dedicated to common use where the main economic activities are the cultivation of beans and cattle raising.

Ethnobotany of Opuntia and the Nopalera System in the Case Study In order to characterize use and management of Opuntia species and varieties and the nopalera system in LG, we carried out 13 semistructured interviews to Opuntia managers, during August–November 2019, when prickly pear fruits and tender cladodes of most of the varieties are available. The questions were oriented to characterize the management of the nopaleras (their location, dimensions, management practices, and the Opuntia varieties occurring there), the uses and traditional knowledge of the system, especially those aspects linked to the recognition of varieties, their local classification, the phenological stages, and harvesting times. Additionally, we carried out ethnobotanical walks (Albuquerque et al. 2014) with managers of the main areas where fruits and cladodes are gathered, managed, and cultivated. With the information provided by the managers during the walks, we compiled free listings of the traditional varieties of Opuntia, their local names, their abundance perceived by people, and descriptions of their attributes of fruit (color, flavor, consistency, prickly, and skin thickness) and stems (consistency, fiber texture, palatability, and spininess, among others) for each variety.

Participatory Workshop A participatory workshop was organized with people of LG that are involved in the Opuntia management, either for fruits or cladodes. We carried out the workshop using graphic support materials as stimuli to obtain information on the local nomenclature and classification of varieties, their useful parts, their special attributes and uses, the annual cycle of practices, seasons of availability of products, and other cultural aspects of the different varieties of the prickly pears occurring in the area. These activities allowed us to rectify, corroborate, specify, share, and expand the knowledge and experiences of local people in light of the plurality of knowledge, beliefs, and values on the plant species studied. This activity propitiated a space of dialogue and consensus among the attendees (Maxwell 2013). In the workshop, all the traditional varieties named through the free lists (Smith and Borgatti 1997) were taken up and a classification was carried out from the

Traditional Management and Diversity of Opuntia: General. . .

241

ranking of six attributes of the Opuntia fruits used by managers to identify them: flavor, consistency, size, skin thickness, spininess, and color. The ranking had values from 1 to 5. For flavor, the scale was generated with 1 being the sourest flavors and 5 being the sweetest. Consistency, 1 the most clearly sandy texture and 5 the juiciest. For size and skin thickness, 1 smaller and thin and 5 larger and thick, and prickly, 1 without spines and 5 with the highest spininess. For recording the attribute color, we used a table of colors characterizing the pulp color. With the values obtained for 30 fruits of Opuntia traditional varieties, a nonmetric multidimensional scaling analysis (NMDS) was performed to evaluate the similarities and differences between the varieties according to the attributes identified and management practiced. This analysis was carried out using the “Vegan” package (Oksanen et al. 2019) in R (R Core Team 2020). Subsequently, based on the ethnobotanical description, photographs and collection of botanical specimens were carried out in the managed spatial units; the identification was corroborated with the studies by Reyes-Agüero et al. (2009) and Reyes-Agüero and Aguirre-Rivera (2011).

Results Opuntia in Mexican Cultures Nopales is considered in Mexico to be the Spanish term for Opuntia, but these plants receive names in practically all Indigenous languages of this country (more than 290 according to Ethnologue; Eberhard et al. 2021) (Table 1). It is a group of multipurpose plants originated in the Americas, but currently widely distributed throughout the world since some species were deliberately or incidentally brought to the Old World after the European colonization (Palevitch et al. 1993; Casas and Barbera 2002; Caruso et al. 2010; Pinedo-Espinoza et al. 2017; Mazri 2018). Opuntia species are deeply rooted in the human cultures of Mexico and form part of landscapes in most of the territory of this country (Bravo-Hollis 1978; Colunga 1984; Colunga-García Marín et al. 1986; Rzedowski 1978, 1993; Toledo and Ordóñez 1993; Hernández-Xolocotzi 1993; Casas and Barbera 2002; Guzmán et al. 2003; Reyes-Agüero 2005; Reyes-Agüero et al. 2005a, b, 2009, 2011). Taxonomists have described about 200 species for Mexico (Bravo-Hollis 1978; Chávez-Moreno et al. 2009; Illoldi-Rangel et al. 2012), but some authors like Guzmán et al. (2003) recorded 270 species and 384 intraspecific taxa. Ethnobotanical studies have reported that at about 50 species of Opuntia, and numerous intraspecific taxa are currently used in different regions of Mexico. At least 15 species have clear signs of domestication (Colunga-García Marín et al. 1986; Hernández-Xolocotzi 1993; Casas and Barbera 2002; Reyes-Agüero 2005; Reyes Agüero et al. 2005b; Paz-Navarro 2021) (Table 2). Their use and management are ancient, as revealed by archeological studies, which have reported that prickly pears were used since the earliest times of human occupation of the Tehuacan Valley (Smith 1967; MacNeish 1967) and Guilá

242

A. Paz-Navarro et al.

Table 2 Species of Opuntia recorded in Mexico, their use, form of management, and signs of domestication reported in the literature. Uses: 1 ¼ edible fruits, 2 ¼ edible stems, 3 ¼ edible flowers, 4 ¼ fodder, 5 ¼ alcoholic beverages, 6 ¼ house construction, 7 ¼ live fences, 8 ¼ medicine, 9 ¼ ornamental, 10 ¼ adhesive, 11 ¼ fuel, 12 ¼ soap, shampoo, and cream, and 13 ¼ cochineal cultivation Species O. albicarpa

Use 1, 2, 4

Management type W, C

Domestication Yes

O. amyclaea

1, 9

W, C

Yes

O. atrispina

9

W, C

No

O. atropes

1, 2, 9, 11, 12

W, M, C

Yes

O. auberi

2, 3, 8, 9, 13

W, M, C

Yes

O. azurea

1, 4, 9

W, C

No

O. bensonii

1

W

No

O. bravoana

4, 9

W, C

No

O. cantabrigensis

4

W

No

O. cochenillifera

W, M, C

Yes

O. crassa

2, 3, 5, 8, 13 1, 11

W, C

Yes

O. chavena

1, 2, 13

W, C

No

O. chlorotica O. decumbens

1, 7, 9 4, 5, 11

W, C W

No No

O. dejecta O. depressa

2, 5 1, 4, 7, 11

W, C W

No No

References This study, Scheinvar (1999), and Paz-Navarro (2021) Hernández-Xolocotzi (1993) and Soberón et al. (2001) Konings and Konings (2009) Bravo-Hollis (1978), Colunga (1984), Hernández-Xolocotzi (1993), and CornejoDenman and Arreola-Nava (2008) Bravo-Hollis (1978), Casas et al. (2001, 2002), and Blancas et al. (2010) Soberón et al. (2001) and Powell and Weedin (2004) Bravo-Hollis (1978) and Cornejo-Denman and Arreola-Nava (2008) Soberón et al. (2001) and Mercado-Muñoz (2014) Soberón et al. (2001), Paz-Navarro (2021), and this study Pennington (1969) and Bravo-Hollis (1978) Bravo-Hollis (1978), Colunga (1984), and Hernández-Xolocotzi (1993) Soberón et al. (2001) and Cornejo-Denman and Arreola-Nava (2008) Soberón et al. (2001) Casas et al. (2001), Casas and Barbera (2002), and Blancas et al. (2010) Bravo-Hollis (1978) Soberón et al. (2001) and Casas et al. (2001) (continued)

Traditional Management and Diversity of Opuntia: General. . .

243

Table 2 (continued) Species O. dillenii

Use 1, 8

Management type W

Domestication No

O. duranguensis O. excelsa

1, 4, 9 1, 9

W, C W

No No

O. engelmani O. escuintlensis O. ficus-indica

4 1 1, 2, 4, 5, 8, 10, 11, 12, 13

W W C

No No Yes

O. fuliginosa

1, 2, 11

W, M, C



O. huajuapensis

1, 2, 3, 4, 5, 8, 9 1, 2, 4, 11

W

No

W, M, C

Yes

O. hyptiacantha

O. imbricata

1, 2, 4, 5, 8

W

No

O. jaliscana

1, 2, 11

W, M



O. joconostle

1, 11

W, M, C

Yes

O. karwinskiana O. kleiniae O. lagunae

8 7 1, 4, 8

W W, C W

No –

O. lasciacantha

1, 2, 4, 9, 11

W, M, C

Yes

References Soberón et al. (2001) and Shirazinia et al. (2019) Soberón et al. (2001) Soberón et al. (2001) and Cornejo-Denman and Arreola-Nava (2008) Soberón et al. (2001) Casas and Barbera (2002) Bravo-Hollis (1978), Colunga (1984), Casas et al. (2001), Casas and Barbera (2002), Reyes Agüero et al. (2005), and Hernández-Xolocotzi (1993) Colunga (1984) and Cornejo-Denman and Arreola-Nava (2008) Colunga (1984)

Bravo-Hollis (1978), Colunga (1984), Hernández-Xolocotzi (1993), Paz-Navarro (2021), this study, and Cornejo-Denman and Arreola-Nava (2008) Bravo-Hollis (1978), Sánchez-Mejorada (1982), this study, and CornejoDenman and Arreola-Nava (2008) Bravo-Hollis (1978), Colunga (1984), and Cornejo-Denman and Arreola-Nava (2008) Martínez (1993), Hernández-Xolocotzi (1993), Paz-Navarro (2021), and this study Bravo-Hollis (1978) Soberón et al. (2001) and Mercado-Muñoz (2014) Bravo-Hollis (1978) and Cornejo-Denman and Arreola-Nava (2008) (continued)

244

A. Paz-Navarro et al.

Table 2 (continued) Management type W

Domestication No

1, 2, 4, 11

W, M, C

Yes

O. linheimeri O. littoralis O. lutea O. macrorhiza O. megarhiza O. megacantha

4 1, 7, 9 1 1, 2, 7 1, 7 1, 2, 4, 11

W W, C W W W W, M, C

No No No No No Yes

O. microdasys

7, 8, 9

W, C

No

O. nerpicolor O. oricola O. pachona O. phaeacantha O. pilifera

1, 5 1, 7, 9 1, 2, 4 4 1, 2, 3, 4, 11

W W, C W, M, C W W, M

No No Yes No No

O. puberula O. pubescens

4 7

W W, C

No No

O. pumila O. pycnantha O. rastrera

2 7 4

W W, C W

No No No

O. rileyi O. robusta

9 1, 2, 4, 5, 11

W W, M, C

No Yes

Species O. leptocaulis

Use 1, 2, 4, 5, 8

O. leucotricha

References Sánchez-Mejorada (1982), Felger and Moser (1985), and Hernández-Xolocotzi (1993) Bravo-Hollis (1978), Colunga (1984), Soberón et al. (2001), Paz-Navarro (2021), and this study Soberón et al. (2001) Soberón et al. (2001) Casas and Barbera (2002) Chahdoura et al. (2014) Colunga (1984), Hernández-Xolocotzi (1993), Paz-Navarro (2021), this study, and Cornejo-Denman and Arreola-Nava (2008) Soberón et al. (2001), Cornejo-Denman and Arreola-Nava (2008), and Chahdoura et al. (2014) Sánchez-Mejorada (1982) Molina-Velázquez (2001) Soberón et al. (2001) Bravo-Hollis (1978), Casas et al. (2001), Blancas et al. (2010), and Ojeda-Linares et al. (2020) Casas et al. (2001) Cornejo-Denman and Arreola-Nava (2008) Casas et al. (2001) Benavides-Ríos (2016) Paz-Navarro (2021) and this study Guzmán et al. (2003) Bravo-Hollis (1978), Colunga (1984), Hernández-Xolocotzi (1993), Reyes Agüero (2005), Paz-Navarro (2021), and this study (continued)

Traditional Management and Diversity of Opuntia: General. . .

245

Table 2 (continued) Management type W, C W, M, C W, C

Domestication No No No

1 4, 8 1, 2, 3, 4, 5, 11

W, C W W, M, C

– No Yes

W W W W, M, C

No No No No

O tunicata

4 1 1, 5, 8 1, 2, 4, 5, 11, 13 7, 9

W, C

No

O. undulata

1, 7, 11

W, C

Yes

O. velutina O. violacea O. wilcoxi

1, 8, 11 4 1, 8, 9

W, M W W

– No No

Species O. rufida O. spinulifera O. spraguei

Use 9 1, 2, 4 7

O. spinulifera O. stenopetala O. streptacantha

O. stricta O. tapona O. tehuantepecana O. tomentosa

References Guzmán et al. (2003) Scheinvar et al. (2011) Guzmán et al. (2003) and Mercado-Muñoz (2014) Bravo-Hollis (1978) Sánchez-Mejorada (1982) Bravo-Hollis (1978), Colunga (1984), Hernández-Xolocotzi (1993), Reyes-Agüero (2005), Paz-Navarro (2021), and this study Soberón et al. (2001) Soberón et al. (2001) Bravo-Hollis (1978) and Cornejo-Denman and Arreola-Nava (2008) Cornejo-Denman and Arreola-Nava (2008) Bravo-Hollis (1978), Colunga (1984), Hernández-Xolocotzi (1993), and CornejoDenman and Arreola-Nava (2008) Colunga (1984) Soberón et al. (2001) Mercado-Muñoz (2014)

Naquitz, Oaxaca (Flannery 1986), among other sites (MacNeish 1992) more than 10,000 years ago. Together with Agave, maize, squashes, avocado, chili peppers, and beans, Opuntia are among the most iconic plants used by the Mesoamerican people since prehistory (Casas and Barbera 2002). Remains of prickly pears are abundant in archaeological records, and they were seemingly key plant resources used by the first humans that occupied the area that is currently Mexico (MacNeish 1967, 1992; Bravo-Hollis 1978; Colunga-García Marín et al. 1986; Casas and Barbera 2002; Scheinvar et al. 2011). Together with Agave spp. and Prosopis spp., Opuntia species formed what some authors call “semi-desert triad” (Nava-Martínez 2019) that was a basis for the sustenance of the nomadic groups of those regions (Anaya-Pérez and Bautista-Zane 2008). Opuntia species play an outstanding role in Mexican cultures, economy, and history (Barros and Buenrostro 1998); their relevance might be consequence but also a cause of the high diversification that has been mentioned above. Among all cactus

246

A. Paz-Navarro et al.

pear species, Opuntia ficus-indica is the most economically important, with the highest cladode and fruit production. It is considered a domesticated plant, and the identity of its wild ancestors has been under debate (Colunga-García Marín et al. 1986; Palevitch et al. 1993; Labra et al. 2003; Griffith 2004; Reyes-Agüero et al. 2009). However, there are several wild semidomesticated and domesticated species gathered in forests or cultivated in different systems, which provide food and incomes to different sectors of Mexican people (Colunga-García Marín et al. 1986; Reyes-Agüero 2005; Sáenz 2013; Yahia and Saenz 2011). Fruits of wild species are consumed in different regions of Mexico; however, their small size, sour flavor, thick peel, and spininess have motivated human selection in favor of plants producing larger fruits with sweeter (in some cases sourer) pulp, thinner peel (in some cases thicker), and low spininess. Similarly, the young tender stems are edible, raw or cooked in many ways; stems of all species are edible, but people prefer those species and intraspecific varieties producing tasty stems with good consistency and fewer spines (Colunga-García Marín et al. 1986; Casas and Barbera 2002; Reyes-Agüero 2005; Reyes-Agüero et al. 2011). These attributes, among others, have been favored through processes of human selection practiced by two main management types. One of them is silvicultural, which involves management of wild individuals in forests and agroforestry systems. Through this management, people procure increasing phenotypes with favorable attributes by letting them stand when diturbing the forest for different purposes, sometimes also deliberately propagating the favorable phenotypes in situ (in the original place in forests and agroforestry systems) (Casas et al. 1997; Casas 2001). The other main management type is cultivation, which consists in moving propagules from forests to crop fields, including home gardens, and among anthropogenic environments (Casas et al. 1996, 1997). In all these practices, it is possible to identify that human selection operates actively diversifying varieties of Opuntia, according to purposes guided by human culture (Colunga-García Marín et al. 1986; Casas et al. 1997, 2007; Reyes-Agüero et al. 2011). Table 2 summarizes ethnobotanical information reported about use and management types of Opuntia species in Mexico, identifying those species with signs of domestication. The Altiplano Central or Meridional Central Plateau of Mexico (MCPM hereinafter) is placed in the central-northern area of the country, comprising localities of the states of Zacatecas, San Luis Potosí, Aguascalientes, Jalisco, Guanajuato, Querétaro, Hidalgo, Michoacán, and Mexico City. It is a remaining area of the cultural region called “Tunal Grande” or “Gran Tunal,” because of the abundance of Opuntia, which are dominant species in forests (Branniff-Cornejo 1999; Rivera-Villanueva et al. 2020). Throughout this region, previous studies have documented the occurrence of 126 variants associated with 18 species (Reyes-Agüero et al. 2005a, b, 2011; Scheinvar and Gallegos 2011). Such diversity highlights the relevance of this region as one of the main centers of human-caused diversification of the genus throughout Mexico and perhaps the continent. The interaction between humans and Opuntia through management and adaptations to the local environments has influenced a high diversification and arising of varieties of prickly pears with differentiated attributes. These varieties are named and

Traditional Management and Diversity of Opuntia: General. . .

247

classified by the managers according to their traditional knowledge and criteria. These varieties have been recognized by ethnobotanists as ethnovarieties, local or traditional varieties. The traditional systems of classification are based on perceptible, morphological, physiological, or sensorial attributes and are strongly associated with the local worldview, knowledge, uses, practices, and purposes (Velásquez-Milla et al. 2011; Figueredo-Urbina et al. 2021). In the MCPM, Opuntia species and their traditional varieties are widely distributed and located in different types of forests described as crassicaule scrub or thornscrub. There, it is commonly possible to find areas with high density of prickly pears, called “nopaleras” (Bravo-Hollis 1978; Casas and Barbera 2002). Nopaleras are commonly found on slopes, canyons, depressions, alluvial fans, and plain lands. These have been characterized to harbor a remarkable diversity of species and a high density of Opuntia individuals. Some studies have recorded up to 412 plant species associated in these areas (Rzedowski 1965; Reyes-Agüero et al. 1996; del Castillo 2000; Reyes-Agüero and Vallejo 2019). Compared with other xerophytic bushes, nopaleras have structural complexity with several strata (González-Espinosa 1999; del Castillo 2000). For instance, the first stratum may be composed by individuals like Joshua tree or izote species (Yucca filifera, Y. decipiens, among others) that surpass the height of the tallest Opuntia plants (Rzedowski 1965). A second stratum is composed of shrubby and arboreal cacti species from 1 to 4 m in height. In the boundaries of the states of Guanajuato and Querétaro, frosts are uncommon, and there the nopaleras are mainly composed with individuals of O. streptacantha. In contrast, in the northern region comprising the states of Zacatecas and Durango, frosts are frequent and intense and there is a lower density of O. streptacantha individuals, which are combined with or replaced by species like O. leucotricha (Flores-Flores and Yeaton 2003). It has also been recorded that in disturbed sites, O. streptacantha is replaced by O. robusta populations (Rzedowski 1965; del Castillo 2000; Flores-Flores and Yeaton 2000). Also in this stratum, legume trees such as mesquites (Prosopis juliflora) or huizaches (Acacia tortuosa, A. farnesiana, among others) are present, varying from scarce to abundant. Although sometimes the density of Opuntia plants constrains or limits the development of lower strata, a third stratum can be identified. This is mainly composed of shrubby species with heights ranging from 0.4 to 1.0 m, represented in some places by the “gatuño” or “garabatillo” (Mimosa biuncifera), Dalea bicolor, Agave salmiana ssp. crassispina, and O. robusta (Reyes-Agüero and Aguirre-Rivera (2011). In addition to their biological and ecological importance, the nopaleras have been recognized as reservoirs of human cultures. These areas have been the setting of construction of management strategies and techniques, based on ecological experiences and knowledge developed by Indigenous peoples since prehistory. The modern peasant societies have learned the ancient practices and have continued innovating as long as the production systems and their culture have changed. Traditionally, these systems have the capacity of promoting and ensuring biodiversity conservation while procuring subsistence, similar to what has been recorded in other areas of biocultural relevance in Mexico (Toledo and BarreraBassols 2008; Berkes 2012; Casas et al. 1997, 2014).

248

A. Paz-Navarro et al.

In Mexico, Opuntia fruits are consumed fresh or processed into different traditional dishes and products such as juices, jams, candies, and fermented beverages. The cladodes are also prepared in numerous food dishes, for pickles, pills, capsules, body lotions, shampoos, and creams, among other products. Their seeds could be used as agents of flavoring, and in local communities these are commercialized to companies to produce lipsticks and oils for health skin treatments (Pareek et al. 2003; Kaur et al. 2012; Valiente-Banuet et al. 1997; Casas 2002; Dávila-Aranda et al. 2016; Ojeda-Linares et al. 2020, 2021). Stems of Opuntia are collected and used fresh or prepared as silage to feed the livestock, while the dead dried stems are collected as firewood. In such diverse contexts, Opuntia species have been under management and processes of human selection. Some species have been under incipient or advanced domestication, which actively promote the diversification of ethnovarieties (Colunga-García Marín et al. 1986, Reyes-Agüero et al. 2005a, b, 2009, 2011). The MCPM is one of the most relevant centers of diversification of Opuntia associated with management and human selection in Mexico. It is therefore an ideal region to characterize the diversity of the Opuntia varieties, their uses and management, the way human selection is practiced, and the strategies that are carried out for their maintenance. To address these relevant topics, we conducted studies in the community of Laguna de Guadalupe in the state of Guanajuato (Ojeda-Linares et al. 2020; Paz-Navarro 2021). There, several Opuntia species were recorded to be used as food, fodder, and to produce the traditional fermented beverage called colonche or nochoctli and sweets called melcocha and queso de tuna.

Management of Spatial Areas of Opuntia A high diversity of Opuntia species and varieties occur in Laguna de Guadalupe. During our field trips, 30 traditional varieties belonging to 10 Opuntia species were recorded as: O. albicarpa, O. cantabrigensis, O. ficus-indica, O. hyptiacantha, O. joconostle, O. leucotricha, O. megacantha, O. rastrera, O. robusta, O. streptacantha, and Cylindropuntia imbricata (for some taxonomists, Cylindropuntia is a subgenus of Opuntia). Through the free list, local varieties were registered by common name, management systems, and ethnobotanical attributes (Table 3). We also identified that the varieties are distributed in two main systems: (1) the monte (mount) or nopaleras, and (2) home gardens.

The Monte or Nopalera The spatial unit considered by the community as monte is a semitransformed forest area where crassicaule shrub grows, several useful species cohabit there, among them: Agave spp., Yucca spp., Jatropha spp., Prosopis sp., and several species and varieties of Opuntia. This type of vegetation is widely distributed in the nearby plains around Laguna de Guadalupe (Fig. 2), and it is the space that gives identity to the community and provides resources that allow sustenance to local people.

Home garden

Home garden Home garden

Home garden

Monte Home garden Home garden

Monte

O. albicarpa

O. albicarpa O. albicarpa

O. albicarpa

O. cantabrigensis O. ficus-indica O. ficus-indica

O. hyptiacantha

O. hyptiacantha O. hyptiacantha

O. hyptiacantha

Hartona Rebusca

Negrita

Monte Monte/home garden Monte

Home garden

Management space Monte (crassicaule shrubs) Home garden

Blanca Cristalina Blanca Espinuda Burra Blanca Chapeada Blanca Reina Cuija Amarilla Colorada or chapeada Cascarona

Species Cylindropuntia imbricata

Opuntia albicarpa O. albicarpa

Blanca

Traditional variety Clavillina

Queso de Tuna, Melcocha

Colonche, Queso de Tuna, Melcocha Edible stems Fresh fruit

Forraje Fresh fruit Fresh fruit

Fresh fruit

Fresh fruit Fresh fruit

Fresh fruit

Fresh fruit

Fresh fruit

Main uses Medicinal

2

2 2

2

1 5 5

4

5 4

4

5

4

Flavor 1

2

1 3

3

3 5 5

5

5 5

5

5

5

Consistency 1

2

3 3

3

1 5 5

5

5 5

5

5

5

Size 1

2

2 5

5

1 1 1

1

1 1

1

1

1

Skin thickness 5

3

5 3

5

5 1 1

5

4 1

5

1

3

Prickly 5

Carmine Light red Dark carmine

Pink

Purple Yellow Red

Red Light red Green

Green

Green

Green

Color Light green

(continued)

4

4 3

4

4 1 1

1

1 1

1

1

2

Abundance 4

Table 3 Traditional varieties of Opuntia, management spaces, attributes, and abundance perceived in the case study in Laguna de Guadalupe, Guanajuato

Traditional Management and Diversity of Opuntia: General. . . 249

Monte

O. megacantha

O. rastrera

Sangre de Toro Arrastradilla

Tapona Castilla Tapona Blanca Tapona Silvestre Ballita

Home garden

O. megacantha

Home garden

Monte

Monte

Monte

O. robusta

O. robusta

O. robusta

O. streptacantha

Monte

Monte/home garden Monte/home garden Home garden Monte

Jarrita

Management space Monte

O. leucotricha O. megacantha

O. joconostle

O. joconostle

Species O. joconostle

Traditional variety Huevo de Gato Xoconostle Blanco Xoconostle Rojo Duraznillo Memela

Table 3 (continued)

Fresh fruit, edible stems Colonche, queso de tuna, melcocha

Fresh fruit

Fresh fruit

Forraje

Fresh fruit Colonche, Queso de Tuna, Melcocha Queso de Tuna, Melcocha Fresh fruit

Fresh fruit

Fresh fruit

Main uses Fresh fruit

2

2

1

3

1

5

4

2 3

1

1

Flavor 1

1

3

3

3

1

5

5

2 2

4

4

Consistency 4

3

4

4

4

1

5

4

1 4

3

3

Size 3

2

4

4

4

5

1

2

4 2

5

5

Skin thickness 5

5

4

4

1

5

2

3

5 5

3

3

Prickly 3

Dark carmine Light carmine

Light carmine Dark carmine Dark carmine Dark carmine Green

Color Light pink Light green Light red Yellow Carmine

4

5

1

1

4

1

3

1 4

3

3

Abundance 3

250 A. Paz-Navarro et al.

Monte

Monte

O. streptacantha

O. streptacantha

Cardona Blanca Jocotilla or jocoquilla

Monte/home garden

O. streptacantha

Cardona

Monte

O. streptacantha

Charola

Forraje

Colonche, Queso de Tuna, Melcocha Colonche, Queso de Tuna, Melcocha, fresh fruit, edible stems Fresh fruit 2

4

4

3

4

2

2

1

1

3

3

4

5

2

2

4

3

2

2

5

Carmine

Green

Carmine

Carmine

1

1

5

4

Traditional Management and Diversity of Opuntia: General. . . 251

252

A. Paz-Navarro et al.

Fig. 2 General aspect of the landscape of the site studied: (a) showing the areas of “monte” or “nopaleras,” which are forest or secondary forest areas where populations of Opuntia species are the dominant components; (b) higher areas of monte, where individuals of Joshua trees share spaces with Opuntia species. (Photos by the authors)

There, we identified 19 traditional varieties of eight Opuntia species cohabiting in the monte system. The species with the highest number of varieties (five) was O. streptacantha, all of them recognized by the local managers of the monte. Some aspects of practices by the gatherers were recorded, for instance, they prefer O. robusta cladodes or nopalitos for their consumption as vegetables. On the other hand, O. streptacantha fruits are the most valuable to be consumed as fresh fruit. In recent years, an increasing demand for fruits and seeds of this species has generated a growing microindustry dedicated to the collection of fruits of this cactus. The commercial exploitation began transporting nearly 2 tons of prickly pear fruits approximately every third day from September to October. As well, Opuntia dead logs are collected to use as firewood; in fact, there are specific ovens for the use of this firewood. The monte are places where extensive ranching techniques are performed; animals like cows, sheep, and goats graze freely, which consequently have an impact on vegetation by browsing, trampling on the ground, and accumulation of manure in particular areas locally named majadas. The majadas are located in the monte system and are composed by high-density clusters of Opuntia plants with several species forming patches (Fig. 3). The composition of these areas is the result of a continual silvicultural management performed in situ by the prickly pear fruit collectors. The constant gathering of fruits, for instance, those of O. streptacantha, involves a process through which the cladodes with fruits are removed and then the fruits are stored, leaving the cladodes and some fruits on the ground (Fig. 4). Later, because of the vegetative propagation of Opuntia species, new individuals emerge from the remains of cladodes dispersed and associated with gathering. The Opuntia plants collected through this mechanism are promoted, sometimes incidentally, sometimes deliberately. Therefore, as part of this process dense clusters of Opuntia species are formed; then a competition for light and space begins among the resulting plant populations, the people encouraging differentially the growth of some individuals (which may reach 4–5 m), with more bifurcations and a high number of

Traditional Management and Diversity of Opuntia: General. . .

253

Fig. 3 A close-up view of the high population density that Opuntia species may reach in the “majadas”: (a) which are patches of secondary forest where abundance of prickly pears is promoted incidentally or deliberately due to human actions; (b) the inside structure of “majada”; and (c) activities performed inside the “majadas”. (Photos by the authors)

Fig. 4 Aspect of the remains of gathering Opuntia streptacantha: (a) showing the gathering of cladodes and fruits resulting from this activity, which eventually become entire plants associated to the “majadas”; (b) a similar principle promotes deliberately or incidentally the abundance of Opuntia in patches of forest. (Photos by the authors)

cladodes (870  100). Therefore, through silvicultural and deliberate cultivation activities, local people mold the landscape of the monte. As mentioned above, the majadas characteristically have high densities of population of Opuntia species; this limits the growth of other species, except the pirul (Schinus molle), which is commonly associated with the majadas. The average life span of the majadas is more than 50 years, as informed by local people, but such form of management is much more ancient. According to local people, as the majadas reach this age, gatherers stop frequenting them, mainly because it is more difficult to harvest the fruits due to the height of the individuals and because they perceive changes in the flavor and texture, becoming softer fruits (Fig. 5). The lack of harvest in the majada reduces the process of dispersal of cladodes, and the

254

A. Paz-Navarro et al.

Fig. 5 (a) An old or “mature” majada, with ancient individuals of Opuntia. The general landscape is modeled through the management of composition and abundance of prickly pears species in the area. (b) A young “majada” promoted by human activities. (Photos by the authors)

reduced light availability prevents the propagation of new individuals, which impacts the life span of the majadas and gives way to another vegetation type dominated by S. molle. The majadas provide an optimal microclimate and a physical structure that favors the cattle can stay there generating an accumulating manure or *majada*, which gives the name to this system. The majadas have been used for hundreds of years by ancient cultural groups which used them as shelter and to satisfy food needs. Currently, the majada play a similar role, and the incorporation of new techniques has allowed the diversification of products obtained from the cactus pear fruits such as melcocha, queso de tuna, and colonche, products that are described ahead.

Home Gardens The home gardens of LG have an area of approximately 400–1000 m2, which are commonly located next to the house where the families live. In these spaces, people cultivate species and varieties of Opuntia, and other plants, including medicinal herbs, vegetables, quelites, and ornamental flowers, and chickens are raised. The products of these spaces are mainly intended for self-consumption. People cultivate Opuntia for obtaining nopalitos, fresh fruit, and fuelwood for the direct consumption by the household. However, these spaces are being reduced because the expansion of family buildings is progressively more common, increasing the number of rooms of the house by occupying areas of the home garden. We registered 15 traditional varieties of prickle pears in homegardens, which correspond to 8 species, most of them of the species O. albicarpa (6 varieties). Abundance, in terms of number of individuals, of Opuntia in home gardens, is generally lower than that of the monte system, since commonly one or two individuals are found per variety in each home garden and each home garden has between four and nine varieties. In these spaces, it is possible to find species that are part of the forest or secondary forest vegetation. These are individuals that were there before building the house and that are therefore varieties from the monte system that they decided not to remove

Traditional Management and Diversity of Opuntia: General. . .

255

from the home garden. This is the practice called tolerance by Casas et al. (1997, 2007), Blancas et al. (2010, 2013), and other authors. This practice was identified to occur on varieties such as cardona (O. streptacantha), rebusca (O. hyptiacantha), and xoconostles (O. joconostle). In addition, there are varieties that have been brought from other localities and that are transplanted, such as the blanca (O. albicarpa), amarilla, and colorada (O. ficus-indica) varieties. It is rare that varieties from the monte system are transplanted to home gardens since it is perceived that these varieties are abundant in the surroundings, and it is therefore unnecessary to have them in the garden. Even so, it may occur when an interesting variety is identified in the monte, and when a cactus is transplanted, it is carried out by using cladodes between the months of January to March, so that it manages to generate roots before the rain season starts, thus preventing rotting of the selected individual’s propagules. One of the most common maintenance practices is pruning, which is usually annual and is carried out with the objective of preventing the Opuntia growing too much, no more than 2.5–3 m in height, as this complicates the collection of nopalitos and fruits. Pruning is practiced in both the monte systems and home gardens.

Traditional Products of Opuntia Species Prickly pears are used in LG in different ways, as referred to above. In this section, we describe some of the most important uses that are scarcely known and poorly documented. Fruits are harvested from wild, silvicultural managed and cultivated Opuntia plants without peeling, their peel is used as fodder and to prepare fertilizers, the seeds are transformed for the production of oils and the pulp for cosmetic dyeing. These new uses have intensified harvesting Opuntia fruits. We still do not know the impacts that the change in harvest intensity have, especially on populations from forest and agroforests; therefore, the investigations should be continued to document these aspects. Similarly, harvesting cladodes have been increased associated to their use as fodder, and the impact should be studied. These are relatively recent uses of Opuntia, but some others are ancient and scarcely documented. Some of these forms of use have motivated for centuries management of prickly pears and such management have had implications as mobiles of domestication and diversification of several species, among them, outstandingly, O. streptacantha. But both processes and their results require more research. In the following subsections, we describe some of these traditional products in LG and the needs of research for a deeper understanding of their biocultural value.

Prickly Pears and Sweet Products Through the workshop and semistructured interviews we identified about 30 families that are dedicated to collecting Opuntia fruits to transform them into the traditional food products colonche, melcocha, queso de tuna and miel de tuna. These foods are commercialized within the town or in the neighboring markets of the municipality of Ocampo, in the state of Guanajuato and in Villa de Arriaga, in the state of San Luis Potosí. The production of these issues generated during the availability season of

256

A. Paz-Navarro et al.

prickly pears is relevant, as it contributes to the economic security of the producers for several more months, complementing the main economic activities (rainfed agriculture and cattle arising). But these products may be the main source of income in areas where there is no agricultural production.

Queso De Tuna and Melcocha: Opuntia Sweets Queso de tuna and melcocha are traditional Mexican sweets, mainly prepared with O. streptacantha fruits. However, when the availability of these fruits decreases, other Opuntia fruits are used, mainly O. hyptiacantha and O. megacantha. The main attributes for the selection of other fruits are a high-density consistency of the pulp, a recognized sweet flavor, a medium-size fruit, few spines, and red color of the pulp. For producing these issues, the prickly pears are collected in a plastic container and peeled in the collecting site or transported to the house and peeled there. Almost ten plastic containers of 20 l of fruits are required to produce approximately 20 kg of queso de tuna or melcocha. After peeled, the fruits are moved to a mill called arnero in order to squash the fruit pulp and obtain their juice. Then, seeds are removed, and the juice placed in a copper pan, where it is cooked for almost 9 h. It is important for producers to carefully remove all the seeds, because otherwise the cooked product will display a smoky flavor. Opuntia logs previously collected in the monte are used to maintain the fire for cooking prickly pears juice. When the pulp starts boiling, the producers (called melcocheros) remove the content from the copper pan and places it into a stone or a hardwood container known as batea in order to beat this dense paste. This current action will give the paste a harder consistency and a brown color similar to a jam, indeed, melcocha is a kind of traditional jam. On the other hand, the queso de tuna must be beaten for longer time and more vigorously. After it has been strengthened, the paste is placed in a table and cut into a square or circular shape and stored under cold conditions until it is commercialized in the local and regional markets. Through the current ethnobiological assessment we identified only three elderly melcocheros with an age interval of 65–85 years old in the community, who have maintained this family trade. In fact, this could be one of the last generation of producers, because there is a low interest to continue with these products among young people. Also, because of the current competition generated by the junk food industry, which has displaced traditional foods from the diet and sweets are no the exception. In addition, these products hardly can compete with industrialized facilities in the market, or it is just a product not having a good reception by the consumers. Colonche: A Traditional Fermented Juice Colonche is a traditional fermented beverage that can be prepared with fruits of at least 17 cacti species. Colonche is the common name of this beverage in the region of the Southern Mexican Highlands, where it is mainly prepared with several Opuntia species, although the most common and favored by the producers and consumers is that prepared with O. streptacantha fruits (Ojeda-Linares et al. 2020, 2021). It has a low alcoholic content with an intense magenta color, a thick texture, and a sweet flavor. It is produced during the fruit production season of O. streptacantha from August to November. Local

Traditional Management and Diversity of Opuntia: General. . .

257

Fig. 6 Colonche production: (a) Producer of the fermented beverage “colonche,” pouring the fermented product from a clay pot, which is the main recipient where the fermentation takes place; (b) another producer with plastic containers for colonche production; and (c) a producer of melcocha with his frinder for extracting the Opuntia streptacantha juice. (Photos by the authors)

people have organized the “Colonche Fair,” which is carried out in September. The fermentation to prepare colonche is mostly spontaneous and occurs in clay pots that have been maintained by several generations of the producers (Fig. 6). These pots are extraordinarily important since recent studies demonstrate that they are reservoirs of microorganisms responsible for colonche fermentation; thus, when one of these pots is broken, people make use of its pieces in new fermentation (Ojeda-Linares et al. 2020, 2021). In addition, some people use to inoculate their colonche by using a portion of fermented juice that remained from previous production. This ferment is called xinaiste, pie, or semilla, which hints the concept existing in people’s mind that this material is responsible for starting (and directing) the fermentation process. Colonche can also be prepared with fruits from other species such as O. hyptiacantha or O. megacantha which have ripening time later than that of O. streptacantha, but since they have similar qualities to the latter species, their use allows extending the season of colonche production. This beverage with cacti fruit, receives different names among distinct cultures. It has been reported to be produced in northern Mexico by using fruit of saguaro (Carnegiea gigantea), cardón (Pachycereus pringlei), and pitaya (Stenocereus thurberi). Also, it is produced in southern-central Mexico, using fruits of several species of Opuntia and columnar cacti like Stenocereus stellatus, S. pruinosus, Escontria chiotilla, and most importantly, Pachycereus weberi (Ojeda-Linares et al. 2020).

Traditional Varieties of Opuntia, Classification, and Attributes In LG, we identified 30 local varieties belonging to 11 species, and we ranked them based on free lists provided by local people. Opuntia streptacantha, locally known as nopal cardón and its fruit, called tuna cardona, is the most valued species in the community. It is the one that has the highest number of uses and that is perceived to have the better attributes.

258

A. Paz-Navarro et al.

A classification of the prickly pear fruits based on the phenotypic attributes that people use to identify them, together with the abundance perceived, the management spaces where they are maintained and managed, and the uses they make to the Opuntia species and varieties are reported in Table 1. The ordering of varieties through the NMDS (stress value: 0.0757) shows two groups (Fig. 7). The group at the left of the plot includes the varieties that are clustered according to their attributes, mainly flavor, size, and consistency, also those that are found mainly in home gardens; these are the cases of the blanca (O. albicarpa), amarilla, or chapeada (O. ficus-indica) varieties, which produce prickly pears much sweeter, larger, and juicier than the rest. We also identified in this cluster three varieties whose provenance is from the monte but that are present in home gardens, because they have fruits of large size, sweet taste, and juicy consistency. Jarrita (O. megacantha) and cardona (O. streptacantha) varieties have characteristics that make them highly valued and used; therefore, they are present close to people’s homes.

Fig. 7 Nonmetric multidimensional scaling (NMDS) of the attributes used to identify varieties of prickly pears present in monte (triangles), home gardens (squares), and in both (circles) in Laguna de Guadalupe

Traditional Management and Diversity of Opuntia: General. . .

259

At the right side of the plot, it is possible to see a group of varieties from the monte, except the duraznillo variety (O. leucotricha), which was found in the home garden because it is small, acidic, and has a high number of spines. This group is characterized by having a greater number of spines and a thicker skin than the varieties grouped on the left. All these features are characteristic of the wild Opuntia species. The group of the monte prickly pears is more dispersed, with respect to those of home garden, due to their greater variability in attributes of the varieties and species that are part of this group. Another relevant case is O. joconostle, which has acid flavor and high skin thickness and is present in the monte, but because of its importance in traditional medicine and food, this species is maintained in home gardens.

Conclusion Traditional knowledge is composed by the complex interactions of collective knowledge; it is based on the characteristics of the group, the type of interactions, and practices performed to maintain the biological resources and ecosystems (Berkes et al. 2000). The body of knowledge is maintained, replicated, and continually innovated by the local human groups. It is stratified, where subgroups due to its characteristics may have a greater depth or specificity on some components of the ecosystems. And it is also dynamic not only with the intrinsic processes of increasing knowledge of all human cultures, but also because of the influence of the changing cultural and ecological contexts where people live and interact with. In our study, traditional knowledge over Opuntia species in LG shows much of the general pattern delineated above. Most of it is generalized among the local population, the inhabitants refer to that knowledge, and practices were inherited from their parents and grandparents. Nevertheless, the group of people called coloncheros and those who are involved in the gathering and other uses related to the prickly pear fruits exhibit a specialized knowledge on attributes of species and varieties of this genus. The local producers of melcocha and colonche are key members of the community due to their extensive knowledge about the diversity of prickly pears, their attributes, and the classification of each one of them. Thus, they recognize a greater number of varieties and specific traits associated with the fruits and also with the Opuntia plants than other members of the community. When classification cannot be easily described, the cardona prickly pear is used as a point of reference (for comparison of attributes), because it is the most used and appreciated. But outstandingly, they possess specialized knowledge in relation to management of plants in the two main management systems recognized (the monte and homegardens), and about the details of techniques used for preparing the products referred to. To summarize, we identified that traditional knowledge over Opuntia species goes through a specialization process in Laguna de Guadalupe. These results are similar to what was reported in other cultural groups as the Ixcatec in the state of Oaxaca (Rangel-Landa et al. 2016) and the Purépechas in Michoacán state (Farfán et al. 2018). So far, the traditional knowledge of Opuntia species in Laguna de Guadalupe is maintained and shared by the entire community, which makes it resistant and resilient.

260

A. Paz-Navarro et al.

Acknowledgments The authors thank the Posgrado en Ciencias Biológicas, UNAM, and the CONACYT, Mexico, for financial support of postgraduate studies. Moreover, to CONACYT, research project A1-S-14306, DGAPA UNAM IN 206520 and IN 224023, and CONABIO/GEF/ FAO ID project 9380 770, research project RG023 for financial support of the field and technical work. We mainly thank the locals of Laguna de Guadalupe for all the support and the maintenance of several Opuntia species.

References Albuquerque UP, da Cunha LVFC, De Lucena RFP, Alves RRN, editors. Methods and techniques in ethnobiology and ethnoecology. New York: Springer; 2014. Anaya-Pérez MA, Bautista-Zane R. El nopal forrajero en México: del siglo XVI al siglo XX. Agric Soc Desarrollo. 2008;5(2):167–83. Arba M, Falisse A, Choukrallah R, Sindic M. Biology, flowering, and fruiting of the cactus Opuntia spp.: a review and some observations on three varieties in Morocco. Braz Arch Biol Technol. 2017;60:e17160568. Barros C, Buenrostro I. El maravilloso nopal: sus propiedades alimenticias y curativas. Grijalbo, Mexico City, Mexico; 1998. Beccaro GL, Bonvegna L, Donno D, Mellano MG, Cerutti AK, Nieddu G, Chessa I, Bounous G. Opuntia spp. biodiversity conservation and utilization on the Cape Verde Islands. Genet Resour Crop Evol. 2015;62:21–33. Benavides-Ríos EC. Patrones de sensibilidad de las cactáceas ante el cambio climático en la Peníncula de Baja California, México (Master in science dissertation). La Paz: Centro de Investigaciones Biológicas del Noroeste S. C.; 2016. Berkes F. Implementing ecosystem-based management: evolution or revolution? Fish Fish. 2012;13(4):465–76. Berkes F, Colding J, Folke C. Rediscovery of traditional ecological knowledge as adaptive management. Ecol Appl. 2000;10(5):1251–62. Blancas J, Casas A, Rangel-Landa S, Torres I, Pérez-Negrón E, Solís L, Moreno AI, Delgado A, Parra F, Arellanes Y, Caballero J, Cortés L, Lira R, Dávila P. Plant management in the TehuacánCuicatlán Valley. Econ Bot. 2010;64(4):287–302. Blancas J, Casas A, Pérez-Salicrup D, Caballero J, Vega E. Ecological and sociocultural factors influencing plant management in Náhuatl communities of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013;9:39. Branniff-Cornejo B. La región septentrional mesoamericana. In: Rojas Rabiela T, Murra JV, editors. Historia general de América Latina. Vol. I: Las sociedades originarias. Madrid: Organización de las Naciones Unidas para la Educación, la Ciencia y la Cultura; 1999. p. 229–60. Bravo-Hollis H. Las cactáceas de México, vol. 1. Mexico City: Universidad Nacional Autónoma de México; 1978. Caruso M, Curró S, Las Casas G, La Malfa S, Gentile A. Microsatellite markers help to assess genetic diversity among Opuntia ficus indica cultivated genotypes and their relation with related species. Plant Syst Evol. 2010;290:85–97. Casas A. Silvicultura y domesticación de plantas en Mesoamérica. In: Rendón B, Rebollar S, Caballero J, Martínez MA, Pérez MA, editors. Plantas, cultura y sociedad. Estudio sobre la relación entre seres humanos y plantas en los albores del Siglo XXI. México City: Universidad Autónoma Metropolitana-Iztapalapa/SEMARNAP; 2001. p. 123–57. Casas A. Uso y Manejo de Cactáceas Columnares Mesoamericanas. CONABIO. Biodiversitas. 2002;40:18–23. Casas A, Barbera G. Chapter 9. Mesoamerican domestication and diffusion of cacti. In: Nobel PS, editor. Cacti: biology and uses. Los Angeles: University of California Press; 2002. p. 143–62. Casas A, Vázquez MC, Viveros JL, Caballero J. Plant management among the Nahua and the Mixtec of the Balsas River Basin: an ethnobotanical approach to the study of plant domestication. Hum Ecol. 1996;24(4):455–78.

Traditional Management and Diversity of Opuntia: General. . .

261

Casas A, Caballero J, Mapes C, Zárate S. Manejo de la vegetación, domesticación de plantas y origen de la agricultura en Mesoamérica. Bol Soc Bot Méx (Bot Sci). 1997;61:31–47. Casas A, Caballero J, Valiente-Banuet A. Use, management and domestication of columnar cacti in south-central Mexico: a historical perspective. J Ethnobiol. 1999;19(1):71–95. Casas A, Valiente-Banuet A, Viveros JL, Dávila P, Lira R, Caballero J, Cortés L, RodríguezArévalo I. Plant resources of the Tehuacán Valley, Mexico. Econ Bot. 2001;55(1):129–66. Casas A, Otero-Arnaiz A, Pérez-Negrón E, Valiente-Banuet A. In situ management and domestication of plants in Mesoamerica. Ann Bot. 2007;100(5):1101–15. Casas A, Blancas JJ, Pérez Negrón E, Torres García I, Vallejo Ramos M, Rangel Landa S, FarfánHeredia B. Manejo sustentable de recursos naturales: naturaleza y cultura. In: Arias G, Hernández F, Zurita-Zafra M, Uranga López ML, editors. Sustentabilidad e interculturalidad Paradigmas entre la relación cultura y naturaleza. Pátzcuaro: Universidad Intercultural Indígena de Michoacán; 2014. p. 10–9. Chahdoura H, Barreira JCM, Barros L, Santos-Buelga C, Ferreira ICFR, Achour L. Phytochemical characterization and antioxidant activity of the cladodes of Opuntia macrorhiza (Engelm.) and Opuntia microdasys (Lehm.). Food Funct. 2014;5(9):2129–36. Chávez-Moreno CK, Tecante A, Casas A. The Opuntia (Cactacea) and Dactylopius (Hemiptera: Dactylopiidae) in Mexico: a historical perspective of use, interaction and distribution. Biodivers Conserv. 2009;18:3337–55. Colunga P. Variación morfológica, manejo agrícola y grados de domesticación de Opuntia spp. en el Bajío guanajuatense (Master thesis). Texcoco: Colegio de Postgraduados Montecillos; 1984. Colunga-García Marín P, Hernández-Xolocotzi E, Castillo A. Variación morfológica, manejo agrícola y grados de domesticación de Opuntia spp. en el Bajío Guanajuatense. Agrociencia. 1986;65:7–49. Cornejo-Denman A, Arreola-Nava H. Usos actuales y potenciales de las cactáceas de Jalisco. XIX Semana de la Investigación Científica. Avances en la Investigación Científica del CUCBA, Zapopan; 2008. p. 201–8. Dávila-Aranda P, Rodríguez-Arévalo I, García-Rojas L, Lecona-Rodríguez A. Ethnobotany and ex situ conservation of plant genetic resources in México. In: Lira R, Casas A, Blancas JJ, editors. Ethnobotany of Mexico. New York: Springer; 2016. p. 475–89. Del Castillo RF. Composición y estructura de una nopalera bajo situaciones contrastantes de exposición de ladera y herbivoría. Bol Soc Bot Méx (Bot Sci). 2000;65:5–22. Eberhard DM, Simons GF, Fennig CD, editors. Ethnologue: languages of the world. 24th ed. Dallas: SIL International; 2021. Farfán B, Casas A, García-Frapolli E, Moreno-Calles AI, Castilleja A. Ethnoecology of the interchange of wild and weedy plants and mushrooms in Phurépecha markets of Mexico: economic motives of biotic resources management. J Ethnobiol Ethnomed. 2018;14(5):1–19. Felger R, Moser MB. People of the desert and sea. Ethnobotany of the Seri Indians. Tucson: University of Arizona Press; 1985. Figueredo-Urbina CJ, Álvarez-Ríos GD, García-Montes MA, Octavio-Aguilar P. Morphological and genetic diversity of traditional varieties of agave in Hidalgo State, Mexico. PLoS One. 2021;16(7):e0254376. Flannery KV. Guilá Naquitz. New York: Academic Press; 1986. Flores-Flores JL, Yeaton RI. La importancia de la competencia en la organización de las comunidades vegetales en el Altiplano Mexicano. Interciencia. 2000;25(8):365–71. Flores-Flores JL, Yeaton RI. The replacement of arborescent cactus species along a climatic gradient in the southern Chihuahuan Desert: competitive hierarchies and response to freezing temperatures. J Arid Environ. 2003;55(4):583–94. Glimn-Lacy J, Kaufman PB. Botany illustrated: introduction to plants, major groups, flowering plant families. New York: Springer; 2006. González-Espinosa M. Interacciones entre fenología, elementos bióticos y disturbio por pastoreo en las nopaleras del centro de México. In: Aguirre-Rivera JR, Reyes Agüero JA, editors. Memoria del VIII congreso nacional y VI internacional sobre el conocimiento y aprovechamiento del nopal. San Luis Potosí: Instituto de Investigación de Zonas Desérticas, Universidad Autónoma de San Luis Potosí; 1999. p. 342–59.

262

A. Paz-Navarro et al.

Griffith MP. Experimental hybridization in northern Chihuahuan desert region Opuntia. Aliso. 2001;20:37–42. Griffith MP. The origins of an important cactus crop, Opuntia ficus-indica (Cactaceae): new molecular evidence. Am J Bot. 2004;91(11):1915–21. Guzmán U, Arias S, Dávila P. Catálogo de cactáceas mexicanas. Mexico City: Universidad Nacional Autónoma de México (UNAM), Comisión Nacional para el conocimiento y uso de la Biodiversidad (CONABIO); 2003. Hernández-Xolocotzi E. Aspects of plant domestication in Mexico: a personal view. In: Ramamoorthy TP, Bye R, Lot A, Fa J, editors. Biological diversity of Mexico. Origins and distribution. Oxford, UK: Oxford University Press; 1993. p. 733–56. Heuzé V, Tran G. Prickly pear (Opuntia ficus-indica). In: Feedipedia, a programme by INRA, CIRAD, AFZ, Rome; 2017. Illoldi-Rangel P, Ciarleglio M, Sheinvar L, Linaje M, Sánchez-Cordero V, Sarkar S. Opuntia in México: identifying priority areas for conserving biodiversity in a multi-use landscape. PLoS One. 2012;7(5):e36650. Kaur M, Kaur A, Sharma R. Pharmacological actions of Opuntia ficus indica: a review. J Appl Pharm Sci. 2012;2(7):15–8. Konings A, Konings G. Cacti of Texas in their natural habitat. Austin: Cichlid Press; 2009. Kumar K, Singh D, Singh RS. Cactus pear: cultivation and uses. Bikaner: ICAR–Central Institute for Arid Horticulture Bikaner; 2018. Labra M, Grassi F, Bardini M, Imazio S, Guiggi A, Citterio S, Banf E, Sgorbati S. Genetic relationships in Opuntia Mill. genus (Cactaceae) detected by molecular marker. Plant Sci. 2003;165(5):1129–36. MacNeish RS. A summary of the subsistence. In: Byers D, editor. The prehistory of the Tehuacan Valley. Volume 1. Environment and subsistence. Austin: University of Texas Press; 1967. p. 290–331. MacNeish RS. The origins of agriculture and settled life. London: University of Oklahoma Press; 1992. Martínez M. Catálogo de nombres vulgares y científicos de plantas mexicanas. Mexico City: Fondo de Cultura Económica; 1993. Martínez-González CR, Gallegos-Vázquez C, Scheinvar L. Re-evaluation of Opuntia matudae (Cactaceae). Phytotaxa. 2019;423(3):158–70. Maxwell JA. Qualitative research design. An interactive approach. 3rd ed. Thousand Oaks: Sage; 2013. Mazri MA. Cactus pear (Opuntia spp.) breeding. In: Al-Khayri J, Jain S, Johnson D, editors. Advances in plant breeding strategies: fruits. New York: Springer; 2018. Mercado-Muñoz F. Diversidad y sistemática del género Opuntia s. s., en la región de Los cabos, Baja California Sur, México (Master in sciences dissertation). La Paz: Centro de Investigaciones Biológicas del Noroeste; 2014. Molina-Velázquez, M. Caracterización, ordenación y clasificación numérica en nopal (Opuntia spp.) mediante atributos morfológicos y físco-químicos (Master in sciences dissertation). Marín: Universidad Autónoma de Nuevo León; 2001. Monteiro A, Cheia VM, Vasconcelos T, Moreira I. Management of the invasive species Opuntia stricta in a Botanical Reserve in Portugal. Weed Res. 2005;45(3):193–201. Nava-Martínez LH. La Triada del Semidesierto. Rev Chicomoztoc. 2019;1(2):146–80. Nefzaoui A, Ben Salem H. Cacti: efficient tool for rangeland rehabilitation, drought mitigation and to combat desertification. Acta Hortic. 2002;581:295–315. Nefzaoui A, Louhaichi M, Ben SH. Cactus as a tool to mitigate drought and to combat desertification. J Arid Land Stud. 2014;24:121–4. Ojeda-Linares CI, Vallejo M, Lappe-Oliveras P, Casas A. Traditional management of microorganisms in fermented beverages from cactus fruits in Mexico: an ethnobiological approach. J Ethnobiol Ethnomed. 2020;16(1):1–12. Ojeda-Linares CI, Álvarez-Ríos GD, Figueredo-Urbina CJ, Islas LA, Lappe-Oliveras P, Nabhan GP, Casas A. Traditional fermented beverages of Mexico: a biocultural unseen foodscape. Foods. 2021;10(10):2390.

Traditional Management and Diversity of Opuntia: General. . .

263

Oksanen J, Guillaume F, Friendly M, Kindt R, Legendre P, McGlinn D, Minchin P, O’Hara R, Simpson G, Solymos P, Stevens H, Szöcs E, Wagner H. Vegan: Community Ecology Package. Ordination methods, diversity analysis and other functions for community and vegetation ecologists, version 2.4-3; 2019. Palevitch D, Earon G, Levin I. Treatment of benign prostatic hypertrophy with Opuntia ficus-indica (L.) Miller. J Herbs Spices Med Plants. 1993;2:45–9. Pareek O, Singh R, Vashishtha B. Performance of cactus pear (Opuntia ficus-indica (L.) Mill.) clones in hot arid region of India. J Prof Assoc Cactus Dev. 2003;5:121–30. Paz-Navarro A. Nohpalli: El nopal tunero. Manejo Tradicional de Opuntia spp. (Cactaceae) en la localidad de Laguna de Guadalupe, Guanajuato (Bachelor dissertation). Morelia: Licenciatura en Ciencias Ambientales, Escuela Nacional de Estudios Superiores, Universidad Nacional Autónoma de México; 2021. Pennington C. The Tepehuan of Chihuahua: their material culture. Salt Lake City: University of Utah Press; 1969. Pinedo-Espinoza JM, Aguirre-Mancilla CL, Jiménez-Alvarado R, Raya-Pérez JC, Iturriaga G, Ramírez-Pimentel JG, Hernández-Fuentes AD. Bioactive compounds and antioxidant activity evolution during the ripening process of 12 Opuntia spp. fruit accessions. Emir J Food Agric. 2017;29:138–46. Powell AM, Weedin JF. Cacti of the Trans-Pecos and adjacent areas. Austin: Texas Tech University Press; 2004. Prat L, Franck N. Morphology and anatomy of Platyopuntiae. In: Inglese P, Mondragon C, Nefzaoui A, Sáenz C, editors. Crop ecology, cultivation and uses of cactus pear. Rome: FAO; 2017. p. 21–8. R Core Team. R: a language and environment for statistical computing. Vienna: R Foundation for Statistical Computing; 2020. Rangel-Landa S, Casas A, Rivera-Lozoya E, Torres I, Vallejo M. Ixcatec ethnoecology: biocultural principles of plant management in Oaxaca, Mexico. J Ethnobiol Ethnomed. 2016;12:30. Reyes-Agüero JA. Variación morfológica de Opuntia (Cactaceae) y la relación con su domesticación en la Altiplanicie Meridional de México (Doctoral dissertation). México City: Facultad de Ciencias, Universidad Nacional Autónoma de México; 2005. Reyes-Agüero JA, Aguirre-Rivera JR. Agrobiodiversity of cactus pear (Opuntia, Cactaceae) in the Meridional Highlands Plateau of Mexico. J Nat Resour Dev. 2011;1:1–9. Reyes-Agüero JA, Vallejo M. Matorral crasicaule. In: La biodiversidad en San Luis Potosí, vol. II. Mexico City: Secretaría de Ecología y Gestión Ambiental de San Luis Potosí, Comisión Nacional para el Conocimiento y Uso de la Biodiversidad (CONABIO); 2019. p. 47–54. Reyes-Agüero JA, González-Medrano F, García-Pérez JD. Flora vascular de la Sierra Monte Grande, municipio de Charcas, San Luis Potosí, México. Bol Soc Bot Méx. 1996;58:31–42. Reyes-Agüero JA, Aguirre-Rivera JR, Hernández HM. Systematic notes and a detailed description of Opuntia ficus-indica (L.) Mill. (Cactaceae). Agrociencia. 2005a;39:395–408. Reyes-Agüero JA, Aguirre-Rivera JR, Flores-Flores JL. Variación morfológica de Opuntia (Cactaceae) en relación con su domesticación en la Altiplanicie Meridional de México. Interciencia. 2005b;30(8):476–84. Reyes-Agüero JA, Aguirre-Rivera JR, Valiente-Banuet A. Reproductive biology of Opuntia: a review. J Arid Environ. 2006;64:549–85. Reyes-Agüero JA, Aguirre-Rivera JR, Carlín-Castelán F, González-Durán A. Cátalogo de las principales variantes silvestres y cultivadas de Opuntia en la Altipanicie Meridional de México (no. 634.7750216 C3); 2009. Rivera-Villanueva JA, Riojas-López ME, Mellink E. The Tunal Grande and associated tuna groves. Habitat of hunter gatherers. Rev Col San Luis. 2020;9(19):151–76. Rzedowski J. Relaciones geográficas y posibles orígenes de la flora de México. Bot Sci. 1965;29: 121–77. Rzedowski J. Vegetación de México. Mexico City: Limusa; 1978. Rzedowski J. Diversity and origins of the phanerogamic flora of Mexico. In: Ramamoorthy TP, Bye R, Lot A, Fa J, editors. Biological diversity of Mexico. Origins and distribution. Oxford, UK: Oxford University Press; 1993. p. 129–46.

264

A. Paz-Navarro et al.

Sáenz C. Chapter 7. Industrial production of non-food products. In: Agro-industrial utilization of cactus pear. FAO Rural Infrastructure and Agro-Industries Division, in collaboration with the International Technical Cooperation Network on Cactus (FAO-CACTUSNET). Rome: FAO; 2013. Samah S, De Teodoro Pardo CV, Serrato-Cruz MA, Valadez-Moctezuma E. Genetic diversity, genotype discrimination, and population structure of Mexican Opuntia sp., determined by SSR markers. Plant Mol Biol Report. 2016;34:146–59. Sánchez-Mejorada H. Mexico’s problems and programmes monitoring trade in common and endangered cacti. Cactus Succul J Great Brit. 1982;44(2):36–8. Scheinvar L. Opuntia albicarpa Scheivar una nueva especie para la ciencia del estado de México. Rev Jardín Bot Nacl. 1999;20:267. Scheinvar L, Gallegos C. Estado del conocimiento de las especies del nopal (Opuntia spp.) productoras de xoconostles silvestres y cultivadas. Informe final. Mexico City: UNAMCONABIO; 2011. Scheinvar L, Olalde G, Sule D. Especies silvestres de nopales mexicanos, proyecto no. GE005. Informe final. México City: Autónoma de México, Instituto de Biología, SNIB-CONABIO, Universidad Nacional; 2011. Shackleton RT, Witt ABR, Piroris FM, van Wilgen BW. Distribution and socio-ecological impacts of the invasive alien cactus Opuntia stricta in eastern Africa. Biol Invasions. 2017;19:2427–41. Shirazinia R, Rahimi VB, Kehkhaie VR, Sahebkar A, Rakhshandeh H, Askari VR. Opuntia dillenii: a forgotten plant with promising pharmacological properties. Journal of Pharmacopunct. 2019;22(1):16–27. Smith E. Plant remains. In: Byers D, editor. The prehistory of the Tehuacan Valley. Volume 1. Environment and subsistence. Austin: University of Texas Press; 1967. p. 220–5. Smith JJ, Borgatti SP. Salience counts and so does accuracy: correcting and updating a measure for free-list-item salience. J Linguist Anthropol. 1997;7:208–9. Soberón J, Golubov J, Sarukhán J. The importance of Opuntia in Mexico and routes of invasion and impact of Cactoblastis cactorum (Lepidoptera: Pyralidae). Fla Entomol. 2001;84(4):486–92. Tesfay YB, Kreyling J. The invasive Opuntia ficus-indica homogenizes native plant species compositions in the highlands of Eritrea. Biol Invasions. 2021;23:433–42. Toledo VM, Barrera-Bassols N. La memoria biocultural: la importancia ecológica de las sabidurías tradicionales. Barcelona: Icaria Editorial; 2008. Toledo VM, Ordóñez MJ. The biodiversity scenario of Mexico: a review of terrestrial habitats. In: Ramamoorthy TP, Bye R, Lot A, Fa J, editors. Biological diversity of Mexico. Origins and distribution. Oxford, UK: Oxford University Press; 1993. p. 757–78. Valadez-Moctezuma E, Samah S, Luna-Paez A. Genetic diversity of Opuntia spp. varieties assessed by classical marker tools (RAPD and ISSR). Plant Syst Evol. 2015;301:737–47. Valiente-Banuet A, Casas A, Pickersgill B, Caballero J. Ethnobotany and domestication in Xoconochtli Stenocereus stellatus (Cactaceae), in the Tehuacán Valley and La Mixteca, Baja México. Econ Bot. 1997;52:279–92. Velásquez-Milla D, Casas A, Torres-Guevara J, Cruz-Soriano A. Ecological and socio-cultural factors influencing in situ conservation of crop diversity by traditional Andean households in Peru. J Ethnobiol Ethnomed. 2011;7(1):1–20. Wang X, Felker P, Paterson A, Mizrahi Y, Nerd A, Mondragon-Jacobo C. Cross hybridization and seed germination in Opuntia species. J Prof Assoc Cactus Dev. 1996;1:49–60. Yahia EM, Saenz C. Cactus pear (Opuntia species). In: Yahia EM, editor. Postharvest biology and technology of tropical and subtropical fruits. Cambridge, UK: Woodhead Publishing; 2011. p. 290–331.

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico Luis Herna´ndez-Sandoval and Hugo Castillo-Go´mez

Abstract

The northeastern mountain chain of the state of Querétaro, locally known as Sierra Gorda, is part of the Sierra Madre Oriental and constitutes a geographically, ecologically, and culturally diverse region. Tropical, temperate, and semiarid vegetation are the dominant biological communities, but riparian forests and grasslands are also present. Human colonization started with Chichimec groups of hunters and gatherers, and Mesoamerican groups arrived in the region later. Before the arrival of Europeans to Mexico, most of the Chichimec groups had abandoned the area, allowing the Mesoamerican’s expansion. Among the known ethnicities located in the mountains were the Nyaxu (Northern Pame), Teenek (Huastec), Ximpeces, Uzá’ (Chichimeco-jonaz), and Hñöñho (Semidesert Otomí). After the Spanish conquest, some Xi’oi, Teenek, and Hñöñho ethnic groups survived. Recent history shows a cultural mix associated with the diverse environments that generated different forms of plant uses, ecosystem management, and agroecosystems. The goal of this chapter is to provide an overview of the traditional knowledge on plant uses from natural and disturbed vegetation and from local agroecosystems. We registered 739 useful plant species from 19 vegetation types and five agroecosystems. The families Cactaceae, Asteraceae, Fabaceae, Solanaceae, Pteridaceae, and Asparagaceae have the highest number of useful species, whereas the genera with the most useful species are Mammillaria, Opuntia, Agave, Salvia, Physalis, and Ipomoea. Within the ethnobotanical categories, medicinal, edible, ornamental, environmental services, and forage plants were the most important for local people. Most useful pants are obtained from tropical deciduous forest, oak forest, and piedmont scrub, but many species are gathered from the secondary vegetation of all types. The agrosystemic rank is based on the benefits they obtain and number of L. Hernández-Sandoval (*) · H. Castillo-Gómez Facultad de Ciencias Naturales, Universidad Autónoma de Querétaro, Querétaro, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_22

265

266

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

species. The milpas, homegardens (or “solares”), and induced grasslands or pastures (also called “potreros”) are the most important managed areas. The altered vegetation and agroforests have a diffuse border, with many useful species occurring in both areas and their transition zone. This chapter shows a general panorama of the ethnobotany of this important mountainous region of Mexico.

Introduction The highest linguistically diverse regions of the World coincide with the presence of high biodiversity (Sutherland 2003; Gorenflo et al. 2012). Biological, linguistic, and cultural diversity are adaptative responses to the environment with similar processes and dynamics (Sutherland 2003; Hua et al. 2019) Mexico is one of the countries with the highest biocultural diversity (Boege 2008). For instance, the Mexican vascular flora ranks fourth with 23,314 species, 2,854 genera, and 297 families (Villaseñor 2016) and second in endemism (ca. 50%), with additional approximately 700 naturalized, introduced species (Espinoza-García and Villaseñor 2017). Mexico has also a remarkable cultural richness, being fifth with more than 360 recognized language variants from 68 indigenous linguistic groups belonging to 11 linguistic families (Boege 2008; De Ávila 2008; Diario Oficial de la Federación 2008). The state of Querétaro has 4,411 species of vascular plants, 1,289 genera, and 215 families (Villaseñor 2016), representing 18.92%, 45.16%, and 72.39% of the Mexican taxa referred to, respectively. In addition, the state is home to three indigenous groups, two of them in the Sierra Gorda: Xi’oi and Teenek (Vázquez-Estrada 2010). Some important but undervalued native plant properties provide, however, environmental services. An extensive list of such services includes ecological succession process, ecological conservation indicators, erosion control, soil recovering, watershed stabilization, rainwater capture and infiltration, wetlands bioremediation, wind barriers, aquatic weeds prevention, nurse plants for vulnerable species, mining wasteland recovery, and pollen and nectar sources for insect, bird, and bat pollinators, among others (Smith et al. 2012; Martínez et al. 2017). Ethnobotany is the study of human-plant relationships through time and space, and includes plant traditional knowledge, its role on group cosmovision, and plant uses and management as part of social and cultural life (Boege 2008; Martínez et al. 2017). Ethnobotanical knowledge is a fundamental biocultural heritage; the current native people’s traditional knowledge and biocultural memory approaches enlighten plant domestication processes and the state of agrobiodiversity, the different forms of species management, plant habitats, and processes generating and influenced by genetic diversity (Boege 2008; Toledo and Barrera-Bassols 2008; Carándula and Ávila 2013; Casas et al. 2016). Agrobiodiversity also considers edible native plants, traditional medicine, and herbs, as well as forage and ornamental plants. Some less frequent uses, such as plants for constructing tools, honey production, fuel, house construction, and art crafts are also included (Hernández 1985). Not only useful or economically valuable plants are considered, but also those with toxic, ritual, and supernatural properties (Hernández 1985). If plants have a name, people consider

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico

267

them to be part of their culture. Plant uses and management have been developed over time and space through many productive activities (praxis) organized under a traditional knowledge, and both are related to their interpretation of nature (beliefs, rituals, and myths) (Valle-Esquivel et al. 2012).

The Sierra Gorda Environment The Sierra Gorda is located in northeastern Querétaro and is part of the Sierra Madre Oriental. Río Verde and the Huastec fertile flats are the northern limit of the Sierra Gorda, while the edge in the west is the state of Guanajuato, the Quéretaro semiarid area is the southern margin, and the flats of the state of Hidalgo and the Huastec high mountains are the eastern border (Nieto-Ramírez 2010) (Fig. 1). The Sierra Gorda is a series of folded mountains with an extent of approximately 3,500 km2. Elevations range from 500 to 3,500 m, mainly with sedimentary rocks with karstic processes forming caves and sinkholes. Geologically, the late Jurassic (140 MYO) and Cretaceous rocks (60 to 80 MYO) originated at the sea bottom, and there are abundant marine fossils (Nieto-Ramírez 2010; De la Garma 2015). Climate ranges from warm subhumid (A)C1(w0)(w), temperate subhumid C(w2) and C(w2)(w), to semiarid (BS1hw y BS0hw) (Zamudio et al. 1992; Instituto Nacional de Ecología 1999). There are three watersheds, the Santa María, Extórax, and Moctezuma, which drain to the Pánuco River. The main soils in the Sierra are black or dark grey, nutrient-rich lithosols, acid reddish to yellow-brown rocky luvisols, rendzines, and faeozem. Calcaric cambisols, regosoles, eutric fluvisols, and pelic vertisols are also present (Nieto-Ramírez 2010; Instituto Nacional de Ecología 1999).

Biological Diversity Natural conditions of the Sierra Gorda have promoted a remarkable biodiversity, and at least 19 vegetation types are present in the region (Zamudio et al. 1992; Instituto Nacional de Ecología 1999; Bayona et al. 2006): Tropical deciduous forest (TDF). With an extension of 72,409 ha, this is a community dominated by trees 10 to 15 m in height that lose their leaves during the dry season. Common species are Bursera simaruba (chaca), Esenbeckia berlandieri (“jopoy”), Lysiloma microphylla (“palo de arco”), Phoebe tampicensis (“laurel”), Psidium sartorianum (“guayabillo”), Mariosousa coulteri (“guajillo”), and Guazuma ulmifolia (“guácima,” “aquiche”). The authorities of scientific names are cited in the species list of the chapter supplementary material. Tropical subdeciduous forest (TSF). This is a forest with trees 18 to 22 m tall with 50–75% of the species losing their leaves in the dry season. It covers an area of 4,873 ha. The most common species are Brosimum alicastrum (“ojite,” “ramón”), Cedrela odorata (“cedro rojo”), Enterolobium cyclocarpum (“orejón”), Ficus pertusa (“higuerón”), and Bursera simaruba (“chaca”).

268

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Fig. 1 Sierra Gorda, Querétaro, Mexico map

Cloud forest (CF). This is a mixed forest present in the most-humid areas. The common species are Liquidambar styraciflua, (“somerio” o “quirámbaro”), Dalbergia palo-escrito (“palo escrito”), Ulmus mexicana (“petatillo”), Taxus globosa (“granadillo”), Tilia mexicana, Clethra pringlei, Quercus affinis (“encino escobillo”), Q. germana (“encino bellotón”), and Magnolia schiedeana. Oak forest (OF). With an extension of 81,993 ha, the representative oaks are Quercus mexicana, Q. castanea, Q. polymorpha, Q. crassifolia, Q. greggii, and Q. affinis.

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico

269

Pine forest (BP). Occupying 3,357 ha, the representative species include Pinus greggii, P. patula (“pino lacio”), P. montezumae, P. ayacahuite, P. teocote, P. rudis, and P. oocarpa. Pinyon pine forest (PPF). This is a variant of the pine forest with an area of ca. 60 ha, recognized by the presence of the pinyon pines P. cembroides and P. pinceana. Pine-oak forest (POF). Occasionally recognized as oak-pine forest, this has an extension of 40,044.83 ha. The most common species are Pinus greggii, P. patula, Quercus laurina, and Q. obtusata. Fir forest (FF). This coniferous forest is dominated by Abies guatemalensis and A. religiosa, and in some areas Pseudotsuga menziesii occurs. Cypress forest (CyF). This small forest has an area of 400 ha, with Cupressus lusitanica (“ciprés”) as the dominant species. Juniper forest (JF). With an area of 2,750 ha, its common species are Juniperus flaccida, J. deppeana (“nebrito”), and Arbustus xalepensis (“madroño”). Riparian forest (RF). This tree community occurs along rivers and streams with species such as Taxodium mucronatum (“sabino”), Platanus mexicana (“álamo”), Salix bonplandiana (“sauce”), Juglans mollis (“nogal”), Ficus padifolia (“higuerón”), and Psidium guajava (“guayaba”). Thorn scrub (TS). This vegetation type is dominated by spiny shrubs like Calliandra eriophylla, Condalia velutina, Cylindropuntia imbricata, C. leptocaulis, Fouquieria splendens, Mimosa depauperata, M. texana, Vachellia farnesiana, and V. schaffneri. Unarmed species as Bursera fagaroides, Croton ciliatoglandulifer, and Iresine schaffneri are also found. Cactus scrub (CS). With an extension of 17,597.32 ha, it is represented by succulent plants, such as Stenocereus dumortieri (“órgano”), S. queretaroensis (“pitayo”), Myrtillocactus geometrizans (“garambullo”), and Opuntia imbricata (“cardenche”). Sclerophyll scrub or chaparral (SC). This vegetation type occurs in a small area of ca. 200 ha, with shrubby species of oaks or “encinos” (Quercus spp.), Arctostaphylos pungens (“pingüica”), and Litsea glaucescens (“laurel”). Desert scrub (DS). Present only in the western or southern part of the Sierra Gorda covering 3,750 ha. Some characteristic species are Larrea tridentata (“gobernadora”), Condalia mexicana (“granjero prieto”), Fouquieira splendens (“ocotillo”), Koeberlinia spinosa (“junquillo”), and Prosopis laevigata (“mezquite”). Rosette scrub (RS). Occupying 16,849.2 ha, this harbors species of Dasylirion parryanum (“sotol”), D. longissimum (“junquillo”), Agave garcia-mendozae, and Hechtia glomerata (“guapilla”). Piedmont scrub (PS). It is the most extensive scrub in the Sierra Gorda with 76,936.54 ha. Characteristic species are Acaciella angustissima (“barba de chivo”), Acacia berlandieri (“guajillo”), Rhus virens (“zumaque”), Cigarrilla mexicana (“San Pedro”), and Cordia boissieri (“trompillo”). Grassland (G). This is a plant community dominated by grasses and other herbaceous species with 123.3 ha, common species are Aristida divaricata, Bouteloua curtipendula (“banderita”), Muhlenbergia rigida, and some herbaceous

270

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

forbs such as Calochorthus barbatus, Dalea prostrata (“engordacabra”), Dyssodia pinnata (“rosella”), and Oxalis decaphylla (“agrito” o “quelite agrio”). Aquatic vegetation (AV). This includes wetlands with plants either floating or rooted within the water, such as Hydrocotyle ranunculoides, Lemna gibba, Ludwigia peploides, Nymphaea ampla, Nymphoides fallax, Nuphar luteum, and Sagittaria demersa. Plant diversity of the Sierra Gorda is estimated at 2,308 vascular plant species, belonging to 195 families and 943 genera (Zamudio 2003). The most diverse families (species/genera) are the Asteraceae (95/274), Fabaceae (54/130), Orchidaceae (49/117), Poaceae (29/106), Euphorbiaceae (21/87), and Solanaceae (13/81).

Historical and Cultural Context Throughout the history of the Sierra Gorda, different local cultures have converged. Hunter-gatherers inhabited the area for 7,000 years BP. From this time until 300 years BP, people established occasional agricultural experiences and trade with the inhabitants of the plains of the Gulf of Mexico and those of the Mexican Central Highlands (De la Garma 2015; Viramontes Anzures 2010). Archaeological evidence shows the presence of some sedentary groups in localities such as Tancama, Purísima, San Francisco Concá, Tancoyol, Tilaco, and Tanchanaco in the municipalities of Arroyo Seco, Jalpan de Serra, and Landa de Matamoros. However, they belonged to the “serrano culture” which developed because of cinnabar exploitation and commercialization (Viramontes Anzures 2010; StresserPéan 2008; Herrera-Muñoz 2010; Mejía-Pérez Campos and Herrera Muñoz 2013). Some other groups practiced agriculture in the Huastec and Rio Verde areas as Tancama, Jalpan, Concá, Tancoyol, Tangojó, Puginguía, and Tanchanaquito (Viramontes Anzures 2010; Stresser-Péan 2008; Herrera-Muñoz 2010). With the abandonment of the region in the twelfth century, attributed to the social and environmental imbalance, these places were occupied by nomadic warriors such as the Ximpeces, Southern Pames, Jonaces, and Guachichiles, groups of peoples called pejoratively Chichimeca by the Aztecs and “nok” (savages) by the Huastec (Chemin-Bässler 1984; De La Garma 2015; Stresser-Péan 2008). At the time of the Spaniards’ arrival in the region, these nomad groups were dominated and finally disappeared when they were exterminated or due to miscegenation processes in the nineteenth century (Nieto Ramírez 2010; De la Garma 2015; Stresser-Péan 2008; Meade 1951; Viramontes Anzures 2000; Gallardo-Arias 2011). However, these people had an important ethnobotanical knowledge. For instance, they used a large variety of plants as food, such as the “chamal” (Dioon edule), “chochas” or “samandoque flowers” (Yucca treculeana), and “pemoches” (Erythrina americana), as well as many wild animals, some of which are still used (Stresser-Péan 2008; Gallardo-Arias 2011). The current human groups and languages within the Sierra Gorda are the Xi’oi or Central Pame in the municipalities of Jalpan de Serra and Arroyo Seco, the Teenek at

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico

271

the limits of Jalpan de Serra with San Luis Potosí (SLP), and the Hñöñho or Otomí at Cadereyta in a semiarid area (Ordoñez 2004; Mendoza-Rico et al. 2006; VázquezEstrada and Prieto-Hernández 2013; Núñez-López 2014). These cultural groups are recognized in three indigenous micro-regions: (1) Las Flores, Las Nuevas Flores, and Tancoyol (as the result of a Santa María Acapulco, SLP Xi’oi migration early in the twentieth century); (2) Purísima de Arista (Xi’oi) in Arroyo Seco; and (3) Tanchanaquito (Teenek) and Tanzozob (Xi’oi) in Jalpan de Serra (VázquezEstrada 2010; Valle-Esquivel et al. 2012; Vázquez-Estrada and Prieto-Hernández 2013). Currently, the Sierra Gorda has a population of 183,219 inhabitants, of which 4% recognize themselves as indigenous people. However, only 428 people still speak an indigenous language, mainly Xi’oi (Vázquez-Estrada and PrietoHernández 2013; Instituto Nacional de Estadística, Geografía e Informática 2020), and only 1,105 inhabitants are self-recognized as afro descendants, mainly in Pinal de Amoles (Instituto Nacional de Estadística, Geografía e Informática 2020). However, the data of the census of population must be interpreted cautiously since many mestizo people have indigenous or afro relatives (López-Ugalde et al. 2013).

Anthropogenic Ecosystems: Agroecosystems Several authors have recognized different agroecosystems practiced in the region (Moreno-Calles et al. 2013, 2017). However, in this chapter we consider only five general agroecosystems present in the Sierra Gorda, namely: (1) agroforests or humanized forests, (2) “solares” or homegardens, (3) milpas, (4) agroforestry orchards, and (5) “potreros,” pastures or induced grasslands. Further studies are still necessary to describe other systems in the area. Agroforests. Also known as humanized forests, intermediate forests, artificial forests, altered forests, or managed forests. These are recognized to be practiced in Mexico by different cultures, such as the kuojtakiloyan by the Nahua of the Sierra Norte de Puebla, cacaotales in Tabasco, and the tel’om practiced by the Huastec or Teenek. These forests are natural spaces where humans intentionally manage or select the structure and composition of tree species according to their necessities, mainly for subsistence, but also with the intention to preserve the natural tree structure and ecological processes (Alcorn 1990; Moreno-Calles et al. 2013). Agroforestry uses simple technologies, in some cases the swidden agriculture, rotating the milpa with the forest. In the Sierra Gorda, the mestizo community of Agua Zarca, municipality of Landa de Matamoros, and the Teenk community of Jalpan de Serra, near Aquismón, SLP, commonly manage agroforests. “Solares” or homegardens (“huertos familiares”). This system occurs near the family house and contains diverse plant species for different purposes. Fruit trees, spices, herbs, and ornamental plants are frequent. Women are in charge, but most family members participate to take care of and maintain the solar (Hernández-Ruiz et al. 2013).

272

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Milpas (sensu lato). The term milpa comes from the Náhuatl term ¨milpan,” composed by two vocables, “milli” meaning sown plot and “pan,” meaning over it. The milpa is a polyculture characterized by maize associated with other domesticated species, such as beans, squashes, chili, tomatoes, and some other semidomesticated plants as “quelites” (leafy greens) which appear spontaneously and have been managed and protected in this system (Hernández-Ruiz et al. 2013). This agroecosystem with maize as the principal component originated in Mesoamerica and later expanded to all Mexico and North and South America. Agroforestery orchard or semi-commercial orchard contains one to many cultivated rosettes, shrubs, or trees in more temperate or semiarid areas (Lozada-Aranda et al. 2017). Márquez-Sánchez (1976) consider this agroecosystem as temperate semi-commercial orchard. In the Sierra Gorda, they are commonly found to have some species of Agave (“magueyales”), Opuntia (“nopaleras”), or fruit trees. Induced grasslands (potreros). The potreros are grasslands developed at the expense of forested areas devoted for cattle management or abandoned crop fields. Plant diversity usually consists of many non-native species. However, it is possible to find plants (trees and shrubs besides the herbaceous species) with other uses than animal forage as part of the vegetation succession.

Ethnobotanical Categories Eighteen ethnobotanical categories were registered, based on the consulted literature, and here are listed in alphabetical order. (1) Beverages, comprising plants prepared for alcoholic or non-alcoholic drinks. (2) Ceremonial plants, which include species used in rituals and festivities. (3) Construction and timber plants that are used for house building, furniture manufacturing, or railway ties. (4) Edible plants, those that are consumed as food, although sometimes people mix them with the plants used as spices. (5) Ecosystem services, including plants that help to maintain a sustainable environment. (6) Fences, which consist of plant wood from trunks and branches that are employed to construct fences. (7) Fibers, including the species used for textiles, straps, sacks, or brushes. (8) Forage are those plants eaten by cattle and domestic animals. (9) Fuel, which are plants used as firewood. (10) Honey production, including plants whose flowers are visited by bees to produce honey. (11) Horticultural tutors are plants used to support others in horticultural practices. (12) Living fences are plants cultivated to make fences. (13) Medicinal plants, which are those species used in traditional or popular medicine to cure ailments. (14) Ornamentals, including those species currently used to adorn houses, gardens, or open spaces, as well as those considered as potentially ornamental plants. (15) Poisonous and toxic plants that produce human or animal illness or intoxication. (16) Staining plants consider species used to dye clothes, ropes, and other items. (17) Tools are plant species used to make handles or other utensils. (18) Toys include plants that amuse children.

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico

273

Ethnobotanical Knowledge Ethnobotanical information was obtained from literature, databases, and personal interviews and fieldwork. The references reviewed are diverse and come mainly from taxonomical-floristic and ethnobotanical sources of information: Agavaceae (Magallán-Hernández and Hernández Sandoval 2000), Cactaceae (Scheinvar 2004), ethnobotany of the Xi’oi (Miranda-Perkins 2003), chile piquín (Capsicum annuum var. glabriusculum) ethnobotany (Martínez 2007), chilcuague (Castro-González 2009), medicinal plants of Querétaro (Fernández-Nava et al. 2001), traditional medicine of the Xi’oi (Martínez-Spinoso and Miranda 2008), edible plants (Lira and Casas 1998; Cilia-López et al. 2008; Aguilar-Rincón et al. 2010; Espitia-Rangel et al. 2010; Santiaguillo et al. 2010; Castro et al. 2011; Cruz et al. 2012; Lobato-Ortiz et al. 2012; Mera-Ovando et al. 2013), ornamental (Pérez-Nicolás and FernándezNava 2007), timber and firewood species (López et al. 2005), and weeds (Suárez et al. 2004). Additional information comes from the QMEX herbarium database of useful plants, the Germplasm Bank of the Universidad Autónoma de Querétaro (UAQ), and interviews conducted before 2020. Most of the authors cited above deposited herbarium vouchers in the herbaria IEB, MEXU, and/or QMEX (acronyms according to Thiers (2022) in the Index Herbariorum. The information on useful species of the Sierra Gorda cited in the Appendix 1 includes taxonomic data, common names, general plant habit, and ethnobotanical categories. Appendix 2 includes information on vegetation types and agroecosystems. Family names are based on APG IV (Stevens 2017), and species were reviewed in Tropicos.org (MOBOT 2022) for Angiosperms, and in the study of linear sequence of extant families and genera of lycophytes and ferns (Christenhusz et al. 2011) for ferns and fern allies.

Plant Uses We registered 739 plant species with traditional uses, which corresponds to 32% of the flora (Zamudio 2003) of the Sierra Gorda. Out of the total, medicinal plants had the highest number of species (327), followed by actual ornamental plants (203, with both actual and potentially ornamental plants being 405), and edible plants (141) (Figs. 1 and 2). The most representative families are Cactaceae (85), Asteraceae (70), Fabaceae (50), Solanaceae (33), Pteridaceae (29), Asparagaceae (27), Lamiaceae (19), Euphorbiaceae (18), Malvaceae (18), and Poaceae (18). Species of these families constitute 49.8% of all the useful plants in the Sierra Gorda (Fig. 2). Among the richest useful genera, Mammillaria (Cactaceae) stands out with 23 species, Opuntia (Cactaceae) with 21, Agave (Asparagaceae) with 15, Salvia (Lamiaceae) with 14, Physalis (Solanaceae) with 12, Ipomoea (Convolvulaceae) with 11, Pinus (Pinaceae) with nine, and finally Euphorbia (Euphorbiaceae), Pinguicula (Lentibulariaceae), Adiantum (Pteridaceae), and Cheilanthes (Pteridaceae) each with eight. Species of these genera constitute 18.6% of all the useful plants in the Sierra Gorda (Fig. 3).

274

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Fig. 2 Botanical families with most useful species in the Sierra Gorda

Fig. 3 Genera with most useful species in the Sierra Gorda

The most important ethnobotanical categories in terms of diversity are the medicinal plants (329), followed by the ornamental (204), which may increase the species number to 414 if we consider the potentially ornamental plants, and

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico

275

then the edible plants (148) (Figs. 6, 7, and 8). Useful categories having 30 to 100 species were the forage plants (82), with many species in such families as Asteraceae, the genera of Fabaceae (Acacia, Dalea, Desmodium, Mimosa, Senna, and Vachellia), Poaceae (Bouteloua and Paspalum), and others such as Opuntia (Cactaceae) and Solanum (Solanaceae). Species used to manufacture tools (45) belong to the genera Agave, Fraxinus, Quercus, Senna, and Vachellia, as well as species as Brahea berlandieri (Arecaceae), Dasylirion parryanum (Asaparagaceae), and Prosopis laevigata (Fabaceae). Plants clearly recognized by local people for honey production are represented by 34 species of the families Fabaceae, Asteraceae, and Solanaceae. For fuel, firewood, and charcoal our study identified 31 species, the families Fabaceae and Fagaceae are mostly used. Among the fewer than 30 species, 24 are used for toys mainly belonging to the Asteraceae, Phyllanthus, and Opuntia. Plant with good wood (23) are species of Cedrela, Prosopis, Lysiloma, Ehretia, and Juniperus. For rural construction (22) are Prosopis laevigata, Cedrela, Abies, and Condalia. Species for fences (18), include living trees like Bursera simaruba, Erythrina coralloides, and some rosette succulents (Agave americana), cacti (Lophocereus marginatus or Opuntia robusta), and spiny shrubs (Mimosa texana and Vachellia schaffneri), and trunks or branches of other species such as Vachellia pennatula, V. schaffneri, Rhus aromatica, and Celtis. Plants for Ceremonial or ritual uses (17) include the resin (“copal”) of several species of the genus Bursera, Tagetes (“cempasúchil”) flowers are used to celebrate the dead, and some species of Datura (toloaches) and Dasylirion (“sotoles”) are employed in religious or ritual ceremonies. Toxic and poisonous species (15) known to harm people when ingested include Datura ceratocaula and D. inoxia (“toloache”), Karwinskia humboldtiana (“tullidora”), Croton ciliatoglandulifer (“solimán” or “palillo”), and Dioon edule (“chamal”). However, the poisonous seeds of “chamal” become edible once cooked as “tamales” and “atole” (a local beverage). Species for art crafts (11) are those with soft wood like Erythrina for masks, Cedrela flowers and fruits for still life ornaments, and Brahea berlandieri for basketry. In the category “Others” are included plants with rare or specific uses and have less species registered. Some of these are Brickellia veronicifolia used as fermentation catalyst for “pulque” preparation; Cylindropuntia imbricata as ornamental; Clinopodium mexicanum as an aphrodisiac; Agave mapisaga var. mapisaga and Agave salmiana as nutritious and stimulant beverages; Senecio salignus and Ambrosia cordifolia for fireworks; Vachellia farnesiana, Bursera galeottiana, and Eysenhardtia polystachya for dying clothes; Cucurbita foetidissima for cosmetics; E. polystachya for tanning; Agave americana for underground meat cooking (“barbacoa”); Yucca filifera for hard fibers; Zea mays subsp. mays for industrial uses; Manfreda scabra as a soap substitute; Zaluzania augusta to clean prickly pear spines; Dodonaea viscosa as a pesticide; Lysiloma microphyllum as a shade tree; Agave funkiana and Gossypium hirsutum for textile fibers; and Celtis caudata as plant support for horticultural practices (Fig. 4).

276

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Fig. 4 Useful species by ethnobotanical category in the Sierra Gorda

Relevant Plant Uses Medicinal plants are the most numerous species (329) in the Sierra Gorda and account for almost half of all the useful plant species. Besides, ca. 50% of the medicinal species have no additional use. Most of the medicinal plants belong to the families Asteraceae with 51 species, the Fabaceae family with 20, Solanaceae with 15, Euphorbiaceae with 13, and Lamiaceae with 12. The plants are primarily herbaceous with essential oils in their leaves. Among the most frequently used are the “epazotes” Dysphania ambrosioides and D. graveolens, “veintiunilla” Asclepias linaria, Gomphrena serrata, Tetramerium nervosum, and Acourtia reticulata. Some of the medicinal plants are grown in homegardens, but most of them gathered from wild populations from different environments. These plants are used by most people but mainly by traditional healers, shamans, midwives, and bonesetters. Locally, illness can be considered as natural or benign, and supernatural, considered by Chemin-Bässler (Christenhusz et al. 2011) as psychosomatic or cultural. Some of the latter include “espanto” (fright), “mal de ojo” (evil eye), “mal aire” (bad air), and several spells supposedly caused by several living animals (snakes, felines), natural phenomena (accidents, landslides, thunder), or supernatural or divine entities (Miranda-Perkins and Castillo-Gómez 2010). Shamen or specialists can detect and cure these depending on the type of sickness. However, healers are rare to find in the Sierra Gorda, and locals usually travel to Xi’oi areas in SLP Huasteca. FernándezNava et al. (2001) found 57 illness categories, 16 of cultural type, 13 signs and symptoms, nine respiratory system problems, and eight skin and dermal problems. Other common ailments are “ascos” (nausea), “boca amarga” (bitter mouth), “calentura” (fever), “calambres” (cramps), cancer, colic, “corajes” (anger), diabetes, dehydration, diarrhea, dysentery, headaches, body pain, stomach pain, throat pain,

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico

277

teeth pain, “empacho” (kind of indigestion) injuries, pimples, heart disease, “mal de orín” (bad urine), “pulmonía” (pneumonia), cysts, rheumatism, coughing, and vomits, among others (Martínez-Spinoso and Miranda 2008). Medicinal plants might be applied, depending on the hot or cold illness categories, to restore natural equilibrium. Medicinal plant part preparation forms are by fresh or dry leaf and plant shoots boiling (infusion); tubers, roots, stems, barks, or fruit cooked or liquefied; thick plant structures grilled; some plant parts dried and macerated; or branches tying (“manojos”). The medicinal application forms are by ingestion, skin spreading, herb bathing, plasters, vaporizations, or body cleansing by branches sweeping (“limpias” or “barridas”). The second category is ornamental plants with 204 species. People like to grow species with nice flowers as orchids, Dahlia spp. (“dalias”), Ipomoea spp. (“quiebraplatos”), Justicia aurea (“pavón amarillo”), Malvaviscus arboreus (“tulipán capullo,” “malvarisco”), Mirabilis jalapa (“maravillla”), Tagetes erecta (“cempasúchil”), and Zephyranthes spp. (“mayitos”). Ferns of several species are present as well as many cacti (Mammillaria spp., and Crassulaceae (Echeveria spp., Sedum spp.). Common ornamental trees are Bauhinia mexicana (“pata de vaca”), Bursera simaruba (“chaca”), Ceiba pentandra (ceiba), Cercis canadensis (“cuaresma”), Ehretia anacua (“muñeco,” “roble prieto”), Eysenhardtia polystachya (“palo dulce,” “palo azul”), Fraxinus uhdei (“fresno”), Gossypium hirsutum (algodón), Plumeria rubra (“jacalosúchil”), Sambucus nigra (“saluco”), Yucca guatemalensis (“izote”). Among the potentially ornamental plants are species of Agave, Alnus, Ariocarpus, Aristolochia, Begonia, Cactaceae, Chamaedorea, Cobaea, Justicia, Echeandia, Eustoma, Lobelia, Loeselia, Lopezia, Magnolia, Maurandya, Milla, Morkillia, Pavonia, Pinguicula, Pisoniella, Populus, Salvia, Telosiphonia, Tigridia, Tillandsia, and Yucca. These waiting to be grown. The third most important group of useful plants are the edible species with 148 records, many prepared as part of the daily diet, depending on plant availability, such as flowers or fruits. Some important species are the “chilitos” (Mammillaria fruits) and “tunas” (Opuntia prickly pear fruits), different kinds of chili peppers (Caspsicum spp. - Solanaceae), “tomates verdes” (Physalis spp. tomatillos), fruits such as “ciruelas” and “jobos” (Spondias, Anacardiaceae), guayabas (Psidium guajava) and guayabillos (P. sartorianum) both belonging to Myrtaceae, “aguacates” and “paguas” (Persea spp. avocados), and “sapotes” from different species of Rutaceae and Sapotaceae. During the rainy season, while people work on the milpa, “quelites” (herbaceous plants tender leaves and stems) are consumed. The most common “quelite” species belong to Amaranthaceae and Piperaceae, which at the same time, might be used as spices. Several species with multiple uses are considered as polyvalent, with many plant structures (tubers, stems, shoots, flowers, fruits, or seeds) consumed, such as in some domesticated species of “calabazas” (Cucurbita spp.) “chayotes” (Sechium edule), and “frijoles” Phaseolus. Several species of the genera Agave, Yucca, Cucurbita, Cnidoscolus, and Erythrina have edible flowers. The most important cereal and seeds used in the Sierra Gorda are maize (Zea mays L.) and beans (at least three Phaseolus species: P. coccineus, P. heterophyllus, and P. vulgaris). There are many corn and bean preparations, including “tortillas,” normal size “tamales” and the huge “tamal zacahuil” (sometimes larger than 1 m diameter), beverages (“atoles”), and soups, among others.

278

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Currently, the Sierra Gorda gastronomy has been enriched through the inclusion of species that enhance local tastes. However, it is still possible to find important native species for the local inhabitants as Dioon edule (“chamal”), Oxalis spp. (agritos), Piper auritum (acoyo, San Pedro), and Cercis canadensis (cuaresma). Common mestizo dishes are: “rejalgar (Xanthosoma sp..) with “acoyo” (Piper auritum); “tomate coyol salsa” (Solanum lycopersicon naturally grown in the milpas, similar to the cherry tomato); “calabaza and chilacayote” sweets (Cucurbita spp. varieties); “atole de guayabilla” (Psidium spp.); cooked “quelites” or “quelitas” (Amaranthus spp.); “jacubes” (cooked Acanthocereus tetragonus shoots) with “garbanzo” (chickpea); cooked “tunas de nopal chamacuero” (prickly pears) (Opuntia cochenillifera and O. auberi); “papatla” leaves (Heliconia schiedeana) for zacahuil” wrapping; “revoltillo verde” fried chili sauce (Capsicum annuum) with scrambled eggs; “calabaza borrada” (tender Cucurbita pepo) tamale wrapped with corn leaves; “chivo tapeado,” many chili varieties (Capsicum annum) spices, and black beans (Phaseolus vulgaris); “mamanxas” tender corn patties in “mole de olla” with “chilcuague” (Heliopsis longipes roots) (Espinoza-García and Villaseñor 2017). Typical Xi’oi dishes are “tamales” of “camote” (Ipomoea batatas tubers) and “mala mujer” or “ortiga” flowers (Cnidoscolus multilobus); “bolime,” a different kind of ritual and large “tamal” of “chibeles” (“lechuguilla” or “maguey Agave spp. flowers); “chochas de palma samandoca” (Yucca treculeana flowers); salty “pinole” (ground corn) with chili, consumed during religious celebrations; and “tamales de chamal” (Dioon edule), prepared with dough similar to corn (Miranda-Perkins and Castillo Gómez 2010).

Plants from Natural Vegetation and Agroecosystems The main vegetation types where people gather or manage the native plants are tropical deciduous forest with 308 species, oak forest (301), piedmont scrub (230), thorn scrub (177), and cactus scrub (175). It is worth mentioning that many useful species occur in altered vegetation, such as Ambrosia cordifolia, Cnidoscolus multilobus, Commelina erecta, Eryngium carlinae, Gomphrena serrata, Melothria pendula, Parthenium hysterophorus, and Vachellia farnesiana (Appendix 2). For agroecosystems, the homegardens or “solares” with 238 species, “potreros” or induced grasslands (114) and agroforests (46) were the most diverse (Fig. 5). However, “milpas” are very important for people. A detailed explanation for each agroecosystem is as follows (Figs. 6, 7, and 8). Homegardens or “solares.” In this rich agroecosystem, is possible to find a mix of cultivated or spontaneous useful plants. Important medicinal species include Agastache mexicana, Calea ternifolia, Jacobinia incana, Montanoa leucantha, Parmentiera aculeata, Salvia spp., and Tagetes spp. Also very common are trees for different purposes (Casimiroa edulis, Cnidoscolus acutilobus, Crataegus mexicana, Plumeria rubra, Pithecellobium dulce, Prunus serotina subsp. capuli, Spondias mombin, and S. purpureum) and ornamental plants (Adiantum tenerum, Calochorthus barbatus, Cestrum nocturnum, Mammillaria spp., Maurandya barclayana, Mirabilis jalapa, Sedum spp., Petrea volubilis, Zephyranthes spp.).

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico

279

Fig. 5 Useful species gathered or managed in ecosystems and agroecosystems in the Sierra Gorda

Opuntia spp. for nopales and fruits (prickly pears, “pitayas,” and “pitahayas”) are frequent in the homegardens. Many edible plants and spices in the “solar” are gathered just before cooking, instead of buying them in the store (Amaranthus spp., Capsicum annuum, Cucurbita spp., Dysphania ambrosioides, Erythtina coralloides, Lycopersicum esculentum, Physalis spp., Phaseolus spp., Piper auritum, Portulaca oleracea, Yucca spp.). People grow many species with multiple uses as medicinal, edible, and ornamental purposes (Bursera simaruba, Carica papaya, Clinopodium mexicanum, Costus pulverulentus, Dahlia coccinea, Malacomeles denticulata, Psidium guajava). Induced grasslands have many useful species, such as Argemone ochroleuca, Astrolepis sinuata, Bouteloua scorpioides, Cenchrus echinatus, Datura spp., Glandularia elegans, Leptochloa dubia, Muhlenbergia rigida, Physalis spp., Portulaca oleracea, Setaria grisebachii, Solanum spp., and Talinum lineare, among others. Agroforests have 46 useful species recorded, although the amount seems biased due to the few studies carried out in this agroecosystem. The main cultivated or managed species are fruit trees like Persea americana, Psidium guajava, and Spondias mombin. Some species had different management types, e.g., Cedrela odorata, Tabebuia rosea, Cupressus lusitanica, Lonchocarpus rugosus, and Clethra kenoyeri are used for timber and their harvest is associated with the new moon lunar phase. Bark, branches, leaves, and fruits of Sambucus nigra subsp. canadensis, Guazuma ulmifolia, Hamelia patens, and Cestrum oblongifolium are used for medicinal purposes, either for physical or psychosomatic diseases. Vines such as Hylocereus undatus, Syngonium podophyllum and Melothria pendula produce edible fruits. Herbaceous plants like Dahlia coccinea and Begonia gracilis have ornamental purposes; flowers of Erythrina coralloides, Bauhinia chapulhuacania, and Cercis canadensis are edible and, finally, Protium copal produces resin (“copal”) used in rituals.

280

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Fig. 6 Edible plants. A. Capsicum annuum L. var glabriusculum “chile piquín”. B. Carica papaya L. “papaya”. C. Cnidoscolus aconitifolius (Mill.) I.M. Johnston “chaya mansa”. D. Dahlia coccinea Cav. “dalia”. E. Helianthus annuus L. “girasol”. F. Juglans mollis Engelm. “nogal”. G. Physalis cinerascens (Dunal) Hitchc. “jaltomate”. H. Porophyllum macrocephalum DC. “papaloquelite”. I. Vitis tiliifolia Humb. & Bonpl. ex Schult. “bejuco de uva”. (Photos by the authors)

Milpas harbor 43 species, most of them useful, commonly Amaranthus spp., Capsicum annuum, Cucurbita spp., Dysphania ambrosioides, Helianthus annuus, Jaltomata procumbens, Phaseolus spp. Physalis spp., Solanum lycopersicum, Tithonia diversifolia, and Zea mays. Agroforestry orchards contain 22 species, and the agroecosystem has a range of one-to-many cultivated plants. Among the most common species are Agave spp., Helianthus annuus, Opuntia spp., Stenocereus stellatus, Psidium guajava, and Yucca spp.

Discussion and Projections The current knowledge of useful plant species within the Sierra Gorda seems to be the result of relationships among native plants and human cultures that have lived there. Traditional uses and names are present, as well as others given more recently.

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico

281

Fig. 7 Medicinal plants. A. Cobaea scandens Cav. “quebraplato”. B. Ehretia anacua (Terán & Berland.) I.M. Johnst. “muñeco”. C. Lopezia speciosa Cav. “perilla”. D. Proboscidea louisiana (Mill.) Thell. “toritos”. E. Physalis cinerascens (Dunal) Hitchc. “jaltomate”. F. Psidium guajava L. “guayaba”. G. Salvia microphylla Kunth “irto”. H. Taxus globosa Schltdl. I. Zanthoxylum fagara (L.) Sarg. “limoncillo”. (Photos by the authors)

A detailed analysis on the continuity or change response of these cultural uses might enlighten the history of these useful plants. But the ethnobotanical survey presented here is still incomplete. We believe that it is very important to continue conducting research on the traditional knowledge of useful plants and their associated processes. Given the high rate of environmental alteration, such as change in land use, soil depletion, migration of youth, plant overcollection, contamination, and global warming, it is urgent to document the extraction rates of the most important useful plants and to assess their ecological conservation status. To preserve native plant species and their cultural importance must be a priority for local managers, scholars, and government, and for that goal understanding the socio-ecological systems’ structure and function is crucial. That would also be an opportunity for people of the Sierra Gorda to find an integrated way to preserve their culture, environment, and plants.

282

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Fig. 8 Ornamental plants. A. Adiantum braunii Meet. ex Kuhn. B. Lobelia laxiflora Kunth var. angustifolia A. DC., plant C. flower. D. Marsilea mollis B.L. Rob. & Fernald “trébol de agua”. E. Morkillia mexicana (DC.) Rose & Painter F. Maurandya barclayana Lindl. G. Plumeria rubra L. “jacalosuchil”. H. Pisoniella arborescens (Lag. & Rodr.) Standl. “jazmincillo”. I. Tigridia pavonia (L. f.) DC. “cocomite”. (Photos by the authors) Acknowledgments To Mahinda Martínez, Victor Steinmann, and two anonymous reviewers for the critical review of the English version of the manuscript. Yolanda Pantoja elaborated the map of the Sierra Gorda. This study was partially supported by the GEF Project ID 9380 CONABIO-GEFFAO/RG057 “Agrobiodiversidad de la Región Huasteca, México”.

Appendix See Appendices 1 and 2.

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Family Acanthaceae

Scientific name Anisacanthus pumilus (F. Dietr.) Nees Jacobinia incana (Nee.) Hemsl Justicia aurea Schltdl. Justicia brandegeeana Wassh. & L.B. Sm. Justicia candicans (Nees) L.D. Benson Justicia fulvicoma Schltdl. & Cham. Justicia spicigera Schltdl. Odontonema callistachyum (Schltdl. & Cham.) Kuntze Odontonema cuspidatum (Nees) Kuntze Tetramerium nervosum Nees

Olotillo, olotillo blanco

Guardia del cardenal

Muicle, muitle, chalahuite

Camarón, cresta de gallo

Pavón amarillo

Muicle

Common name



– – – –



– – 1 –



1

1



1

1

1

1









1

















1

1

















– –





















































































































































































– 1

OT –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – – – – – – – – – – – – –



















(continued)

Other –

Appendix 1 Useful plants list. Keys in alphabetical order. AR ¼ Artcrafts; CE ¼ Ceremonial; CO ¼ Construction and timber; ED ¼ Edible; ES ¼ Ecosystem services; FE ¼ Fences, including living fences; FO ¼ Forage; FU ¼ Fuel; HP Honey production; ME ¼ Medicinal; OR ¼ Ornamental; OT ¼ Other (various uses); TO ¼ Tools; TX ¼ Poisonous and toxic plants; TY ¼ Toys

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 283

Amaranthus hybridus L. Amaranthus hypochondriacus L. Chenopodium berlandieri Moq. Dysphania ambrosioides L.

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthus acutilobus Uline & W.L. Bray Amaranthus cruentus L.

Scientific name Sambucus nigra L. ssp. canadensis (L.) Bolli Liquidambar styraciflua L. Alternanthera caracasana Kunth Alternanthera pungens Kunth

Amaranthaceae

Amaranthaceae

Amaranthaceae

Altingiaceae

Family Adoxaceae

Appendix 1 (continued)

Epazote, de comer, cenizo, quelite cenizo, qui’bish cu’sié (Southern Pame)

Amaranto, quelite morado, colorado, rojo Quelite, q. de pollo, quintonil, amaranto Amaranto, quelite morado Quelite cenizo

San Pedro, cascabelillo y sonajita Quelite de pollo

Somerio, liquidámbar Tianguis

1 1 1

– –

1 – – 1

1 – – – – – – – – 1

1

1

1

1









1





1

















1



























































































































































Common name OR MD ED ES FO CO TO HP FU TY FE CE TX AR Sauco, sauco blanco 1 1 – – – – – – – – – – – –

















OT –

















Other –

284 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Iresine cassiniiformis S. Schauer Iresine schaffneri S. Watson Allium glandulosum Link & Otto Sprekelia formosissima (L.) Herb. Zephyranthes concolor (Lindl.) Benth. & Hook. f. Zephyranthes fosteri Traub Zephyranthes verecunda Herb. Rhus aromatica Ait. var. trilobata (Nutt.) A. Gray Rhus virens Lindh. ex A. Gray Spondias mombin L.

Amaranthaceae

Anacardiaceae

Anacardiaceae

Anacardiaceae

Amaryllidaceae

Amaryllidaceae

Amaryllidaceae

Amaryllidaceae

Amaryllidaceae

Amaranthaceae

Amaranthaceae

Dysphania graveolens (Willd.) Mosyakin & Clemants Gomphrena serrata L.

Amaranthaceae

1 –

– –



Jobo, ciruelo

Lantrisco

Agrillo, agrito

1

– 1

– – –

1

– 1

1







1

Mayito, flor de mayo Mayito









1

Mayito, flor de mayo

1

1



1

1

1

Gallitos, pata de gallo

Tabardillo, pie de paloma Cebollín, cebollita

1





– –































1







1





Sanguinaria, amor seco, amor de soltero

1

1

Epazote de zorrillo, – epazote, pazote de zorrillo



























































































































1

































1





























































1





















(continued)

tanning, beberage –





















Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 285

Apocynaceae

Apocynaceae

Apiaceae

Apiaceae

Apiaceae

Apiaceae

Apiaceae

Apiaceae

Anemiaceae

Anemiaceae

Family Anacardiaceae

Scientific name Spondias purpurea L. Anemia adiantifolia (L.) Sw. Anemia phyllitidis (L.) Sw. Arracacia aegopodioides (Kunth) J.M. Coult. & Rose Arracacia tolucensis (Kunth) Hemsl. Eryngium carlinae F. Delaroche Eryngium heterophyllum Engelm. Eryngium nasturtiifolium Juss. ex F. Delaroche Eryngium serratum Cav. Asclepias curassavica L. Asclepias linaria Cav.

Appendix 1 (continued)

1 1 1

1

1 1 1

Apio, hierba del hoz – – – – – – –

Talayote de perro, venenillo, veneno, romerillo, solimán, algodoncillo

Hierba del niño

Hierba del sapo

Hierba del sapo

Hierba del sapo, mosquitos Hierba del sapo

















1



Hierba del oso





1

1



1







































































1



















































































































































OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – 1 – – – – – – – – – – – –

Common name Ciruela, ciruela colorada





















OT –





















Other –

286 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Arecaceae

Araliacaeae

Araliacaeae

Araceae

Apocynaceae

Apocynaceae Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Codo de Fraile

Codo de Fraile

Talayote, chayote manso Hierba de la cucaracha, hierba de San Pedro Matelea pedunculata Borreguitos, (Decne.) Woodson puerquito Matelea pilosa Suspiro (Benth.) Woodson Plumeria rubra L. Jacalosuchil Telosiphonia hypoleuca (Benth.) Henrickson Vallesia glabra Pasita (Cav.) Link Syngonium Huevo de burro podophyllum Schott Aralia regeliana Hierba del oso Marchal Dendropanax Nixtamal arboreus (L.) Decne. & Planchon Acrocomia aculeata Corozo (Jacq.) Lodd. ex Mart.

Cascabela ovata (Cav.) Lippold Cascabela thevetia (L.) Lippold Gonolobus chloranthus Schltdl. Mandevilla foliosa (Müll. Arg.) Hemsl.

– –

1 –

– –

– –

1 – 1 – –

1 1

– 1 – –

1







1

1

1













– –







1

1







– –





1



1





1











– –















1







– –























– –























– –























– –























– –























– –























– –









1













– 1























– –









1













– –























– –













(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 287

Asparagaceae

Agave applanata Lem. ex Jacobi

Scientific name Brahea berlandieri Bartlett Arecaceae Chamaedorea elegans Mart. Arecaceae Chamaedorea microspadix Burret Aristolochiaceae Aristolochia orbicularis Duch. Asparagaceae Agave angustifolia var. rubescens (Salm-Dyck) Gentry Asparagaceae Agave albomarginata Gentry Asparagaceae Agave americana L. var. expansa (Jacobi) Gentry Asparagaceae Agave americana L. var. marginata Trel. Asparagaceae Agave americana L. var. mediopicta Trel. Asparagaceae Agave americana L.

Family Arecaceae

Appendix 1 (continued)

Maguey cenizo

1

1

Maguey, maguey pinto, fino





1

Maguey, maguey fino





1

1

Maguey, maguey pinto, fino







1

Maguey, maguey fino





1

Lechuguilla



1



1



1





1



Hierba del indio



1



1











1

Palmilla







1



1



















1

– –































































































1





















– –





















































OR MD ED ES FO CO TO HP FU TY FE CE TX AR – – – – – – 1 – – – – – – 1

Common name Palma, sh’ijuá (Southern Pame) Palmilla

1

1







1









OT –

beberage, to wrap “barbacoa” fibers







fibers









Other –

288 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Asparagaceae

Asparagaceae Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Agave desmetiana Jacobi Agave filifera Salm.- Maguey, palma, Dyck cucharitas, maguey pencón Agave Lechuguilla heteracantha Zucc. Agave mapisaga Maguey, maguey Trel. var. mapisaga fino, xilote, verde, manso, fino, agave pulquero, Agave mitis Mart. Maguey de peña, g’uruá (Southern Pame) Agave salmiana Otto Maguey, maguey ex Salm.-Dyck ssp. cenizo, moro, verde, crassispina (Trel. ex manso L.H. Bailey) Gentry Agave salmiana Otto Maguey, maguey ex Salm.-Dyck verde, blanco, mexicano, manso, guacamelo, mezote, agave pulquero, Agave striata Zucc. Agave xylonacantha Salm-Dyck Dasylirion Cucharilla, sotol acrotrichum (Schiede) Zucc. 1

1 1

1

– – –

– – –





– –





– –

1

















1





1

1



1

1

1

1

1

1



1



1



1









1



– –

1















– –

















– –

















– –

















– –

















– –

















– –

















– –

















– –

















– –

1

1



1









– –

(continued)

beberage

beberage, to wrap “barbacoa”



beberage







Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 289

Yucca guatemalensis 1 Baker Yucca queretaroensis Estoquillo, lonja, 1 Piña palma, toquillo, flor de yuca, junco, junquillo Yucca treculeana 1 Carrière

Asparagaceae

Asparagaceae

Palma, yuca, yuca del desierto, izote, palma china

1

1

1

1

1

1

1

Asparagaceae

Asparagaceae

Estrella, estrellita, flor de queso

Amole

Cebollitas

Cucharilla, sotol

Varacuete

Nolina robusta L. Hern. Yucca filifera Chabaud.

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

– –

– –



1

1 1

1

– –





– 1

1

1

1



1







1















– –























1



























1



































































































1











































OR MD ED ES FO CO TO HP FU TY FE CE TX AR – – – – – – – – – – – 1 – –

Asparagaceae

Common name Sotol

Scientific name Dasylirion berlandieri S. Watson Dasylirion longissimum Lem. Dasylirion parryanum Trel. Echeandia nana (Baker) Cruden Manfreda scabra (Ortega) McVaugh Milla biflora Cav.

Family Asparagaceae

Appendix 1 (continued)

1





1





1



1



OT –







fibers





soap



beberage



Other –

290 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Aspleniaceae

Aspleniaceae

Aspleniaceae

Aspleniaceae

Aspleniaceae

Aspleniaceae

Asplenium exiguum Bedd. Asplenium monanthes L. Asplenium praemorsum L. Asplenium sessilifolium Desv. Holodyctium ghiesbreghtii (E. Fourn.) Maxon Schaffneria nigripes Feé Acmella repens Rich. ex Pers. Acourtia reticulata (Lag. ex D. Don) Reveal & R.M. King Adenophyllum cancellatum (Cass.) Villarreal Ageratina deltoidea (Jacq.) R.M. King & H. Rob. Ageratina espinosarum (A. Gray) R.M.King & H.Rob.

Malvavisco

Rosita amarilla, berro Cola de zorra, cola de coyote

Lecho de cerro, helecho

– –

– – – –

– – 1 1

– – –

1 – –

1

1

1

1







1























1









1













1

1









1







1











































































































































































































































(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 291

Asteraceae Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Family Asteraceae

Scientific name Ageratina glabrata (Kunth) R.M. King & H. Rob. Ageratina petiolaris (Moc. ex DC.) R.M. King & H. Rob. Ageratum corymbosum Zuccagni Ambrosia cordifolia (A. Gray) W.W. Payne Artemisia ludoviciana Nutt. Aster spinosus Benth. Baccharis heterophylla Kunth Baccharis pteronioides DC. Baccharis salicifolia (Ruiz & Pav.) Pers. Bidens aurea (Aiton) Sherff Bidens odorata Cav. Bidens pilosa L.

Appendix 1 (continued)

Aceitilla blanca Aceitilla

Té de huerto

Jarilla

Jara china, china

Barredora

Junquillo

Istafiate

Vara de cuete

Peshthó

Common name Paloma

1 1

– –

1



1

1





1



1

1



1

1



– –

– –



1

– –



























1



1 –

1











1













































1 –



1















– – –

– –



































– –



















– –



















– –



















– –



















– –



1









1







– 1

OT –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR – 1 – – – – – – – – – – – –

– –



fierworks









fierworks





Other –

292 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Brickellia cavanillesii (Cass.) A. Gray Brickellia veronicifolia (Kunth) A. Gray Calea ternifolia Kunth Calyptocarpus vialis Less. Cosmos bipinnatus Cav. Cosmos diversifolius Otto ex Knowles & Westc. Cosmos sulphureus Cav. Dahlia coccinea Cav. Dyssodia pinnata (Cav.) B.L. Rob. Dyssodia tagetiflora Lag. Eutetras pringlei Greenm. Grindelia inuloides Willd. Gymnosperma glutinosum (Spreng.) Less. – – – –

– 1 1 – 1 1

– 1 –

Escobilla, pegajosa, – zarzal

Árnica

Cinco llagas









Aceitilla, flor de San – Francisco Dalia 1 Hierba del torsón –









– –









1

1 –

1

1 –

Mirasol

Hierba del negro, prodigiosa Tirisia 1





1









1







1





Pextó

Prodigiosa









– –

1





















– –















1







– –











1











– –























– –























– –























– –























– –























– –























– –























– –











1











– –









(continued)

accelerate “pulque” fermenation –



Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 293

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Family Asteraceae

Scientific name Helenium mexicanum Kunth Helianthus annuus L. Helianthus laciniatus A. Gray Heliopsis annua Hemsl. Heliopsis longipes (A. Gray) S.F. Blake Heterotheca inuloides Cass. Melampodium divaricatum DC. Melampodium perfoliatum (Cav.) Kunth Montanoa leucantha (Lag.) S.F. Blake Montanoa tomentosa Cerv. Parthenium hysterophorus L. Parthenium incanum Kunth Pinaropappus roseus (Less.) Less.

Appendix 1 (continued)

Hierba de leche o hierba del gallo

Hierba blanca

Amargoso

Ziguapatle

Tronadora

Tinaja, tinajilla

Rosa amarilla

Árnica

Muela de caballo, rama amarilla Chilcuague

Girasol

Girasol gindo

Common name Cabezona



– – 1 1 – –

1 – – – – –

1







1



– –

1

















1

1















1





1



1

1







1







1





1













1































































































1















































































































OR MD ED ES FO CO TO HP FU TY FE CE TX AR – 1 – – – – – – – – – – – –

















1







OT 1

















spice







Other pesticide

294 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Piqueria trinervia Cav. Pittocaulon praecox (Cav.) H. Rob. & Brettell Porophyllum linaria (Cav.) DC. Porophyllum macrocephalum DC. Porophyllum viridiflorum (Kunth) DC. Pseudognaphalium canescens (DC.) Anderb. Pseudognaphalium roseum (Kunth) Anderb. Roldana angulifolia (DC.) H. Rob. & Brettell Roldana aschenborniana (S. Schauer) H. Rob. & Brettell Roldana ehrenbergiana (Klatt) H. Rob. & Brettell

Hierba del perro

Gordolobo

Guardalobo

Gordolobo

Gordolobo

Tepehua

Pápaloquelite

Hierba del venado

Tabardillo blanco, tabardillo rojo Candelón

1

1

1

1

1

– – – –















1

1



1





1

1







1





























































1



















1



































































































1

















































































(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 295

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae Asteraceae Asteraceae Asteraceae Asteraceae

Asteraceae Asteraceae

Family Asteraceae

Scientific name Sanvitalia procumbens Lam. Senecio salignus DC. Simsia amplexicaulis Pers. Stevia lucida Lag. Stevia serrata Cav. Tagetes erecta L. Tagetes lucida Cav. Tagetes lunulata Ortega Tagetes micrantha Cav. Thymophylla setifolia Lag. Tithonia diversifolia (Hemsl.) A. Gray Viguiera dentata (Cav.) Spreng. Viguiera linearis (Cav.) Sch. Bip. ex Hemsl. Wedelia acapulcensis Kunth var. hispida (Kunth) Strother

Appendix 1 (continued)

1 1 1 1 1 1 1 1 – 1

1

– – 1 1 1 – –

Chayo/chaya macho – – – –

Llora sangre

Romerillo

Parraleña

Anisillo

Cabezona Santa teresita Cempazuchil Pericón Cinco llagas

1











1

– – 1 – –

– –













– – – – –

– –













– – – – –

– 1



1









– – – – –

– –



1









– – – – –

1 –





1







– – – – –

– –













– – – – –

– –













– – – – –

– –













– – – – –

– –



1









– – 1 – 1

– –













– – – – –

– –













– – – – –

– –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 1 – – 1 – – – – – – – – – 1 –

Jara cuetera Shotol delgado

Common name Ojo de pollo











1

– – – – –

– –

OT –











spice

– – – – –

– –

Other –

296 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Bignoniaceae

Bignoniaceae

Bignoniaceae

Bignoniaceae

Betulaceae

Begoniaceae

Basellaceae

Basellaceae

Athyriaceae

Athyriaceae

Asteraceae

Asteraceae

Asteraceae

Alnus jorullensis Kunth Crescentia alata Kunth Parmentiera aculeata (Kunth) Seem. Tabebuia rosea (Bertol.) DC. Tecoma stans (L.) Juss. ex Kunth

Xanthium strumarium Ell. Zaluzania augusta (Lag.) Sch. Bip. Zinnia peruviana (L.) L. Diplazium franconis Liebm. Woodsia mollis (Kaulf.) J. Sm. Anredera ramosa (Moq.) Eliasson Anredera vesicaria (Lam.) C.F. Gaertn. Begonia gracilis Kunth

San pedro

Palo de rosa

Chote

Coco, guaje cirial

Begonia, ala de ángel, cebollita, doncella Aile

Suelda-consuelda

Mata de caracol

Cadillo blanco, de burro shihuite, limpia tunas Mal de ojo – –

– – 1

1 –



1

– – 1

1 –

1

1

1

1

– – –

– – –

1

– 1













1

1









1

1



1



1





1





1

















1









1

















































1



















































1



















































1

























1













1













1

















































1























(continued)

prickly pear cleansing –



Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 297

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Brassicaceae

Boraginaceae

Boraginaceae

Boraginaceae Boraginaceae

Boraginaceae

Boraginaceae

Family Blechnaceae

Scientific name Woodwardia spinulosa Mart. & Gal. Cordia boissieri A. DC. Ehretia anacua (Terán & Berland.) I.M. Johnst. Ehretia tinifolia L. Heliotropium angiospermum Murray Nama origanifolia Kunth Wigandia urens (Ruiz & Pav.) Kunth Lepidium virginicum L. Hechtia glomerata Zucc. Hechtia podantha Mez Tillandsia erubescens Schltdl. Tillandsia ionantha Planch.

Appendix 1 (continued)

Gallitos, magueysitos Gallitos

Guapilla

Guapilla china

Ortiga de tierra caliente Lentejilla

Toronjil Cola de alacrán, alacrancillo

Muñeco, roble prieto, meón

Trompillo

Common name

– –

– 1

– – 1 1

1 – –

1

1

1

1



– – –

1 – –







1





– –

1



1



– –

1



1 –



1

1









1

1





– –



















– –

1

















– –

1

















– –

1

















– –

1

















– –



















– –



















– –



















– –



















– –





OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – – – – – – – – – – – – –















– –





OT –















– –





Other –

298 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Cactaceae

Cactaceae

Cactaceae

Burseraceae

Burseraceae

Burseraceae

Burseraceae

Burseraceae

Burseraceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Tillandsia karwinskyana Schult. & Schult. f. Tillandsia recurvata (L.) L. Tillandsia usneoides (L.) L. Bursera fagaroides (Kunth) Engl. var. fagaroides Bursera galeottiana Engl. Bursera palmeri S. Watson Bursera schlechtendalii Engl. Bursera simaruba (L.) Sarg. Protium copal (Schltdl. & Cham.) Engl. Acanthocereus tetragonus (L.) Hummelinck Ariocarpus kotschoubeyanus (Lem.) K. Schum. Astrophytum ornatum (DC.) Britton & Rose 1

– – – –



1

– –

1

– – – – 1

– – –

1 – –

1

1

1

1









1













1

1







1







1

Pata de venado, 1 peyote, falso peyote, biznaga-maguey 1

Jacube

Oshite

Chaca

Palo cuchara

Xixote verde

Heno

Heno

Gallito









































1



















1



























1



















1

1

1



















1



















1







1



















1

1

1











































1

























1























(continued)

staining









Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 299

Echinocactus horizontalonius Lem.

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Scientific name Corypahntha erecta (Lem) Lem. Coryphantha cornifera (DC.) Lem. Coryphantha octacantha (DC.) Britton & Rose Cylindropuntia imbricata (Haw.) Knuth Cylindropuntia kleiniae (DC.) Knuth Cylindropuntia leptocaulis (DC.) Knuth Disocactus flagelliformis (L.) Barthlott Echinocactus grusonii Hildm.

Family Cactaceae

Appendix 1 (continued)

1

1

1

1

1

1

1

– –

1



1

– – – –



1



1

1













1

























– –









































































































































OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – – – – – – – – – – – – –

Espina de oro, 1 biznaga de bola, biznaga tonel dorada, asiento de suegra Biznaga meloncillo, 1 manca caballo, manca mula, biznaga de dulce

Cardón, cardenche, cholla, abrojo, xoconostle Tasajillo, tasajo, cordoncillo Afarerillo, catalinaria, cholla, tasajo, tasajillo Pitajaya

Biznaga

Chichitas

Common name

















OT –

















Other –

300 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Echinocereus cinerascens (DC.) Lem. Echinocereus pectinatus (Scheidw.) Engelm. ssp. pectinatus Echinocereus pentalophus (DC.) J.N. Haage Echinocereus schmollii (Weing.) N. P. Taylor Ferocactus echidne (DC.) Britton & Rose Ferocactus glaucescens (DC.) Britton & Rose Ferocactus histrix Lindsay Ferocactus latispinus (Haw.) Britton & Rose Isolatocereus dumortieri (Scheidw.) Backeb.

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Echinocactus platyacanthus Link & Otto

Cactaceae

– –

1 1

1 1



– – – – – – –



Organito de víbora, – órgano pequeño; fruto: pitahita Biznaga 1

1

1

Biznaga, biznaga 1 ganchuda, uña de águila; fruto: pochas 1

Huamishe

Biznaga de bola, biznaga

Biznaguilla

1





1















1





1



Biznaga, biznaga de 1 dulce, biznaga gigante, biznaga tonel grande Organito 1









1

1







1





























































































































































































































(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 301

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Family Cactaceae

Scientific name Lophocereuss marginatus(DC.) S. Arias & Terrazas Lophophora diffusa (Croizat) Bravo Mammillaria candida (Scheidw.) Buxbaum Mammillaria compressa DC. Mammillaria crinita DC. ssp. painteri (Rose ex Quehl) W. & B. Fitz Maurice Mammillaria geminispina Haw. Mammillaria hahniana Werderm. Mammillaria herrerae Werderm. Mammillaria longimamma DC. var. longimamma Mammillaria magnimamma Haw.

Appendix 1 (continued)

1

1

1

1

Biznaguita de chilito

1

1

1











1



1

1















1





















































































































































































































– 1

OT –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR – 1 – – – – – – – – – – – –

Biznaga vieja de la 1 sierra de Jalpan Biznaga bola de hilo 1

Biznaga de chilito

Biznaga cabeza de viejo, biznaga de chilito Biznaguita de chilito Biznaguita painteri

Peyote

Common name Órgano, guad’a (Southern Pame)



















Other –

302 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Mammillaria perbella Hildm. ex K. Schum. Mammillaria pettersonii Hildm. Mammillaria polythele Mart. ssp. polythele Mammillaria prolifera (Mill.) Haw. ssp. multiceps (Salm-Dyck) U. Guzmán Mammillaria rhodantha Link & Otto ssp. aureiceps (Lem.) D.R. Hunt Mammillaria sartorii J. A. Purpus

Mammillaria mathildae Kraehenb. & Krainz Mammillaria microhelia Werderm. Mammillaria muehlenpfordtii C.F. Först. Mammillaria parkinsonii Ehrenb.

Biznaguita de chilito

Biznaga dorada

Biznaguita de chilito

Biznaga de chilito

Biznaga de San Onofre, biznaga de chilito Biznaguita de chilito

Biznaga de La Cañada, biznaguita de chilito Biznaguita de chilitos Biznaga

1

1

1

1

1

1

1

1

1

1

1

1 – – – – – –





– – – – – – – –









































































































































































































































































(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 303

Opuntia auberi Pfeiff.

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Scientific name Mammillaria schiedeana Ehrenb. ssp. schiedeana Mammillaria sempervivi DC. Mammillaria uncinata Zucc. ex Pfeiff. Mammillaria wiesingeri Boed. Mammillaria zephyranthoides Scheidw. Myrtillocactus geometrizans (Mart. ex Pfeiff.) Console Neobuxbaumia polylopha (DC.) Backeb. Neolloydia conoidea (DC.) Britton & Rose ssp. conoidea Opuntia albicarpa Scheinvar

Family Cactaceae

Appendix 1 (continued)



Biznaga de flor 1 occidental, biznaguita de chilito Garambuyo 1

1

Nopal de – Alfajayucan, blanca de Alfajayucan, reina Nopal de lengüita, 1 lengua de vaca 1





1

Biznaguita

1

1



1



Órgano de piedra

1

1

1



Biznaga



1

1

– 1

1

1

















– –









– –

















































1

















1

































































































OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – – – – – – – – – – – – –

1

Biznaguita de chilito Biznaga de chilito, uña de gato

Common name Biznaga de Metztitlán



















OT –



















Other –

304 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Nopal

Tapona

Lengua de vaca, nopal de lengüita

Nopal tapona, tuna tapona

Nopal aguamielo, amarillo, blanco, chinito, hartón, mancaño, memelo, negrito, redondo, zarco Opuntia icterica Joconostle, D. A. Griffiths xoconostle Opuntia incarnadilla Tasajillo D.A. Griffiths Opuntia joconostle Joconostle, F.A.C. Weber ex xoconostle Diguet Opuntia lasiacantha Tasajillo Pfeiff. Opuntia leucotricha Duraznilo DC.

Opuntia cochenillifera (L.) Mill Opuntia cantabrigiensis Lynch Opuntia engelmannii Salm-Dyck ex Engelm. Opuntia ficus-indica (L.) Mill. Opuntia hyptiacantha F.A.C. Weber

1

– – 1

– –

– – – – 1

1

1



1

1

1

1

1





1





1

1



1





1









1













1





– –

1





























































































1

1































































































































(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 305

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Opuntia microdasys (Lehm.) Pfeiff. Opuntia oligacantha Xoconostle C. F. Först corriente, xoconostle manzano, xoconostle de burro Opuntia pachona Nopal hartón, nopal D. A. Griffiths barroso Opuntia pubescens Perrilla J.C. Wendl. ex Pfeiff. Opuntia robusta Nopal camueso, J.C. Wendl. hartón Opuntia Nopal cardón streptacantha Lem. Opuntia tomentosa Nopal Salm-Dyck Opuntia zamudioi Xoconostle, frutos: Scheinvar duraznillo Lophocereus Órgano marginatus (DC.) Britton & Rose

Cactaceae

1

1

1







1 –

1



1

1

1

1

1



1

1



1









1

1





1

1









1



1















































































1







1







1























































Common name OR MD ED ES FO CO TO HP FU TY FE CE TX AR Xoconostle – – 1 – – – – – – – – – – – cuaresmeño, xoconostle colorado con espinas, joconostle 1 – – – – – – – – – – – – –

Scientific name Opuntia matudae Scheinvar

Family Cactaceae

Appendix 1 (continued)





1













shade













– –

Other –

OT –

306 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Stenocereus huastecorum Alvarado-Sizzo, Arreola-Nava & Terrazas Stenocactus lamellosus (A. Dietr.) A. Berger ex A.W. Hill Stenocactus Tepenexcomitl obvallatus (DC.) A. Berger Stenocactus Biznaga, biznaguita ochoterenaus Tiegel

Cactaceae

Cactaceae

Cactaceae

Flower: reina de la noche, dama de la noche; fruit: pitahaya

Dama de la noche

Órgano, cabeza de viejo, pitayo viejo, pitayón

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Alfilerillo, cola del diablo

Pereskiopsis diguetii (F. A. C. Weber) Britton y Rose Pilosocereus leucocephalus (Poselg.) Byles & G. D. Rowley Rhipsalis baccifera (Mill.) Stearn Selenicereus spinulosus (DC.) Britton & Rose Selenicereus undatus (Haw.) D.R. Hunt

Cactaceae



– 1

1



– –



– –

1





– –

1





1

1

1

1

1

1





1











































































































































































































































(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 307

Caprifoliaceae

Caprifoliaceae

Cannabaceae

Cannabaceae

Campanulaceae

Calceolariaceae

Cactaceae

Cactaceae

Cactaceae

Family Cactaceae

Lonicera mexicana (Kunth) Rehder Lonicera pilosa (Kunth) Spreng.

Scientific name Stenocereus queretaroensis (F. A. C. Weber) Buxb. Thelocactus hastifer (Werderm. & Boed.) F.M. Knuth Thelocactus leucacanthus (Zucc.) Britton & Rose ssp. leucacanthus Turbinicarpus pseudomacrochele Backeb. ssp. pseudomacrochele Calceolaria mexicana Benth. Lobelia laxiflora Kunth var. angustifolia A. DC. Celtis caudata Planch. Celtis pallida Torr.

Appendix 1 (continued)

– –



– –

1 1 – –





1 – – – – –

1

– 1



1

1



1





1

1















1

















1

















1















1

















1

1

































1

1

































































OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – 1 – – – – – – – – – – –

Granjeno, cimarrón, 1 blanco, granjenillo 1

Zorro

Botita amarilla

Turbinita de Querétaro

Biznaga-pezón hastada

Common name Órgano, pitayo de Querétaro





1

1











OT –





plant tutor

plant tutor











Other –

308 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Symphoricarpos microphyllus Kunth Caricaceae Carica papaya L. Caryophyllaceae Paronychia mexicana Hemsl. Caryophyllaceae Silene laciniata Cav. Cistaceae Helianthemum glomeratum (Lag.) Lag. Clethraceae Clethra kenoyeri Lundell Clethraceae Clethra mexicana DC. Commelinaceae Commelina coelestis Willd. Commelinaceae Commelina erecta L. Commelinaceae Commelina tuberosa L. Commelinaceae Tradescantia crassifolia Cav. Convolvulaceae Dichondra argentea Humb. & Bonpl. ex Willd. Convolvulaceae Ipomoea capillacea (Kunth) G. Don Convolvulaceae Ipomoea cardiophylla A. Gray Convolvulaceae Ipomoea carnea Jacq.

Caprifoliaceae



1 – –

– 1 1 1 1 1

– – –

1 – – – – 1

– 1

Hierba del pollo

Ojito de pollo

Maguey de cerro

Oreja de ratón

Piñitas

1





1

Escoplo

Quiebraplatos

– –

– 1

1 –

Nanajuana

1











– –

















– –







– –

– –

– – 1 –







1 –



1 –

1 1

1 –



Papaya Chinilla

1



Perlilla











– –







– –

– –













– –







– –

– –

1











1 –







– –

– –













– –







– –

– –













– –







– –

– –













– –







– –

– –













– –







– –

– –













– –







– –

– –













– –







– –

– –













– –







– –

– –













– –







– –

– –



(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 309

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Convolvulaceae Convolvulaceae

Convolvulaceae

Convolvulaceae

Convolvulaceae

Convolvulaceae

Convolvulaceae

Convolvulaceae

Family Convolvulaceae

Scientific name Ipomoea hederifolia L. Ipomoea indica (Burm.) Merr. Ipomoea jalapa (L.) Pursh Ipomoea murucoides Roem. & Schult. Ipomoea nil (L.) Roth Ipomoea orizabensis (G. Pelletan) Ledeb. ex Steud. var. orizabensis Ipomoea purpurea (L.) Roth Ipomoea stans Cav. Merremia dissecta (Jacq.) Hallier f. Echeveria mucronata Schltdl. Echeveria secunda Booth. ex Lindl. Echeveria semivestita Moran Echeveria subrigida (B.L. Rob. & Seaton) Rose

Appendix 1 (continued)

Conchita, florecitas 1







1



1



Oreja de burro



1





– –

Conchita, florecita



1 –







– –







Espanta vaqueros Pata de gallo

1



1









– –















1

1

1

1

1

1









– 1

1



















– –





















– –





















– –







1













– –







1













– –





















– –





















– –





















– –





















– –













OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – – – – – – – – – – – – –

1

Quiebra plato

Palo bobo, cazahuate

Common name









– –



1









OT –









– –



staining









Other –

310 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Cucurbitaceae

Cucurbitaceae

Cucurbitaceae

Cucurbitaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Echeveria tolimanensis Matuda Graptopetalum Dedito de Dios pachyphyllum Rose Pachyphytum Dedo de Dios, compactum Rose dedito de Dios, dedos Pachyphytum viride Dedos E. Walther Sedum bourgaei Chisme Hemsl. Sedum dendroideum Siempreviva DC. Sedum greggii Chisme Hemsl. Sedum moranense Chisme Kunth Sedum praealtum A. DC. Apodanthera Calabacilla undulata A. Gray hedionda Cucurbita Calabaza pipiana argyrosperma K. Koch Cucurbita ficifolia Chilacayote Bouché Cucurbita Calabaza hedionda foetidissima Kunth lisa – –



– – – –

– 1 1 – 1 – 1 – – 1

1 – –

1 – – – –

1

1

1

1

1

1







1







1

















1







1





















































1













































































1

































































































































1

























(continued)

cosmetic

























Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 311

Scientific name Cucurbita moschata Duchesne Cucurbitaceae Cucurbita pepo L. Cucurbitaceae Melothria pendula L. Cucurbitaceae Sechium edule (Jacq.) Sw. ssp. sylvestre Lira & Castrejón Cucurbitaceae Sicyos deppei G. Don Cupressaceae Cupressus lusitanica Mill. Cupressaceae Juniperus deppeana Steud. Cupressaceae Juniperus flaccida Schltdl. Crysopteridaceae Cystopteris fragilis (L.) Bernh. Dennstaediaceae Dennstaedtia distenta (Kunze) T. Moore Dennstaediaceae Dennstaedtia globulifera (Poir.) Hieron. Dicksoniaceae Lophosoria quadripinnata (J.F. Gmel.) C. Chr.

Family Cucurbitaceae

Appendix 1 (continued)

– – – – –

1 – – –

– 1

1

1

Helecho de soros





























1

















































– –













– –













– –













– –













– –













– –



1

– –

– – –

1



– –

– – –

Táscate



– –

1 – –



– –

– – –



1 –

– – –



– –

– – –

Táscate

– –

– – –

– –

– – –

– 1

– – –

Chayotillo

– – –

– – –

– – –

1 – 1

– – –

Calabaza Sandiita, meloncito Chayote pelón

1 1 –

OT –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR – – 1 – – – – – – – – – – –

Common name Calabaza













– –

– – –

Other –

312 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Ericaceae

Ericaceae

Equisetaceae

Equisetaceae

Dryopteridaceae

Dryopteridaceae

Dryopteridaceae

Dryopteridaceae

Dryopteridaceae

Dryopteris pseudofilix-mas (Fée) Rothm. Dryopteris wallichiana (Spreng.) Hyl. Elaphoglossum petiolatum (Sw.) Urb Phanerophlebia nobilis (Schltdl. & Cham.) C. Presl. Polystichum distans E. Fourn. Equisetum hyemale L. var. affine (Engelm.) A.A. Eaton Equisetum myriochaetum Schltdl. & Cham. Arbutus xalapensis Kunth Arctostaphylos pungens Kunth Acalypha adenostachya Cav. Acalypha infesta Poepp. Acalypha monostachya Cav.

Nanajuana

Hierba del cáncer

Hierba del cáncer

Pingüica

Madroño

Cola de caballo

Cola de caballo

– – – –

– – – – 1

1



1 –

– – 1 1 1 1

1 – – –

1

1

1







1

1









1

1





1



























1

– –

1







































1













































1

1























































































































































































(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 313

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Family Euphorbiaceae

Scientific name Bernardia mexicana (Hook. & Arn.) Müll. Arg. Cnidoscolus multilobus (Pax) I.M. Johnst. Cnidoscolus aconitifolius (Mill.) I.M. Johnst. Croton ciliatoglandulifer Ortega Croton cortesianus Kunth Croton morifolius Willd. Euphorbia anychioides Boiss. Euphorbia bracteata Jacq. Euphorbia dentata Michx. Euphorbia graminea Jacq. Euphorbia prostrata Aiton Euphorbia radians Benth.

Appendix 1 (continued)



1 1 1 – 1 – 1

– – – 1 – – – 1

Celedonia

Sangrinaria

Lechocilla

Periquito

Golondrina roja

Pusgual

Hierba del moro

1



1



Picosa, solimán

1



1

1







1

1

1



Mala mujer, ortiga

Chaya mansa



– –



1

1

























1



1















































































































































































1



























OR MD ED ES FO CO TO HP FU TY FE CE TX AR – – – 1 – – – – 1 – – – – –

1

Common name Oreja de ratón























OT –























Other –

314 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbia tanquahuete Sessé & Moc. Jatropha dioica Sessé Manihot angustiloba (Torr.) Müll. Arg. Acaciella angustissima (Mill.) Britton & Rose Albizia occidentalis Brandegee Albizia plurijuga (Standl.) Britton & Rose Bauhinia chapulhuacania Wunderlin Bauhinia coulteri J.F. Macbr. var. coulteri Bauhinia divaricata L. Bauhinia ramosissima Benth. ex Hemsl. Caesalpinia pulcherrima (L.) Sw. Calliandra eriophylla Benth. – – – – – – – – –

1

– – – – – – – –



1

1

1

1





1

Tabachín de monte 1

1

Pata de vaca, pata de 1 cabra 1

Pata de vaca

Palo blanco

Palo blanco

Timbre

1



1













1





1

1

1



Lantrisco, palo – lechón, palo amarillo, borreguillo Sangregado 1

1















































1



















































































1













































































1















































1

1





















(continued)

industry

cosmetic



Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 315

Dalea lutea (Cav.) Willd.

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae Fabaceae

Scientific name Calliandra grandiflora (L’Hér.) Benth. Cercis canadensis L. Chamaecrista nictitans (L.) Moench var. jaliscensis (Greenm.) H.S. Irwin & Barneby Cojoba arborea (L.) Britton & Rose Crotalaria longirostrata Hook. & Arn. Dalbergia paloescrito Rzed. & Guridi-Gómez Dalea bicolor Humb. & Bonpl. ex Willd. Dalea foliosa (A. Gray) Barneby

Family Fabaceae

Appendix 1 (continued)



– 1



1 – –

1 –



– –



1 –

1







– –





– –

1 –

– 1



1

1







– –







1





– –

1











– –













– –













– –













– –













– –













– –













– –













– –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – – – – – – – – – – – – –

Escobilla, 1 engordacabras Limoncillo, – almaraduz, hierba de la hormiga, mezquitillo, motita, sonajilla Escoba 1

Palo escrito

Garbancillo

Cojoba, acacia roja

Cuaresma Tiricia

Common name







1





– –

OT –







violins





– –

Other –

316 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Desmodium grahamii A. Gray Desmodium molliculum (Kunth) DC. Enterolobium cyclocarpum (Jacq.) Griseb. Erythrina coralloides DC. Eysenhardtia polystachya (Ortega) Sarg. Harpalyce arborescens A. Gray Havardia pallens (Benth.) Britton & Rose Indigofera suffruticosa Mill. Leucaena esculenta (Moc. & Sessé ex DC.) Benth. Leucaena leucocephala (Lam.) de Wit Leucaena pulverulenta (Schltdl.) Benth.

Efez, tepeguaje

Efez, teucaena

Efez, leucaena, guaje

Añil

Gatillo amarillo

Chicharrillo

Palo dulce

Colorín

Orejón, parota, guanacaste

Pegarropa

Pegarropa

– – –

1

1 – –

1

– –

1



1





1





1

1

1

1

1

1





1









1















1

1







1











1

1



1

1

























1

1

– 1

























1

1



















1

































– –

1







1





















1











































1



















1





















(continued)

tan, staining









Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 317

Macroptilium gibbosifolium (Ortega) A. Delgado Mimosa depauperata Benth. Mimosa leucaenoides Benth. Mimosa texana (A. Gray) Small Parkinsonia aculeata L. Phaseolus coccineus L. Phaseolus heterophyllus Humb. & Bonpl. ex Willd. Phaseolus vulgaris L. Pithecellobium dulce (Roxb.) Benth.

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Scientific name Lonchocarpus rugosus Benth. Lysiloma acapulcense (Kunth) Benth. Lysiloma microphyllum Benth.

Family Fabaceae

Appendix 1 (continued)

– –

– – – – – –

1

Gatillo, uña de gato – –

1 – – – 1

Frijol

Guamuchil

Frijol chimate, ayocote

Palo verde

Uña de gato

Quebramachete 1

1







– –







1

1



1



1



1

1

1



1





1

1

Tepeguaje, palo de arco, quibracho, palo fierro Jicamita/Jícama

Tepeguaje











1



1

1



1













1





1











1





















1





















1



1

1



1































1

1







































































OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – – – – – – – – – – – – – –

Common name Chicharrillo corral



1













1



OT –



other













shade



Other –

318 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Vachellia pennatula (Schltdl. & Cham.) Seigler & Ebinger

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Prosopis laevigata (Humb. & Bonpl. ex Willd.) M.C. Johnst. Rhynchosia prostrata Brandegee Senna atomaria (L.) Irwin et Barneby Senna polyantha (Moc. & Sessé ex Collad.) H.S. Irwin & Barneby Senna septemtrionalis (Viv.) H.S. Irwin & Barneby Senna wislizeni (A. Gray) H.S. Irwin & Barneby var. painteri (Britton) H.S. Irwin & Barneby Vachellia farnesiana (L.) Willd. y Arn.

Fabaceae

Huizache

Huizache

Chaparro prieto

1

1

1

1

1

Palo fierro

Café

1

1

1

Palo hediondo

Cacahuatill

Mezquite

1

1





1

1







1







– 1







1

1



1

1

1

1

1



1

1



1

1

1





1

1





1



1





1





1

1

1





1





1

1

1



1

1





1









1







1

1



1

1









































1











1



1



1







1

(continued)

industry, cosmetic, staining aromatic –



reforest







other

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 319

Gentianaceae

Gentianaceae

Garryaceae

Fagaceae Fourqueriaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fabaceae

Family Fabaceae

Scientific name Vachellia schaffneri (S. Watson) Seigler & Ebinger Zornia diphylla (L.) Pers. Quercus crassipes Bonpl. Quercus crassifolia Humb & Bonpl Quercus greggii (A. DC.) Trel. Quercus grisea Liebm. Quercus potosina Trel. Quercus rugosa Née Fouquieria splendens Engelm. Garrya laurifolia Hartw. ex Benth. ssp. laurifolia Eustoma exaltatum (L.) Salisb. ex G. Don Gentiana spathacea Kunth

Appendix 1 (continued)





1

1

Juana mipila, flor de – hielo







1







1 –

– –

1 –

Roble, encino Chiquiñá, ocotillo

1 1



Encino blanco







– –











– –











1 –



1









– –













– –



1











– –







1







– –

















– –

















– –

















– –

















– –











1

1





















1



















1















1











1











1





OT .

OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 1 – 1 1 . 1 . 1 1 1 . . 1



Encino

Encino

Encino rojo

Encino palo blanco

Cascabelillo

Common name Huizache







– –













Other .

320 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Lamiaceae

Lamiaceae

Lamiaceae

Lamiaceae Lamiaceae

Lamiaceae

Lamiaceae

Koeberliniaceae

Juglandaceae

Iridaceae

Iridaceae

Hydrangeaceae

Geraniaceae

Geranium seemannii Peyr. Philadelphus coulteri S. Watson Tigridia pavonia (L. f.) DC. Tigridia vanhouttei Roezl ex Van Houtte Juglans mollis Engelm. Koeberlinia spinosa Zucc. var. spinosa Agastache mexicana (Kunth) Lint & Epling Clinopodium mexicanum (Benth.) Govaerts Hyptis albida Kunth Lepechinia caulescens (Ortega) Epling Salvia amarissima Ortega Salvia coccinea Buc’hoz ex Etl. Salvia elegans Vahl

Mirto

Mirto

Mirto

Salvia Bahthá o salve

Poleo

Nogal silvestre, de caballo Junco, corona de cristo Toronjil blanco

Carita de víbora

Cocomite, carcoma

Shampoo o hierba de la rozadura Jazmín blanco

1







1 1

– 1 1

– –

1 – –



1

– –

1

1







– –







1





1

1











1













1



1





1









– –























– –























– –























– –























– –























– –























– –























– –























– –























– –























– –

1





















– –

(continued)

aphrodisiac















Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 321

Lauraceae

Lauraceae

Lamiaceae Lamiaceae Lamiaceae

Lamiaceae Lamiaceae

Lamiaceae Lamiaceae

Lamiaceae

Lamiaceae Lamiaceae Lamiaceae

Family Lamiaceae

Scientific name Salvia helianthemifolia Benth. Salvia hispanica L. Salvia keerlii Benth. Salvia leucantha Cav. Salvia melissodora Lag. Salvia mexicana L. Salvia microphylla Kunth Salvia patens Cav. Salvia polystachia Cav. Salvia regla Cav. Salvia tiliifolia Vahl Teucrium cubense Jacq. ssp. chamaedrifolium (Mill.) Epling Listea glaucesens Kunth Persea americana Mill.

Appendix 1 (continued)

Aguacate

Laurel

Rama apestosa Epazotillo

Mirto Saponaria

Mirto

Cordón de San Francisco

Chía

Common name Manto de la virgen

– – – – –

– – – 1 1

1

1 – –

– 1

1 1

1

– –

– –

1 –

1







1

– –



– – –

– –

– –



– – –



– – –

– –

– –



– – –





– – –

– –

– –



– – –





– – –

– –

– –



– – –





– – –

– –

– –



– – –





– – –

– –

– –



– – –





– 1 –

– –

– –



– – –





– – –

– –

– –



– – –





– – –

– –

– –



– – –





– – –

– –

– –



– – –





– – –

– –

– –



– – –



1

– – –

– 1

– –



– – –

– – –

– 1 1 1 – 1

OT –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR – 1 – – – – – – – – – – – –



spice

– – –

– soap

– –



– – –

Other –

322 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Lentibulariaceae Pinguicula agnata Casper Lentibulariaceae Pinguicula calderoniae Zamudio Lentibulariaceae Pinguicula elizabethiae Zamudio Lentibulariaceae Pinguicula esseriana B. Kirchn. Lentibulariaceae Pinguicula lilacina Schltdl. & Cham. Lentibulariaceae Pinguicula macrophylla Kunth Lentibulariaceae Pinguicula moctezumae Zamudio & R.Z. Ortega Lentibulariaceae Pinguicula moranensis Kunth var. moranensis Liliaceae Calochortus barbatus (Kunth) J.H. Painter Loasaceae Mentzelia hispida Willd. Loranthaceae Psittacanthus calyculatus (DC.) G. Don 1





Injerto





1









1

1





1







1

Pega ropa





1







1

1





1

Café

Perritos, pichueca





1

























































































1





































































































































































































(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 323

Malvaceae

Malvaceae

Malvaceae

Malpighiaceae

Malpighiaceae

Magnoliaceae

Magnoliaceae

Lythraceae

Lythraceae

Lythraceae

Lygodiaceae

Family Lycopodiaceae

Scientific name Lycopodium clavatum L. Lygodium venustum Sw. Cuphea aequipetala Cav. Cuphea lanceolata W.T. Aiton Heimia salicifolia Link Magnolia grandiflora L. Magnolia schiedeana Schltdl. Byrsonima crassifolia (L.) Kunth Gaudichaudia cynanchoides Kunth Allowissadula holosericea (Scheele) D.M. Bates Anoda cristata (L.) Schltdl. Ayenia jaliscana S. Watson

Appendix 1 (continued)

Violeta, malva violeta Hierba del cáncer

Hierba del zorro

Nanche

Magnolia

Escoba de río

Hierba de la cuerdilla Hierba del pollo

Common name

– – – – –

– –

– 1 – – – – –

1 –

1 –

1 – – 1

1

1





1



1



1

1



1

1

1



















1























































1



























































































































1







1





























OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – – – – – – – – – – – – –























OT –























Other –

324 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Sida rhombifolia L. Sphaeralcea angustifolia (Cav.) G. Don

Malvaceae Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Ceiba aesculifolia (Kunth) Britten & Baker f. Ceiba pentandra (L.) Gaertn. Gossypium hirsutum L. Guazuma ulmifolia Lam. Kearnemalvastrum lacteum (Ait.) D.M. Bates Malvaviscus arboreus Cav. Pavonia candida (DC.) Fryxell Phymosia umbellata (Cav.) Kearney Pseudobombax ellipticum (Kunth) Dugand var. ellipticum Sida abutifolia Mill.

Malvaceae

– – 1

– – –

1

Mocoque (from the 1 Teenek: mokok)

Hierba del pastor, la – pastora Malvarisco – Hierba del negro 1

1

1

1 1





Tulipán capullo, malvarisco Chilitos

1



1

– – 1

– –



















Malvón

Aquiche

1













1



1

1

1

1

1

Algodón

Ceiba

Caibilla

1 –



















1

– –



















1

– –





















– –



















1

– –



















1

– –



















1

– –





















– –





















– –





















– –



















1

– –















1



1

– –















(continued)

fibers



fibers

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 325

1



1

Pata de cabra

Cedro, cuatal, cuaternari, nogal, nogal corriente, nogalillo, nogalillo cimarrón Cedro Cucharillo 1 – 1

1 1 –

1

– –

– –

– –



1









– –



– –





– –



1 –

1



– –





– –





– –





– –

1



– –





– –

1



– –





– –

1



– –



– –



– –





– –



– –



– –





– –



– –



– –





– –



– –



– –

1



– –



– –



– –





– –



– –

– 1



– –



1 –

Torito, toritos, vaquitas



– –





– –





– –



1

– –



Trébol de agua, trébol

– –



– –



1 –



– 1



Tapacola







1



Cadillo negro



OT –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR – 1 – – – – – – – – – – – –

Common name Tila

Cedrela odorata L. Trichilia havanensis Jacq. Menispermaceae Cissampelos pareira Hierba del peso L.

Meliaceae Meliaceae

Scientific name Tilia americana Schltdl. var. mexicana Malvaceae Triumphetta semitriloba Jacq. Malvaceae Waltheria indica L. Marattiaceae Marattia weinmanniifolia Liebm. Marsileaceae Marsilea mollis B.L. Rob. & Fernald Martyniaceae Martynia annua L. Martyniaceae Proboscidea louisiana (Mill.) Thell. Melastomataceae Tibouchina longifolia (Vahl) Baill. Meliaceae Cedrela dugesii S. Watson

Family Malvaceae

Appendix 1 (continued)



– –





– –



– –



Other –

326 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Menispermaceae Cocculus diversifolius DC. Menispermaceae Menispermum canadense L. Moraceae Ficus cotinifolia Kunth Moraceae Morus celtidifolia Kunth Myrtaceae Myrcianthes fragrans (Sw.) McVaugh Myrtaceae Psidium guajava L. Nyctaginaceae Mirabilis glabrifolia (Ortega) I.M. Johnst. Nyctaginaceae Mirabilis jalapa L. var. jalapa Nyctaginaceae Mirabilis longiflora L. Nyctaginaceae Pisonia aculeata L. Nyctaginaceae Pisoniella arborescens (Lag. & Rodr.) Standl. Nymphaeaceae Nymphaea ampla (Salisb.) DC. Oleaceae Forestiera phillyreoides (Benth.) Torr. Oleaceae Fraxinus rufescens Lingelsh. – – 1

– 1 – 1

1 – 1 –

1 1 1

Higuerón

Mora

Guayabilla

Guayabo

Maravilla

– –

– – –

1 – –

1 1

1

1



Acebuche, azibuche, panalero

Lantrisco



1







Maravilla, maravilla 1 blanca Furica – Jazmincillo 1



1







1



1



– –





– –

– –

1 –

1 –

























1



Hierba del ojo



1



– –





– –











1

1



– –





– –













1



– –





– –













1



– –





– –











1

1



– –





– –















– –



– –





– –













– –





– –

















– –





– –

















– –





– –

















– –





– –

















– –





– –











(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 327

Orobanchaceae

Orobanchaceae

Orobanchaceae

Orchidaceae

Orchidaceae

Onagraceae

Onagraceae

Onagraceae

Onagraceae

-Onagraceae

Family Oleaceae

Scientific name Fraxinus uhdei (Wenz.) Lingelsh. Gaura coccinea Nutt. ex Pursh. Lopezia racemosa Cav. Oenothera deserticola (Loes.) Munz Oenothera kunthiana (Spach) Munz Oenothera rosea L’Hér. ex Ait. Dichromanthus cinnabarinus (La Llave & Lex.) Garay Mesadenus polyanthus (Rchb. f.) Schltr. Castilleja arvensis Schltdl. & Cham. Castilleja tenuiflora Benth. Lamourouxia dasyantha (Cham. & Schltdl.) W.R. Ernst

Appendix 1 (continued)





1

1

San francisquito, – machetona de milpa Machetona, – garañona 1 –

– –

1 –































































































































































































1





















1



















1









Hierba del golpe, cólica rosa Hierba del golpe







1







Cólica





1





Perilla







1



Hierba del golpe

OT –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 1 – – – 1 1 1 – 1 1 – – –

Common name Fresno





















Other –

328 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Passiflora foetida L. var. lanuginosa Killip Passiflora Bejuco de granada serratifolia L. Passiflora subpeltata Ortega Turnera diffusa Hierba del pastor Willd.

Passifloraceae

Passifloraceae

Passifloraceae

Bolsa de gato, pasiflora y té deinsomnio. Granadilla

Mano de león

Chicalote

Coyol

Agritos

Chivivi Jaboncillo

Passifloraceae

Passifloraceae

Passifloraceae

Papaveraceae

Papaveraceae

Papaveraceae

Oxalidaceae

Oxalidaceae

Oxalidaceae Oxalidaceae

Lamourouxia multifida Kunth Oxalis corniculata L. Oxalis decaphylla Kunth Oxalis latifolia Kunth Oxalis divergens Benth. ex Lindl. Argemone ochroleuca Sweet Bocconia frutescens L. Hunnemannia fumariifolia Sweet Passiflora bryonioides Kunth Passiflora exsudans Zucc.

Orobanchaceae

– –

1 –

– 1 – – 1 1

1

– – 1

– –

1 – – – – 1 –

1









1





1

1

– v

1 1

– –





1

















1

























1



1































































































































































































































































1

























(continued)

pesticide









Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 329

Family Scientific name Pentaphylacaceae Ternstroemia sylvatica Schltdl. & Cham. Petiveriaceae Rivina humilis L. Phyllantaceae Phyllanthus compressus Kunth Phyllantaceae Phyllanthus grandifolius L. Phytolaccaceae Agdestis clematidea Moc. & Sessé ex DC. Phytolaccaceae Phytolacca icosandra L. Phytolaccaceae Phytolacca rivinoides Kunth & C.D. Bouché Pinaceae Abies religiosa (Kunth) Schltdl. & Cham. Pinaceae Pinus ayacahuite C. Ehrenb. ex Schltdl. Pinaceae Pinus cembroides Zucc. Pinaceae Pinus devoniana Lindl. Pinaceae Pinus greggii Engelm. ex Parl.

Appendix 1 (continued)



– –

1 – –

– – – –

1

1

1 1 1

Oyamel, guayame

Piñas

Piñonero, piñón

Pino michoacano

Pino

1





1

Venenosa o cóngora –



1



Jabonera





1





















– –

1 1

– –



















– –

1

1

1

1

1









– –



















– –



















– –



















– –

















1

– –



















– –



















– –



















– –



















– –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR – 1 – – – – – – – – – – – –

1 –

Bajatripa Hierba de la golondrina Sonaja, sonajillla

Common name Tila, trompillo



















– –

OT –



















– –

Other –

330 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Plantaginaceae

Plantaginaceae

Plantaginaceae

Plantaginaceae

Plantaginaceae

Piperaceae Plantaginaceae

Piperaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinus montezumae Lamb. Pinus patula Schltdl. & Cham Pinus pinceana Gordon & Glend. Pinus pseudostrobus Lindl. var. pseudostrobus Pinus teocote Schltdl. & Cham. Peperomia hispidula (Sw.) A. Dietr. Piper auritum Kunth Bacopa monnieri (L.) Wettst. Maurandya antirrhiniflora Humb. & Bonpl. ex Willd. Maurandya barclayana Lindl. Penstemon barbatus (Cav.) Roth Penstemon hartwegii Benth. Penstemon hidalgensis Straw

Hoja santa, acoyo Verdolaguilla, baraima

C’aní uala

Piñas

Pino

Piñonero

Piñas

– –

1 – – – 1 – –

– – – –

– 1

– 1 – 1 –

– – – –

1 –

1

1

1

1

1

1 1

1

1











– 1











1







1











– –























– –



1

1

1

1

1











– –





1

















– –























– –























– –























– –























– –























– –























– –























– –























– –













(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 331

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae Poaceae

Plumbaginaceae

Platanaceae

Family Plantaginaceae

– –

– – – – – – – –

– – – – – –

Chondrosum – hirsutum (Lag.) Sweet Eragrostis mexicana Zacate llorón, amor, – (Hornem.) Link bayal, de agua, cola de ardilla, pelo de conejo

Pata de gallo mexicano, paragüitas Banderita

Cadillo

Pasto cortalón

Pasto pluma

Tres barbas









– –



1



Tianguiz

Alamillo



1



1



1

1

– 1





1

1

1

1

1

1

1 1

















– –

















– –

















– –

















– –

















– –

















– –

















– –

















– –

















– –





OR MD ED ES FO CO TO HP FU TY FE CE TX AR – 1 – – – – – – – – – – – – –

Common name Pastito 1

Scientific name Plantago nivea Kunth Platanus mexicana Moric. Plumbago pulchella Boiss. Aristida laxa Cav. Bouteloua curtipendula (Michx.) Torr. Bouteloua scorpioides Lag. Bromus carinatus Hook. & Arn. Cenchrus echinatus Cav. Chloris submutica Kunth

Appendix 1 (continued)













– –





OT –













– –





Other –

332 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Polemoniaceae

Polemoniaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

– – – – – –

– – – – – –

1

1

1





Zacate gigante, – desparramado dubiano Muhlenbergia rigida Grama, liendrilla – (Kunth) Kunth morada Paspalum notatum Zacate grama dulce – Flüggé Paspalum Zacate huilotero – prostratum Scribn. & Merr. Pennisetum crinitum Pasto popotón – (Kunth) Spreng. Setaria grisebachii Pegarropa – E. Fourn. Urochloa fasciculata – (Sw.) R. Webster Zea mays L. ssp. Maíz, maíz dulce, – mays cabellos de elote Cobaea scandens Quebraplato – Cav. Loeselia coerulea Guachichile – (Cav.) G. Don – –

– –





1



1

1

1

1

1

1

1



1

1



1

1



1

1



1





1



Hopia obtusa (Kunth) Zuloaga & Morrone Leptochloa dubia (Kunth) Nees

Poaceae





Rizado, toboso menudo, zacate galleta, chino, mezquite Grama



Hilaria belangeri (Steud.) Nash

Poaceae











1

















1











1











1







































































































































































1



















– (continued)

beberage / industry –



















Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 333

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polygonaceae

Polygonaceae

Polygonaceae

Family Polemoniaceae

Scientific name Loeselia mexicana (Lam.) Brand Persicaria hydropiperoides (Michx.) Small Persicaria segetum (Kunth) Small Rumex mexicanus Meisn. Campyloneurum angustifolium (Sw.) Fée Phlebodium areolatum (Humb. & Bonpl. ex Willd.) J. Sm. Pleopeltis macrocarpa (Bory ex Willd.) Kaulf. Pleopeltis polylepis (Roem. ex Kunze) T. Moore Polypodium subpetiolatum Hook. Polypodium thyssanolepis A. Braun ex. Klotzsch

Appendix 1 (continued)

Helecho

Helecho

Lengua de ciervo

Lengua de ciervo

Canahuala

Lengua de vaca

Sanguinaria

Common name Espinosilla





1

1

1





1

1

1

1





– – –

– – –

1

– 1



















1







































































1



































































































































– 1

OT 1

OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 1 – – – – – – – – – – – –



















Other cosmetic

334 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Proteaceae

Portulacaceae

Portulaca oleracea Verdolaga L. Roupala montana Aubl. Adiantum andicola Liebm. Adiantum braunii Meet. ex Kuhn Adiantum capillusveneris L. Adiantum concinnum Humb. & Bonpl. ex Willd. Adiantum poiretii Wikstr. Adiantum tenerum Sw. Adiantum trapeziforme L. Adiantum tricholepis Fée Aleuritopteris farinosa (Forssk.) Fée Argyrochosoma incana (C. Presl) Windham Astrolepis sinuata Helecho (Lag. ex Sw.) 1

1

1

1

1

1

1

1

1

1

1

1

– 1 – – – – – – – – – – – –

1 – – – – – – – – – – – –



















































































































































































































































































































































(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 335

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Family

Common name

D.M. Benham & Windham Cheilanthes angustifolia Kunth Cheilanthes Doradilla bonariensis (Willd.) Proctor Cheilanthes candida M. Mart. & Gal. Cheilanthes kaulfussii Kunze Cheilanthes lendigera (Cav.) Sw. Cheilanthes leucopoda Link Cheilanthes meifolia DC. Eaton Cheilanthes Helecho myriophylla Desv. Cheiloplecton rigidum (Sw.) Fée var. rigidum Hemionitis palmata L. Llavea cordifolia Lag.

Scientific name

Appendix 1 (continued)

– – – – – – – – –

1

– – – – – – – – –



1

1

1

1

1

1

1

1

1





1





















































































































































































































































OR MD ED ES FO CO TO HP FU TY FE CE TX AR























OT























Other

336 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Rhamnaceae

Rhamnaceae

Rhamnaceae

Rhamnaceae

Rhamnaceae

Ranunculaceae Rhamnaceae

Pteridaceae Pteridaceae Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pellaea cordifolia (Sessé & Moc.) A.R. Sm. Pellaea ovata (Desv.) Weath. Pellaea ternifolia (Cav.) Link Pityrogramma calomelanos (L.) Link var. calomelanos Pteris cretica L. Pteris longifolia L. Pteris quadriaurita Retz. Clematis dioica L. Ceanothus caeruleus Lag. Colubrina elliptica (Sw.) Brizicky & W.L. Stern Condalia mexicana Schltdl. Condalia velutina I.M. Johnst. Frangula capreifolia (Schltdl.) Grubov Gouania polygama (Jacq.) Urb.

Bejuco espumoso

Palo de zorra

Granjeno, granjeno prieto Granjeno

Palo mole

Barba de chivo

Perlilla

– – –

– – –

– – 1

– – – 1 1 1

1

1 –

1 1 1 – 1 – – 1 – 1

1 – –

1





1

1



– 1





1



1

1

1



– –

– – –





1

1







1

– –

1



– –

– – –











1

– –

– – –















1



– 1

– – –



















– –

– – –













1





– –

– – –















1



– –

– – –



















– –

– – –



















– –

– – –



















– –

– – –



















– –

– – –



















– –

– – –



















– –

– – –









(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 337

Rubiaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rhamnaceae

Family Rhamnaceae

Scientific name Karwinskia humboldtiana (Schult.) Zucc. Karwinskia mollis Schltdl. Crataegus gracilior Phipps Crataegus mexicana DC. Crataegus rosei Eggl. Duchesnea indica (Andrews) Teschem. Fragaria vesca L. ssp. bracteata (Heller) Staudt Lindleya mespiloides Kunth Malacomeles denticulata (Kunth) Decne. Prunus serotina Ehrh, ssp. capuli (Cav.) McVaugh Purshia mexicana (D. Don) S.L. Welsh Bouvardia longiflora (Cav.) Kunth

Appendix 1 (continued)

Cedrillo romerillo, cedro cimarrón Huele de noche

Capulín

Membrillo

Fresa

Fresa de monte

Manzanillo

Tejocote

Tejocote

Zarabullo

1 –

– – – –

1

– –

– – 1

– – – – 1

– 1 –

1 1

1

1

1

1

1





1

















1





1

1









1











































1







1

























































1















































































1







1















Common name OR MD ED ES FO CO TO HP FU TY FE CE TX AR Tullidora, sarabullo 1 1 1 1 – – 1 – – – – – 1 –























OT –























Other –

338 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Sapindaceae

Sapindaceae Sapindaceae

Salicaceae

Salicaceae

Salicaceae

Rutaceae

Rutaceae

Rutaceae

Rutaceae

Rubiaceae Rutaceae

Rubiaceae

Rubiaceae

– – –

1 – 1 1

1





1







1

– –







1

– –

1

Palo hediondo, palo – zorrillo Zanthoxylum fagara Limoncillo – (L.) Sarg. Neopringlea Palillo, vidrioso – integrifolia (Hemsl.) S. Watson Populus tremuloides Alamillo 1 Michx. Salix bonplandiana Sauce 1 Kunth Acer negundo L. Arce 1 Cardiospermum Bolsas de coyote 1 halicacabum L. Dodonaea Ocotillo 1 viscosa Jacq.







1



Chongua



– 1

– 1

1











1



1



1 1

Madura plátano Zapote, zapote blanco Ciruelillo

Trompetilla roja

1 1

Bouvardia ternifolia (Cav.) Schltdl. Cigarrilla mexicana (Zucc. & Mart. ex DC.) Aiello Hamelia patens Jacq. Casimiroa edulis La Llave Casimiroa greggii (S.Watson) F. Chiang Decatropis bicolor (Zucc.) Radlk. Ptelea trifoliata L.

– –















– –





1

– –





1









– –





1

– –





1



1





– 1





– –









1





– –





1

– –





1









– –





– 1















– 1





– –















– –





1

– –















– –





– –















– –





1

– –













– –





1

– –















– –





(continued)

pesticide

– –















– –





Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 339

Solanaceae

Selaginellaceae

Selaginellaceae

Selaginellaceae

Scrophulariaceae Selaginellaceae

Scrophulariaceae

Scrophulariaceae

Scrophulariaceae

Sapotaceae

Sapotaceae

Family Sapindaceae

Scientific name Sapindus saponaria L. Sideroxylon palmeri (Rose) T.D. Penn. Sideroxylon salicifolium (L.) Lam. Buddleja cordata Kunth Buddleja scordioides Kunth Buddleja sessiliflora Kunth Capraria biflora L. Selaginella harrisii Underw. & Hieron. Selaginella lepidophylla (Hook, & Grev.) Spring Selaginella sellowii Hieron Selaginella stenophylla A. Braun Capsicum annum L. var. glabriusculum (Dunal) Heiser & Pickersgill

Appendix 1 (continued)

Chile piquín

Musgo





1







1







1

1





1

Doradilla, flor de peña

– –

1

– – –





1



1

– –

– 1

Jarilla

1

– 1

1

1



1









– –





1











– –

















1 –



















– –



















– –



















– –



















– –





1













– –



















– –



















– –



















– –





1















1





OT –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR 1 – – – – – – – – – – – – –

1

Lengua de vaca

Hierba del perro

Tepozán

Capulín

Capulín prieto

Common name Jaboncillo









– –



to wrap sweets –





Other –

340 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Solanaceae

Solanaceae

Solanaceae

Solanaceae Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Capsicum pubescens L. Capsicum rhomboideum (Kunth) Kuntze Cestrum dumetorum Schltdl. Cestrum nitidum Mart. & Gal. Cestrum nocturnum L. Cestrum oblongifolium Schltdl. Cestrum thyrsoideum Kunth Datura ceratocaula Ortega Datura inoxia Mill. Datura stramonium L. Jaltomata procumbens (Cav.) J.L. Gentry Lycianthes moziniana (Dunal) Bitter Nicotiana trigonophylla Dunal 1

1 1 1 1 1



1

– –

Toloache, tolbache, – atlinán, tornaloca Toloache, tolbache – Toloache – – – –

San juana

Jaltomate

Hierba de la mula

1

– –





– –

– – 1

1

















1

Huele de noche, dama de noche Hierba del chupil







1

1



Horca judas

Chile miracielo





1





1 –

1

– 1

Chile manzano







– –























– –























– –























– –























– –



















1



– –























– –























– 1

1





















1 –

1





















– –























– –























– –

















(continued)

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 341

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae Solanaceae

Family Solanaceae

Scientific name Physalis ampla Waterf. Physalis angulata L. Physalis cinerascens (Dunal) Hitchc. Physalis coztomatl Dunal Physalis gracilis Miers Physalis nicandroides Schltdl. Physalis patula Mill. Physalis philadelphica Lam. Physalis pubescens L. Physalis solanacea (Schltdl.) Axelius Physalis sordida Fernald Physalis virginiana Mill. Solandra maxima (Sessé & Moc) P.S. Green.

Appendix 1 (continued)

1 –

– – – –

– – –



1

1





















– –











– –











– –







1

1



1 1

1











– –















– –















– –

















– –

















– –

















– –

















– –









– –







– –



– 1







– –

– 1







– –

– –







– –







– –







– –

1



1 1





– –

1

– –



– –

– –

– –

– 1

– – 1 1

OT –

OR MD ED ES FO CO TO HP FU TY FE CE TX AR – – – – – – – 1 – – – – – –

Copa de oro, 1 trompeta, trompetero gigante, solandra

Tomatillo

Tomatillo

Tomate verde

Tomate borracho

Zalcomate Costomate, jaltomate Costomate

Common name











– –







– –

Other –

342 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Solanum americanum Mill. Solanaceae Solanum dulcamaroides Dunal Solanaceae Solanum elaeagnifolium Cav. Solanaceae Solanum lycoporsicum L. Solanaceae Solanum nigrescens Mart. & Gal. Solanaceae Solanum rostratum Dunal Styracaceae Styrax glabrescens Benth. var. glabrescens Talinaceae Talinum lineare Kunth Taxaceae Taxus globosa Schltdl. Tectariaceae Tectaria heracleifolia (Willd.) Underw Thelypteridaceae Thelypteris pilosa Mart. & Gal.) Crawford Thelypteridaceae Thelypteris pilosula (Klotzsch &

Solanaceae

– – – – – – –

1 – 1 1 – – – – – –

– – – – 1

1 1 1

1

1

Doña Juana, nana Juana Coyol

Hierba mora

Abrojo

Sabinillo





1

1

1



1

1















1























1

1





































































1

































































































































1









































1























(continued)

cosmetic







Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 343

Scientific name

Common name

H. Karst. ex Mett.) R.M. Tryon Thelypteridaceae Thelypteris puberula (Baker) C.V. Morton var. puberula Tropaeloaceae Tropaeolum majus L. Mastuerzo ó mantuerzo Urticaceae Pilea microphylla Rusby Urticaceae Urtica Ortiga chiquita chamaedryoides Pursh Verbenaceae Aloysia gratissima Huele de noche, (Gillies & Hook.) santanilla, vara Tronc. dulce, varaduz Verbenaceae Bouchea prismatica Verbena (L.) Kuntze Verbenaceae Citharexylum Pasilla berlandieri B.L. Rob. Verbenaceae Citharexylum caudatum L.

Family

Appendix 1 (continued)



1



1 – 1

1

1 – –

– 1 – – – –

1











1









1































































1









































































































OR MD ED ES FO CO TO HP FU TY FE CE TX AR

















OT

















Other

344 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Petrea volubilis L. Verbena carolina L. Verbena litoralis Kunth Verbena menthifolia Verbena Benth. Cissus tiliacea Kunth Tripa de vaca Cissus verticillata (L.) Nicolson & C.E. Jarvis

Vitaceae Vitaceae

Verbenaceae

Hierba de la hormiga, hierbabuenilla, hierba dulce Raspa sombrero Verbena Verbena

Hierba del aire Orégano

Moradilla

Verbenaceae Verbenaceae Verbenaceae

Verbenaceae

Verbenaceae Verbenaceae

Verbenaceae

Glandularia bipinnatifida (Nutt.) Nutt. Glandularia elegans (Kunth) Umber Lantana camara L. Lippia graveolens Kunth Lippia queretarensis Kunth

Verbenaceae

Duranta, corona de novia

Duranta erecta L.

Verbenaceae

1 1 1 1 1 –

– – 1

1



1 – –

1 1

1 1

– –

– –



– – –

– – – –

1



1 –

– –

1



– –





1









1

– –



– – –



1 –







– –



– – –



– –







– –



– – –



– –







– –



– – –



– –







– –



– – –



1 –







– –



– – –



– –







– –



– – –



– –







– –



– – –



– –







– –



– – –



1 –







– –



– – –



– –







– –



– – –



– 1







– –



– – –



(continued)

– spice







Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 345

Zygophyllaceae

Zygophyllaceae

Zamiaceae Zamiaceae Zingiberaceae

Zamiaceae

Zamiaceae

Family Vitaceae

Scientific name Vitis tiliifolia Humb. & Bonpl. ex Schult. Ceratozamia hildae G.P. Landry & M.C. Wilson Ceratozamia microstrobila Vovides & J.D. Rees Dioon edule Lindl. Zamia fischeri Miq. Costus pulverulentus C. Presl Larrea tridentata (DC.) Coville Morkillia mexicana (DC.) Rose & Painter

Appendix 1 (continued)

Gobernadora

Caña de puerco

Chamal

Common name Bejuco de uva





1



1



1 – –





1 – 1





1 1 –

1

1







– – –

– – – –













– – –









– – –









– – –









– – –









– – –









– – –









– – –









1 – –









– – –





OR MD ED ES FO CO TO HP FU TY FE CE TX AR – 1 – – – – – – – – – – – –





– – –





OT –





– – –





Other –

346 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Adoxaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Acanthaceae

Family Acanthaceae

Scientific name Anisacanthus pumilus (F. Dietr.) Nees Jacobinia incana (Nee.) Hemsl Justicia aurea Schltdl. Justicia brandegeeana Wassh. & L.B. Sm. Justicia candicans (Nees) L.D. Benson Justicia fulvicoma Schltdl. & Cham. Justicia spicigera Schltdl. Odontonema callistachyum (Schltdl. & Cham.) Kuntze Odontonema cuspidatum (Nees) Kuntze Tetramerium nervosum Nees Sambucus nigra L. ssp. canadensis (L.) Bolli – –





1

1









1





1

1

1

1







1









– –



























1

1

– –

1



– –



1

1























1









1













































































































1

1







1



















1















































































1













































1



1



1







1

1





















1



1



1

1



1













































M AF – –

(continued)

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 – – – – – – – – – – – – – – – – – – – 1 – –

Appendix 2 Useful species per vegetation type and agrosystem type. Keys in alphabetical order. Natural vegetation: AV ¼ Aquatic vegetation; CF ¼ Cloud forest; CS ¼ Cactus scrub; CyF ¼ Cypress forest; DS ¼ Desert scrub; G ¼ Grassland; JF ¼ Juniper forest; RF ¼ Riparian forest; OF ¼ Oak forest; PF ¼ Pine forest; PPF ¼ Pinyon pine forest; POF ¼ Pine-Oak forest; PS ¼ Piedmont scrub; RS ¼ Rosette scrub; SS ¼ Ssclerophyll scrub (chaparral); SV ¼ Secondary (altered) vegetation; TS ¼ Thorn scrub; TDF ¼ Tropical deciduous forest; TsDF ¼ Tropical subdeciduous forest. Agrosystems: AF ¼ Agroforests (humanized forests); AFH ¼ Agroforestry orchards; HG ¼ Home garden (“solar”); IG ¼ Induced grassland (“potrero”); M ¼ Milpa

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 347

Amaryllidaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Family Altingiaceae

Scientific name Liquidambar styraciflua L. Alternanthera caracasana Kunth Alternanthera pungens Kunth Amaranthus acutilobus Uline & W.L. Bray Amaranthus cruentus L. Amaranthus hybridus L. Amaranthus hypochondriacus L. Chenopodium berlandieri Moq. Dysphania ambrosioides L. Dysphania graveolens (Willd.) Mosyakin & Clemants Gomphrena serrata L. Iresine cassiniiformis S. Schauer Iresine schaffneri S. Watson Allium glandulosum Link & Otto

Appendix 2 (continued)

– –



1

1

– 1



1

















1

1































































































































































































































1







1

1



1













1

1

1





1

1

1



























1





1



1

1

1

1

1







1



1

































































1













1

























1































1



































1



1























1























1



















1

























1















M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH – – – – – – 1 – – – – – – – – – – – – – – – –

348 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Apiaceae

Apiaceae

Apiaceae

Apiaceae

Apiaceae

Anemiaceae

Anemiaceae

Anacardiaceae Anacardiaceae

Anacardiaceae

Anacardiaceae

Amaryllidaceae

Amaryllidaceae

Amaryllidaceae

Amaryllidaceae

Sprekelia formosissima (L.) Herb. Zephyranthes concolor (Lindl.) Benth. & Hook. f. Zephyranthes fosteri Traub Zephyranthes verecunda Herb. Rhus aromatica Ait. var. trilobata (Nutt.) A. Gray Rhus virens Lindh. ex A. Gray Spondias mombin L. Spondias purpurea L. Anemia adiantifolia (L.) Sw. Anemia phyllitidis (L.) Sw. Arracacia aegopodioides (Kunth) J.M. Coult. & Rose Arracacia tolucensis (Kunth) Hemsl. Eryngium carlinae F. Delaroche Eryngium heterophyllum Engelm. Eryngium nasturtiifolium Juss. ex F. Delaroche – –

– –



– – 1

– – – –

– –

1



1



1

1



1







1



1

1 1

1









– –



1





1











– 1

1

1

1















– –







1





1

1

1



1

1

1







1

1



1

1

– –























1

1

– –



















1







– –











1





1







1

– –

















1





1



– –

















1

1







– –



























– –

1

























– –













1













1 1



















1







– –

1

























– –



















1







– –







1



















– –













1













– –



























– –



























1 1





1

1

1

1















– –



























1 –



1























1 1













(continued)















– –













Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 349

Araliaceae

Araliaceae

Araceae

Apocynaceae

Apocynaceae Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Family Apiaceae

Scientific name Eryngium serratum Cav. Asclepias curassavica L. Asclepias linaria Cav. Cascabela ovata (Cav.) Lippold Cascabela thevetia (L.) Lippold Gonolobus chloranthus Schltdl. Mandevilla foliosa (Müll. Arg.) Hemsl. Matelea pedunculata (Decne.) Woodson Matelea pilosa (Benth.) Woodson Plumeria rubra L. Telosiphonia hypoleuca (Benth.) Henrickson Vallesia glabra (Cav.) Link Syngonium podophyllum Schott Aralia regeliana Marchal Dendropanax arboreus (L.) Decne. & Planchon

Appendix 2 (continued)

– –

– – – – – – – – – –











1

– –

1





1

1 1



1



1

1

1

1

1

1

1

– –

1 –





– –











1

1









1









– –













1



1







– –

















1







– –

























– –



















1





– –













1











– –

























– –

























– –

















1







– –























1

– –

























– –

























– –



















1





– –

























– –

























– –

























– –

























1 1

1





1



1



1









– –

















1



1



– –

















Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH – 1 – 1 – – – – – 1 – – – – – – – – – – – – –









– –







1







1









– –

















M AF – –

350 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Acrocomia aculeata (Jacq.) Lodd. ex Mart. Arecaceae Brahea berlandieri Bartlett Arecaceae Chamaedorea elegans Mart. Arecaceae Chamaedorea microspadix Burret Aristolochiaceae Aristolochia orbicularis Duch. Asparagaceae Agave angustifolia Haw. var. rubescens (Salm-Dyck) Gentry Asparagaceae Agave albomarginata Gentry Asparagaceae Agave americana L. var. expansa (Jacobi) Gentry Asparagaceae Agave americana L. var. marginata Trel. Asparagaceae Agave americana L. var. mediopicta Trel. Asparagaceae Agave americana L. Asparagaceae Agave applanata Lem. ex Jacobi Asparagaceae Agave desmetiana Jacobi Asparagaceae Agave filifera Salm.Dyck Asparagaceae Agave heteracantha Zucc.

Arecaceae

– –

1

1













1 –











1











– –











1

– –













1

1















– 1



























– –



























1 1









1

















– –











1

1

1





1

1

1

1 1







1

1

















1 –



























– –













1













1 –



























– –



























– –















1



1







– –













1













– –



























– –



























– –



























– –



























– –



























– –























1

1

1 –

1

1

1



1









1







– –



















1







– –













1

1











1 –



1

1



1









(continued)







– –



















Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 351

Asparagaceae

-Asparagaceae Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae Asparagaceae

Asparagaceae

Asparagaceae Asparagaceae

Family Asparagaceae

Scientific name Agave mapisaga Trel. var. mapisaga Agave mitis Mart. Agave salmiana Otto ex Salm-Dyck ssp. crassispina (Trel. ex L.H. Bailey) Gentry Agave salmiana Otto ex Salm-Dyck Agave striata Zucc. Agave xylonacantha Salm-Dyck Dasylirion acrotrichum (Schiede) Zucc. Dasylirion berlandieri S. Watson Dasylirion longissimum Lem. Dasylirion parryanum Trel. Echeandia nana (Baker) Cruden Manfreda scabra (Ortega) McVaugh Milla biflora Cav. Nolina robusta L. Hern. Yucca filifera Chabaud.

Appendix 2 (continued)







1 – –

– – –

– –



1

1 1

1



1 1



1



– –







1 –



1





– –









– –











– –











– –













– 1













– –















– –















– –













– –



– –













– –

1

– –











1

– –

1



– –













– –





– –













– –





– –













– –





– –





1







– –



1

– –













– –

1

– –



1 –

1











– –



– –

1 –

1











– –



– –

– –





1







– –



1 1

1











– –



– –









1



– –



– –

1









– –



– –





1

1

1

– –



– –

1





1

– –



– –









– –



– –







1 1



– –







– –



– –





– –

1

– –





– –



– –

1

– –

1

– 1



– –



– 1

– –



1 –

– –



– –



– –



– –

– –

– –

– –

– –

– –

M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH – – – – – – – – – – – – – – – – – – – – 1 – –

352 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Aspleniaceae

Aspleniaceae

Aspleniaceae

Aspleniaceae

Aspleniaceae

Asparagaceae

Asparagaceae

Asparagaceae

Yucca guatemalensis Baker Yucca queretaroensis Piña Yucca treculeana Carrière Asplenium exiguum Bedd. Asplenium praemorsum L. Asplenium sessilifolium Desv. Holodyctium ghiesbreghtii (E. Fourn.) Maxon Schaffneria nigripes Feé Acmella repens Rich. ex Pers. Acourtia reticulata (Lag. ex D. Don) Reveal & R.M. King Adenophyllum cancellatum (Cass.) Villarreal Ageratina deltoidea (Jacq.) R.M. King & H. Rob. Ageratina espinosarum (A. Gray) R.M.King & H.Rob. Ageratina glabrata (Kunth) R.M. King & H. Rob. – 1





1



1



1









1

1









1

1

1









1





– –













1





1



1

1



1

1



1 –

















1

1

























1





1































1



1

1











1

1















1





1



































1

























1































































































































































































































1



1



















































1















1



1





1

1



1









































































(continued)





























Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 353

Asteraceae

Asteraceae

Asteraceae

Asteraceae Asteraceae Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae Asteraceae

Asteraceae

Asteraceae

Asteraceae

Family Asteraceae

Scientific name Ageratina petiolaris (Moc. ex DC.) R.M. King & H. Rob. Ageratum corymbosum Zuccagni Ambrosia cordifolia (A. Gray) W.W. Payne Artemisia ludoviciana Nutt. Aster spinosus Benth. Baccharis heterophylla Kunth Baccharis pteronioides DC. Baccharis salicifolia (Ruiz & Pav.) Pers. Bidens aurea (Aiton) Sherff Bidens odorata Cav. Bidens pilosa L. Brickellia cavanillesii (Cass.) A. Gray Brickellia veronicifolia (Kunth) A. Gray Calea ternifolia Kunth Calyptocarpus vialis Less.

Appendix 2 (continued)

1





– – 1

– 1 1







1





1

1





1

1

– – 1





1

1 1

– 1

1 –

1

– 1

1

1

1

1

1

1



1



– – 1







1 –

1

1





1



– – –







– –

1









– 1



– – –





1





1

– – –





1

– –

1

– – –











1 1 –







– –

1

1







1

– – –







– –

1



1



1



– – –







– –













– – –





1

– –

1











– – –







– –

1











– – –







– –













– – –



1



– –





1



1



– – –







– –













– – –







– –













– – –







– –













– – –







– –





1







– – –

1





– –













– – –







– –







1

1



1 – –

1





– –





1

1





1 1 –

1





– –













– – –







– –







Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 1 1 – – – – – – 1 – – – 1 – – – – – – – – –

1

1



1 – –

1





– –













– – –







– –







M AF – –

354 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae Asteraceae

Asteraceae

Asteraceae

Asteraceae

Cosmos bipinnatus Cav. Cosmos diversifolius Otto ex Knowles & Westc. Cosmos sulphureus Cav. Dahlia coccinea Cav. Dyssodia pinnata (Cav.) B.L. Rob. Dyssodia tagetiflora Lag. Eutetras pringlei Greenm. Grindelia inuloides Willd. Gymnosperma glutinosum (Spreng.) Less. Helenium mexicanum Kunth Helianthus annuus L. Helianthus laciniatus A. Gray Heliopsis annua Hemsl. Heliopsis longipes (A. Gray) S.F. Blake Heterotheca inuloides Cass. Melampodium divaricatum DC. Melampodium perfoliatum (Cav.) Kunth 1

– – 1 – – – 1

1





– –



1



1





1



– –

1



1



1



1





1





1

– 1 1

1







1 1

1



1 –

1













– –







1











– –







1



1 1

– 1 –

















– –











1 –







1

1 –



1 –



1

– 1

1

1

1



1

1 –

1



1











– –



1







1 1









1







– –











1 –

1

1

1

1

1

1





– –



1







1 1

1

1

1



1





1

– –

1



1





1 1

1



1











– –











– –

















– –











– –

















– –

1









– –

















– –



1







1 –

















– –





1





– –



1













– –











– –

















– –











1 –





1





1





– –

1









– –

















– –











– –

















1 1











1 –







1

1

1



1

1 1

1

1

1





1 –

1



1











– –











1 –

















1 –











– –







(continued)











1 –











– –





1

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 355

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Family Asteraceae

Scientific name Montanoa leucantha (Lag.) S.F. Blake Montanoa tomentosa Cerv. Parthenium hysterophorus L. Parthenium incanum Kunth Pinaropappus roseus (Less.) Less. Piqueria trinervia Cav. Pittocaulon praecox (Cav.) H. Rob. & Brettell Porophyllum linaria (Cav.) DC. Porophyllum macrocephalum DC. Porophyllum viridiflorum (Kunth) DC. Pseudognaphalium canescens (DC.) Anderb. Pseudognaphalium roseum (Kunth) Anderb. Roldana angulifolia (DC.) H. Rob. & Brettell

Appendix 2 (continued)

1

1 1 1







1







1

1



1





1







1











1

1









1

1

1





1







1-

1







1





1

1



1











1









1







1



1

1





























1

1



1















1



























1























1

1



















1













1

































































1





































































1











1

























































1











1











1











1

1

























Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 1 1 1 1 1 – – 1 1 – – – – – – – – – – 1 1 –









1











1



























M AF – –

356 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae Asteraceae Asteraceae Asteraceae Asteraceae

Asteraceae Asteraceae

Asteraceae

Asteraceae

Asteraceae

Roldana aschenborniana (S. Schauer) H. Rob. & Brettell Roldana ehrenbergiana (Klatt) H. Rob. & Brettell Sanvitalia procumbens Lam. Senecio salignus DC. Simsia amplexicaulis Pers. Stevia lucida Lag. Stevia serrata Cav. Tagetes erecta L. Tagetes lucida Cav. Tagetes lunulata Ortega Tagetes micrantha Cav. Thymophylla setifolia Lag. Tithonia diversifolia (Hemsl.) A. Gray Viguiera dentata (Cav.) Spreng. Viguiera linearis (Cav.) Sch. Bip. ex Hemsl. Wedelia acapulcensis Kunth var. hispida (Kunth) Strother Xanthium strumarium Ell. – 1 –

– –

– 1 – 1 –







– –

– – – 1 1

1



1

1

1



1







1

– – – – –







1

1

1

– –





1





1





1

1

1



1



– – – – –

– –









1

1

1







– – – – –

– –







1



1

1

1



1

– – – 1 1

– 1

1







– –

1



1

– –



1



– – – – –

– –













1 – – – –

– –











1

1







1 – – – –

– –



















1

– – – 1 1

– –

















1



– – – – –

– –









1











– – – – –

– –





















– – – – –

– –





















– – – – –

1 –





















– – – – –

– –





















– – – – –

– –





















– – – – –

– –





















– – – – –

– –







1













– – – – –

– –





















– – – – –

– –















1





– – 1 1 1

– –







1





1

1





– – – – –

– 1

1



















– – – – –

– –





1















– – – – –

– –







(continued)









1





– – 1 – –

– –







Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 357

Boraginaceae

-Blechnaceae

Bignoniaceae

Bignoniaceae

Bignoniaceae

Bignoniaceae

Betulaceae

Begoniaceae

Basellaceae

Basellaceae

Athyriaceae

Athyriaceae

Asteraceae

Family Asteraceae

Scientific name Zaluzania augusta (Lag.) Sch. Bip. Zinnia peruviana (L.) L. Diplazium franconis Liebm. Woodsia mollis (Kaulf.) J. Sm. Anredera ramosa (Moq.) Eliasson Anredera vesicaria (Lam.) C.F. Gaertn. Begonia gracilis Kunth Alnus jorullensis Kunth Crescentia alata Kunth Parmentiera aculeata (Kunth) Seem. Tabebuia rosea (Bertol.) DC. Tecoma stans (L.) Juss. ex Kunth Woodwardia spinulosa Mart. & Gal. Cordia boissieri A. DC.

Appendix 2 (continued)

– –





1

1

1





1



1

1

1





1

1



1







1













1















1

















1

– –





















1





1

1









1









1







1

1



















1



































1





























1















1





















1









1



















































































































1



1

1

1







1

















1























1

1





1





























































1













1

1





1

















1









1

1











1





1













1



1













1



1







M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH – – – 1 1 – – – – – – – – – – – – – – – – – –

358 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Burseraceae

Burseraceae

Burseraceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Brassicaceae

Boraginaceae

Boraginaceae

Boraginaceae Boraginaceae

Boraginaceae

Ehretia anacua (Terán & Berland.) I.M. Johnst. Ehretia tinifolia L. Heliotropium angiospermum Murray Nama origanifolia Kunth Wigandia urens (Ruiz & Pav.) Kunth Lepidium virginicum L. Hechtia glomerata Zucc. Hechtia podantha Mez Tillandsia erubescens Schltdl. Tillandsia ionantha Planch. Tillandsia karwinskyana Schult. & Schult. f. Tillandsia recurvata (L.) L. Tillandsia usneoides (L.) L. Bursera fagaroides (Kunth) Engl. var. fagaroides Bursera galeottiana Engl. Bursera palmeri S. Watson 1 – –







1

1

1

1

1



1

1

1

1



1

1

1





1



1





1











1

1

1









1

1



– –



1

1







– –





1











1







– –

– –

1 –

1







1

1

1 1

1











1

1



– –

















– –













1

1



– 1











1







1









– –





























– –









1







1











– –





























– –











1







1

1







– –





























– 1





























– –











1

















– –





























– –





























– –





























– –





























– –





























– –









1







1





1

1

1

1 –

1























1



– 1





























1 –

1



























– –



(continued)





















1





– –



Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 359

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Burseraceae

Burseraceae

Family Burseraceae

Scientific name Bursera schlechtendalii Engl. Bursera simaruba (L.) Sarg. Protium copal (Schltdl. & Cham.) Engl. Acanthocereus tetragonus (L.) Hummelinck Ariocarpus kotschoubeyanus (Lem.) K. Schum. Astrophytum ornatum (DC.) Britton & Rose Corypahntha erecta (Lem) Lem. Coryphantha cornifera (DC.) Lem. Coryphantha octacantha (DC.) Britton & Rose Cylindropuntia imbricata (Haw.) Knuth Cylindropuntia kleiniae (DC.) Knuth Cylindropuntia leptocaulis (DC.) Knuth Disocactus flagelliformis (L.) Barthlott

Appendix 2 (continued)

– 1

1 1







1





1





1







1







1





1



1

1

1

1











1





1

1

1

1

1













1

1

1

1

1

1



1









– 1







– –















1











1







1



1

1

1

1

1

1

































































1

1













1



1

1

1

1

1

1



























1

1

1









































































































































































1

1

1

1

1

1



1

1



1











1

































1

1

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 – – – – – – – – – – – 1 – – – – – – – – – –

















































M AF – –

360 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Echinocactus grusonii Hildm. Echinocactus horizontalonius Lem. Echinocactus platyacanthus Link & Otto Echinocereus cinerascens (DC.) Lem. Echinocereus pectinatus (Scheidw.) Engelm. ssp. pectinatus Echinocereus pentalophus (DC.) J.N. Haage Echinocereus schmollii (Weing.) N. P. Taylor Ferocactus echidne (DC.) Britton & Rose Ferocactus glaucescens (DC.) Britton & Rose Ferocactus histrix Lindsay Ferocactus latispinus (Haw.) Britton & Rose Isolatocereus dumortieri (Scheidw.) Backeb. Lophocereuss marginatus (DC.) S. Arias & Terrazas



1

1



































1

1

1





1









1

1

1

1

1

1

1

1

1

1







1

1

1

1

1

1

1

1

1

1

1







1

1

1

1

1

1

1

1

1

1

1

1

1

























– –

















– –











1

1

1

1

1

1

1

1

1

1

1

1

1































































1





















1





1



1

1

1

1

1

1

1















































































































































































































































1

1





1

1

1

1

1





















































































(continued)





























Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 361

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Family

Scientific name Lophophora diffusa (Croizat) Bravo Mammillaria candida (Scheidw.) Buxbaum Mammillaria compressa DC. Mammillaria crinita DC. ssp. painteri (Rose ex Quehl) W. & B. Fitz Maurice Mammillaria geminispina Haw. Mammillaria hahniana Werderm. Mammillaria herrerae Werderm. Mammillaria longimamma DC. var. longimamma Mammillaria magnimamma Haw. Mammillaria mathildae Kraehenb. & Krainz Mammillaria microhelia Werderm. Mammillaria muehlenpfordtii C.F. Först. Mammillaria parkinsonii Ehrenb.

Appendix 2 (continued)

1

1

1

















1





1







– 1

1





1

1



1

1

1



1

1

1

1

1





1

1





1

1



1

1

1

1

1

1

1

1

1

1

1

1

1

















































1

1





1

1





1

1



1









1



1

1

























































1

1





1

1





1

1













































































































1























































































1

1

1

1

1



1

1

1

1

1

1

































































































1





M AF

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH

362 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Mammillaria perbella Hildm. ex K. Schum. Mammillaria pettersonii Hildm. Mammillaria polythele Mart. ssp. polythele Mammillaria prolifera (Mill.) Haw. ssp. multiceps (SalmDyck) U. Guzmán Mammillaria rhodantha Link & Otto ssp. aureiceps (Lem.) D.R. Hunt Mammillaria sartorii J. A. Purpus Mammillaria schiedeana Ehrenb. ssp. schiedeana Mammillaria sempervivi DC. Mammillaria uncinata Zucc. ex Pfeiff. Mammillaria wiesingeri Boed. Mammillaria zephyranthoides Scheidw. Myrtillocactus geometrizans (Mart. ex Pfeiff.) Console Neobuxbaumia polylopha (DC.) Backeb. 1

– 1











1





1





1









1

1

1 –

1







1

1



1



1







1







1





1

1

1





1

1



1



1

1

1

1

1

1

1

1

1

1

1

1





– –















































1



1

1

1

1







1



1





1















1

























1













1

1











1





1





1

1







1

1



1



















































































1





























































































































1

1

1

1

1

1

1

1

1

1

1

1

1















































































(continued)



























Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 363

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Family Cactaceae

Scientific name Neolloydia conoidea (DC.) Britton & Rose ssp. conoidea Opuntia albicarpa Scheinvar Opuntia auberi Pfeiff. Opuntia cochenillifera (L.) Mill Opuntia cantabrigiensis Lynch Opuntia engelmannii Salm-Dyck ex Engelm. Opuntia ficus-indica (L.) Mill. Opuntia hyptiacantha F.A.C. Weber Opuntia icterica D. A. Griffiths Opuntia incarnadilla D.A. Griffiths Opuntia joconostle F.A.C. Weber ex Diguet Opuntia lasiacantha Pfeiff. Opuntia leucotricha DC.

Appendix 2 (continued)





























1





1

1

1

1



















– –







1









1



1

1

1

1

1

1

1

1

1

1

1





































– –







1

















1













































































1

















1































































1





































































































































1



1





1

1





1

1















































1





































1

















1











M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH – – 1 1 1 – – 1 – – 1 – – – – – – – – – 1 – –

364 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Opuntia matudae Scheinvar Opuntia microdasys (Lehm.) Pfeiff. Opuntia oligacantha C. F. Först Opuntia pachona D. A. Griffiths Opuntia pubescens J.C. Wendl. ex Pfeiff. Opuntia robusta J.C. Wendl. Opuntia streptacantha Lem. Opuntia tomentosa Salm-Dyck Opuntia zamudioi Scheinvar Lophocereus marginatus (DC.) Britton & Rose Pereskiopsis diguetii (F. A. C. Weber) Britton y Rose Pilosocereus leucocephalus (Poselg.) Byles & G. D. Rowley Rhipsalis baccifera (Mill.) Stearn Selenicereus spinulosus (DC.) Britton & Rose Selenicereus undatus (Haw.) D.R. Hunt – –

– – –



















1



















1



1

1

1









































1





























1

1

1

1

1

1

1

1

1

1

1





1



























1

























































































1

























































































1

1





























1













































































































































































































1



1

1

1

1





1

1









































1



1







































1















(continued)































Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 365

Campanulaceae

Calceolariaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Family Cactaceae

Scientific name Stenocereus huastecorum Alvarado-Sizzo, Arreola-Nava & Terrazas Stenocactus lamellosus (A. Dietr.) A. Berger ex A.W. Hill Stenocactus obvallatus (DC.) A. Berger Stenocactus ochoterenaus Tiegel Stenocereus queretaroensis (F. A. C. Weber) Buxb. Thelocactus hastifer (Werderm. & Boed.) F.M. Knuth Thelocactus leucacanthus (Zucc.) Britton & Rose ssp. leucacanthus Turbinicarpus pseudomacrochele Backeb. ssp. pseudomacrochele Calceolaria mexicana Benth. Lobelia laxiflora Kunth var. angustifolia A. DC.

Appendix 2 (continued)



– .

– 1

– –

1







1

1





























1



































1

1



















































1

































1

1

1

1





1









































































1





















1















1















1

1









1



1





1

1





1

1





1





1











1

1









1



1







1

















1





1





1









1





1

– 1

M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 – – – 1 – – – – – – – – – – – – – – – 1 – –

366 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Celtis caudata Planch. Cannabaceae Celtis pallida Torr. Caprifoliaceae Lonicera mexicana (Kunth) Rehder Caprifoliaceae Lonicera pilosa (Kunth) Spreng. Caprifoliaceae Symphoricarpos microphyllus Kunth Caricaceae Carica papaya L. Caryophyllaceae Paronychia mexicana Hemsl. Caryophyllaceae Silene laciniata Cav. Cistaceae Helianthemum glomeratum (Lag.) Lag. Clethraceae Clethra kenoyeri Lundell Clethraceae Clethra mexicana DC. Commelinaceae Commelina coelestis Willd. Commelinaceae Commelina erecta L. Commelinaceae Commelina tuberosa L. Commelinaceae Tradescantia crassifolia Cav. Convolvulaceae Dichondra argentea Humb. & Bonpl. ex Willd. Convolvulaceae Ipomoea capillacea (Kunth) G. Don Convolvulaceae Ipomoea cardiophylla A. Gray

Cannabaceae – –

– 1

1

1

– –

1 1





1

1 1

1 –





– 1

– –







1 –

1

1 1 –

– 1 – – – –

1



1



1



1







1





– –

– 1





1 –

1

1





– –

– 1







1

1





1



1 1

1





– –

– 1





1 –











1 1







– –

– 1





– –











– 1



1

1

1 1

– –



1

– 1









1

– 1







– –

– 1





1 –











– 1







– 1

– –





– 1









1

1 1

1

1

1

1 –

– –





– –





1



1

1 1

1





1 1

– 1





– –











1 1







– –

– 1





1 –











– –







– –

– –





– –











– –







– –

– –





– –









1

– 1







– –

– –





– –











– –



1



– –

– –





– –











– –







1 –

– 1





– –











– –







– –

– 1





– –











– 1







– –

– –





– –











– –







– –

– –





– –







1



– –







– –

1 –



1

– 1









1

1 1







– –

– 1





– –











– –





1

– –

1 –





– –











– –







– –

1 –





– –



(continued)









1 –

1





– –

– –





– –



Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 367

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Convolvulaceae Convolvulaceae

Convolvulaceae

Convolvulaceae Convolvulaceae

Convolvulaceae

Convolvulaceae

Convolvulaceae

Convolvulaceae

Family Convolvulaceae

Scientific name Ipomoea carnea Jacq. Ipomoea hederifolia L. Ipomoea indica (Burm.) Merr. Ipomoea jalapa (L.) Pursh Ipomoea murucoides Roem. & Schult. Ipomoea nil (L.) Roth Ipomoea orizabensis (G. Pelletan) Ledeb. ex Steud. var. orizabensis Ipomoea purpurea (L.) Roth Ipomoea stans Cav. Merremia dissecta (Jacq.) Hallier f. Echeveria mucronata Schltdl. Echeveria secunda Booth. ex Lindl. Echeveria semivestita Moran Echeveria subrigida (B.L. Rob. & Seaton) Rose Echeveria tolimanensis Matuda Graptopetalum pachyphyllum Rose

Appendix 2 (continued)

– – –



– 1

1



1 1





1





1

– 1



1



1

1



1

1

1

1 –

– 1

1 1



– 1





1

– –



1

– –



– –



1





1



– –



– –



1





1









1

1









1













– –



1 1















1

1



– –



– 1





1









1

1

1

– –



– –













1





1

– –



– 1















1





– –



– –



















1

– –



– 1





















– –



– –











1









– 1



– –





















– –



– –



















1

– –



– –



















1

– –



– –





















– –



– –





















– –



– –



















1

– –



– –





















– –



– –









1



1





1

– –



– –



















1

– –



– –





















– –



– –









Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 – – – – – – – – – – – – – – – – – – – 1 – –













– –



– –





















– –



– –









M AF – –

368 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Cupressaceae

Cucurbitaceae Cupressaceae

Cucurbitaceae Cucurbitaceae Cucurbitaceae

Cucurbitaceae

Cucurbitaceae

Cucurbitaceae

Cucurbitaceae

Cucurbitaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Pachyphytum compactum Rose Pachyphytum viride E. Walther Sedum bourgaei Hemsl. Sedum dendroideum DC. Sedum greggii Hemsl. Sedum moranense Kunth Sedum praealtum A. DC. Apodanthera undulata A. Gray Cucurbita argyrosperma K. Koch Cucurbita ficifolia Bouché Cucurbita foetidissima Kunth Cucurbita moschata Duchesne Cucurbita pepo L. Melothria pendula L. Sechium edule (Jacq.) Sw. ssp. sylvestre Lira & Castrejón Sicyos deppei G. Don Cupressus lusitanica Mill. Juniperus deppeana Steud. – – –







1

1













– 1 –

1



1

















– 1 1

1 –

1

1 –





1 –

– 1 –



1





1

1

1







1



1 –

– 1 –



1





1

1

1













1 –

– 1 –



1





1

1

1







1





1 –

1 1 –



1

1

1

1

















1 –

– 1 –

– 1 –



1





1

1

1







1

1

– 1 –



























1 –

– – –













1













– 1

– – –













1

1











– –

– 1 –



1





1



1













– –

– 1 –



1





1



1













– –

– – –



























– –

– – 1



























– –

– – –



1























– 1

– – –



















1





1



– – –













1











1

– –

– – –



























– –

– – –



























– 1

– – –



























1 1

1 1 –

1

1

1

1

1



1

1





1

1



1 –

1 1 –



1

1

1



















– 1

– 1 –



























1 –

– – –

























(continued)



– –

1 – –

1

1

1

1

















Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 369

Scientific name Juniperus flaccida Schltdl. Cystopteridaceae Cystopteris fragilis (L.) Bernh. Dennstaedtiaciae Dennstaedtia distenta (Kunze) T. Moore Dennstaedtiaciae Dennstaedtia globulifera (Poir.) Hieron. Dicksoniaceae Lophosoria quadripinnata (J.F. Gmel.) C. Chr. Dryopteridaceae Dryopteris pseudofilix-mas (Fée) Rothm. Dryopteridaceae Dryopteris wallichiana (Spreng.) Hyl. Dryopteridaceae Elaphoglossum petiolatum (Sw.) Urb Dryopteridaceae Phanerophlebia nobilis (Schltdl. & Cham.) C. Presl. Dryopteridaceae Polystichum distans E. Fourn. Equisetaceae Equisetum hyemale L var. affine (Engelm.) A.A. Eaton Equisetaceae Equisetum myriochaetum Schltdl. & Cham. Ericaceae Arbutus xalapensis Kunth

Family Cupressaceae

Appendix 2 (continued)

1





– –

– –





1





1





1















1

1

1

















































1

1

1

1





















1









1











































1

1













































































































































































1

1







































1

1











1























1











1





















1





























1

























1





1















1





























1













1





M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH – – – – – – – – – – – – – – – – 1 1 – – – – –

370 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Ericaceae

Arctostaphylos pungens Kunth Acalypha adenostachya Cav. Acalypha infesta Poepp. Acalypha monostachya Cav. Bernardia mexicana (Hook. & Arn.) Müll. Arg. Cnidoscolus multilobus (Pax) I.M. Johnst. Cnidoscolus aconitifolius (Mill.) I.M. Johnst. Croton ciliatoglandulifer Ortega Croton cortesianus Kunth Croton morifolius Willd. Euphorbia anychioides Boiss. Euphorbia bracteata Jacq. Euphorbia dentata Michx. Euphorbia graminea Jacq. Euphorbia prostrata Aiton –

– – –

1









1





1

1

1





1





1



– –



































1

1







1









1

1

1

1

1









1 –





1







1









1





1











1

1

1

1

1

1

1

1







1









1









1





1











1







1

1

































1

1











1















1









1













1





























1























1

1















1

































1



















1





























1





























1





























































































1



1











1













1

1





















































































(continued)



1



























Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 371

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Family Euphorbiaceae

Scientific name Euphorbia radians Benth. Euphorbia tanquahuete Sessé & Moc. Jatropha dioica Sessé Manihot angustiloba (Torr.) Müll. Arg. Acaciella angustissima (Mill.) Britton & Rose Albizia occidentalis Brandegee Albizia plurijuga (Standl.) Britton & Rose Bauhinia chapulhuacania Wunderlin Bauhinia coulteri J.F. Macbr. var. coulteri Bauhinia divaricata L. Bauhinia ramosissima Benth. ex Hemsl. Caesalpinia pulcherrima (L.) Sw. Calliandra eriophylla Benth.

Appendix 2 (continued)





1 – 1 – –





1







1





1

1

1

1

1







1



1



1

1















1

1

1





– 1



1

1

1

1



















1



1





1



















1

1











– –























1

1











1







































1











































1









1





























































1











1

















































































































1









1









































1



1













Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH – – 1 1 1 – – – – – – – – – – – – – – – – – –

















































M AF – –

372 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae Fabaceae

Fabaceae

Calliandra grandiflora (L’Hér.) Benth. Cercis canadensis L. Chamaecrista nictitans (L.) Moench var. jaliscensis (Greenm.) H.S. Irwin & Barneby Cojoba arborea (L.) Britton & Rose Crotalaria longirostrata Hook. & Arn. Dalbergia paloescrito Rzed. & Guridi-Gómez Dalea bicolor Humb. & Bonpl. ex Willd. Dalea foliosa (A. Gray) Barneby Dalea lutea (Cav.) Willd. Desmodium grahamii A. Gray Desmodium molliculum (Kunth) DC. Enterolobium cyclocarpum (Jacq.) Griseb. Erythrina coralloides DC. Eysenhardtia polystachya (Ortega) Sarg. – –

– –



1



1

1



– –





1

1







1

1

1

1

1

1



1

1



1 1

1

1

1







1

1

1

1

1



– –

– –

1 1

1





1

1

1







– –





1























– 1





1

1 –









– 1



1

1

1

1





– –













1

1

1







– –









1

1









1



– –

1







1

1-













1 –

1







1



1

1

1







– –













1

1

1







– –







1















1

– –























1

1 –













1

1

1







– –

























– –

























– –

















1







– –

























– –

























– –



1

1

1















1

– –















1









– –





1

















1

1 –























1

– –



(continued)























– –



Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 373

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Family Fabaceae

Scientific name Harpalyce arborescens A. Gray Havardia pallens (Benth.) Britton & Rose Indigofera suffruticosa Mill. Leucaena esculenta (Moc. & Sessé ex DC.) Benth. Leucaena leucocephala (Lam.) de Wit Leucaena pulverulenta (Schltdl.) Benth. Lonchocarpus rugosus Benth. Lysiloma acapulcense (Kunth) Benth. Lysiloma microphyllum Benth. Macroptilium gibbosifolium (Ortega) A. Delgado Mimosa depauperata Benth. Mimosa leucaenoides Benth. Mimosa texana (A. Gray) Small

Appendix 2 (continued)

1



1

– 1

1 1











– –





1

1





1







1

1





1

1





– –













1

















1

1



1

1

1

1

1

1

1

1











1



1





1















1

1











































































































































1

1

1-







1

























1



1













































1







































































































1

1



1

1

























1















1



1







Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 – – – – – – – – – – – – – – – – – – – – – –









































1







M AF – –

374 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

-Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Parkinsonia aculeata L. Phaseolus coccineus L. Phaseolus heterophyllus Humb. & Bonpl. ex Willd. Phaseolus vulgaris L. Pithecellobium dulce (Roxb.) Benth. Prosopis laevigata (Humb. & Bonpl. ex Willd.) M.C. Johnst. Rhynchosia prostrata Brandegee Senna atomaria (L.) Irwin et Barneby Senna polyantha (Moc. & Sessé ex Collad.) H.S. Irwin & Barneby Senna septemtrionalis (Viv.) H.S. Irwin & Barneby Senna wislizeni (A. Gray) H.S. Irwin & Barneby var. painteri (Britton) H.S. Irwin & Barneby Vachellia farnesiana (L.) Willd. y Arn. Vachellia pennatula (Schltdl. & Cham.) Seigler & Ebinger

1

1

1

1





1 –



1-

1

1



1

1

– –



1

1







1

1





1

1











1

1

1

1









1













1

1



1











1

1

1

1





1



1

1

1



1

1



1















1























1















1





1





1



1







1



1

























1



1













1















1









1





















1





1

























1

























































1

























1

























1





































































1

1



1

1

1













1

1

1





1























1





























(continued)



















1



1



Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 375

Iridaceae

Hydrangeaceae

Geraniaceae

Gentianaceae

Gentianaceae

Garryaceae

Fagaceae Fouquieriaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fabaceae

Family Fabaceae

Scientific name Vachellia schaffneri (S. Watson) Seigler & Ebinger Zornia diphylla (L.) Pers. Quercus crassipes Bonpl. Quercus crassifolia Humb & Bonpl Quercus greggii (A. DC.) Trel. Quercus grisea Liebm. Quercus potosina Trel. Quercus rugosa Née Fouquieria splendens Engelm. Garrya laurifolia Hartw. ex Benth. ssp. laurifolia Eustoma exaltatum (L.) Salisb. ex G. Don Gentiana spathacea Kunth Geranium seemannii Peyr. Philadelphus coulteri S. Watson

Appendix 2 (continued)

1

1

1

1

1

1

1 –

1





1

1

1

1











– –

























– –

1

1

1



















– 1

























– 1















1









– –













1











– –

























– 1















1









1 –



















1



1

1 –









1















– –



1





















– 1

























– –





















1



– –

























– –



















1





– –

























– –







1

















– –





















1



– –















1









– –













1











– 1

























– –

























– –













Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH – – – 1 1 1 – – – – 1 – – – – – – – – – – 1 –













– –

























– –













M AF – –

376 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Lamiaceae

Lamiaceae Lamiaceae

Lamiaceae

Lamiaceae Lamiaceae Lamiaceae

Lamiaceae Lamiaceae

Lamiaceae

Lamiaceae

Lamiaceae Lamiaceae

Lamiaceae

Lamiaceae

Koeberlinaceae

Juglandaceae

Iridaceae

Tigridia pavonia (L. f.) DC. Tigridia vanhouttei Roezl ex Van Houtte Juglans mollis Engelm. Koeberlinia spinosa Zucc. var. spinosa Agastache mexicana (Kunth) Lint & Epling Clinopodium mexicanum (Benth.) Govaerts Hyptis albida Kunth Lepechinia caulescens (Ortega) Epling Salvia amarissima Ortega Salvia coccinea Buc’hoz ex Etl. Salvia elegans Vahl Salvia helianthemifolia Benth. Salvia hispanica L. Salvia keerlii Benth. Salvia leucantha Cav. Salvia melissodora Lag. Salvia mexicana L. Salvia microphylla Kunth Salvia patens Cav. 1



1 1

1

1

– –





– –

– 1 1

1 – –



– 1 1

1



– –

– –

– 1

– –



– 1

1





1 –

1





1









1

1



1







– –



– – –



– –



– 1 –

– –



– – –



– –













– –





1







– –



– – –

– –





– –







1





– –



1 – –

1 –





– –

1







1



– –

1

– – –

– –





– –











1

– 1



– – –

1 –



1

– –

1











1 1



– – –

1 –

1



– –







1





1 –



– 1 –

– –





– –













– –



– – –

– –





– –





1







– –



– – –

– –





– –













1 –



– – –

– –

1



1 –













– –



– 1 –

– –





– –











1

1 1



– – –

1 –





– –













1 1



– – –

– –





– –













– 1



– – –

– –



1

– –













– –



– – –

– –





– –













1 1



– – –

– –





– –













– 1



1 – 1

– –

1

1

1 –

1

1









– –



– – –

– –





– –













– –



– – –

– –





– –







1





– –



– – –

– –





– –











(continued)



– 1



– – –

– –





– –











Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 377

Lentibulariaceae

Lentibulariaceae

Lentibulariaceae

Lentibulariaceae

Lentibulariaceae

Lentibulariaceae

Lentibulariaceae

Lentibulariaceae

Lauraceae

Lauraceae

Lamiaceae Lamiaceae Lamiaceae

Family Lamiaceae

Scientific name Salvia polystachia Cav. Salvia regla Cav. Salvia tiliifolia Vahl Teucrium cubense Jacq. ssp. chamaedrifolium (Mill.) Epling Listea glaucesens Kunth Persea americana Mill. Pinguicula agnata Casper Pinguicula calderoniae Zamudio Pinguicula elizabethiae Zamudio Pinguicula esseriana B. Kirchn. Pinguicula lilacina Schltdl. & Cham. Pinguicula macrophylla Kunth Pinguicula moctezumae Zamudio & R.Z. Ortega Pinguicula moranensis Kunth var. moranensis

Appendix 2 (continued)



1

1

1













1





1

1

1































– –











1

1



1



1





















1





1





1



1











1









1







1









1

1

1 – –

1



1















– – –





















1 – –





















– – –





















– – –



1

















– – –



















1

– – –





















– – –









1





1





1 – –









1











1 – –





















– – –





















1 – –

















1



– – 1





















– – –

















1

1

– – –









































– – –

1

1

– – –



1

– – –

1

– – –



1 1 –

– – 1

– – –

1 1 –

– – 1

– – 1

M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH – 1 – – 1 – 1 – 1 – – – – – – – 1 1 – 1 – – –

378 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malpighiaceae

Malpighiaceae

Magnoliaceae

Magnoliaceae

Lythraceae

Lythraceae

Lythraceae

Lygodiaceae

Lycopodiaceae

Loranthaceae

Loasaceae

Liliaceae

Calochortus barbatus (Kunth) J.H. Painter Mentzelia hispida Willd. Psittacanthus calyculatus (DC.) G. Don Lycopodium clavatum L. Lygodium venustum Sw. Cuphea aequipetala Cav. Cuphea lanceolata W.T. Aiton Heimia salicifolia Link Magnolia grandiflora L. Magnolia schiedeana Schltdl. Byrsonima crassifolia (L.) Kunth Gaudichaudia cynanchoides Kunth Allowissadula holosericea (Scheele) D.M. Bates Anoda cristata (L.) Schltdl. Ayenia jaliscana S. Watson Ceiba aesculifolia (Kunth) Britten & Baker f. – –





















1

1

1

1

1







1 – –

1











1



1







1













1







1







1

1



1

1



1

1

1

1

1

1

1

1









– –































1









– –

1



1















1

1





1

1





















1









1

– –









1















1





























































1









1































1













1











1

























1

1































1





























































































































1













































































1

































































































(continued)

































Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 379

Malvaceae Marattiaceae

Malvaceae

Malvaceae

Malvaceae Malvaceae Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Family Malvaceae

Scientific name Ceiba pentandra (L.) Gaertn. Gossypium hirsutum L. Guazuma ulmifolia Lam. Kearnemalvastrum lacteum (Ait.) D.M. Bates Malvaviscus arboreus Cav. Pavonia candida (DC.) Fryxell Phymosia umbellata (Cav.) Kearney Pseudobombax ellipticum (Kunth) Dugand var. ellipticum Sida abutifolia Mill. Sida rhombifolia L. Sphaeralcea angustifolia (Cav.) G. Don Tilia americana Schltdl. var. mexicana Triumphetta semitriloba Jacq. Waltheria indica L. Marattia weinmanniifolia Liebm.

Appendix 2 (continued)

1 – 1 – – – –

– – –





– –

1 1 1



1 –

1 – –





1 – 1



1





1

1





– –





– – –

















1













1

1

1

– 1





1



1 –





1 1 1



1

1

1





1

– 1





– 1 –



1

1

1

1





– –





1 – –















– –





– – –







1







– –





– – –

1













– –





1 – –









1





– –





1 – –















– –



1

– – –

1













– –





– – –







1







– –





– – –















– –





– – –















– –





– – –



1











– –





– – –















– –





– – –















– –





– – –















– –





– 1 1

1

1



1



1

1

– –





1 1 –













1

– –





– – –











1



– –





– 1 –













– –





– – –























M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 – – – – – – – – – – – 1 – – – – – – – 1 1 –

380 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Nyctaginaceae

Nyctaginaceae

Nyctaginaceae

Myrtaceae Nyctaginaceae

Myrtaceae

Moraceae

Moraceae

Menispermaceae

Menispermaceae

Menispermaceae

Meliaceae Meliaceae

Meliaceae

Melastomataceae

Martyniaceae Martyniaceae

Marsileaceae

Marsilea mollis B.L. Rob. & Fernald Martynia annua L. Proboscidea louisiana (Mill.) Thell. Tibouchina longifolia (Vahl) Baill. Cedrela dugesii S. Watson Cedrela odorata L. Trichilia havanensis Jacq. Cissampelos pareira L. Cocculus diversifolius DC. Menispermum canadense L. Ficus cotinifolia Kunth Morus celtidifolia Kunth Myrcianthes fragrans (Sw.) McVaugh Psidium guajava L. Mirabilis glabrifolia (Ortega) I.M. Johnst. Mirabilis jalapa L. var. jalapa Mirabilis longiflora L. Pisonia aculeata L. – 1 – – – –

– – –

– –

1



– –

1











1 1



1 1









– –





1

1







1 1

1 –



1

1 –

1

1

1







1



– –





– 1

– 1 –













– –





– –















– –





– 1



– –



1

– –







– –









1







1 –







1



1 1

– –





– –









– 1













– –





– –









1 –











1

– –





– –









1 –











1

– –





– –









– –













– –





– –









– 1













– –





– –









1 –

1









1

1 –





1 –



1





1 –





1







– –





– –

1







– –













– –





– –









– –













– –





– –









– –













– –





– –









– –













– –





– –









– –













– –





– –

1







– –













– –





– –







1

1 –











1

1 1





– –









– –









1



– –





– –









1 –





1







1 –





– –









– –













– –





– –



(continued)







– –













– –





– –



Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 381

Orobanchaceae

Orobanchaceae

Orchidaceae

Orchidaceae

Onagraceae

Onagraceae

Onagraceae

Onagraceae

Onagraceae

Oleaceae

Oleaceae

Oleaceae

Nymphaeaceae

Family Nyctaginaceae

Scientific name Pisoniella arborescens (Lag. & Rodr.) Standl. Nymphaea ampla (Salisb.) DC. Forestiera phillyreoides (Benth.) Torr. Fraxinus rufescens Lingelsh. Fraxinus uhdei (Wenz.) Lingelsh. Gaura coccinea Nutt. ex Pursh. Lopezia racemosa Cav. Oenothera desertícola (Loes.) Munz Oenothera kunthiana (Spach) Munz Oenothera rosea L’Hér. ex Ait. Dichromanthus cinnabarinus (La Llave & Lex.) Garay Mesadenus polyanthus (Rchb. f.) Schltr. Castilleja arvensis Schltdl. & Cham. Castilleja tenuiflora Benth.

Appendix 2 (continued)









1

1

1





1

1







1

1











1



1















































1

1

1



















1

1

1

















1













1







































1













1





























1



1















1







1















































1

















1































































































1

























































































1

1





1





1











1

1









1















1









1







1











1







1









1

1



1



1



1















M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 – – – – – – – – – – – – 1 – – – – – – – – –

382 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Lamourouxia dasyantha (Cham. & Schltdl.) W.R. Ernst Orobanchaceae Lamourouxia multifida Kunth Oxalidaceae Oxalis corniculata L. Oxalidaceae Oxalis decaphylla Kunth Oxalidaceae Oxalis latifolia Kunth Oxalidaceae Oxalis divergens Benth. ex Lindl. Papaveraceae Argemone ochroleuca Sweet Papaveraceae Bocconia frutescens L. Papaveraceae Hunnemannia fumariifolia Sweet Passifloraceae Passiflora bryonioides Kunth Passifloraceae Passiflora exsudans Zucc. Passifloraceae Passiflora foetida L. var. lanuginosa Killip Passifloraceae Passiflora serratifolia L. Passifloraceae Passiflora subpeltata Ortega Turneraceae Turnera diffusa Willd. Pentaphylacaceae Ternstroemia sylvatica Schltdl. & Cham. Petiveriaceae Rivina humilis L. Phyllantaceae Phyllanthus compressus Kunth

Orobanchaceae

1 1

– – –

1





1

1

1

1





– 1





1



1

1

1





1 1

– –

– –

– –



1





1













1



1

1

– –

– –

1 –

1 –

1 1

– –

1 1

1 1



1



1



1

1

1

– –



1















1

– –

– 1





– –





1

1









1

1

– –

– –





– 1

1



1

1









1



1 –

– –





– –



1









1





1

– –

1 1





– –

















1



1 –

– 1

1

1

– –





1





1





1



– –

– –

1



– –













1







– –

– –



1

– –



1

















– –

– –





– –







1

1











– –

– –





– –





















– –

– –





– –





















– –

– –



1

– –





















– 1

– –





– –





















– –

– –





– –





















– –

– –





– –





















– –

– –





– –





















– –

– –





– –





















– –

– –





– –



















1

– –

– –





– –





















– –

– –





– –





















– –

– –





(continued)

– –





















– –

– –





Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 383

Piperaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Phytolaccaceae

Phytolaccaceae

Phytolaccaceae

Family Phyllantaceae

Scientific name Phyllanthus grandifolius L. Agdestis clematidea Moc. & Sessé ex DC. Phytolacca icosandra L. Phytolacca rivinoides Kunth & C.D. Bouché Abies religiosa (Kunth) Schltdl. & Cham. Pinus ayacahuite C. Ehrenb. ex Schltdl. Pinus cembroides Zucc. Pinus devoniana Lindl. Pinus greggii Engelm. ex Parl. Pinus montezumae Lamb. Pinus patula Schltdl. & Cham Pinus pinceana Gordon & Glend. Pinus pseudostrobus Lindl. var. pseudostrobus Pinus teocote Schltdl. & Cham. Peperomia hispidula (Sw.) A. Dietr.

Appendix 2 (continued)

1

1







1

1



1









1

























1

1

















































































































1

1





1



1















































1



1

1

1





1

1

















1



1





























































































































































1















1









































1

1







1



























































1









1









1

1

1

1

































































Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 1 – – – – 1 – 1 1 – – – – – – – – – – – – –













































1

1









M AF – –

384 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae Poaceae

Plumbaginaceae

Platanaceae

Plantaginaceae

Plantaginaceae

Plantaginaceae

Plantaginaceae

Plantaginaceae

Plantaginaceae

Piperaceae Plantaginaceae

Piper auritum Kunth Bacopa monnieri (L.) Wettst. Maurandya antirrhiniflora Humb. & Bonpl. ex Willd. Maurandya barclayana Lindl. Penstemon barbatus (Cav.) Roth Penstemon hartwegii Benth. Penstemon hidalgensis Straw Plantago nivea Kunth Platanus mexicana Moric. Plumbago pulchella Boiss. Aristida laxa Cav. Bouteloua curtipendula (Michx.) Torr. Bouteloua scorpioides Lag. Bromus carinatus Hook. & Arn. Cenchrus echinatus Cav. Chloris submutica Kunth –



– –

1 –





1

1

1





1

1 1

1

1





1









1



– 1





1



1



1

– 1

1







– –

1 –

1 1





1

1

– 1

1















– –





1

1

– –

1



1











– –

1

1

1



– –

















– –









– –



1





1







1 –





1

1

– 1

1



1











– –





1

1

– –

1







1

1





– –

1



1



– –

1





1

1

1





– –

1

1

1

1

– –

1



1











– –





1

1

– –

1





1









– –









– –

















– –

1







– –



1













– –





1



– –

1















– –





1

1

– –

1









1





– –





1

1

– –

















– –





1



– –

1















– –









– –

















– 1





1



– –

















– –









– –













1



1 1

1

1

1

1

– –

















– –









– –

















1 –









– –

















– –

(continued)

1







– –

















– –

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 385

Polemoniaceae

Polemoniaceae

Polemoniaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

Poaceae

-Poaceae

Family Poaceae

Scientific name Chondrosum hirsutum (Lag.) Sweet Eragrostis mexicana (Hornem.) Link Hilaria belangeri (Steud.) Nash Hopia obtusa (Kunth) Zuloaga & Morrone Leptochloa dubia (Kunth) Nees Muhlenbergia rigida (Kunth) Kunth Paspalum notatum Flüggé Paspalum prostratum Scribn. & Merr. Pennisetum crinitum (Kunth) Spreng. Setaria grisebachii E. Fourn. Urochloa fasciculata (Sw.) R. Webster Zea mays L. ssp. mays Cobaea scandens Cav. Loeselia coerulea (Cav.) G. Don Loeselia mexicana (Lam.) Brand

Appendix 2 (continued)



1









1





1

1

1

1

1







1

1



1



1



1

1

1

1

1

1







1

1



1

1

1

1





1

1





1

1

– –

1



1

1

– –

1

1





1 –

















1

























1













1









1













1

1









1





1

1





1

1

1

1

1



1



















1





1





































1











1















1







1















1

















































































1









1





































1

1



1





























1





1

1

1

1

















1 –

1

1

1



1







1



1



1







































1













1









1

1



1



1

1







1



1



1







1



1



1



– 1

M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH – 1 1 1 1 – – 1 1 1 1 1 – – – – 1 1 – – – 1

386 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Pteridaceae

Pteridaceae

Pteridaceae

Proteaceae

Portulacaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polygonaceae

Polygonaceae

Polygonaceae

Persicaria hydropiperoides (Michx.) Small Persicaria segetum (Kunth) Small Rumex mexicanus Meisn. Campyloneurum angustifolium (Sw.) Fée Phlebodium areolatum (Humb. & Bonpl. ex Willd.) J. Sm. Pleopeltis macrocarpa (Bory ex Willd.) Kaulf. Pleopeltis polylepis (Roem. ex Kunze) T. Moore Polypodium subpetiolatum Hook. Polypodium thyssanolepis A. Braun ex. Klotzsch Portulaca oleracea L. Roupala montana Aubl. Adiantum andicola Liebm. Adiantum braunii Meet. ex Kuhn Adiantum capillusveneris L.



1









1





1

1























1

1





1

– –



























1







1

















1-











1

























1













1







1

1

1





1



1

1

1







































1





1

1

1





1







1



1











1



































































1















1

1









1





















1



1



1





























1









1



























































































1



1





































1



























1













1





























































(continued)









1



















Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 387

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Family Pteridaceae

Scientific name Adiantum concinnum Humb. & Bonpl. ex Willd. Adiantum poiretii Wikstr. Adiantum tenerum Sw. Adiantum trapeziforme L. Adiantum tricholepis Fée Aleuritopteris farinosa (Forssk.) Fée Argyrochosoma incana (C. Presl) Windham Astrolepis sinuata (Lag. ex Sw.) D.M. Benham & Windham Cheilanthes angustifolia Kunth Cheilanthes bonariensis (Willd.) Proctor Cheilanthes candida M. Mart. & Gal. Cheilanthes kaulfussii Kunze Cheilanthes lendigera (Cav.) Sw.

Appendix 2 (continued)

1 –

– –



1

1



1

1

1

1

1



1

1

1



1

1

1

1



1



1

1

1



1





1

1

1



1











1





1

1







1

– –





– 1





1

1







1



1

1

1

















1







































1



1

1

1

1

1







1







1



1

1









1



1



1





1



















































1

1

1



















1



1

1

1



1

1

























1





1







1

1





1



1



















1

















































1









1

























1













1

1



































Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 1 1 – – 1 1 – – – – – 1 1 – – – – – – – – –

















































M AF – –

388 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Rhamnaceae

Rhamnaceae

Ranunculaceae Rhamnaceae

Pteridaceae Pteridaceae Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Cheilanthes leucopoda Link Cheilanthes meifolia DC. Eaton Cheilanthes myriophylla Desv. Cheiloplecton rigidum (Sw.) Fée var. rigidum Hemionitis palmata L. Llavea cordifolia Lag. Pellaea cordifolia (Sessé & Moc.) A.R. Sm. Pellaea ovata (Desv.) Weath. Pellaea ternifolia (Cav.) Link Pityrogramma calomelanos (L.) Link var. calomelanos Pteris cretica L. Pteris longifolia L. Pteris quadriaurita Retz. Clematis dioica L. Ceanothus caeruleus Lag. Colubrina elliptica (Sw.) Brizicky & W.L. Stern Condalia mexicana Schltdl.

1

1

1 –

1 – 1

1 –



– 1

1

– –

– –

1 1

1 – 1

– – –

1

– 1 –







1





– –

– – –





1

1





1



1

1





1

1



1







1

1

1

1

1

– 1





1

1

1

– 1



1



1

1

1

1

1



1

1

1





– –

– 1 –







– –

1 1 1

1



1

– –



1





1



– –



1













– –

– – –







1





1











– –

– – –









1





1









– –

1 – –



1

1

1







1

1

1





– –

– – –



1











1









– –

– – –



















1





– –

– – –









1

1













– –

1 – 1

























– 1

– – –









1





1









– –

– – –

























– –

– – –









1





1









– –

– – –

























– –

– – –























– –

– – 1

























– –

– – –

























– –

– – –

























– –

– – –

























– –

– – –





















(continued)





– –

– – –





















Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 389

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rhamnaceae

Rhamnaceae

Rhamnaceae

Rhamnaceae

Family Rhamnaceae

Scientific name Condalia velutina I.M. Johnst. Frangula capreifolia (Schltdl.) Grubov Gouania polygama (Jacq.) Urb. Karwinskia humboldtiana (Schult.) Zucc. Karwinskia mollis Schltdl. Crataegus gracilior Phipps Crataegus mexicana DC. Crataegus rosei Eggl. Duchesnea indica (Andrews) Teschem. Fragaria vesca L. ssp. bracteata (Heller) Staudt Lindleya mespiloides Kunth Malacomeles denticulata (Kunth) Decne. Prunus serotina Ehrh, ssp. capuli (Cav.) McVaugh Purshia mexicana (D. Don) S.L. Welsh

Appendix 2 (continued)

1





1

1

1



1

1

1















1

1



1









1

1











1





1





1















1











1

– –



1

1



1

1



1



















1







1





1







1







1





















1







1

1



1





1

























1

1





















1































1





























1















1













1



1

1



























1



















1





























1









































































1

1









1





















1















































































1



















M AF – –

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 – – 1 – – – – – – – – – – – – – – – – – – –

390 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Sapotaceae

Sapindaceae

Sapindaceae

Sapindaceae Sapindaceae

Salicaceae

Salicaceae

Salicaceae

Rutaceae Rutaceae

Rutaceae

Rutaceae

Rubiaceae Rutaceae

Rubiaceae

Rubiaceae

Rubiaceae

Bouvardia longiflora (Cav.) Kunth Bouvardia ternifolia (Cav.) Schltdl. Cigarrilla mexicana (Zucc. & Mart. ex DC.) Aiello Hamelia patens Jacq. Casimiroa edulis La Llave Casimiroa greggii (S. Watson) F. Chiang Decatropis bicolor (Zucc.) Radlk. Ptelea trifoliata L. Zanthoxylum fagara (L.) Sarg. Neopringlea integrifolia (Hemsl.) S. Watson Populus tremuloides Michx. Salix bonplandiana Kunth Acer negundo L. Cardiospermum halicacabum L. Dodonaea viscosa Jacq. Sapindus saponaria L. Sideroxylon palmeri (Rose) T.D. Penn.

1

1

– –



– 1



– 1

– 1



1











1

– 1









1

– –

– 1







– –





– –



1



1 1







– –

1



1 –



1

– –





– –







– –

– –

1–

– – 1





1

1





1

1

1 1

1

1

1 1

1

– –



1

1







– –



1

1



– –

– –

1 1 1

1





1



1

1

1

1







– –







– –





– –













– –



1



– –





– –



1







1

– –



1



– –





– –



1









– –







– –





– –













– –







– –





– –







1





– –







– –





– –







1





– –

1





– –





– –











1

– –







– –





– –













– –







– –





– –













– –







– –





– –













– –







– –





– –













– –







– –





– –













– –







– –





– –









1



1 1



1



– –





1 1





1







– –







– –





– –









1-



– –







– –





1 –













– –







– –





– –







(continued)







– –







– –





– –







Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 391

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae Solanaceae

Selaginellaceae

Selaginellaceae

Selaginellaceae

Scrophulariaceae Selaginellaceae

Scrophulariaceae

Scrophulariaceae

Scrophulariaceae

Family Sapotaceae

Scientific name Sideroxylon salicifolium (L.) Lam. Buddleja cordata Kunth Buddleja scordioides Kunth Buddleja sessiliflora Kunth Capraria biflora L. Selaginella harrisii Underw. & Hieron. Selaginella lepidophylla (Hook, & Grev.) Spring Selaginella sellowii Hieron Selaginella stenophylla A. Braun Capsicum annum L. Capsicum frutescens L. Capsicum rhomboideum (Kunth) Kuntze Cestrum dumetorum Schltdl. Cestrum nitidum Mart. & Gal. Cestrum nocturnum L.

Appendix 2 (continued)

– – – 1

1 – – 1 –



– 1







– 1



1

1

1

1 1



1







1 1

1

1

1











1









– 1





1

– –















– –





1

– –















– –







– –







1

1

1



– –

1





– 1















– –





1

– –







1





– –







– –













1-

– –







– –















– –





1

– –



1











– –





1

– –















– –

1





– –















– –







– –















– –





1

– –



1











– –







– –















– –







– –















– –







– –



1











– –







– –















– –







– –







1







1 –



1

1

– –





1









– –







– –















– –







– –















1 –







– –













1 –







– –









1



1

M AF – 1

Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 – – – – – – – – – – – – – – – – – – – 1 – –

392 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Solanaceae

Solanaceae Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae Solanaceae

Solanaceae

Solanaceae

Solanaceae

Cestrum oblongifolium Schltdl. Cestrum thyrsoideum Kunth Datura ceratocaula Ortega Datura inoxia Mill. Datura stramonium L. Jaltomata procumbens (Cav.) J.L. Gentry Lycianthes moziniana (Dunal) Bitter Nicotiana trigonophylla Dunal Physalis ampla Waterf. Physalis angulata L. Physalis cinerascens (Dunal) Hitchc. Physalis coztomatl Dunal Physalis gracilis Miers Physalis nicandroides Schltdl. Physalis patula Mill. Physalis philadelphica Lam. Physalis pubescens L.

– 1





– 1





1

– –

– –

1

1

1

1

1 –

1





– 1





1

1

– –

1





– 1



1



1



1

– –

1

1



– –





– –

– –





1



















1



1 1 1



1

– – –





1 1





1



1 1



1







– 1









– –









– –



1

1

– –









– –





1



– –







– 1









– –



1





– –





1

– 1







1

– –



1





– –

1



1

– 1









– –



1





1 –







– 1

1







– –









1 –







– –









– –









– –







– –









– –







1

– –







– –









– –

1







– –







– –









– –









– –





1

– –









– –



1





– –







– –









– –









– –







– –









– –









– –







– –









– –









– –







– –









– –









– 1







1 –

1



1

1

– –







1

1 1

1





1 1





1



– 1



1





– –







– –









– –





1



– –







– –









– –







(continued)



– 1







– –





1



– –







Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 393

Tectariaceae

Taxaceae

Talinaceae

Styracaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Family Solanaceae

Scientific name Physalis solanacea (Schltdl.) Axelius Physalis sordida Fernald Physalis virginiana Mill. Solandra maxima (Sessé & Moc) P.S. Green. Solanum americanum Mill. Solanum dulcamaroides Dunal Solanum elaeagnifolium Cav. Solanum lycoporsicum L. Solanum nigrescens Mart. & Gal. Solanum rostratum Dunal Styrax glabrescens Benth. var. glabrescens Talinum lineare Kunth Taxus globosa Schltdl. Tectaria heracleifolia (Willd.) Underw

Appendix 2 (continued)



– – 1 – – –

– 1



1-









1

1





1



1







1

1

1

1









1

1



1



1





1



1





1

1

1

1







1



1















1





1



1

1



1



1









1















1

1









1



















1

























1











1













1

























1





1



















1





















1































































































































































































1







1







1









1



1

1

1

1



1





1



























Natural vegetation Agrosystems TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP HG IG AH 1 – 1 1 1 1 – 1 – 1 1 – 1 – – – – – – – 1 –













1

























1













M AF – –

394 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Thelypteridaceae Thelypteris pilosa Mart. & Gal.) Crawford Thelypteridaceae Thelypteris pilosula (Klotzsch & H. Karst. ex Mett.) R.M. Tryon Thelypteridaceae Thelypteris puberula (Baker) C.V. Morton var. puberula Tropaeolaceae Tropaeolum majus L. Urticaceae Pilea microphylla Rusby Urticaceae Urtica chamaedryoides Pursh Verbenaceae Aloysia gratissima (Gillies & Hook.) Tronc. Verbenaceae Bouchea prismatica (L.) Kuntze Verbenaceae Citharexylum berlandieri B.L. Rob. Verbenaceae Citharexylum caudatum L. Verbenaceae Duranta erecta L. Verbenaceae Glandularia bipinnatifida (Nutt.) Nutt. Verbenaceae Glandularia elegans (Kunth) Umber Verbenaceae Lantana camara L. Verbenaceae Lippia graveolens Kunth Verbenaceae Lippia queretarensis Kunth Verbenaceae Petrea volubilis L. – 1 – – 1

– – – – 1

– – – – – 1 – –

– 1







1



– –

1

1 1





– 1

1

1

1

1



1 1

1-

1 1

1

1

1

1



1



– –













1

1





– 1



– –





– –

1

– –





– –

1





– –







1

1 –



– –







1







1







1 –

1

– –

1









– –

1

1

1





– –



– –





1





– –











– –

1

– –











– –



1

1





– –



– –











– –



1







– –



– –











– –











– –



– –



1

1





– –







1



– –



1 –









1

– –











– –



– –







1



– –

1

1

1





– –



– –











– –











– –



– –











– –





1





– –



– –











– –





1





– –



– –











– –









1

– –



– –











– –











– –



– –











– –





1

1



1 –

1

1 –







1



1 1











– –

1

– –











1 –











– –



– –











– –











– –



– –











– –







(continued)





– –



– –











– –







Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico 395

Zygophyllaceae

Zygophyllaceae

Zamiaceae Zamiaceae Costaceae

Zamiaceae

Zamiaceae

Vitaceae

Vitaceae Vitaceae

Verbenaceae

Family Verbenaceae Verbenaceae

Scientific name Verbena carolina L. Verbena litoralis Kunth Verbena menthifolia Benth. Cissus tiliacea Kunth Cissus verticillata (L.) Nicolson & C.E. Jarvis Vitis tiliifolia Humb. & Bonpl. ex Schult. Ceratozamia hildae G.P. Landry & M.C. Wilson Ceratozamia microstrobila Vovides & J.D. Rees Dioon edule Lindl. Zamia fischeri Miq. Costus pulverulentus C. Presl Larrea tridentata (DC.) Coville Morkillia mexicana (DC.) Rose & Painter

Appendix 2 (continued)

1 – –





1 – –



– 1



1 – –





1



1







– – –

– – – –















– –

– –

1 – –







1

1 1

1

1

1 1

1





– – –





1

– –

1





– 1 1



1

1

– 1







– – –







– –







– – –





1

– –

1





1 – –

1





– –







– – –







– –



1

1

– – –







– –







1 – –





1

– 1







– – –







– –







– – –







– –







– – –

1





– –







– – –







– –







– – –







– –







– – –







– –







– – –







– –



Natural vegetation TDF OF PS TS CS SV CF RS PF POF G DS TsDF RF SS AF JF PPF AV CyP – 1 – – – – – – 1 – – – – – – 1 – 1 – – 1 – – – – – – – – – – – – – – – – – – –





– – 1







– –







– – –







– –

1





– – –







– –



Agrosystems HG IG AH – – – – – –





– – –







– –







– – –







– –



M AF – – – –

396 L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico

397

References Aguilar-Rincón VH, Corona Torres T, López López P, Latournerie Moreno L, Ramírez Meraz M, Villalón Mendoza H, Aguilar Castillo JA. Los chiles de México y su distribución. México: Sistema Nacional de Recursos Fitogenéticos para la Alimentación y la Agricultura, Colegio de Postgraduados, Instituto Nacional de Investigaciones Forestales, Agrícolas y Pecuarias, Instituto Tecnológico-Conkal, Universidad Autónoma de Nuevo León, Universidad Autónoma de Nayarit; 2010. Alcorn JB. Indigenous agroforestry systems in the Latin American tropics. In: Altieri MA, Hecht SB, editors. Agroecology and small farm development. Boca Raton: CRC Press; 1990. Bayona A, Rivera Sánchez E, Valtierra JG, Martínez Reséndiz WE, Montoya Martínez A. Caracterización de los Ecosistemas, Cambios en el Uso del Suelo y Unidades Paisajísticas en la Reserva de la Biosfera “Sierra Gorda” de Querétaro. Centro Queretano de Recursos Naturales Tomo XII. Consejo de Ciencia y Tecnología del Estado de Querétaro; 2006. Boege E. El patrimonio biocultural de los pueblos indígenas de México. México: INAH/Comisión Nacional para el Desarrollo de los Pueblos Indígenas; 2008. Carándula M, Ávila LE, coordinators. Patrimonio Biocultural, Territorio y Sociedades Afroindoamericanas en Movimiento. Argentina: Consejo Latinoamericano de Ciencias Sociales; 2013. Casas A, Torres-Guevara J, Parra F. Domesticación en el continente americano, Manejo de biodiversidad y evolución dirigida por las culturas del Nuevo Mundo, Volumen 1. México/ Perú: Universidad Nacional Autónoma de México - Universidad Nacional Agraria La Molina; 2016. Castro D, Bye RA, Mera LM. Diagnóstico del pápaloquelite en México Porophyllum ruderale (Jacq.) Cass. var. macrocephalum (DC.) Cronq. México: Universidad Autónoma Chapingo; 2011. Castro-González V. Monografía de la raíz de oro, chilcuague, Heliopsis longipes A. Gray. México: Tesis del diplomado: Medicina tradicional de México y plantas medicinales, TlahuiEdu A.C.; 2009. Chemin-Bässler H. Los pames septentrionales de San Luis Potosí. México: Instituto Nacional Indigenista; 1984. Christenhusz JM, Xian-Chun Z, Schneider H. A linear sequence of extant families and genera of lycophytes and ferns. Phytotaxa. 2011;19:7–54. Cilia-López VG, Aguirre-Rivera JR, Reyes-Agüero JA, Juárez-Flores BI. Etnobotánica de Heliopsis longipes (Asteraceae: Heliantheae). Bol Soc Bot. 2008;83:81–7. Cruz A, Pita A, Rodríguez B. Jocotes, jobos, abales o ciruelas mexicanas. Mexico: Universidad Autónoma Chapingo; 2012. De Ávila A. La diversidad lingüística y el conocimiento etnobiológico. In: Soberón J, Halffter G, Llorente-Bousquets J, compiladores. Capital natural de México, vol. I: Conocimiento actual de la biodiversidad. México: Comisión nacional para el conocimiento y uso de la biodiversidad; 2008. De la Garma E. Museo Histórico de la Sierra Gorda, retrato documental del patrimonio cultural serrano. Querétaro: Museo Histórico de la Sierra Gorda; 2015. DOF (Diario Oficial de la Federación). Catálogo de las Lenguas Indígenas Nacionales: Variantes Lingüísticas de México con sus auto denominaciones y referencias estadísticas. México; 2008. Espinoza-García FJ, Villaseñor JL. Biodiversity, distribution, ecology, and management of non-native weeds in México: a review. Rev Mex Biodivers. 2017;88:76–96. Espitia-Rangel E, Mapes Sánchez C, Escobedo López D, De la O Olán M, Rivas Valencia P, Martínez Trejo G, Cortés Espinoza L, Hernández Casillas JM. Conservación y uso de los recursos genéticos de Amaranto en México. Instituto Nacional de Investigaciones Forestales, Agrícolas y Pecuarias, Centro de Investigación Regional Centro, Celaya, Guanajuato, México; 2010. Fernández-Nava R, Ramos D, Carranza E. Notas sobre plantas medicinales del estado de Querétaro. Polibotánica. 2001;12:1–39.

398

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Gallardo-Arias P. Los pames coloniales: un grupo de fronteras. México: Centro de Investigaciones y Estudios Superiores en Antropología; 2011. Gorenflo LJ, Romaine S, Mittermeier RA, Walker-Painemilla K. Co-occurrence of linguistic and biological diversity in biodiversity hotspots and high biodiversity wilderness areas. PNAS. 2012;109:8032–7. Hernández E. Exploración etnobotánica y su metodología. In: Xolocotzia, Obras de Efraím Hernández Xolocotzi. Revista de Geografía Agrícola. México: Universidad Autónoma Chapingo; 1985. Hernández-Ruiz JR, Juárez-García A, Hernández-Ruiz N, Hernández-Silva N. Uso antropocéntrico de especies vegetales en los solares de San Pedro Ixtlahuaca, Oaxaca México. Ra Ximhai. 2013;9(1):99–108. Herrera-Muñoz A. Los enfoques básicos en la Arqueología de la Sierra Gorda, un balance desde sus dimensiones principales. En: Nieto Ramírez J (coordinador). Sierra Gorda de Querétaro, la tierra y el hombre. Ediciones Universidad Autónoma de Querétaro – Ediciones Universitarias Maristas; 2010. Hua X, Greenhill SJ, Cardillo M, Schneemann H, Bromham H. The ecological drivers of variation in global language diversity. Nat Commun. 2019;10:2047. INE (Instituto Nacional de Ecología). Programa de manejo Reserva de la Biosfera Sierra Gorda, México. a Secretaría de Medio Ambiente, Recursos Naturales y Pesca. México; 1999. INEGI (Instituto Nacional de Estadística, Geografía e Informática). Censo general de población y vivienda. México; 2020. https://www.inegi.org.mx/programas/ccpv/2020/default.html. Accessed 28 February 2022. Lira R, Casas A. Usos y manejo en Ibervillea millspaughii (Cogn.) C. Jeffrey, Melothria pendula L. y otras especies silvestres de la familia Cucurbitaceae: Posibles procesos de domesticación incipiente. Bol Soc Bot. 1998;62:77–89. Lobato-Ortiz R, Rodríguez-Guzmán E, Carrillo-Rodríguez JC, Chávez-Servia JL, Sánchez-Peña P, Aguilar Meléndez A. Exploración, colecta y conservación de recursos genéticos de jitomate: avances en la Red de Jitomate. Sistema Nacional de Recursos Fitogenéticos para la Alimentación y la Agricultura, Secretaría de Agricultura, Ganadería, Desarrollo Rural, Pesca y Alimentación y Colegio de Postgraduados. Texcoco, México; 2012. López C, Chanfón S, Segura G, editors. La riqueza de los bosques mexicanos: Más allá de la madera. Experiencias en comunidades rurales. Secretaría de Medio Ambiente y Recursos Naturales - Center for International Forestry Research. México; 2005. López-Ugalde R, Vázquez Estrada A, Prieto Hernández D. Los indios de Querétaro, Perfil socio demográfico de la presencia india en la entidad. In: Vázquez Estrada A, Prieto Hernández D, coordinadores. Los pueblos indígenas del estado de Querétaro: compendio monográfico. México: Comisión Nacional para el Desarrollo de los Pueblos Indígenas, Delegación Estatal Querétaro – Universidad Autónoma de Querétaro; 2013. Lozada-Aranda M, Rojas Barrera I, Mastretta Yanes A, Ponce-Mendoza A, Burgeff C, Orjuela-R MA, Oliveros Galindo O. La Milpa. Oikos. 2017;17:10–2. Magallán-Hernández F, Hernández Sandoval L. La familia Agavaceae en el estado de Querétaro. Bol Soc Bot Méx. 2000;66:103–12. Márquez Sánchez F. Agroecosistemas de México: contribuciones a la enseñanza, investigación y divulgación agrícola. E. Hernández-Xolocotzi ed. 2ª ed. Colegio de Postgraduados Chapingo, México; 1976. Martínez HL. Etnobotánica del chile piquín (Capsicum annuum var. glabriusculum) en la Sierra Gorda y el Semidesierto Queretano. Tesis de Maestría en Ciencias. Colegio de Postgraduados, Montecillo; 2007. Martínez M, Hernández-Sandoval L, Pantoja Hernández Y, Gómez Sánchez M, Bárcenas R, Cabrera A. Guía ilustrada de la flora del valle de Querétaro. Editorial Universitaria. Universidad Autónoma de Querétaro – Comisión nacional para el conocimiento y uso de la biodiversidad. Querétaro, Qro; 2017.

Ethnobotanical Knowledge Within the Sierra Gorda, Quere´taro, Mexico

399

Martínez-Spinoso MY, Miranda K. Etnobotánica de los Xi’úi de la Sierra Gorda de Querétaro. In: Curanderos y Medicina Tradicional en la Huasteca, P. Gallardo Arias Coordinadora. Ediciones del programa de desarrollo cultural de la huasteca. México; 2008. Meade J. La Huasteca Queretana. Memorias de la Academia Mexicana de la Historia. 1951;X 4. Mejía-Pérez Campos E, Herrera Muñoz AJ. Mines and miners: metals in sediments and human remains from southern Sierra Gorda, Querétaro, México. Chungara Rev Antrop Chil. 2013;45(1):161–76. Mendoza-Rico M, Ferro Vidal LE, Solorio Santiago E. Otomíes del semidesierto queretano, Pueblos indígenas del México contemporáneo. México: Comisión Nacional para el Desarrollo de los Pueblos Indígenas; 2006. Mera-Ovando LM, Bye BR, Villanueva VC, Castro LD, Basurto PF. El cultivo de la verdolaga (Portulaca oleraceae L.) ejemplo en la promoción, producción y el comercio de alimentos sanos y de calidad. In: Mera O LM, Castro L D, Bye B RA, editors. Especies Vegetales, Poco Valoradas: Una Alternativa para la Seguridad Alimentaria. Instituto de Biología, Universidad Nacional Autónoma de México. México; 2013. Miranda-Perkins K. Estudio etnobotánico de las comunidades Pame (Xi'ui): Las Nuevas Flores, Las Flores y El Rincón del estado de Querétaro. Querétaro: Tesis de Licenciatura en Biología, Universidad Autónoma de Querétaro; 2003. Miranda-Perkins K, Castillo Gómez HA. La medicina tradicional en los pueblos xi’oi de la Sierra Gorda. En: Vázquez Estrada A (coordinador). Xi’oi, los verdaderos hombres. Atlas etnográfico, pames de la Sierra Gorda Queretana. Universidad Autónoma de Querétaro. Querétaro, México; 2010. MOBOT, Missouri Botanical Garden; 2022. www.tropicos.org. Moreno Calles AI, Casas A, Toledo VM, Vallejo Ramos M. Etnoagroforestería en México. Universidad Nacional Autónoma de México; 2017. Moreno-Calles AI, Toledo VM, Casas A. Los sistemas agroforestales tradicionales de México: una aproximación biocultural. Bot Sci. 2013;91(4):375–98. Nieto Ramírez J, coordinador. Sierra Gorda de Querétaro, la tierra y el hombre. Querétaro: Ediciones UAQ, Ediciones Universitarias Maristas; 2010. Núñez-López RA. Fitonimia hñäñho: una aproximación a la etnotaxonomía de la flora útil del pueblo ñäñho de Amealco, Querétaro. Tesis de Maestría. Querétaro: Universidad Autónoma de Querétaro; 2014. Ordoñez G. Pames. Pueblos Indígenas del México Contemporáneo. México: Comisión Nacional para el Desarrollo de los Pueblos Indígenas; 2004. Pérez-Nicolás ML, Fernández-Nava R. Plantas del estado de Querétaro, México, con potencial para uso ornamental. Polibotánica. 2007;24:83–115. Santiaguillo JF, Cedillo E, Cuevas JA. Distribución Geográfica de Physalis spp. en México. Mexico: Universidad Autónoma Chapingo; 2010. Scheinvar L. Flora cactológica del estado de Querétaro: diversidad y riqueza. Mexico: Fondo de Cultura Económica; 2004. Smith J, Döring T, Pearce BD, Wolfe MS. Agriculture and the environment IX, valuing ecosystems: policy, economic and management Interactions. 2012;257–262. Stevens PF. In: Angiosperm phylogeny website. Versión 14, july 2017. http://www.mobot.org/ MOBOT/research/APweb/. Stresser-Péan G. Viaje a la Huasteca. Mexico: Fondo de Cultura Económica, Centro de Estudios Mexicanos; 2008. Suárez G, Serrano-Cárdenas V, Balderas P, Pelz R. Atlas de malezas arvenses del estado de Querétaro. Querétaro: Universidad Autónoma de Querétaro, Comisión nacional para el conocimiento y uso de la biodiversidad, Instituto Tecnológico de Estudios Superiores de Monterrey-Querétaro, Departamento de Agricultura y Tecnología de Alimentos-Instituto Nacional Politécnico; 2004. Sutherland WJ. Parallel extinction risk and global distribution of languages and species. Nature. 2003;423:276–9.

400

L. Herna´ndez-Sandoval and H. Castillo-Go´mez

Thiers B. [continuously updated] Index Herbariorum: a global directory of public herbaria and associated staff. New York Botanical Garden’s Virtual Herbarium. 2022. Available at http:// sweetgum.nybg.org/ih/. Accessed 28 Feb 2022. Toledo VM, Barrera-Bassols N. La memoria biocultural. La importancia ecológica de las sabidurías tradicionales. México: Icaria editorial; 2008. Valle-Esquivel J, Prieto Hernández G, Utrilla Sarmiento B, Coordinadores. Los pueblos indígenas de la Huasteca y el Semidesierto Queretano: Atlas etnográfico. Mexico: Instituto Nacional de Antropología e Historia, Instituto Nacional de Lenguas Indígenas, Universidad Autónoma de Querétaro, Instituto Queretano de la Cultura y las Artes; 2012. Vázquez-Estrada A, coordinador. Xi’oi los Verdaderos Hombres, Atlas etnogŕafico, Pames de la Sierra Gorda Queretana. Querétaro: UAQ; 2010. Vázquez-Estrada A, Prieto-Hernández D, coordinadores. Los pueblos indígenas del estado de Querétaro: compendio monográfico. México: Comisión Nacional para el Desarrollo de los Pueblos Indígenas, Delegación Estatal Querétaro – Universidad Autónoma de Querétaro; 2013. Villaseñor JL. Checklist of the native vascular plants of México. Rev Mex Biodivers. 2016;87(3): 559–902. Viramontes Anzures C. De chichimecas, pames y jonaces, los recolectores cazadores del semidesierto de Querétaro. Serie Arqueología, Instituto Nacional de Antropología e Historia; 2000. Viramontes Anzures C. Los antiguos pobladores de la Sierra Gorda. In: Vázquez Estrada A (coordinador). Xi’oi Los Verdaderos Hombres, Atlas etnográfico, Pames de la Sierra Gorda Queretana. Querétaro: Universidad Autónoma de Querétaro; 2010. Zamudio S. Consideraciones acerca de la flora y vegetación de la Reserva de la Biósfera Sierra Gorda, en el Estado de Querétaro, México. Querétaro: Memorias del Primer encuentro de Investigación sobre la Sierra Gorda; 2003. Zamudio S, Rzedowski J, Carranza E, Calderón G. La vegetación del estado de Querétaro. Consejo de Ciencia y Tecnología del Estado de Querétaro – Instituto de Ecología A.C. Querétaro: Centro Regional del Bajío; 1992.

Ethnobotany of the Sierra Norte de Puebla Francisco Basurto, Cristina Mapes, Tania Escobar, and Juan Carlos Gonza´lez

Abstract

The Sierra Norte de Puebla is an environmentally, biologically, and culturally diverse region comprising 60 municipalities in the north of the state of Puebla. Vegetation in the lowlands is the tropical evergreen forest and tropical oak forest established on particular edaphic conditions like sandy soils with fast runoff. Between 700 to 1800 m of elevation, cloud mountain forests are found with several plant associations. In the highlands, conifer forests mainly of pines (Pinus spp.), and temperate oak (Quercus spp.) forests grow. Fagus forests were registered in the municipality of Teziutlán. These types of vegetation do not have sharp boundaries among each other, rather presenting broad ecotone zones between them. In this region humans have been present since over 4000 years and it is currently inhabited by more than one million people, one third of which are indigenous. The original inhabitants of the region were probably of Totonac and Tepehua identities, but several Pre-Columbian migrations from central and western Mexico have been recorded, mainly of Nahua and Otomi or Yuhu people, which settled in the region in different periods. In the past four decades, ethnobotanical research has been conducted in the region focused on documenting the local knowledge and management of useful flora, especially non-timber forest products, traditional agricultural systems, and phytogenetic resources, always with the community members’ consent and collaboration. A summary of the results obtained along that time period are presented here. We report and discuss the observed changes in the use and management of plant resources and the

F. Basurto (*) · C. Mapes · T. Escobar · J. C. González Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Coyoacán, Ciudad de México, Mexico e-mail: [email protected]; [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_13

401

402

F. Basurto et al.

composition of the traditional agricultural systems to cope with new economic, social, and cultural paradigms. We analyze their causes, and consequences for the preservation of the biocultural heritage, and their contribution to the conservation of germplasm and sustainability.

a la memoria y a la obra del Profesor Miguel Ángel Martínez Alfaro maestro y amigo generoso pionero y pilar de la etnobotánica en México

Introduction: The Territory and the People The Sierra Norte de Puebla (SNP) region is at present inhabited by a population of nearly one million people, over 30% of which are in indigenous communities that form part of 60 municipalities in the north of the state of Puebla. The region has an ample diversity of environmental, biological, and cultural components (Anónimo 1987; Fuentes 1972; Martínez 2007; Masferrer and Martínez 2010). It extends over 70 km in a north-south direction and 60 km from east to west, sloping down over 2000 m in elevation in three physiographic regions. On of them around 2300 m in the summit of the Cerro Cozolépetl in the Sierra Madre Oriental, another through the middle elevations of the Trans-Mexican Volcanic Belt, and the thrid one about elevetions 100 m in the coastal lowlands of the Gulf of Mexico Coastal Plain (Fig. 1). In the Gulf of Mexico Coastal Plain and the Trans-Mexican Volcanic Belt, the topography is of smooth-sloped hills and plains, and in the Sierra Madre Oriental, it is abrupt with steep slopes and numerous ravines and canyons. Regarding the landforms, the Sierra Madre Oriental includes a high mountain range with mature erosion processes that become rejuvenated toward the NE by runoff from the highlands in the Mexican Plateau, a low mountain range, and intermontane valleys (Anónimo 1970; Fuentes 1972; Saldaña 2011). Climates in the region are distributed according to the elevation range, going from warm and humid subtemperate in the lowlands to humid temperate, all of them showing year-round rainfall. The mean annual temperature in the higher elevations is of 12–15  C and in the lower zones it ranges from 18–24  C (Fig. 2). Located in the intertropical zone, the SNP region receives the direct influence of tropical easterlies carrying humidity from the Gulf of Mexico. This fact causes strong summer precipitations that add to the humidity delivered by the tropical cyclones that impact the region – mostly from the late summer to the early fall. During the winter, the cold north winds blowing from the Arctic regions in the continent, known as Norte, pick up humidity from the Gulf of Mexico and deliver it on the outer watershed of the SNP. All these phenomena make this one of the zones with higher precipitation in Mexico, with a mean annual precipitation of up to 4500 mm recorded in the Cuetzalan-Hueytamalco region. The rivers crossing the SNP region have their

Ethnobotany of the Sierra Norte de Puebla

403

Fig. 1 The Sierra Norte of Puebla Region

headwaters in the highlands of the states of Hidalgo and Tlaxcala and their mouths in the Gulf of Mexico, forming the Nautla River, Tecolutla River, Cazones River, and Tuxpan River basins (Anónimo 1970; Arrequín 2011). The vegetation also corresponds with the elevation gradient and climates (Fig. 3), so that the potential climax vegetation in the lowlands is the tropical evergreen forest, but in particular edaphic conditions, like those with sandy soils with fast runoff, tropical oak (Quercus) forests are established. Between 700 to 1800 m of elevation, cloud mountain forests are found with several plant associations, and in the highlands, conifer forests mainly of pines (Pinus spp.), and temperate oak (Quercus spp.) forests grow. In addition, Fagus forests exist in the municipality of

404

F. Basurto et al.

Fig. 2 The climates of the Sierra Norte of Puebla

Teziutlán. These types of vegetation do not have sharp boundaries among each other, rather presenting broad ecotone zones between them (Basurto 1982, 2000; Castro 1988; Evangelista and Mendoza 1987; Fox and Sharp 1954; López and Cházaro 1995; Martínez et al. 2000; Miranda and Hernández-X 1963; Miranda and Sharp 1950; Paray 1946; Puig 1991; Rzedowski 1978; Vázquez 1990). In general, the inhabitants of the SNP region recognize three zones defined by elevation and climate: The Bocasierra zone, located between the high plateau and the slopes toward the Gulf of Mexico; the Tierra Fría or temperate zone, at

Ethnobotany of the Sierra Norte de Puebla

405

Fig. 3 Vegetation types of the Sierra Norte of Puebla

elevations from 900 to 2000 m, and the warm Tierra Caliente zone, below the 900 m a.s.l. The characteristics of the relief define different meso and microclimatic conditions in relatively close areas, which allows ecological complementarity in which a farmer can crop in different agroclimatic zones using an assortment combination of plants. It also allows the access to plant products from different climatic zones, thus broadening the range of plant resources available to the population.

406

F. Basurto et al.

The archaeological recors show that at least since the Formative Horizon (2500 BCE–200 CE), the SNP region was populated by groups of Huastec (at present living in the states of San Luis Potosí and Veracruz), Totonac, Tepehua, Nahua, and Otomi people, the latter of which continue inhabiting in the region. After the Spanish invasion, the indigenous population was accompanied by American-European mestizo population (Diez 2017, 2019; Faulhaber 1994; García 1987; Ochoa 1984). At least since the Classic Period (300–900 CE), the SNP region has supported relatively dense populations, as it is shown by the large number of archaeological sites in the region. This is for instance the case of Yohualichan, which was contemporaneous with Tajin in the state of Veracruz (García 1987). The original inhabitants of the region were probably of Totonac and Tepehua identities, but several Pre-Columbian migrations from central and western Mexico are recorded, mainly of Nahua and Otomi or Yuhu people, which settled in the region in different periods (García 1987). Historically, the SNP region was linked to the ruling centers on the highlands and the Gulf of Mexico (Teotihuacan, Tula, El Tajin), and there is evidence of the existence, since Teotihuacan times, of commercial exchange between the Gulf of Mexico and the highlands through several trade routes crossing the SNP region (García 1987). The long period of occupation in the region has left a deep imprint on the landscape, which is at present observed as a mosaic where relicts of primary vegetation alternate with acahuales (secondary vegetation) in several stages of succession, and cropland planted with milpa (maize-bean-squash association), cafetales (coffee plantations), pastures, home gardens, and many other minor cropland types like frijolares (land used for cultivating bean), chilares (for cultivating chili), platanares (for cultivating banana), peanut, sugarcane, vanilla, broad bean, or sesame. Also, in the lowlands of the SNP region in the municipalities of F. Z. Mena, Ayotoxco, and Tenampulco, there is at present an extensive citrus cultivation zone (Martínez 2007). From the economic and social perspective, most of the municipalities in the SNP region are considered of high and very high marginality index (CONAPO 2015; Pérez and Romero 2011). The main economic activity is subsistence agriculture and the main crops are maize (Zea mays), bean (Phaseolus spp.), squash (Cucurbita spp.), coffee (Coffea arabica), as well as temperate (prune Prunus domestica), pear Pyrus communis, apple Malus domestica, peach Prunus persica, avocado Persea americana) and tropical fruit trees (mamey Pouteria sapota, black sapote Diospyros nigra, and banana Musa spp.). Maize, beans and squash are part of the basic diet of the population, while the latter crops are harvested for the market. The predominant land tenure form in the region is smallholding. Because of historical reasons (mainly the mid-nineteenth-century laws of confiscation), a process of privatization and parceling of the land took place. Because of this fact smallholding has been the dominant land tenure modality for over one and a half centuries, which has a strong influence on land use and the management of natural plant resources, mainly seen in the land cover-land use conversions in agroecosystems; a process which can be intensely dynamic (González 2014). Another factor influencing the forms of management of plant resources is the differential development observed between the western and eastern parts of the SNP region. In the western portion, the Mexico City-Tuxpan-Tampico highway – enabled

Ethnobotany of the Sierra Norte de Puebla

407

since the 1950s – and the building of the Necaxa hydroelectric dam, acted in their time, and continue acting at present, as development poles. In contrast, in the eastern portion roads were built much later and the only communication available until the 1990s was the Amozoc-Teziutlán-Martínez de la Torre highway with a branch to Zacapoaxtla and Cuetzalan. Until the present, the building of roads has been poorly developed in the central portion of the SNP region. The influence of the abovementioned highways is evident when relating their traces with the degree of marginality in municipalities (Masferrer 2006).

Ethnobotanical Exploration in the Sierra Norte de Puebla Ethnobotanical research in the SNP region started in the 1970’s mainly focused on documenting the useful flora, agricultural systems, traditional agriculture, and non-timber forest products (NTFP) (Basurto et al. 2009, 2014; Escobar 2012, 2016; Evangelista et al. 2010a; López 2008; Mapes et al. 2011, 2012, 2013; Martínez et al. 2007, 2008; Molina 2000; Peralta 2007; Vaylón 2012). Work in communities always counted with the community members’ consent through official letters to the civil authorities asking for their permission and collaboration. Also, efforts were always made to return the results from research to the communities through presentations during the celebration of the town’s patron saint fairs, talks to local elementary and high school students about the relevance of plant resources, and establishing links with producer associations. More recently, students from educational institutions interested in several research projects have been involved in the region. Among the more significant activity of results return were the presentations delivered to three communities in the SNP region titled De quelites me como un taco (Mera et al. 2003, 2005), the title in Spanish making reference to a common saying that compares the ease of eating tacos with learning about quelites (edible greens), for which didactic material was elaborated including posters (Mera et al. 2003), a quelite recipes compilation (Castro et al. 2005), and memory and Mexican bingo (in Spanish, lotería) board games (Basurto et al. 2005). Another similar contribution was the creation of a communitarian squash seed bank in Zoatecpan, municipality of Xochitlán de V. Suárez as part of the work made by the Red Calabaza (Squash Network) of the National System of Phytogenetic Resources for Food and Agriculture (SINAREFI in its Spanish abbreviation), which has been functioning from 2012 to the present. As a strategy of linkage with the community and renewal of the genetic material, we contributed to implement, four contests of diversity and gastronomy of native squashes, and a raffle of agricultural inputs. In these contests, squashes were processed to obtain the seeds and deposited as accessions of the seed bank, which are also available for the community.

408

F. Basurto et al.

The Useful Flora and Non-timber Forest Products The inventory of the useful flora of the SNP has been continuously growing; in 1995, the Catalogue of Useful Plants of the SNP (Martínez et al. 1995) listed 603 species. By the year 2011, 750 useful plants were counted and at present, 900 useful plant species have been recorded (Evangelista et al. 2011), grouped into 13 anthropocentric categories and into 17 subcategories (Table 1). While the total number of useful species has increased, the percentages by category have maintained similar values, the first places corresponding to the medicinal, food, and ornamental uses. Also, 30% of the recorded plants have more than one use. Among the medicinal plants, there are species used for diverse ailments of the human body’s apparatuses and systems, and for all the categories posed by Aguilar et al. (1994) and López et al. (2017) for the recorded pathologies. The inhabitants of the SNP make ample use of medicinal plants, but also commercialize them, mainly in the municipalities of Pahuatlán and Tlaxco, where medicinal plants are gathered to be traded in the national market (Zurita 2004). Food plants include basic, secondary, and seasonal foodstuffs. Most of these species are in the secondary and seasonal food categories, mostly fruits, with Table 1 Anthropocentric categories of plants in Sierra Norte of Puebla Category Medicine Edible

Ornamental Firewood Agricultural use

Timber and building materials Households utensils and handicrafts Working tool Ritual Living fences Forage Wrap food Industrial uses

Subcategory Food Beverage Spices

Shade Soil cover Trellis Timber Building material Households utensils Handicrafts

Dye Soap Glue Gums and resins Oil Tanning

No. species 390 235 13 32 192 90 70 5 5 13 48 25 20 15 35 35 29 10 11 4 2 2 1 1

Ethnobotany of the Sierra Norte de Puebla

409

107 species (Martínez 2007) and quelites with nearly 100 species (Basurto et al. 1998, 2003). Also included in this category are plants used to prepare beverages and flavoring and aromatic plants. Among the quelites, the quintoniles (Amaranthus spp.) are notable because of their high foliage production, palatability, and management characteristics, which is evidence of a prolonged period of human selection in the region (Mapes et al. 2012). Besides being numerous, the quelites are amply and commonly used, the recorded consumption being of every 2–3 days per week. Many of the species of food plants are of current interest due to their potential as nutraceuticals (Morales 2016). The ornamental category is third in number of species, and although most are used as home ornaments, an important number of live plants and foliage are traded in the SNP, mostly in the Huauchinango-Xicotepec region (Evangelista et al. 2011). In recent times, in the tianguis (local open-air markets) in the SNP numerous orchids and bromeliads are sold, mostly obtained by extraction from natural populations, which together with habitat reduction, threatens them. The way in which this commerce of orchids and bromeliads is currently carried out, is a direct consequence of the social and economic marginality in which most of the population lives, people selling these plants to earn income. The plants used for firewood (Martínez 1992) and as coffee shade are also numerous (Cruz 2004). Firewood continues being a main source of energy in the SNP region and the coffee plantations provide a significant part of the required firewood supplies. Although few, the ceremonial plants have high cultural importance, notably among which are the sempoalxochitl (Tagetes erecta), chamaqui (Heliconia bihai), tepejilote (Chamaedorea spp.), and tepecintle (Ceratozamia mexicana, C. fuscoviridis, and the recently described C. totonacorum) (Escobar 2016; Martínez et al. 2017). The cucharilla (Dasylirion acrotrichum) is also a relevant ceremonial plant used to make the ornaments for the front of churches during local patron site fairs. This plant grows in arid zones and is absent in the SNP region, but it is brought from Tepeyahualco, in the state of Puebla, where its sustainable use, therefore, involves its use in the SNP region (Torres 2016). A task pending to be completed is the study of the melliferous flora in the SNP region, used both by the European bees (Apis mellifera) and by the stingless native bees locally called pisilnekmej (Scaptotrigona mexicana) that are currently gaining relevance as a resource contributing to the monetary income of the population (Guzmán 2018; Sánchez et al. 2016).

Non-timber Forest Products (NTFPs) The NTFPs in the SNP include an ample spectrum of plant products, among them spices (Martínez 2009; Martínez et al. 2004), fruits (Escobar 2012; Martínez 2007), latex (Vaylón 2012), food plants (Basurto et al. 2003; Evangelista and Basurto 2003; Evangelista et al. 2010a; Mapes et al. 1997), medicinal plants (Zurita 2004),

410

F. Basurto et al.

ceremonial plants (Escobar 2016), and plant species used for making handcrafts (Eleuterio and Perez 2006; López 2004). The ways of local use and management of NTFPs derive from traditional knowledge that at present has been modified or adapted to the new conditions of demand and commercialization by having access to new markets. This access is possibly due to the action and initiative of producers themselves, who take advantage of the recently opened roads and highways. The PFNMs contribute, in higher or lower degree, to the economy of the producers and in cases like the pimienta (allspice; Pimenta dioica) (Martínez et al. 2004) or amate paper (López 2004), they provide a large percentage of their total income. However, the commercial exploitation of some of these resources in the SNP region involves social-environmental and economic issues that might compromise their sustainability (Eleuterio and Perez 2006; López 2004; Vaylón 2012). The use of NTFPs as natural resources has been posed to be a way to contribute to the wellbeing of rural communities and the sustainability of forested areas (Alexiades and Shanley 2004). But the numerous factors and processes involved in the value chains of NTFP in the SNP make the number of links between the producers and consumers greater, thus, more asymmetries exist, and the producers or gatherers of these resources receive the lower percentage of the benefit. Some of the NTFPs from the SNP region reach international markets, as does the pimienta (Martínez et al. 2004), others are commercialized at the national or regional markets like medicinal plants, mamey, zapote negro (black sapote) (Escobar 2012), and other products that are sold locally in the regional markets or door-to-door. The latter represent the shortest value chains, even without intermediaries, because the products are commercialized by the producers or gatherers themselves, as in the case of the xocoyoli (Begonia spp.), tequelite (Peperomia peltata), tepezintle, and tepejilote (Chamaedorea tepejilote). The latter cases are of high interest because several of them are currently being introduced into cultivation and are no longer only gathered in forests. The management of these NTFPs agrees with that described for other similar resources (Ruiz-Pérez et al. 2004) because, due to their demand, the gatherers are beginning to cultivate the xocoyoli or the tequelite, while tepejilote is now favored in homegardens and cafetales.

Agricultural Systems and Traditional Agriculture Agriculture is practiced in the SNP mostly according to what Hernández-Xolocotzi (1988) defines as traditional agriculture technique. That is, that whose comprehensive study can only be understood from three axes involving socio economic constraints, ecological limitations, and resulting technologies (Arias and Ortega 2013). That is characterized as derived from the empirical knowledge of agricultural practices developed through trial and error methods, which is transmitted from fathers to sons, by having a low energy input and the quality of the energy used in the agroecosystem, and by including physical and metaphysical components determined by the worldview of the farmers (Hernández 1988).

Ethnobotany of the Sierra Norte de Puebla

411

Fig. 4 Agroecosystems of the Sierra Norte of Puebla

In general, the agroecosystems in the SNP region are polyculture, which is practiced using domesticated and cultivated plants together with a complex of wild to domesticated plant species at different stages of management, including gathered, tolerated, favored, and cultivated species (Mapes and Basurto 2016). These agroecosystems, as the natural vegetation, are distributed along the different elevation levels, and in the case of milpas, which are planted throughout the elevation range, different combinations of species or varieties are used depending on elevation (Fig. 4). The agricultural calendar, that is, the dates in which the agricultural practices are performed, depend largely on the elevation at which cultivation takes place. Notably, cultivation might occur on high-steep slopes, sometimes with over 100% slope. As part of the inputs used in these agroecosystems, the seeds are crucial, and native seeds are used in the SNP region, which the producers keep and obtain from their own harvest.

The Milpa In the SNP region, the milpas contain maize, beans, and squashes in combinations that depend on the elevation (Fig. 4). Six races of maize are reported, although Tuxpeño predominates in four seed colors – white, yellow, red, and black – and in two types recognized by the thickness of the corncob (olote in Spanish): olopitza

412

F. Basurto et al.

(thin corncob) and kuaugzi (thick corncob). This maize race is distributed in the region at elevations from 100 to 2000 m and is cultivated in association with four bean species – common bean (Phaseolus vulgaris), ayocote (P. coccineus), exoyema (P. dumosus), and navajita (P. lunatus), and the torito or torojet bean (Vigna unguiculata) of African origin – and four species of squash – pipiana (Cucurbita argyrosperma), chilacayote (C. ficifolia), tamalayota (C. moschata), and huitzayota (C. pepo). Other races of maize reported in the SNP are Mushito, Elotes Cónicos, Arrocillo, Cacahuacintle, Olotillo, and Ratón. At elevations between 1400 and 1800 m, the maize race Tuxpeño presents genetic infiltration of the Elotes Cónicos and Arrocillo races (Evangelista 1999). The milpas in the SNP can be divided according to the fallow period into the variants described below.

Permanent Parcels (without fallow period) The non-fallow, or permanent parcel systems in the higher parts of the SNP region are cultivated in a single annual cycle beginning in January or April – depending on the region – and ending between September and December (Basurto 2000; Cruz 1995; González 2014; Vázquez 1990). In the lower parts of the Tierra Fría, it is possible to begin sowing in November or December because they are frost-free areas. In the Tierra Caliente zone, at elevations below 600 m, planting is made twice a year. The first cycle is called tonalmil or tonalmile and takes place between December-February to June-July, and the second cycle called xopalmil or xopalmile takes place between June and December. When both cycles are practiced in the same plot, they overlap, because the sowing of the xopalmil is made when maize of the tonalmil is still in the field. Sowing is made “al piquete,” in Spanish meaning by pricking the soil without turning over and using a coa or planting stick.

Milpa in Acahuales or “Coamil” (short, 2 to 5 years fallow periods) The short fallow coamiles are planted in land where natural vegetation that has been allowed to grow for 2–5 years is slashed and burned before soil preparation and sowing. These plots are used only during one or two agricultural cycles, after which they are abandoned for leaving the natural vegetation to grow again.

Milpa with Ilite (Alnus spp.) (short, 4 to 5 years fallow) The maize-Alnus system is located in the elevation range corresponding to the cloud forest, being interesting for using the biological characteristics of trees of the genus Alnus, locally known as ilite, which have a high capability for resprouting and fixing atmospheric nitrogen. The system works by allowing the ilite to grow during a

Ethnobotany of the Sierra Norte de Puebla

413

period of 4–5 years, after which it is felled and burned to prepare the land for planting the milpa. Cultivation is made during 1 year after which the land is abandoned for a monospecific ilite grove to develop from the sprouts in the stumps left when the trees were felled. After 5 years, the ilite saplings are between 10 and 20 cm in diameter at breast height and it is then when the cycle restarts. Farmers say this is the ideal time to cut them because it is when they resprout easier. Another advantage of this system is that the trunks of the ilite –purposefully allowed to grow up to 2 m in heigh t– are used as support for exoyema bean plants, of which three to four seeds are sown around each trunk, which results in a high production of this bean. No fertilizers are used, and a low incidence of pests and diseases is reported (Basurto 2000).

Slash-and-Burn Milpas (long, 20 to 25 years fallow period) The slash-and-burn system was observed in the SNP until the 1980s but is not practiced since then because there is no land available for it. In part this is because of the modality of land tenure in most of the region (smallholding), which causes the atomization of property when land inherited from fathers to sons, and also because of the increase of the population and expansion of the agricultural area. The slash-andburn system was mainly applied in both the evergreen tropical forest and the cloud forest zones. Whichever the fallow period used, the milpas have more or less uniform structure in terms of composition and temporal and spatial distribution of the associated species. Similar agricultural technology – tools and inputs – is applied in them, and the sequence of agricultural practices is also similar (Basurto 2000; Cruz 1995; Evangelista and Mendoza 1987; González 2014; Vázquez 1990). Maize is always sowed in furrows, although the distance between plants can vary from 70 to 120 cm and, thus, the density of maize may be between 40,000 and 80,000 plants/ha. This spatial distribution allows plants to receive nearly individual attention, besides making possible to access into the parcel at nearly any moment of the agricultural cycle. This fact enables the harvest of the different species at different moments of their life cycle (tender and mature corncobs; string beans, flowers, tender stems, and dry bean seeds; squash flowers, tender stems, tender squash, and mature squashes), quelites, and other edible or fodder plants associated to this agroecosystem (Basurto 2000; Basurto et al. 1999; González 2014). The species associated in the milpa have different spatial distribution patterns and regarding the time of their sowing. It can be concomitant: sown and harvested at the same time; imbricate: sown and harvested at different times; or intercalated: sown at the same time, but harvested in different dates. The agricultural tasks are performed in the following sequence: • Soil preparation, either only by turning over, or in the cases of acahuales and coamiles, including slash and burn

414

• • • • • • • •

F. Basurto et al.

Sowing and resowing separated by 15–25(30) days Weeding and hilling Harvest of bush beans (when planted) Fertilization, either during sowing, or after weeding Harvest of tender maize ears, string beans, and squash flowers and tender fruits Bending over maize stalks Harvest of mature maize Harvest of dry beans and mature squash

All the tools used in the preparation of the soil are manual and include machete, hook, and hoe. The use of machinery is impossible because of the topography, although oxen-drawn Egyptian plows and horse-drawn moldboard plows – the latter locally called “foreign plows” – are in some cases used. Sowing is made with a handspike or a coa (either with or without an iron tip). The above-described tools and inputs used in the slash-and-burn milpa have not changed for generations, at least since the late nineteenth century. After the Zacapoaxtla Plan implemented by the government in 1974 (a product of the Green Revolution), the Mexican Institute of Coffee (IMECAFE in its Spanish abbreviation) promoted the cultivation of coffee in the SNP region. The opening and improvement of roads and highways in the last decade of the twentieth century brought changes in the cultivation of the milpa. These changes influenced a trend to the disappearance of the oxen-drawn plow, the increased use of herbicides for soil preparation and weeding, and the use of chemical fertilizers, which changed the composition of the milpa through the loss of diversity in broadleaf crop plants (like beans and squashes), but above all, through the decline of spontaneous useful plants like quelites and fodder plants Table 2 (González 2014). The social organization also suffered losses like the decline of reciprocal labor retribution among farmers – known in Spanish as “mano vuelta,” today rarely practiced because, at present, nearly all labor is paid in cash.

The Frijolares The different species of beans in the SNP are sown at different elevations, P. lunatus in the warm climate in the lowlands, P coccineus in the temperate and cold climates, P dumosus in the wet and foggy cloud forest areas, and different local varieties of P vulgaris that are cultivated throughout the region’s elevation range (Fig. 4). The wild or agrestal forms of P. coccineus and P. vulgaris are present in the SNP, P. dumosus can be found as escaped from cultivation, and P. lunatus is only recorded as cultivated crop (Basurto 2000). The four species of beans of the genus Phaseolus cultivated in the SNP are closely linked to maize cultivation and up to three of them (P. vulgaris, P. coccineus, and P. dumosus) might be present in the same milpa. These bean species display high infraspecific diversity, plants differing in growth habit, flower and seed coat color, and phenology. Beans are mostly cultivated in association with maize and squash in

Ethnobotany of the Sierra Norte de Puebla

415

Table 2 Useful weeds associated to milpa in Xiloxochico, Cuetzalan. (Uses: 1 Food; 2 Forage; 3 Medicine; 4 green manure; 5 ritual; 6 Soil coverture: 7 living fence; 8 handicrafts: 9 firewood; 10 household utensils) Scientific name Amaranthus cruentus Amaranthus spinosus Amaranthus hybridus Eryngium foetidum Xanthosoma robustum Xanthosoma violaceum Xanthosoma atrovirens Bidens odorata Galinsoga parviflora Melampodium divaricatum Porophyllum ruderale Tagetes erecta Neurolema lobata Impatiens walleriana Begonia heracleifolia Rorippa nasturtium-acuaticum Busera simarouba Lobelia linearis Chenopodium ambrosioides Tinantia erecta Tripogandra serrulata Cucurbita okeechobeensis martinezii Cyclanthera dissecta Citrullus lanatus Saurauia cana Dioscorea bulbifera Cnidoscolus multilobus Ricinus communis Jatropha curcas Euphorbia lanceolata Phaseolus coccineus Phaseolus vulgaris Phaseolus dumosus Vigna unguiculata Mucuna pririens Pachyrizus erosus Mimosa albida Leucaena leucocephala Hyptis verticillata Allium glandulosum Sida acuta

Current use + + + +

+ + + + +

Past use + + + + + + + + + + + + +

+

+ + + +

+

+ + + + + + +

+ +

+ + + + + + + + + + + + + + + + + + + + + + + + + + +

Uses 1 1, 2, 3 1 1 1, 3 1 1 4, 2, 3 2, 3 2, 4, 3 1, 3 5, 3 3 6 1 1 3, 7 3 1, 3 6, 4, 2, 1 4, 6, 2, 3 8, 5 1 1 1 1, 2 3 9 1, 9 3 1 1 1 1 4, 1, 6 1 3 1 1 1 10, 3 (continued)

416

F. Basurto et al.

Table 2 (continued) Scientific name Maranta arundinacea Heliconia schiedeana Ocimum sellowi Lopezia racemosa Oxalis corniculata Oxalis latifolia Passiflora coriacea Piper auritum Peperomia peltilimba Peperomia maculosa Portulaca oleraceae Crusea hispida Crusea diversifolia Diodia brasilensis Hamelia patens Capsicum annuum var. glabriusculum Lycopersicon esculentum Solanum americanum Physalis gracilis Jaltomata procumbens Helicteres guazumaefolia Lantana camara Verbena carolina

Current use +

+ + +

+ + + + + +

+

Past use + + + + + + + + + + + + + + + + + + + + + + +

Uses 6, 1, 3 10 3 4, 3 2, 1 1, 3 1, 3 1, 3 1 1 1 2 4, 3 10, 3 3 1 1 1, 3 1 1 3 1, 3 1

the milpa, but P. dumosus and P. vulgaris might also be sown in monoculture in permanent milpas or in coamiles (slash-and-burn plots). In the SNP region – as in other regions in Mexico – the processes of selection under domestication are ongoing, as in the case of an early maturing variety of P. dumosus recorded in the region called in Spanish frijol gordo abreviador (thick early bean). Selection was made because this is the most valued bean species in the region and under high demand, because of which the earliest harvests of this bean to reach to the market attain high prices, thus being profitable for the producers (Basurto et al. 1995, 1996).

The Squashes Four species of squashes of the genus Cucurbita are cultivated in the SNP, each within a limited elevation range (Fig. 4), the calabaza pipiana (C. argyrosperma) is sown in the lowlands with warm to semi-warm climate, the calabaza tamalayota (C moschata), the chilacayote (C. ficifolia), and the huitzayota (C. pepo) are sown in the highlands with temperate climate. All these squash species are associated with

Ethnobotany of the Sierra Norte de Puebla

417

maize and beans in the milpas, so the sequence of agricultural practices is similar to that applied for maize cultivation. But squashes can also be found cultivated in homegardens like C. moschata and C. ficifolia, which in these cases behave as climbers. Occasionally, four-pole structures roofed by a lattice are built for trellising C. moschata, and C ficifolia climbs on the trees in the homegardens. Except for the roots, all the parts of the squash plants are used as food, including the flowers, tender stems, immature fruits, mature fruits, and seeds, each harvested at a different stage of the agricultural cycle. The infraspecific diversity of squash species in the SNP region is high, in particular in C. moschata and C. ficifolia, where 11 of the 14 forms of fruit described for the genus Cucurbita are present (Basurto et al. 2014; ECPGR 2008).

The Chilares (chili peppers plots) Two species of chili are cultivated in the SNP region, the native C. annuum with two varieties –the chile verde or chile serrano (C. annuum var. annuum) and the chiltepín (C. annuum var. glabriusculum) – as well as the introduced species chile cera (C. pubescens). The latter species is grown in homegardens in the temperate zone, at elevations above 1400 m, in general, at a small scale for self-consumption and sale in local tianguis, except for some places as the municipality of Hueyapan where it is produced commercially. Varieties are cultivated with red, green, and yellow mature fruits; the last one being most valued in the market. Local seeds are used for cultivation of the chile verde, either obtained from the framers’ own harvest or acquired from neighbors in the community, and its cultivation is carried out in three modalities or types, each having different characteristics as described below. Type I cultivation. Recorded in the municipalities of Zapotitlán de Méndez (Tuxtla), Ixtepec, Atlequizayan, Ignacio Ramírez, and Jonotla. It is a polyculture system practiced in land left to follow for at least 4–5 years, after which the vegetation is fell (but not burned), and the chile verde seeds are directly sown in the unturned soil. Furrows and plants are separated by 40 cm and sowing is made in two cycles, one from May to August-November, and the other from October to March-July, during which the fruits are cut several times during the harvesting period. Manual weeding – by hand or using simple tools like hoe – is made after sowing approximately every month throughout the agricultural cycle, powdered and foliar fertilizer is applied, and sometimes, also fungistatic chemicals (Castro 2000). The harvest is sold in the local and the state markets. Quelites and other plants used for self-consumption or sale are sown in association with the chile verde. Type II cultivation. Practiced in Huahuaxtla, Amatitán Totolchan, and Huapalejcan in the municipality of Xochitlán de Vicente Suárez. It is also a polyculture system. Sowing is made on unturned soil in recently cleared land, weeding is manual, no fertilizers or pesticides are applied, and the harvest is primarily for selfconsumption, surplus being sold locally. The agricultural cycle is annual, or biannual and other cultivated or spontaneous species are also associated in the chilares.

418

F. Basurto et al.

Type III cultivation. Recorded in the municipalities of Naupan, Huauchinango, Tlaola, and Chiconcuautla. It is the most recent of the three types, consisting in a monoculture system with commercial purposes. Management is made as monoculture in permanent parcels that are turned at the beginning of each agricultural cycle, which have a duration of 9 months. Seedbeds are sown, and seedlings are later transplanted to the parcels, which are weeded every 15–30 days, to which pesticides and fertilizers are frequently applied (Evangelista 1999). The harvest is sold in the national markets and no useful spontaneous plants are associated to the crop, except if the chile seedlings fail to develop – described in Spanish as “no prende,” when bean or maize is sown in the void places; which is justified by farmers by in that way taking advantage of the applied fertilizer and avoiding its loss. The more noticeable differences between these cultivation systems is their purpose (for commercial or self-consumption), the duration of the agricultural cycle (6 months to 2 years), the use of agrochemicals (fertilizers and pesticides), and if spontaneous useful plants are associated with cultivation (Table 3). In Tuxtla, the harvesting of associated species – mainly quelites – represents a significant portion of the monetary income generated by the chilar (Castro 2000). The chiltepín began being cultivated about 25 years ago in response to the high demand for the product in markets, which was formerly obtained by gathering from wild or weedy chili plants dispersed by birds in the milpas, home gardens, and cafetales. The seedlings are obtained from seeds sown in fertilizer sacks, disposable cups, and, lately, in unicel or plastic germination trays and are transplanted to the chile plots or chilares. The amount of seeds fitting between the thumb and index finger is sown in each container with substrate prepared with soil and coffee husk compost, or with “suelo de monte” (forest soil). The transplant is made in the cafetales or home gardens when the plants are between 12 and 25 cm tall. Small holes are opened in the soil where seedlings will be transplanted, no fertilizer or pesticide is applied, and weeding is made in the same way as in the cafetales or homegardens, that is, using machete and hook. The fruit is always cut when green because if the plant is allowed to mature, its lifespan is shortened.

The Homegardens Depending on the elevation at which they are located, the homegardens in the SNP region can be divided into tropical and temperate, which differ in composition and structure, and in the importance of their production of basic foodstuffs. The temperate gardens are agroforestry systems where annual plants, like maize and bean, are associated with fruit trees including avocado, apple, pear, prune, and other trees belonging to the rose family. The floristic richness is high, having three or four strata: a herb layer, a shrubby layer, and one or two tree layers, with avocado as the emerging tree in the canopy (Table 4). Maize and beans are sown in furrows and trees may either be distributed in rows, grouped in “islands,” or dispersed throughout

Ethnobotany of the Sierra Norte de Puebla

419

Table 3 Species associated to chilares Especies asociadas Amaranthus cruentus Amaranthus hypochondriacus Solanum americanum Porophyllum ruderale Phaseolus polyanthus Phaseolus vulgaris Pachyrhizus erosus Lagenaria siceraria Physalis philadelphica Jaltomata procumbens Sonchus oleraceus Celosia argentea Cucurbita pepo Coriandrum sativum Brassica oleraceae Brassica rapa Citrullus lanatus Tagetes erecta Gomphrena globosa Allium tuberosum

Naupan

Huahuaxtla favored cultivated tolerated cultivated

cultivated

cultivated

cultivated tolerated tolerated cultivated cultivated cultivated

Tuxtla cultivated cultivated tolerated cultivated cultivated cultivated cultivated cultivated cultivated

cultivated cultivated cultivated cultivated favored cultivated cultivated cultivated cultivated

the land. The agricultural practices applied in home gardens are the same as those for the milpas and might be turned with horse-drawn plows or hoe. Trees rarely receive any treatment except for pruning as required and, after the harvest, limewash is sometimes applied to the trunks as a preventive measure against some pests. These orchards have a high agrodiversity of fruit tree species, with 30 varieties of apple, 40 of avocado, 24 of prune, and 34 of peach recorded in them, all according with the appreciation of farmers, who distinguish them by color, form, and size of the fruits and flowers, but also because by their organoleptic properties and phenology (Martínez 2007; Muñoz 2009; Peralta 2007). A characteristic of the temperate orchards in the SNP region is that they provide a substantial contribution to the supply of basic foodstuffs because they produce maize, beans, squashes and vegetables, pears, besides several quelites. The tropical orchards are equally diverse and multi-stratified as the temperate orchards, together containing over 300 useful plant species (Basurto 1982). Production in these agroecosystems is mostly for self-consumption and dietary supplements and not basic foodstuffs are produced. The agricultural practices are limited to weeding with machete and hook, and tree pruning as required. Plants are not in a specific spatial arrangement; although it is possible to delimit some areas within the orchards. A zone in front of the house commonly has ornamental and medicinal plants. The live fences have a particular floristic composition, and the bulk of the orchard area harbors the tallest plants (Basurto 1982).

420

F. Basurto et al.

Table 4 Species in temperate homegardens Species Amaranthus cruentus Amaranthus hypochondriacus Tagetes erecta

Common name Quintonil Quintonil

Management Cultivated, favored Cultivated, favored

Use Food, forage Food, forage

Cempoalxochitl

Ritual, medicine

Cucurbita ficifolia Porophyllum ruderale Sechium edule Capsicum pubescens Phaseolus vulgaris Phaseolus coccineus Phaseolus dumosus Zantesdechia aethiopica

Chilacayote Pápalo Espinoso Chile cera Frijol Tacuahuaquet Eexoyema, Acalete Alcatraz

Cultivated, tolerated Favored, cultivated Cultivated Cultivated Cultivated Cultivated Cultivated Cultivated

Brassica rapa B. oleracea Phisalys ixocarpa Rumex crispus Chenopodium berlandieri Chenopodium ambrosioides Amaranthus hybridus Amaranthus spinosus Melampodium divaricatun Simsia amplexicaule Sonchus oleraceus Cyclanthera dissecta Cyclanthera ribiflora Tinantia erecta Bidens odorata Tigridia pavonia Amphicarpae bracteata Persea americana Malus domestica Prunus domestica Pyrus communis Capsicum pubescens Prunus persica Sambucus nigra Juglans regia Crataegus mexicana Portulaca oleracea Prunus serotina spp. capuli Cestrum nocturnum

Nabo Col Tomate Xocoquilit Cenizo Epazote Quintonil Quintonil de burro Acahual Acahual Endivia Cincoquelite Huihuila Pata de pollo Mozote Taltoxpi Talet Aguacate Manzana Ciruela Pera Chile cera Durazno Sauco Nuez Tejocote Verdolaga Capulín Huele de noche

Cultivated Favored Favored Favored Favored Favored Favored Tolerated Tolerated Tolerated Tolerated Tolerated Tolerated Tolerated Tolerated Tolerated Tolerated Tolerated Cultivated Cultivated Cultivated Cultivated Cultivated Cultivated Tolerated Cultivated Tolerated, favored Tolerated Cultivated Tolerated

Food Food Food Food Food Food Food Ornamental, ritual Food Food Food Food Food Food Food, forage Food, forage Forage Forage Food Food Food Food Food, forage Food Food Food Food Food Food Food Food Medicine Food Food, medicine Food Food, medicine Medicine

Ethnobotany of the Sierra Norte de Puebla

421

The Coffee Plantations or Cafetales The coffee plantations or cafetales in the SNP region have an upper limit at elevations 1400 m (Fig. 4). They are agroecosystems dominated by the exotic species Coffea arabica. Coffee has possibly been present in the region since the eighteenth century but had a major expansion during the second half of the twentieth century, when the Mexican government policies – the Zacapoaxtla Plan, for example, promoted its cultivation with the goal of acquiring foreign currency. In that period, Mexico became the fourth largest exporter of coffee in the world, the state of Puebla occupying the fourth place of Mexican’s production with over 60,000 ha planted with the crop (Evangelista et al. 2010a). However, this status ended toward 1989 with the extinction of the INMECAFE, leaving producers subject to the “free market.” At present, the coffee producing zones are in crisis – both in the SNP as elsewhere in Mexico – and Mexico is excluded from the top ten coffee producing and exporting countries. Regardless of that, the cafetales remain highly relevant agroecosystems in the region, both in economic and environmental terms, because they maintain a tree cover that contributes to provide environmental services. Among them, control of soil erosion and water catchment areas, besides harboring a diverse fauna and flora (Leyequien and Toledo 2009; Martínez et al. 2007. The five types of coffee plantations described by Moguel and Toledo (1999): (1) traditional rustic, (2) shaded traditional polyculture, (3) shaded monoculture, (4) shaded commercial polyculture (with introduced productive species), and (5) unshaded monoculture (Martínez et al. 2007) can be found in the SNP region. In addition, 12 varieties of coffee are found in them, all belonging to C. arabica – Arabic or criollo, Bourbon, Red Caturra, Yellow Caturra, Mundo Novo, Catimor, Catuai, Caturrillo, Garnica, Marago or Maragogipe, Pacamara, and Oro Azteca. Except for the unshaded coffee plantations – grown by large estate owners in the municipalities of Jopala, Xicotepec, and Zihuateutla in the central western part of the SNP region – the cafetales are agroecosystems with floristic composition and structure depending on the environmental factors, but also on the farmers’ preferences, economic potential, and experience (Basurto 1982; Cruz 2004; Martínez et al. 2007). The most common coffee plantations in the SNP are those shaded either by the woody species in the natural vegetation (the traditional rustic cafetales) or by an increasingly modified in species composition canopy that also becomes less diverse and dense, going from the traditional and commercial polyculture cafetales with exotic trees to the shaded monoculture cafetales. In the monospecific shaded coffee plantations in the warmest zones and up to the transition to the temperate climates at elevations from about 900 and 1000 m, the trees used for shade are in the genus Inga, and above that elevation, the ilite (Alnus acuminata ssp. arguta) is preferably used (Cruz 2004). The canopy in the shaded polyculture cafetales includes the use of over 70 native and exotic tree species, which besides providing shade to the coffee also produce fruit or wood (Cruz 2004; Martínez et al. 2007).

422

F. Basurto et al.

The polyculture coffee plantations shaded with exotic productive species mostly use banana (Musa acuminata x M. balbisiana), which, besides shade, provide fruits and leaves used for producing tamales. In the 1990s, the timberyielding cedro rosado (shingle tree; Acrocarpus fraxinifolius) was introduced as a shade tree for coffee, but the introduction failed over the years, both because its wood is not valued in the region and because of lack of infrastructure for its industrial processing. In this context, attempts to improve the economic return from coffee plantations, the strategies of the coffee growers in the SNP region have been diversified and become more successful. One strategy has been to increase the richness of plants of economic importance in the coffee plantations with allspice and fruit trees like mamey and black sapote, or plant species with medicinal use. Another strategy has been to add value to the harvested coffee grains through their local processing, roasting, and commercialization (Escobar 2012; Martínez et al. 2004; Zurita 2004). The shaded coffee plantations in the SNP have a complex multi-layer structure with three to four vegetation layers and a high plant species diversity, of which 319 are useful (Basurto 1982; Cruz 2004; Martínez et al. 2007). The spatial distribution patterns of the plant species in coffee plantations depend on their type. For instance, the coffee plants in the rustic and polyculture shaded cafetales are planted in rows or triangles, but in the monoculture coffee plantations and in which exotic species are used as shade trees, both the shade trees and the coffee plants are planted in rows. The cafetales continue to be a productive alternative for farmers because, thanks to their diverse floristic composition and versatility of possibilities for introducing new useful species, they provide them products to satisfy several needs (Martínez et al. 2007). At the same time, the cafetales in the SNP have a dynamic process of conversion to other agroecosystems, which is driven by social and economic factors (Fig. 5). In part, this conversion between agroecosystems is caused by the abandonment of the furthest, or of more difficult access, coffee plantations in response to the low market prices of coffee, which resulted in the increase of forested area reported by Evangelista et al. (2010a) in the coffee producing area of the SNP. The ecological succession processes in abandoned plantations make cafetales to be invaded by natural vegetation, which is locally described in Spanish as land becoming enmontada or recovering vegetation of the mount. Within the high plant diversity in cafetales, at least 64 tree species from the primary vegetation are included; however, the importance value indexes (IVI) of these species are low – in the order of 1.1 to 10.2 – compared to species in the shaded polyculture and shaded monoculture plantations – with IVI of 20.0 to 80.2. In general, these are species with economic importance as in the case of fruit trees (mamey, black sapote), timber trees (Spanish cedar Cedrela odorata, mahogany Swietenia macrophylla, carboncillo Ocotea puberula), or spice plants (allspice). Thus, while the coffee plantations in the SNP region together with the home gardens are preserving a high biodiversity, that diversity is mostly composed of economically important species, and contain only few species or individuals of dominant trees or of those that are characteristic of the primary vegetation.

Ethnobotany of the Sierra Norte de Puebla

423

Fig. 5 Dynamics of replacing between coffee plantations and other agroecosystems

Perspectives The SNP region is rich in plant resources and its inhabitants possess vast traditional knowledge about them, both for their management and for their use. The floristic inventory of the region, which currently comprises over 900 species, has not been completed yet and it is likely that this goal will be hardly accomplished; above all, because the use of plants changes and adaptations to new economic, social, and cultural contexts – perhaps in that order – that are generated by local communities with the purpose of integrating them to the national society and market economy. Added to new and improved road networks, which change the contexts, leading to new demands for the use of resources and to the transformation of the traditional forms of resource appropriation – aimed at self-consumption – into commerciallyoriented approaches that, in general, have a stronger impact on plant populations. A similar process takes place in the agricultural practice in northern Puebla, which is based on traditional technology supported by the empirical experience of farmers, and that is not static despite the trends to conserve the proven and verified practices and processes. On the contrary, the experience from several decades of ethnobotanical research made in the region shows that the local agriculture is highly

424

F. Basurto et al.

dynamic, the producers constantly experimenting to adapt their practices and knowledge to the new socio-environmental conditions. Through this process, new techniques and plant species or varieties are incorporated to the established traditional agroecosystems, and the local farmers demonstrate their great interest for innovating their agricultural practices to adapt them to the social, ecological, and technological changes, as much as possible, according to their own interests and needs. Simultaneously, the agricultural practice in the SNP region is highly dynamic regarding the conversion of land use and land cover. Particularly, the milpas and cafetales might change to different agroecosystems in relatively short periods due to the versatility and plasticity of these agroecosystems to satisfy different requirements of farmers, either for self-consumption of households or for provision of monetary income. The plant resources present in the SNP region together with the local inhabitants’ knowledge about them cover the diverse needs of the producers, including food, health, ceremonial, aesthetic, firewood, clothing, shelter, and tool production. But despite the persistence of the traditional knowledge, we consider that there is a current threat of loss of these knowledge and know-hows because young people – both males and females – are abandoning agriculture. Also, because of the increased tendency of this age group to emigrate in search for jobs to cities and urban areas in the country (and in the last decades, also to the USA.). Considering that farmers in the region are about 50 years old on average, and because of the lack of generational exchange, it is possible that local traditional knowledge could be severely eroded in a few years, with that causing the loss of know-hows that could contribute to meeting the national challenge of preserving agrobiodiversity and making a sustainable use of plant resources. Biocultural conservation is a great challenge, and ethnobotany may substantially contribute to face it. Acknowledgments We are pleased to express our gratefulness to the already several generations of inhabitants on the SNP that have collaborated in the ethnobotanical exploration made in the region, sharing with us their knowledge and experience and giving us their generous hospitality, but above all, their invaluable friendship. We also acknowledge the collaboration of many students that have made research in the SNP region to obtain their bachelor’s degree. We thank Dr. Sergio Zárate for the English translation of the manuscript.

References Aguilar A, Camacho JR, Chino S, Jácques P, López E. Herbario medicinal del Instituto Mexicano del Seguro Social. Información etnobotánica. Instituto Mexicano del Seguro Social: México DF; 1994. Alexiades M, Shanley P. Productos forestales, Medios de subsistencia y Conservacion: Estudios de caso sobre sistemas de Manejo de Productos forestales No Maderales. En: Alexiades M, Shanley P, editors. Conservación y medios de subsistencia. Diversos casos sobre productos forestales no maderables América Latina. Volumen 3. Bogor, Indonesia: CIFOR; 2004.

Ethnobotany of the Sierra Norte de Puebla

425

Anónimo. Boletín hidrológico No. 42 Región hidrológica 27 Tuxpan-Nautla. México DF: Secretaría de Recursos Hidráulicos; 1970. Anónimo. Síntesis geográfica, nomenclátor y anexo cartográfico del Estado de Puebla. México DF: INEGI; 1987. Arias L, Ortega R. Semblanza: La milpa del maestro “Xolo,” Efraim Hernández Xolocotzi. In: Arias L, González T, Perez M, editors. Seminario Internacional “Efraím Hernández Xolocotzi” sobre la Milpa Maya. Mérida, Yucatán, México. 31 de Octubre de 2013. Memoria y resúmenes del seminario. 2013. Arrequín VO. Recursos hídricos. En: CONABIO, editor. La biodiversidad en Puebla: estudio de estado. Puebla: CONABIO, Gobierno del Estado de Puebla, BUAP; 2011. Basurto F. Huertos familiares en dos comunidades nahuas de la Sierra Norte de Puebla: Yancuictlalpan y Cuauhtapanaloyan. Tesis profesional. México DF: Facultad de Ciencias, Universidad Nacional Autónoma de México; 1982. Basurto F. Aspectos etnobotánicos de Phaseolus coccineus L. y Phaseolus polyanthus Greenman en la Sierra Norte de Puebla, México. Tesis de Maestría en Ciencias. México DF: Facultad de Ciencias. Universidad Nacional Autónoma de México; 2000. Basurto F, Martínez D, Castellanos A, Martínez MA. Ciclo agrícola y fenología de Phaseolus coccineus L. en sistemas de agricultura tradicional de la Sierra Norte de Puebla. Etnoecológica. 1995; III (4–5): 71–81. Basurto F, Martínez D, Martínez MA, Castellanos A. “Frijol gordo abreviador”, una forma precoz de Phaseolus coccineus L. ssp. darwinianus Hernández X. & Miranda C. en Nauzontla, Puebla, México. Rev Geogr Agrícola. 1996;22–23:95–102. Basurto F, Martínez MA, Villalobos G. Los quelites de la Sierra Norte de Puebla, México: Inventario y formas de preparación. Bol Soc Bot Mex. 1998;62:49–62. Basurto F, Villalobos G, Martínez MA, Sotelo A, Gil L, Delgado A. Use and nutritive value of Talet beans Amphicarpaea bracteata (Fabaceae: Phaseoleae) as human food in Puebla, Mexico. Econ Bot. 1999;53(4):427–34. Basurto F, Castro D, Martínez MA. Edible Begonia from the North of Puebla, Mexico. Econ Bot. 2003;57(1):48–53. Basurto F, Evangelista V, Castro D, Martínez MA. Lotería de quelites. Los quelites de México. México DF: Instituto de Biología, Universidad Nacional Autónoma de México y CONABIO; 2005. Basurto F, Evangelista V, Mendoza M. Manejo, conservación y aprovechamiento de los recursos vegetales de la Sierra Norte de Puebla. En: Ramírez B, Bernal H, coordinators. Investigación multidisciplinaria en la Sierra Norte de Puebla. Puebla, México: Colegio de Postgraduados; 2009. Basurto F, Castro D, Martínez D, Rodríguez T, Peralta L. Uso y manejo de calabazas cultivadas (Cucurbita spp) en el estado de Puebla. Agro Productividad. 2014;7(1):44–9. Castro A. Estudio comparativo del conocimiento sobre plantas medicinales utilizadas por dos grupos étnicos del Municipio de Pahuatlán, Puebla. Tesis profesional. Tlalnepantla, México: Escuela Nacional de Estudios Profesionales Iztacala, Universidad Nacional Autónoma de México; 1988. Castro D. Etnobotánica y papel económico de cuatro especies de quelites en Tuxtla, Zapotitlán, Puebla. Tesis profesional. México DF: Facultad de Ciencias, Universidad Nacional Autónoma de México; 2000. Castro D, Alvarado R, Evangelista V. Recetario de quelites de la Sierra Norte de Puebla. México, DF: Instituto de Biología, Universidad Nacional Autónoma de México y CONABIO; 2005. CONAPO. 2015. Índice de marginación por entidad federativa y municipio 2015. https://www. gob.mx/conapo/documentos/indice-de-marginacion-por-entidad-federativa-y-municipio2015. http://www.conapo.gob.mx/es/CONAPO/Indices_de_Marginacion_2010_por_entidad_ federativa_y_municipio Cruz A. Los sistemas agrícolas de Jilotzingo, Municipio de Zacatlán, Puebla. Tesis profesional. México DF: Facultad de Ciencias, Universidad Nacional Autónoma de México; 1995.

426

F. Basurto et al.

Cruz A. La importancia del hilite blanco (Alnus acuminata ssp arguta (Schlechtdl.) Furlow (Betulaceae) en la sombra de cafetales de Xochitlán de Vicente Suárez, Puebla. Tesis de Maestría en Ciencias. México DF: Facultad de Ciencias, Universidad Nacional Autónoma de México; 2004. Diez A. El proyecto arqueológico Teteles de Avila Castillo; investigaciones en la Sierra Norte de Puebla. Arqueología. 2017;54:167–72. Diez A. Desarrollo regional en la Sierra Norte de Puebla durante la época prehispánica. INAH: México; 2019. ECPGR. 2008. Minimum descriptors for Cucurbita spp., cucumber, melon and watermelon. ECPGR Secretariat. https://www.ecpgr.cgiar.org/fileadmin/templates/ecpgr.org/upload/NW_ and_WG_UPLOADS/Cucurbits_DescriptorLists.pdf Eleuterio A, Perez D. Management of tree ferns (Cyathea spp) for handicraft production in Cuetzalan, México. Econ Bot. 2006;60:182–91. Escobar T. El zapote negro (Diospyros digyna Jacq. Ebenaceae): un PFNM de la Sierra Norte de Puebla. Tesis profesional. Tlalnepantla, México: Facultad de Estudios Superiores Iztacala, Universidad Nacional Autónoma de México; 2012. Escobar T. Importancia cultural y prácticas de manejo de las plantas asociadas a dos rituales totonacos en la Sierra Norte de Puebla. Tesis Maestría. Xalapa, Veracruz: Centro de Investigaciones Tropicales, Universidad Veracruzana; 2016. Evangelista V. Influencia de dos cultivos comerciales en el cultivo de maíz en la comunidad de Naupan, Puebla. Tesis Maestría en Ciencias. México DF: Facultad de Ciencias, Universidad Nacional Autónoma de México; 1999. Evangelista V, Basurto F. Plantas comestibles de la Sierra Norte de Puebla. En: Masferrer E, Vences G, Barranco N, Díaz E, Mondragón J, editors. Etnografía del estado de Puebla. Norte. Puebla: Gobierno del Estado de Puebla, Secretaria de Cultura, Puebla; 2003. Evangelista V, Mendoza M. Calendarios agrícolas en cuatro ejidos del Municipio de Coxquihui, Ver. Tesis profesional. México DF: Facultad de Ciencias, Universidad Nacional Autónoma de México; 1987. Evangelista V, Basurto F, Martínez MA, Mendoza M. La flora alimentaria de la Sierra Norte de Puebla. En: Masferrer E, Mondragón J, Vences G, editors. Los pueblos indígenas de Puebla. Atlas etnográfico. México DF: Gobierno del Estado de Puebla. Instituto Nacional de Antropología e Historia; 2010a. Evangelista V, López J, Caballero J, Martínez MA. Patrones espaciales de cambio de cobertura y uso del suelo en el área cafetalera de la Sierra Norte de Puebla. Bol Inst Geogr, Universidad Nacional Autónoma de México. 2010b;72:23–38. Evangelista V, Mendoza M, Basurto F. Sistemas agrícolas y recursos vegetales en el municipio de Huauchinango. En: Mora L, editor. Huauchinango. El rumor del tiempo. Perspectivas Interdisciplinarias en Red AC. Puebla, México: Gobierno municipal de Huauchinango, Puebla; 2011. Faulhaber J. Antropología biológica de las sociedades prehispánicas. En: Manzanilla L, López L, coord. Historia antigua de México, vol. I. México DF: Miguel Angel Porrúa; 1994. Fox W, Sharp A. La distribución de Fagus en México. Bol Soc Bot Mex. 1954;17:31–53. Fuentes L. Regiones naturales del Estado de Puebla. México DF: Instituto de Geografía, Universidad Nacional Autónoma de México; 1972. García B. Los pueblos de la sierra. El poder y el espacio entre los indios del norte de Puebla hasta 1700. El Colegio de México: México DF; 1987. González JC. Historia ambiental de la milpa como agroecosistema (1970–2009) en Xiloxochico, Cuetzalan, Puebla. Tesis profesional. México DF: Facultad de Ciencias, Universidad Nacional Autónoma de México; 2014. Guzmán F. La abeja prehispánica, mejor que la europea. 2018. https://www.gaceta.unam.mx/laabeja-prehispanica-mejor-que-la-europea/ Hernández XE. La agricultura tradicional en México. Comercio Exterior. 1988;38(8):673–8.

Ethnobotany of the Sierra Norte de Puebla

427

Leyequien E, Toledo V. Floras y aves de cafetales: ensambles de biodiversidad en paisajes humanizados. Biodiversitas. 2009;83:7–10. López C. “Amate” papel de corteza mexicano (Trema micrantha (L.) Blume): Estrategias de extracción de corteza para enfrentar la demanda. En: Alexiades M, Shanley P, editors. Conservación y medios de subsistencia. Diversos casos sobre productos forestales no maderables América Latina. Volumen 3. Bogor, Indonesia: CIFOR; 2004. López E. Inventario de especies frutales en huertos y cafetales en dos comunidades del municipio de Tlatlauquitepec, Puebla: Yoloctzin y Mazatepec. Tesis profesional. Zacapoaxtla, Puebla: Instituto Tecnológico Superior de Zacapoaxtla; 2008. López L, Cházaro M. Plantas leñosas raras del bosque mesófilo de montaña. I. Fagus mexicana Martínez. (Fagaceae). Bol Soc Bot Mex. 1995;57:113–5. López E, Aguilar A, Aguilar S, Xolalpa S. Las Verbenaceae empleadas como recurso herbolario en México: una revisión etnobotánica-médica. Polibotánica. 2017;44:195–216. Mapes C, Basurto F. Biodiversity and edible plants of Mexico. In: Lira R, Casas A, Blancas J, editors. Ethnobotany of Mexico. Interactions of people and plants in Mesoamerica. New York: Springer; 2016. Mapes C, Basurto F, Bye R. Ethnobotany of quintonil. Knowledge, use and management of edible greens Amaranthus spp. in the Sierra Norte de Puebla, México. Econ Bot. 1997;51(3):293–306. Mapes C, Basurto F, Bye R. Importancia biológica, económica y cultural del ‘quintonil’ Amaranthus spp. en la Sierra Norte de Puebla, México. En: L. Mera, D. Castro, R. Bye, compilers. Especies vegetales poco valoradas: una alternativa para la seguridad alimentaria. Instituto de Biología, Universidad Nacional Autónoma de México; 2011. Mapes C, Basurto F, Bautista L. Manejo y cultivo de Amaranthus spp. como quelite en la Sierra Norte de Puebla, México. México DF: Instituto de Biología, Universidad Nacional Autónoma de México, SNICS-SINAREFI; 2012. Mapes C, Basurto F, Díaz A. Diversidad de ‘quintoniles’ (Amaranthus spp.) en la Sierra Norte de Puebla, México. México DF: Instituto de Biología, Universidad Nacional Autónoma de México; 2013. Martínez E. Especies vegetales como recurso energético de uso doméstico, en Zozocolco de Hidalgo, Veracruz. Tesis profesional. México DF: Facultad de Ciencias, Universidad Nacional Autónoma de México; 1992. Martínez MA. Los frutales de la Sierra Norte de Puebla. En: Nieto R, editor. Frutales nativos: un recurso fitogenético de México. Texcoco: Universidad Autónoma Chapingo; 2007. Martínez A. Estudio de la Pimenta dioica (L.) Merril, un recurso forestal no maderable en dos comunidades del municipio de Tuzamapan de Galeana, Sierra Norte de Puebla, México. Tesis profesional. Tlalnepantla, México: Facultad de Estudios Superiores Iztacala, Universidad Nacional Autónoma de México; 2009. Martínez MA, Evangelista V, Mendoza M, Morales G, Toledo G, Wong A. Catálogo de plantas útiles de la Sierra Norte de Puebla, México. Cuadernos 27. México DF: Instituto de Biología, Universidad Nacional Autónoma de México; 1995. Martínez MA, Evangelista V, Basurto-Peña F, Mendoza M. La etnobotánica y los recursos fitogenéticos: El caso de la Sierra Norte de Puebla. Rev Geogr Agrícola. 2000;31:79–88. Martínez MA, Evangelista V, Mendoza M, Basurto F, Mapes C. Estudio de la pimienta gorda (Pimenta dioica (L.) Merrill), un producto forestal no maderable de La Sierra Norte de Puebla, México. En: Alexiades M, Shanley P, editors. Productos forestales, medios de subsistencia y conservación: estudios de caso sobre sistemas de manejo de productos forestales no maderables. Vol. 3 América Latina. Bogor: CIFOR; 2004. Martínez MA, Evangelista V, Basurto F, Mendoza M, Cruz A. Flora útil de los cafetales de la Sierra Norte de Puebla, Mexico. Rev Mex Biodiversidad. 2007;78(1):15–40. Martínez MA, Evangelista V, Mendoza M, Basurto F, Zurita A. Estudios sobre etnobotánica otomí en el municipio de Pahuatlán de Valle, Puebla. En: Heiras C, coordinator. Memoria de papel. Actas del Primer Coloquio sobre Otomíes de la Sierra Madre Oriental y grupos vecinos. México DF: Instituto Nacional de Antropología e Historia; 2008.

428

F. Basurto et al.

Martínez L, Nicolalde F, Stevenson D, Vergara F. A new species of Ceratozamia (Zamiaceae) from the sierra Norte of Puebla, Mexico. Brittonia. 2017;69(4):516–24. Masferrer E. Cambio y continuidad entre los totonacos de la Sierra Norte de Puebla. Tesis Maestría en Antropología social. México DF: Universidad Iberoamericana; 2006. Masferrer E, Martínez MA. Los indígenas, las regiones poblanas y sus agroecosistemas. En: Masferrer E, Mondragón J, Vences G, editors. Los pueblos indígenas de Puebla. Atlas etnográfico. México DF: Gobierno del estado de Puebla. Instituto Nacional de Antropología e Historia; 2010. Mera LM, Alvarado R, Basurto F, Bye R, Castro D, Evangelista V, Mapes C, Martinez MA, Molina N, Zaldívar J. “De quelites me como un taco”. Experiencia en educación nutricional. Rev Jardín Botánico Nacional. 2003;24(1/2):45–9. Mera L, Alvarado R, Basurto F, Bye R, Castro D, Evangelista V, Mapes C, Molina N, Saldívar J. De quelites me como un taco. Ciencias. 2005;77:36–8. Miranda F, Hernández-X E. Los tipos de vegetación de México y su clasificación. Bol Soc Bot Mex. 1963;28:29–179. Miranda F, Sharp A. Characteristics of the vegetation in certain temperate regions of Eastern Mexico. Ecology. 1950;31:313–33. Moguel P, Toledo V. Biodiversity conservation in traditional coffee systems of Mexico. Biol Conserv. 1999;13(1):11–21. Molina N. Etnobotánica de quelites en el sistema milpa en Zoatecpan, una comunidad indígena náhuat de la sierra norte de Puebla. Tesis profesional. México DF: Facultad de Ciencias, Universidad Nacional Autónoma de México; 2000. Morales L. Caracterización y análisis de compuestos bioactivos en quelites de la familia AMA Tesis Maestría en Ciencias. Ciudad de México: Facultad de Química, Universidad Nacional Autónoma de México; 2016. Muñoz A. Diversidad de manzana en el municipio de Zacatlán, Puebla, México. Tesis profesional. Puebla, Puebla: Escuela de Biología, Benemérita Universidad Autónoma de Puebla; 2009. Ochoa L. Historia prehispánica de la Huasteca. México DF: Instituto de Investigaciones Antropológicas, Universidad Nacional Autónoma de México; 1984. Paray L. Exploraciones botánicas en el norte del estado de Puebla. Bol Soc Bot Mex. 1946;4: 10–2. Peralta L. Diversidad de frutales en huertos familiares de Cuacuilco y Las Lomas, Zacapoaxtla, Puebla. Tesis profesional. Instituto Tecnológico Superior de Zacapoaxtla: Zacapoaxtla, Puebla; 2007. Pérez R, Romero S. coordinators. Capítulo 2. Medio socioeconómico en CONABIO editor. La biodiversidad en Puebla: estudio de estado. CONABIO, Gobierno del Estado de Puebla, BUAP: Puebla; 2011. Puig H. Vegetación de la Huasteca, México: estudio fitogeográfico y ecológico. ORSTOM, Instituto de Ecología A.C., CEMCA: México DF; 1991. Ruiz-Pérez M, Belcher B, Achdiawan R, Alexiades M, Aubertin C, Caballero J, Campbell B, Clement C, Cunningham T, Fantini A, de Foresta H, García Fernández C, Gautam K, Hersch P, de Jong W, Kusters K, Govindan Kutty M, López C, Martínez MA, Raghavan Nair T, Ndoye O, Ocampo R, Rai N, Ricker M, Schreckenberg K, Shackleton S, Shanley P, Sunderland T, Youn Y-C. Markets drive the specialization strategies of forest peoples. Ecol Soc. 2004;9(2):4. Rzedowski J. Vegetación de México. LIMUSA: México DF; 1978. Saldaña J. coordinator. El medio físico. En: CONABIO. La biodiversidad en Puebla: estudio de estado. CONABIO, Gobierno del Estado de Puebla, BUAP; 2011. Sánchez G, Santiago R, Caballero Y, Rojas I, De Gyves S. 2016. Flora melífera potencial con propiedades medicinales y nutricionales, colindante a los meliponarios de Scaptotrigona mexicana, Cuetzalan, Puebla. Memorias XX Congreso Mexicano de Botánica. Ciudad de México 4–9 septiembre 2016. Sociedad Botánica de México, A.C. https://www.researchgate. net/publication/320962574_Flora_melifera_potencial_con_propiedades_3es_y_nutricionales_ colindante_a_los_meliponarios_de_Scaptotrigona_mexicana_Cuetzalan_Puebla

Ethnobotany of the Sierra Norte de Puebla

429

Torres G. Manejo tradicional de Dasylirion acrotichum (Schiede) Zucc. (Asparagaceae) para la elaboración de arcos florales en el centro de Veracruz y la evaluación del impacto en sus poblaciones naturales. Tesis Maestría. Xalapa, Veracruz: Centro de Investigaciones Tropicales, Universidad veracruzana; 2016. Vaylón L. Uso y distribución de Castilla elastica (hule) en Zozocolco de Guerrero, Veracruz. Tesis profesional. México DF: Facultad de Ciencias, Universidad Nacional Autónoma de México; 2012. Vázquez H. El conocimiento ecológico en las prácticas agrícolas tradicionales entre los Totonacos de una comunidad de la Sierra Norte de Puebla. Tesis profesional. México DF: Facultad de Filosofía y Letras, Universidad Nacional Autónoma de México; 1990. Zurita A. Estudio de un producto forestal no maderable, el malabar Solanum erianthum D. Don, en el municipio de Pahuatlán, Puebla. Tesis profesional. Tlalnepantla, México: Faculta de Estudios Superiores Iztacala, Universidad Nacional Autónoma de México; 2004.

Preserving Healthy Eating Habits: Quelites in the Food System of a Nahua Mountain Community, Mexico Claudia Sa´nchez-Ramos, Heike Vibrans, María Rivas-Guevara, Edelmira Linares, Edmundo García-Moya, and Alfredo Saynes-Va´squez

Abstract

Quelites (edible tender leaves, stems, and flowers) are important components of the Mesoamerican food system, but, according to the literature, their use has been declining. Sociodemographic factors that influence their consumption are not well-known. We studied these factors in relation to species of quelites and commercial vegetables used, sources, preparation, and quantities in the highland Nahua community of Tetlatzinga, Veracruz, Mexico. Twenty families, randomly selected, and school students were interviewed. Food diaries contributed some data. The population consumed 35 species of quelites; the primary source were maize fields (milpas). During the season, May to July, families consumed about 10 kg per week on average. Gender, age, and occupation influenced knowledge and consumption, but not the socioeconomic level, bilingualism (SpanishNahuatl), or years of schooling. Contrary to expectations, the consumption of quelites was well-regarded with some nuances among students; family cooks innovated, and quantities approached recommended per capita consumption for all vegetables. Talks by the local health clinic apparently had positive effects. We show that quantitative studies help to understand local decisions and can challenge common beliefs.

C. Sánchez-Ramos · H. Vibrans (*) · E. García-Moya · A. Saynes-Vásquez Posgrado en Botánica, Colegio de Postgraduados. Montecillo, Texcoco, Estado de México, Mexico e-mail: [email protected] M. Rivas-Guevara Universidad Autónoma Chapingo, Texcoco, Estado de México, Mexico E. Linares Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_12

431

432

C. Sa´nchez-Ramos et al.

Introduction Mountains are refuges and areas of withdrawal, for plants, animals, and, especially, people. This property is due to the variety of habitats they offer, but mostly to their low attractiveness for many elements of modernization. Mountainous areas are more easily defended, both militarily and culturally (e.g., Scott 2009, its critics and derived literature), infrastructure is expensive to build, and one of the strongest drivers of modern life, the economies of scale, is much less powerful. Thus, traditional ways of life tend to have advantages in these areas. Vegetables are an essential part of balanced human nutrition. Most derive from plant parts other than fruit and seeds. They contribute some calories, proteins, and fat. However, their main value lies in their content of vitamins, minerals, essential oils, antioxidants, fibers, and a large group of other substances that are essential for humans, but only in small quantities (Robertson 2003; Sadler et al. 2005). Also, they add variety to the diet and lower risks of cardiovascular disease (Reddy and Katan 2004). They are much more diverse and subject to chemical changes through processing and spoilage than other food types. Like mountains, they are somewhat less susceptible to economies of scale than other food system components consumed in larger quantities and lower variety, such as carbohydrates, proteins, fats, and oils. However, even in the mountains, rural livelihoods, food systems, and nutrition are changing (Łucza et al. 2012). Today, people have access to industrialized food, as well as economic and social incentives to consume it. In some cases, this leads to better, less deprived lives. However, very commonly, only the cheapest and most attractive (sweet, salty, and fatty) parts of the dominant urban food culture are adopted, with worrisome public health outcomes. Also, as people integrate into modern ways of life, lack of time leads to an abandonment of previous food habits. Ethnobotanists can address basic questions on this transition: the degree to which different factors, cultural, agronomical, or economic, influence this change. The answers to these questions, in turn, help to identify policies that are more effective in mitigating negative consequences of cultural change, and to put common generalizations in context. In this chapter, we describe a study that explores the relationship between various sociodemographic factors and the food use of wild-growing leafy vegetables. The study community was a mountain village, which was nevertheless closely linked to the outside world through its trade of woodworking products. It shows that common expectations, such as that poor people consume more wildgrowing greens, are not necessarily true; also, relatively simple interventions - such as information on the healthiness of these foods - can counteract cultural prejudices. Quelites (from the Nahua term quilitl in singular and quilite in plural) is a term used in Mexico for plants whose leaves, tender stems, flowers, or inflorescences are eaten as vegetables (Bye 2000). The term is sometimes translated as spinach greens, leafy vegetables, weedy or wild-growing vegetables, but none of these translations circumscribe the group adequately. They are mostly, but not always green or leafy, and some are cultivated. For this reason, we use the local term in this chapter. Historically, these plants have been the main vegetable component in the Mesoamerican diet. Most have not been highly domesticated, though several species have

Preserving Healthy Eating Habits: Quelites in the Food System of. . .

433

been subjected to selection (Mapes et al. 1996, 1997; Rendón et al. 2001; Rendón and Núñez–Farfán 2001). Many reproduce spontaneously and do not require much labor input, apart from harvesting them, though some are managed to some extent (Casas et al. 2007; Vibrans 2016). In Mexico, about 250 species of quelites (Basurto 2011) are commonly consumed in different regions. Most are seasonal foods, and many are associated with annual crops, particularly the mixed maize cultivation (milpa). They are a relevant part of the diet of many peasant families until the main crops are harvested (Bye and Linares 2000). Homegardens, coffee and other perennial crop plantations, as well as pastures and natural vegetation also provide quelites (Castro-Lara et al. 2005). Some are planted to extend the growing season or because they do not grow locally (Basurto-Peña et al. 1998). Quelites are usually for family consumption, but some are sold to neighbors, other communities, or in weekly regional markets (Mota-Cruz et al. 2011). For example, in the markets of the Tehuacán-Cuicatlán Valley near our study area, quelites such as hierbamora (Solanum nigrescens M. Martens & Galeotti), quintonil (Amaranthus hybridus L.), pipicha (Porophyllum linaria (Cav.) DC, papaloquelite (Porophyllum macrocephalum DC. ¼ Porophyllum ruderale var. macrocephalum (DC) Cronq.), and chepil (Crotalaria pumila Ortega) are sold (Arellanes and Casas 2011). Several authors have noticed, and everyday observations show that the consumption of quelites is declining in many regions (González 2008; Mera-Ovando et al. 2011). Various reasons have been proposed. Quelites are often associated with poverty; this image problem goes back to colonial times, was taken up by agronomists and early ethnobotanists, and still turns up in present-day writing (Vibrans 2016). Several decades of work by contemporary ethnobotanists and the last two decades with its food movements have provided some pushback; for example, one of the best-known and expensive restaurants in Mexico City is called “Quintonil,” the Nahuatl word for Amaranthus greens; another highly regarded restaurant, Pujol, highlights these greens in its menu. Other observations on the decline of quelite consumption include the abandonment of vegetable consumption in general, the introduction of commercial vegetables, destruction of habitats, change of agricultural practices, economic activities (and the associated lack of time), social class, and urbanization. Additional reasons may be changes in food preferences due to migration to cities and various external influences (Castro-Lara et al. 2005; González 2008). Vázquez-García et al. (2004) consider that the change is due partly to the abandonment of previous fieldwork organization, based on family labor. However, most of these are qualitative observations; there are few quantitative data on the drivers of quelite consumption or abandonment. While numerous aspects of quelites have been studied, particularly floristics, we know little about the sociodemographic factors that influence the knowledge and consumption of these plants in a rural environment. For example, quelite use and knowledge were differentiated by gender and culture in a comparative study of villages in Oaxaca, Mexico, and in Zimbabwe (Madamombe-Manduna et al. 2009). Women always knew more about food plants than men, as they were usually responsible for their preparation. However, this difference was much more

434

C. Sa´nchez-Ramos et al.

pronounced in Zimbabwe (where women also collected them and men did not get involved in their procurement) than in Mexico, where men often take home quelites from their fieldwork. Also, few studies quantify consumption or cultural aspects (but see Basurto et al. 2011). For this report, we studied a community in the mountains of Veracruz: Tetlatzinga in the municipality of Soledad Atzompa. It has an indigenous population that retains the native language, Nahuatl. Local agriculture is traditional with little or no external inputs. This region has integrated into the national economy only in the last 30 years. People are inserted partially into the money economy through an important woodworking tradition (mainly for furniture) that has developed in the region. The trade with these products also leads to much travel and exposure to other cultural norms, as well as access to commercial goods. We analyzed the sociodemographic factors that favor or impede the consumption of quelites in this rural area. Also, we investigated the consumption frequency of these plants and commercial vegetables during the quelite season (May to July), as well as the quantities and variety of recipes. We expected to find an inverse relationship between socioeconomic status and quelite consumption, some amount of cultural disdain for quelites as “poor man’s food” and an ongoing substitution with commercial vegetables such as broccoli or carrots that are important in the mainstream Mexican diet.

Methods Study Area Tetlatzinga belongs to the municipality of Soledad Atzompa, located in the centralwestern part of the state of Veracruz, Mexico, within the mountainous region of the Sierra de Zongolica (Fig. 1; 18  410 1800 N and 97  090 0800 ; altitude 2400–2800 m). The climate belongs to the temperate subhumid subtropics with rainfall in summer (García 1987). The average annual temperature ranges from 15  C to 18  C because of altitude differences (SEFIPLAN 2015). The natural vegetation consists of pineoak forests (Juárez 2007). Nahuatl is the main language spoken, though most people are bilingual with Spanish as their second language. The food system is based on maize tortillas, beans, rice, lentils, pasta soup, potatoes, peas, beans, nopales (prickly pear cactus pads), squash and pumpkins, various quelites and mushrooms, and seasonal fruit (apples, pears, plums, peaches, prickly pears). Meat is not eaten daily, but occasionally (chicken, beef, pork, fish, and lamb). Most people practice traditional intercropped maize agriculture on hillsides. The main crops are maize (Zea mays L.), squash (Cucurbita ficifolia Bouché and Cucurbita pepo L.), runner beans (Phaseolus coccineus L.), peas (Pisum sativum L.), broad beans (Vicia faba L.), potatos (Solanum tuberosum L.), oka (called foreign or red potato in the region, Oxalis tuberosa Molina) (Sánchez 2014), and barley (Hordeum vulgare L.). However, agriculture is not the principal source of income: a traditional activity, carpentry of pine wood furniture, has expanded in the last decades, and most households are involved in both production and commerce of these products.

Preserving Healthy Eating Habits: Quelites in the Food System of. . .

435

Fig. 1 The study area (Tetlatzinga, municipality of Soledad Atzompa, Veracruz, Mexico)

Fieldwork Permission was sought from the local authorities, and approval was given in writing. A random sample of 20 families was chosen by drawing lots from a list of families kept by the local clinic, which had the most complete population data. In May 2015, these families were visited. The first author, who speaks Nahuatl, explained the reason for the study and sought their consent orally. Preliminary quelite collections between November 2014 and April 2015 were used to make a field herbarium and for initial, informal interviews. During the main fieldwork phase (May to June 2015), the first author walked different types of vegetation with members of the selected families. The species mentioned in the initial interviews were collected and identified with literature at CHAPA and MEXU, and the help of some specialists. The voucher specimens were deposited at CHAPA. In this main work phase, 53 individuals of the 20 families over the age of 10 were interviewed individually. First, a list of quelites they knew and used was requested orally and documented by the first author. Then, the field herbarium was shown to the interviewees, to complete the list of quelites known to and consumed by each family. Interviews were about the source of each species mentioned and consumed by the family. These conversations were audio-recorded with consent. Two additional species were reported during informal interviews with other villagers (Piper auritum and Rumex acetosella; full names of the quelites mentioned in this text with author citations can be found in Table 1); we included these in the species list, but do not know the use frequency or sources.

C. Sa´nchez-Ramos et al.

436

Table 1 Quelite species consumed by the families of Tetlatzinga, Soledad Atzompa. The species with an asterisk were purchased at regional markets

Family Amaranthaceae Amaranthaceae Amaranthaceae

Species Amaranthus hybridus L. Beta vulgaris L. Chenopodium berlandieri Moq.

Amaranthaceae Amaranthaceae Apiaceae Asparagaceae

Dysphania ambrosioides L. Spinacia oleracea L. * Coriandrum sativum L. Agave atrovirens Karw. ex Salm-Dyck

Asparagaceae

Beschorneria yuccoides K. Koch

Asparagaceae

Yucca guatemalensis Baker

Asteraceae Asteraceae

Bidens triplinervia Kunth Porophyllum linaria (Cav.) DC. * Porophyllum ruderale (Jacq.) Cass. * Sonchus oleraceus L. Brassica oleracea L. Brassica rapa L. Capsella bursa-pastoris (L.) Medik. Nasturtium officinale W.T. Aiton * Raphanus raphanistrum L.

Asteraceae Asteraceae Brassicaceae Brassicaceae Brassicaceae Brassicaceae Brassicaceae

Common name in Spanish Quintonil Acelga Quelite blanco o cenizo Epazote Espinaca Cilantro Cacaya o flor de maguey

Life form Herb Herb Herb

Herb Herb Herb Succulent rosette without stem – Patahmolkahkaya Succulent rosette without stem Flor de izote Ikzoxochitl Succulent rosette with stem (shrub) – Kuilakochi Herb Pipitza – Herb

Epazotl – Kolanto Kahkaya

Pápalo

Pápaloquilitl

Herb

Lechuguilla Col de hoja Nabo –

Memeya Kolex Kilapox Kilkolex o Koahkolex Ateskilitl

Herb Herb Herb Herb

Berro Quelite de rábano

Caryophyllaceae Stellaria media (L.) Vill.



Cleomaceae

Quelite de cinco hojas

Cleome magnifica Briq. (¼ Andinocleome magnifica (Briq.) Iltis & Chochrane)

Common name in Nahuatl Wahkilitl – Kohkokilitl

Rabanoskilitl, Horrohkilitl o Kiramonoskilitl Momatilana, Mozozowa, Mapisil o Torohkilitl Makuilkilitl

Herb Herb

Herb

Shrub

(continued)

Preserving Healthy Eating Habits: Quelites in the Food System of. . .

437

Table 1 (continued)

Family Cucurbitaceae

Species Cucurbita ficifolia Bouché

Cucurbitaceae

Sechium edule (Jacq.) Sw.

Fabaceae Fabaceae

Erythrina americana Mill. * Phaseolus coccineus L.

Lamiaceae

Mentha spicata L.

Montiaceae

Calandrinia ciliata (Ruiz & Pav.) DC. Oxalis tuberosa Molina

Oxalidaceae

Phytolaccaceae Piperaceae Polygonaceae

Phytolacca rugosa A. Braun & C.D. Bouché Piper auritum Kunth Rumex acetosella L.

Polygonaceae

Rumex crispus L.

Portulacaceae Solanaceae

Portulaca oleracea L. * Jaltomata procumbens (Cav.) J.L. Gentry Solanum nigrescens M. Martens & Galeotti Solanum tuberosum L.

Solanaceae Solanaceae

Common name in Spanish Guías y flores de chilacayote Guías de chayote Gasparito Guías y flores del ayocote Hierba buena Quelite de borrego Hojas de papa extranjera Ñamole Hierba santa Quelite agrio o lengua de vaca chiquito Quelite agrio o lengua de vaca Verdolaga Jaltomate

Common name in Nahuatl Life form Ayohkilitl iwan Herbaceous ayohkilxochitl liana Witzayohkilitl, Witzkilitl o Pinozoskilitl – Esokilitl iwan esokilxochitl

Herbaceous liana Tree Herbaceous liana

Alwaweno

Herb

Ichkakilitl

Herb

Xokokamohkilitl

Herb

Ñamoli

Herb

Tlanekpakilitl Xokokilitl inon pixintzin

Shrub Herb

Xokokilitl

Herb

– Xaxaltohkilitl

Herb Herb

Hierba mora Tomakilitl

Herb

Hojas de papa

Herb

Kamohkilitl

The woman responsible for cooking in each of the 20 families was asked about how she prepared each quelite species. Also, each family cook was asked for a specific quelite recipe, for which the name, number of portions, ingredients and their quantities, and the preparation was registered.

438

C. Sa´nchez-Ramos et al.

Consumption of quelites was estimated through two sets of data – recall interviews and a food diary. For the first set, we asked the family cook about the number of times that each species of quelite mentioned in the interviews was consumed per month during the main season. To estimate quantities, we measured the informally used units (example: manojo, a handful) for one or two species with each family. For the rest of the species, the amounts consumed were estimated based on these measurements. Quelites that were used as condiments, particularly epazote and coriander, were not considered, as only a few leaves were used. Purchased green vegetables were also not considered, as their consumption was relatively rare and estimating quantities more difficult. For the second data set that included purchased non-quelite vegetables, we asked the families to keep food diaries for 3 weeks from the end of June to the first week of July of 2015. Of the 20 families selected, only 19 collaborated because in one family no member was literate. Older school children often kept the food diary, and they received a symbolic compensation (e.g., school supplies); the diaries were revised weekly with the first author. The food diary included all foods consumed by all members of the family: breakfast, lunch, dinner, and food eaten between meals. Quantities were not considered, only frequency. To elucidate the motives that influence the knowledge and attitudes toward quelites among children and young people, we interviewed the students of the last grade of a primary, a secondary and a high school (Ignacio Zaragoza Elementary School, 21 students, Secondary Technical Agriculture and Preparatory School # 137, 66 students, and the College of Scientific and Technological Studies of the State of Veracruz, 44 students), in each case with the permission of the director and the class teacher. Students, both male and female, indicated their age, listed the quelites they knew, their favorites, and explained motives for consumption or non-consumption.

Socioeconomic Data Interviewees were questioned about their age, main occupation, years of schooling and language proficiency (Nahuatl monolingual ¼ 1, speaks Nahuatl but understands Spanish ¼ 2, speaks Nahuatl and Spanish more or less equally ¼ 3; there were no interviewees that were more proficient in Spanish than in Nahuatl). Occupations were also codified. The 20 families interviewed were classified into 3 socioeconomic levels (low, medium, and high), according to local criteria, mainly based on characteristics of their houses (size, material, appliances) and the ownership of vehicles. The categories were adjusted so that each class contained about one-third of the families. This classification was used to analyze the relationship between the use of quelites and socioeconomic status.

Preserving Healthy Eating Habits: Quelites in the Food System of. . .

439

Data Analysis The data collected on consumption frequency, quantities, and sources were first summarized with descriptive statistics. Their relationship with sociodemographic data (gender, age, schooling, occupation, level of linguistic competence, and socioeconomic level) were analyzed using the statistical program SPSS Statistics 21 with different methods. For the relationship between the number, quantities, and frequencies of species used and the socioeconomic level, first, the normality and homogeneity of variances were confirmed. Then, the variation of data was analyzed with an ANOVA. Finally, the relationships between knowledge, gender, age, occupation, socioeconomic level, linguistic competence, and schooling were calculated with a multiple correlation analysis. For the sample of school children, normality and homogeneity of variance of the data on the number of known species and the number of years of schooling were confirmed. After an analysis of variance (ANOVA), we asked if girls or boys knew more species of quelites with a t-test.

Results General Observations on the Food System The families usually had two or three formal meals a day (breakfast, lunch – usually in the afternoon – and dinner). The food for breakfast and lunch was similar, generally consisting of a noodle-tomato soup, eggs prepared in various ways, beans, rice, quelites, cactus pads (nopales), potatoes, lentils, soya, chicharrón (pork rind) in chili sauce, sardines, tuna, zucchini, and other vegetables. Maize tortillas were always present. Beverages included coffee, herbal tea, water, or soft drinks for breakfast, whereas soft drinks usually accompanied lunch. Some families did not have a formal dinner, others had bread or sweet bread with coffee, milk, or atole (a liquid maize gruel with sugar and other condiments), and some ate leftovers from the day. Meat was eaten 5–6 times in the 3 weeks per family, that is, once or twice a week. The most common meat was chicken, followed by beef, pork, fish and, rarely, mutton.

Quelite Species and Sources Families in Tetlazinga knew and consumed 35 species of quelites; they belonged to 29 genera and 17 botanical families (Table 1). Of these, only six were regularly bought in local fixed or weekly markets. Most species were obtained from maize fields (milpa) (Fig. 2). Examples were pigweed (quintonil, Amaranthus hybridus), lamb’s quarter (quelite cenizo,

440

C. Sa´nchez-Ramos et al.

Fig. 2 Sources of 33 species of quelites (two species not used by the selected families are omitted)

Chenopodium berlandieri), turnip (nabo, Brassica rapa), and quelite de borrego (Calandrinia ciliata). The second most common source of quelites were the homegardens, particularly for condiments: coriander (cilantro, Coriandrum sativum; epazote, Dysphania ambrosioides) and spearmint (hierba buena, Mentha spicata). Other plants obtained from homegardens were potato greens (Solanum tuberosum), chayote (Sechium edule), and chard (acelga, Beta vulgaris). The gardens of family and neighbors were another source of quelites, usually as a gift. Examples were chayote, quelite de cinco hojas (Cleome magnifica), and cabbage leaves (Brassica oleracea). Not all family members liked Cleome because of its bitter taste. However, some people reported specific cravings for this plant, despite the fact it requires many hours of cooking to make it edible. Another, but less common source of quelites was the wild or non-cultivated secondary vegetation. Examples of species collected in this type of place were hierba mora (Solanum nigrescens) and agave flower buds (Agave atrovirens). Finally, a few species in high demand or not available in the region were bought from neighbors, or at fixed or weekly markets (tianguis). Examples of these species were: izote flowers (Yucca guatemalensis), chard, pápalo (Porophyllum ruderale), and pipitza (Porophyllum linaria).

Preserving Healthy Eating Habits: Quelites in the Food System of. . .

441

Preparation The plant parts commonly used for food were leaves and stems; flowers and young shoots were less frequent. Preparation was, in order of importance, by boiling, frying, blanching, only washing and consuming raw, and roasting (Fig. 3). Some recipes required two types of preparation, for example, first boiling or blanching, then frying. If the quelites were boiled, salt and a piece of onion or garlic were usually added. They were also cooked with eggs or other foods, as well as in tomato or other sauces. Some recipes called for a combination of species. Fried quelites were usually prepared with onion and chili. Blanched quelites were cooled with water and then formed into balls; these were eaten with salt and lemon juice. Some species were eaten raw directly, outside in the field, or washed if consumed at home. Only the flowers of Beschorneria yuccoides were roasted on a comal (griddle), covered with a maize tortilla. When they started to release the juice, some salt was added, they were left for a few more minutes and then eaten in tacos with maize tortillas. Preparations of similar recipes varied between families. Comments by the interviewees pointed to active learning and instruction between generations, as well as modifications to recipes. We also documented innovation: one woman, who had lost her husband and had to find ways to feed a large family with very little money,

Fig. 3 Dishes. (a) Sowthistle salad; (b) Soup of squash shoots and flowers; (c) Huauzontle in tomato sauce; and (d) Yucca flowers with scrambled eggs. (Photos by the authors)

442

C. Sa´nchez-Ramos et al.

invented a very good-tasting salad of young sowthistle leaves, tomato, onion, and chili with no model for a similar dish within the community. Purchased vegetables fell into two groups. The first group consisted of tomatoes (Solanum lycopersicum), husk tomatoes (Physalis philadelphica), onions, garlic, and chili, which were needed practically daily for sauces and condiment and had a long history, but were rarely the basis of dishes. These were not included in this study because of the difficulties in quantifying their use. The second group included the other commercial vegetables, which were introduced only recently and bought occasionally. They were mostly obtained at the market in Ciudad Mendoza, Veracruz. The commercial vegetables (zucchini, cauliflower, head cabbage, green beans, lettuce, and radishes, mainly) were mostly boiled or cooked in sauces with other foods in different ways, including with meat. For example, cauliflower was boiled with potatoes in a tomato sauce with dry guajillo chili. The zucchinis were eaten a la mexicana, that is, with tomatoes, onion, and chili; green beans and head cabbage was cooked with chicken or beef. Lettuce and radishes were eaten raw with lemon juice and salt.

Frequency and Quantity of Consumption Based on the recall data, each family consumed on average about 10 kg of fresh quelites (untrimmed) in two meals per week during the main season for most species (May–July). That is, each quelite meal consisted of between 3 and 4 kg of these plants (fresh, after trimming). If divided by the number of family members (74 for the 20 families, including children under 10 years old), the per capita consumption was 1.6–2.2 kg per week or 230–310 g per day during the quelite season. Amounts consumed varied strongly between families (Fig. 4), as they depended on family tastes and composition (minimum 1.5 kg and maximum 22 kg per week) with a slight correlation between the number of family members and amounts consumed. The main quelites recorded in the three-week food diary were Chenopodium berlandieri, Amaranthus hybridus, flowers and shoots of squash (Cucurbita ficifolia), Calandrinia ciliata, Brassica rapa, flowers and shoots of runner bean (Phaseolus coccineus), tender shoots of chayote, sowthistle (Sonchus oleraceus), potato leaves, Beta vulgaris, and Cleome magnifica. No new quelite species were recorded in this data set. Quelites were consumed much more frequently than purchased vegetables (Fig. 5) and were the main vegetables during the first half of the rainy season. Quelite consumption and socioeconomic level. Families with fewer economic resources knew and consumed more species on average (X ¼ 15.4) than families with medium (X ¼ 13.0) and relatively high resources (X ¼ 13.2). However, this difference was not statistically significant ( p ¼ 0.442) with an ANOVA, based on the first group of data (recall interviews). The data on consumption frequency (food diary) and quantity (recall) also showed no significant differences.

Preserving Healthy Eating Habits: Quelites in the Food System of. . .

443

Fig. 4 Monthly consumption of 24 quelite species – average and standard deviation. The species used for condiment and purchased species were not included (see methods)

Fig. 5 Consumption frequency of quelites (collected and purchased) and the group of recently introduced vegetables in the three-week food diary of 19 families

Relationship Between Sociodemographic Factors and Quelite Knowledge and Use A multiple correlation analysis of the number of species known to the interviewees found low but significant correlations related to gender (R2 ¼ 0.254, p < 0.001), age (R2 ¼ 0.201, p < 0.001), and occupation (R2 ¼ 0.075, p ¼ 0.007). However, no correlation was found with the socioeconomic level (R2 ¼ 0.003, p ¼ 0.559),

444

C. Sa´nchez-Ramos et al.

understood or spoken languages (R2 ¼ 0.005, p ¼ 0.474), and years of schooling (R2 ¼ 0.000, p ¼ 0.828). The total model, including the statistically non-significant variables, had a corrected R2 of 0.479; that is, it explained about half of the variation found. Most families valued the quelites and thought that they were tasty and healthy. Eleven of the 20 family cooks mentioned unprompted that talks at the clinic (which were obligatory for the recipients of some governmental programs) had emphasized the importance of consuming vegetables and specifically quelites.

Quelite Preferences and Knowledge Among Children and Young People In the comments on preferences, less than 10% (12 of 131) of the interviewed students said they did not like quelites. The main reason given was their strong taste. Their favorite species were Amaranthus hybridus, Chenopodium berlandieri, Brassica rapa, Rumex crispus, Calandrinia ciliata, and Beta vulgaris, in this order, which corresponds well with the consumption data of the families. Primary school students knew a higher number of species on average (X ¼ 8.43) than secondary school (X ¼ 6.62) and high school students (X ¼ 4.66). The difference between the three levels of schooling was significant (F2, 128 ¼ 15.107, p < 0.001) with an ANOVA. The difference had a p ¼ 0.023 between primary and secondary school students, while all the other differences had a p equal to or lower than 0.001. Males had less knowledge of species of quelites on average (X ¼ 5.43) than females (X ¼ 6.97). This difference was confirmed as significant with a t-test.

Discussion Quelite Species The number of quelite species consumed by the population of Tetlatzinga (35 species) was close to that of regions with a similar temperate-humid climate. In Naupan, a village in the humid mountains of northern Puebla, 37 species were recorded as quelites, and in Zoatecpan, Xochitlan de Vicente Suárez, in the same Sierra Norte de Puebla, 36 species (Basurto et al. 2011; Molina 2000). A mountainous area near Mexico City on the slopes of the Sierra Nevada around Ozumba had 35 species (Linares et al. 2017). Mota-Cruz et al. (2011) found a total of 46 species of quelites in Tepepan de Zaragoza and La Guacamaya, a region in Oaxaca south of our study area, but these were two villages with different ethnic groups. In Talea de Castro, also in a very humid area of northern Oaxaca, about half of 70 collected wild food species were leafy vegetables (Manduna 2008). These data appear to indicate that there is some regularity to the number of species used as quelite in this vegetation type.

Preserving Healthy Eating Habits: Quelites in the Food System of. . .

445

The leaves of kuilakochi (Bidens triplinervia), shepherd’s purse (kilkolex; Capsella bursa-pastoris), chickweed (momatilana; Stellaria media), and oka (Oxalis tuberosa) are new records for quelites in Mexico. However, Capsella and Stellaria are known as edible in their region of origin, Europe (Paoletti et al. 1995; Tardío et al. 2006). The high regard for the flowers of various agave relatives (maguey, Agave atrovirens; izote, Yucca guatemaltensis; and patahmolkahkaya, Beschorneria yuccoides) was notable.

Sources of Quelites The milpa is not always the main source of quelites in Mexico; in other studies, the home garden was more important than the maize field (Mota-Cruz et al. 2011). In the study area, home gardens provided cultivated quelites – species that were not easily available otherwise. This division confirms the observation by Larios et al. (2013), that home gardens are spaces to obtain goods that are not available naturally. Perhaps the large number of these plants in maize fields was due to the humid climate. The culinary importance of quelites was underscored by the fact that they were both exchanged and sold. Some were given to or obtained for free from relatives or neighbors; this is a common practice (Blanckaert et al. 2004) and strengthened ties of friendship and family. However, particularly high-value species were sold even within the social network. Quelites, both cultivated and collected, were also acquired in markets. This confirms observations of other authors that these plants are widely traded, for example, in the weekly markets of the Tehuacán-Cuicatlán Valley (Arellanes and Casas 2011; Arellanes et al. 2013), Nanacamilpa (González-Amaro et al. 2009), and Ixtlahuaca (Vieyra-Odilon and Vibrans 2001).

Preparation In the study area, the quelites were usually the main focus of the meal and were cooked alone or in combination with other quelites, vegetables, eggs, and cheese, but seldom with meat, except for one recipe that calls for the combination of purslane and pork. In other regions of Mexico, quelites are considered more of a side dish or additional component of a meat-centered combination. For example, in San Cristobal de las Casas, Chiapas, people almost always cook quelites with chicken or beef (own observation). In San Bartolo de Llano, Mexico State, both pigweed and purslane are prepared with pork (Linares and Aguirre 1992; Vieyra-Odilon and Vibrans 2001). These differences in the role of these plants in the meals are interesting and should be explored comparatively. Perhaps the reason is economic or ecological, but it is also possible that these practices are the result of deeper cultural traits, particularly as introduced vegetables are paired with meat more commonly in the study area. However, some preparation methods, such as frying quelites with onion and chili, are common elsewhere in central Mexico; also, the preferred species were similar to other regions (Linares et al. 2017).

446

C. Sa´nchez-Ramos et al.

In some other regions, quelites are used to make fresh beverages (agua fresca) to drink with meals. We did not record this use in our study area. The experience of the first author in the region indicates that, apparently, there is no agua fresca tradition; historically, coffee or hot herbal tea was prepared in the morning and evening, and plain water accompanied daytime meals. There were a few other preparations, for example, chilacayote (Cucurbita ficifolia, a squash) cooked with water and sugar, or a nutritious beverage (atole) based on maize dough sweetened with agave juice (aguamiel). Currently, commercial soft drinks have replaced plain water.

Amounts Consumed The quantities of consumed quelites were relatively high. Basurto et al. (2011) found a daily per capita consumption of on average only 130 g in an ecologically similar area in the Sierra Norte de Puebla, though the frequency (2–3 times per week) was similar. The Rarámuri of Cuiteco, Chihuahua, eat between 11.5 and 51.8 kg per family of four important species (Camou-Guerrero et al. 2008). Casas et al. (2014) recorded annual consumption of almost 17 kg of watercress (Rorippa nasturtiumofficinale), but only 0.41 kg of the condiment Porophyllum linaria, among the Cuicatecos of San Lorenzo Pápalo, Oaxaca. In San Bartolo de Llano, State of Mexico, the consumption of quelites per month per family averaged 4.5 kg (Vieyra-Odilon and Vibrans 2001). In Santiago Quiotepec, Oaxaca, average annual consumption per household ranged from 1 to 3 kg per species (Pérez-Negrón and Casas 2007). In other parts of the world, for example, in two villages in Zimbabwe, on average, families consumed 210 g of wild plants 1–5 times per week (Manduna 2008). The World Health Organization (2003) and the United States Department of Agriculture recommend the consumption of about 400 g of vegetables and fruit daily. There are few concrete recommendations for leafy green vegetables, but USDA does recommend a minimum of 2 cups (equivalent to 400–500 g) per week (http://www.choosemyplate.gov/vegetables). This means that the average 230–320 g per capita daily consumption in our study area far exceeds the recommended minimum, and even low-consuming families were likely to cover their needs. Indeed, during the season, quelites covered almost the entire vegetable requirement. If the other bought vegetables are considered (tomatoes, onions, cabbages, etc.), the diet of the population of the study area was quite remarkable for the amount of consumed vegetables. Newly introduced, commercial vegetables did not play a large role in the diet in summer and, up to now, did not appear to be preferred or to replace the local quelites. Such replacement has been reported from Costa Rica, for example (González 2008). In Zimbabwe, Africa, some elderly people commented that modern vegetables were replacing traditional wild vegetables because young people prefer the taste of (modern) cultivated vegetables (Manduna 2008). Though we found similar sentiments among some schoolchildren, it was not wide-spread.

Preserving Healthy Eating Habits: Quelites in the Food System of. . .

447

Socioeconomic and Sociodemographic Factors and Quelite Knowledge and Consumption Some of the expected factors played a role in quelite knowledge, such as gender. Both adult and young females knew more species than males, presumably because of their role in meal preparation. Young boys had more contact with their fathers, who spent most of their time building furniture. The role of age was also confirmed: older people generally knew more, similar to, for example, Isthmus Zapotecs’ plant knowledge (Saynes-Vásquez et al. 2016). In Niger, a curvilinear relationship between age and knowledge about herbaceous and woody species was found. The authors suggest that medicinal species knowledge decreases after a certain age (Ayantunde et al. 2008). We did not find this, perhaps because quelites were used more often and also included fewer species than medicinal. Likewise, occupation is known to play a significant role in the knowledge of species. Saynes-Vásquez et al. (2016) found that the farmers, hunters, or collectors of firewood knew more names of plants than storekeepers or professionals, as did Martínez-Ballesté et al. (2006) in a different context. However, unexpectedly, the consumption of quelites was not related to socioeconomic levels in the study area, similar to the results of Vieyra-Odilon and Vibrans (2001). Most interviewees considered them a desirable food source. Any possible negative attitudes had apparently been countered by the recommendations of the local health professionals. So, relatively simple interventions may reverse the negative image of these foods. If quelite use declines in this area in the future, it will probably be due to the abandonment of the milpa system for economic reasons, rather than cultural attitudes. Also, bilingualism – as a proxy for the strength of the indigenous heritage – was not a relevant factor in the knowledge of species. In the Isthmus of Tehuantepec, the opposite was true for general wild plant knowledge, with a positive correlation between Zapotec linguistic competence and knowledge of plants (Saynes-Vásquez et al. 2013; SaynesVásquez et al. 2016). Again, this may be due to the different cognitive effort required to know a few edible plants versus many wild ones. Also, our study area was more uniform culturally – all inhabitants still spoke Nahuatl, and the majority was bilingual. Years of schooling did not influence species knowledge among the families of Tetlatzinga. However, among students, age/years of schooling were negatively related with the knowledge of quelites. Estrada-Castillón et al. (2014) reported a similar phenomenon: children knew more useful species than young people, knowledge then increased in adults and decreased in the elderly. Apart from changing interests, teenagers are sometimes reluctant to admit knowledge of concepts they consider old-fashioned.

Conclusions and Perspectives The tradition of quelite consumption was alive in our study region. The number of species used was similar to earlier studies of comparable regions. The fact that maize fields were the most important source of these plants underscores the role of these

448

C. Sa´nchez-Ramos et al.

fields as a provider not only of macro- but also of micronutrients, often neglected in productivity evaluations. The recipes were diverse, and there was active innovation. We did not find the expected indicators of cultural change in people’s relationship with quelites in the study area. Relatively well-off people consumed just as many as poor people; there were few signs of newly introduced vegetables being preferred to the traditional ones. Quelites were not looked upon as inferior by the general population. We suggest that simple reinforcement by authority figures can play a role in the maintenance of these healthy eating traditions. The fact that quelites were considered a main and not a side dish in the region should be explored more intensively. This study shows that quantitative data are highly informative and provide context for understanding peoples’ decisions. They may also contradict commonly held beliefs of investigators. Acknowledgments We thank the people and the authorities of Tetlatzinga for their time, cooperation, and help. Dr. Abisaí García helped with the identification of Asparagaceae (Agave). This study was funded by the Consejo Nacional de Ciencia y Tecnología, Mexico (CONACYT; grant number 638929) through a grant for living expenses to the first author for her graduate studies.

References Arellanes Y, Casas FA. Los mercados tradicionales del Valle de Tehuacán-Cuicatlán: antecedentes y situación actual. Nueva Antropol. 2011;24(74):93–123. Arellanes Y, Casas A, Arellanes-Meixueiro A, Vega E, Blancas J, Vallejo M, Torres I, Solís L, Pérez-Negrón E. Influence of traditional markets and interchange on plant management in the Tehuacán Valley. J Ethnobiol Ethnomed. 2013;9:38. Ayantunde AA, Briejer M, Hiernaux P, Udo HMJ, Tabo R. Botanical knowledge and its differentiation by age, gender and ethnicity in Southwestern Niger. Hum Ecol. 2008;36(6):881–9. Basurto F. Los quelites de México: especies de uso actual. In: Mera-Ovando LM, Castro-Lara D, Bye RA, editors. Especies vegetales poco valoradas: una alternativa para la seguridad alimentaria. México, D. F.: Universidad Nacional Autónoma de México, Servicio Nacional de Inspección y Certificación de Semillas, Sistema Nacional de Recursos Filogenéticos para la Agricultura y la Alimentación; 2011. p. 23–45. Basurto F, Evangelista V, Molina-Martínez N, Alvarado-Flores R. Frecuencia de consumo de quelites en la Sierra Norte de Puebla. In: Mera-Ovando LM, Castro-Lara D, Bye RA, editores. Especies vegetales poco valoradas: una alternativa para la seguridad alimentaria. México, DF: Universidad Nacional Autónoma de México, Servicio Nacional de Inspección y Certificación de Semillas, Sistema Nacional de Recursos Filogenéticos para la Agricultura y la Alimentación; 2011. p. 23–45. Basurto-Peña F, Martínez MA, Villalobos G. Los quelites de la Sierra Norte de Puebla, México: inventario y formas de preparación. Bol Soc Botánica México. 1998;62:49–62. Blanckaert I, Swennen RL, Paredes Flores M, Rosas López R, Lira SR. Floristic composition, plant uses and management practices in homegardens of San Rafael Coxcatlán, Valley of TehuacánCuicatlán, Mexico. J Arid Environ. 2004;57(2):179–202. Bye RA. Quelites – ethnoecology of edible greens - past, present, and future. In: Minnis PE, editor. Ethnobotany: a reader. Oklahoma: University of Oklahoma Press; 2000. p. 197–213.

Preserving Healthy Eating Habits: Quelites in the Food System of. . .

449

Bye R, Linares E. Los quelites plantas comestibles de Mexico, una reflexion sobre intercambio cultural. Biodiversitas. 2000;31:11–4. Camou-Guerrero A, Reyes-García V, Martínez-Ramos M, Casas A. Knowledge and use value of plant species in a Rarámuri community: a gender perspective for conservation. Hum Ecol. 2008;36(2):259–72. Casas A, Otero-Arnaiz A, Perez-Negron E, Valiente-Banuet A. In situ management and domestication of plants in Mesoamerica. Ann Bot. 2007;100(5):1101–15. Casas A, Camou A, Otero-Arnaiz A, Rangel-Landa S, Cruse-Sanders J, Solís L, et al. Manejo tradicional de biodiversidad y ecosistemas en Mesoamérica: el Valle de Tehuacán. Investig Ambient. 2014;6(2):23–44. Castro-Lara D, Alvarado-Flores R, Evangelista-Oliva V. Recetario de quelites de la Sierra Norte de Puebla. Mexico City: Instituto de Biología, Universidad Nacional Autónoma de Mexico, Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2005. 109 p. Estrada-Castillón E, Garza-López M, Villarreal-Quintanilla J, Salinas-Rodríguez M, Soto-Mata B, González-Rodríguez H, et al. Ethnobotany in Rayones, Nuevo León, México. J Ethnobiol Ethnomed. 2014;10(1):62. García E. Modificaciones al sistema de clasificación climática de Köppen. México City: Universidad Nacional Autónoma de México; 1987. González AR. De flores, brotes y palmitos: alimentos olvidados. Agron Costarric. 2008;32(2): 183–92. González-Amaro RM, Martínez-Bernal A, Basurto-Peña F, Vibrans H. Crop and non-crop productivity in a traditional maize agroecosystem of the highland of Mexico. J Ethnobiol Ethnomed. 2009;5(1):38. Juárez R. Desarrollo histórico de la comunidad de Huitzila, Municipio de Soledad Atzompa, Veracruz. Veracruz: Comisión Nacional para el Desarrollo de los Pueblos Indígenas y H. Ayuntamiento Constitucional Soledad Atzompa; 2007. Larios C, Casas A, Vallejo M, Moreno-Calles A, Blancas J. Plant management and biodiversity conservation in Náhuatl homegardens of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013;9(1):74–89. Linares E, Aguirre J. Los quelites, un tesoro culinario. México City: Instituto de Biología, Universidad Nacional Autónoma de México; 1992. 143 p. Linares HME, Bye BR, Ortega NN, Arce VAE. Quelites: sabores y saberes del sureste del Estado de México [Internet]. Universidad Nacional Autónoma de México, Instituto de Biología; 2017 [citado 12 de febrero de 2020]. Disponible en: http://www.ibiologia.unam.mx/barra/ publicaciones/Recetario%20final-3.pdf Łuczaj Ł, Pieroni A, Tardío J, Pardo-de-Santayana M, Sõukand R, Svanberg I, et al. Wild food plant use in 21st century Europe: the disappearance of old traditions and the search for new cuisines involving wild edibles. Acta Soc Bot Pol. 2012;81(4):359–70. Madamombe-Manduna I, Vibrans H, Vázquez-García V. Género y conocimientos etnobotánicos en México y Zimbabwe. Un estudio comparativo. Soc Rural Prod Medio Ambiente. 2009;6(18): 21–48. Manduna IT. Etnobotánica comparativa de plantas comestibles recolectadas en sistemas de agricultura tradicional de México y Zimbabwe [Internet] Ph.D. thesis. Montecillo, Texcoco, México: Colegio de Postgraduados; 2008 [citado 20 de agosto de 2014]. Disponible en: http:// www.biblio.colpos.mx:8080/xmlui/handle/10521/1521 Mapes C, Caballero J, Espitia E, Bye RA. Morphophysiological variation in some Mexican species of vegetable Amaranthus: evolutionary tendencies under domestication. Genet Resour Crop Evol. 1996;43(3):283–90. Mapes C, Basurto F, Bye R. Ethnobotany of quintonil: knowledge, use and management of edible greens Amaranthus spp. (Amaranthaceae) in the Sierra Norte de Puebla, México. Econ Bot. 1997;51(3):293–306.

450

C. Sa´nchez-Ramos et al.

Martínez-Ballesté A, Martorell C, Caballero J. Cultural or ecological sustainability? The effect of cultural change on sabal palm management among the lowland Maya of Mexico. Ecol Soc. 2006;11(2):27. Mera-Ovando LM, Castro-Lara D, Bye-Boettler R, editores. Especies vegetales poco valoradas: una alternativa para la seguridad alimentaria. México DF: Universidad Nacional Autónoma de México, Servicio Nacional de Inspección y Certificación de Semillas, Sistema Nacional de Recursos Filogenéticos para la Agricultura y la Alimentación; 2011. 213 p. Molina MN. Etnobotánica de quelites en el sistema milpa en Zoatecpan, una comunidad indígena nahuat de la Sierra Norte de Puebla. [Internet] Tesis de licenciatura México, D F: Universidad Nacional Autónoma de México; 2000. Disponible en: http://oreon.dgbiblio.unam.mx/F/? func¼service&doc_library¼TES01&doc_number¼000281314&line_number¼0001&func_ code¼WEB-BRIEF&service_type¼MEDIA Mota-Cruz C, Vibrans H, Ortega-Paczka R, Koch SD. Quelites entre nahuas y mazatecos en una región de bosque mesófilo del sureste de Puebla, México. In: Mera-Ovando LM, Castro-Lara D, Bye-Boettler R, editors. Especies vegetales poco valoradas: una alternativa para la seguridad alimentaria. México DF.: Universidad Nacional Autónoma de México, Servicio Nacional de Inspección y Certificación de Semillas, Sistema Nacional de Recursos Filogenéticos para la Agricultura y la Alimentación; 2011. p. 85–101. Paoletti MG, Dreon AL, Lorenzoni GG. Pistic, traditional food from Western Friuli, N.E. Italy. Econ Bot. 1995;49(1):26–30. Pérez-Negrón E, Casas A. Use, extraction rates and spatial availability of plant resources in the Tehuacán-Cuicatlán Valley, Mexico: the case of Santiago Quiotepec, Oaxaca. J Arid Environ. 2007;70(2):356–79. Reddy KS, Katan MB. Diet, nutrition and the prevention of hypertension and cardiovascular diseases. Public Health Nutr. 2004;7(1a):167–86. Rendón B, Núñez–Farfán J. Population differentiation and phenotypic plasticity of wild and agrestal populations of the annual Anoda cristata (Malvaceae) growing in two contrasting habitats. Plant Ecol. 2001;156(2):205–13. Rendón B, Bye R, Núñez-Farfán J. Ethnobotany of Anoda cristata (L.) Schl. (Malvaceae) in Central Mexico: uses, management and population differentiation in the community of Santiago Mamalhuazuca, Ozumba, state of Mexico. Econ Bot. 2001;55(4):545–54. Robertson CE. McCance and Widdowson’s the composition of foods – sixth summary edition. Nutr Bull. 2003;28(1):81–3. Sadler MJ, Strain JJ, Caballero B. Encyclopedia of human nutrition. Amsterdam, Boston: Elsevier/ Academic; 2005. Sánchez RC. Diversidad de plantas y animales en los huertos familiares de la localidad de Tetlatzinga, Municipio de Soledad Atzompa, Veracruz, México [Tesis de licenciatura]. [Jalapa, Veracruz]: Universidad Veracruzana; 2014. Saynes-Vásquez A, Caballero J, Meave JA, Chiang F. Cultural change and loss of ethnoecological knowledge among the Isthmus Zapotecs of Mexico. J Ethnobiol Ethnomed. 2013;9(1):40. Saynes-Vásquez A, Vibrans H, Vergara-Silva F, Caballero J. Intracultural differences in local botanical knowledge and knowledge loss among the Mexican Isthmus Zapotecs. PLoS One. 2016;11(3):e0151693. Scott JC. The art of not being governed: an anarchist history of upland Southeast Asia, Yale agrarian studies series. New Haven: Yale University Press; 2009. 442 p SEFIPLAN (Secretaria de Finanzas y Planeación del Estado de Veracruz). Sistema de información municipal. Cuadernillos municipales, 2014. Soledad Atzompa [Internet]. consultado Marzo 2015; 2014. Disponible en: http://www.veracruz.gob.mx/finanzas/files/2013/04/SoledadAtzompa.pdf Tardío J, Pardo-De-Santayana M, Morales R. Ethnobotanical review of wild edible plants in Spain. Bot J Linn Soc. 2006;152(1):27–71.

Preserving Healthy Eating Habits: Quelites in the Food System of. . .

451

Vázquez-García V, Godínez Guevara L, Montes-Estrada M, Montes Estrada M, Ortiz Gómez AS. The quelites of Ixhuapan, Veracruz: availability, supply and consumption. Agrociencia. 2004;38(4):445–55. Vibrans H. Ethnobotany of Mexican weeds. In: Lira R, Casas A, Blancas J, editors. Ethnobotany of Mexico [Internet]. New York, Springer New York; 2016. p. 287–317. Vieyra-Odilon L, Vibrans H. Weeds as crops: the value of maize field weeds in the Valley of Toluca, Mexico. Econ Bot. 2001;55(3):426–43. World Health Organization (WHO). Dieta, nutrición y prevención de enfermedades crónicas. Informe de una Consulta Mixta de Expertos OMS/FAO. Geneva, Switzerland; 2003. (Serie de Informes Técnicos # 916.).

Ethnobotany of the Nahua People: Plant Use and Management in the Sierra Negra, Puebla, Mexico Jose´ Juan Blancas Vázquez, Alejandro Casas, Hilda Ramírez-Monjaraz, Andrea Martínez-Balleste´, Ignacio Torres-García, Itzel Abad-Fitz, Leonardo Beltra´n-Rodríguez, Carolina Larios, Aketzalli Olvera-Espinosa, Myriam A. Miranda-Gamboa, Elisa Lotero, and Mariana Vallejo Abstract

The Sierra Negra region is a rich mosaic of ecosystems and cultures in interaction. There, people practice a pattern of multiple use and management of the different plant species and vegetation types occurring in their communitarian territories and, through interchange, those of other communities of the region. Due to a complex socio-ecological history, characterized by a relative isolation of the region for a long time, people inhabiting the area generated strategies of using the diverse environments to mitigate the sources of climate, social, and economic uncertainties. The strategies of risk management included creative forms to J. J. Blancas Vázquez (*) Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico e-mail: [email protected] A. Casas Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] H. Ramírez-Monjaraz Facultad de Estudios Superiores Iztacala – Universidad Nacional Autónoma de México, Tlalnepantla, Estado de México, Mexico e-mail: [email protected] A. Martínez-Ballesté Jardín Botánico-Instituto de Biología, Universidad Nacional Autónoma de México, Coyoacán, Ciudad de México, Mexico e-mail: [email protected] I. Torres-García Escuela Nacional de Estudios Superiores (ENES), Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_19

453

454

J. J. Blancas Vázquez et al.

achieve food security (use of numerous species and ecosystems, different types of management of species, including cultivation and domestication, techniques for food conservation and storing, among others) and, in general, using their own resources to satisfy their basic subsistence needs without depending on external regions. Until the present, ethnobotanical studies by our research group have identified 496 plant species, which constitute the useful flora of the region, but more studies are still needed, and this cipher surely will increase. In the current inventory, the richest plant families providing useful species are Asteraceae, Fabaceae, Asparagaceae, Solanaceae, and Rosaceae, which are used, in order of importance, as ornamental, medicine, food, and fuel. We have also documented a strategy of multiple use of plant resources, each species having 1 to 16 uses (3.48 uses on average per species), and each vegetation type providing on average 161 useful species. Also, we have identified a pattern of eco-symbiotic complementarity among environments and vegetation types, which has been constructed through time as a response to uncertainty in the availability of resources due to ecological and socioeconomic factors. Strategies to deal with ecological and social uncertainty have been identified and documented to include management and incipient domestication of some species in forest and agroforestry systems, as well as interchange among communities living in different ecosystems. This research analyzes the deep knowledge of the regional people about ecological relations of plants, their properties associated with use, the management types operating on them, and requirements and consequences associated with management. We recognize the still limited research conducted in the area and the need of continuing ethnobotanical studies; we identify some priorities. The inventory of useful plants and their management is still limited while significant areas of diversity of ecosystems and cultures of the region remain unexplored. The Sierra Negra constitutes an important refuge of biocultural diversity that contributes valuable lessons on plant and ecosystem management. Such experience deserves

I. Abad-Fitz Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico e-mail: [email protected] L. Beltrán-Rodríguez Laboratorio de Etnobotánica Ecológica, Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico e-mail: [email protected] C. Larios · A. Olvera-Espinosa Instituto de Investigaciones en Ecosistemas y Sustentabilidad – Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected]; [email protected] M. A. Miranda-Gamboa · E. Lotero · M. Vallejo Jardín Botánico, Instituto de Biología – Universidad Nacional Autónoma de México, Ciudad de México, Mexico e-mail: [email protected]; [email protected]; [email protected]

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

455

to be understood for constructing sustainable forms of human-nature interactions for ensuring the future of new generations of local and regional people.

Introduction The Sierra Negra is a mountainous region located in the south-eastern part of the state of Puebla, neighboring the semiarid Tehuacán-Cuicatlán Valley at west and the wet coastal plain of the Gulf of Mexico at east. It is part of the Southern Sierra Madre, an important mountain chain traversing Mexico from east to west in the south of Mexico (Morrone 2017). The Sierra Negra region has an extent of nearly 5000 km2, with some of the highest elevations of Mexico like the Sierra Negra volcano, which reaches 4600 m, and the Zitzintépetl mountain which is 3600 m high (INEGI 2020). It is a zone with high biocultural diversity, which, since pre-Columbian times, has been the settlement area of diverse cultural groups that significantly modified the natural environment. Although several archaeological sites have been registered in the region, these have been poorly studied (Sepúlveda 2005), compared with those of the TehuacánCuicatlán Valley like the Maize Cave and the Purrón Dam, or the fortification of the hill of Cuthá in Zapotitlán Salinas. Records of the pre-Columbian occupation of the zone are inferred from the stone and ceramic anthropomorphic figures, which are called xantiles and that are commonly found by people when preparing land for agriculture. There are, in addition, chronicles and codices from the colonial period referring to historical facts about the main settlement of the region (Sepúlveda 2005). The huge churches in Coyomeapan and Zoquitlán, the main towns of the region, are prints of the importance of the region during the colonial period. Currently, the Sierra Negra is a rich cultural mosaic where Nahua, Mazatec, Mixtec, Popoloca, and Mestizo people live (INEGI 2020). All these cultures have particularities but share similarities in forms of using and managing the local forests. These are based on the multiple use of species, conforming local landscapes characterized by the presence of milpas, homegardens, coffee plantations, several types of agroforestry systems combining fruit trees (mainly apple and quince) with milpa, and patches of forest components, as well as conserved areas of diverse forest types (Mota 2007; Blancas et al. 2013; Larios et al. 2013; Vallejo et al. 2015; Olvera 2016). The region has a high diversity of vegetation types, outstandingly the extended cloud forest, tropical rain forest, and several types of associations of pine-oak, pine, tropical dry, thorn-scrub forests, and riparian vegetation (INEGI 2020). Such diversity of ecosystems has favored the development of different strategies of ecosystem management, their multiple use, and complementarity, which has supported mechanisms of socioecological resilience (Blancas et al. 2013). The relief of the mountainous landscape of the area is rugged, with pronounced slopes determining marked elevation gradients and changes of vegetation types in short distances. The environmental heterogeneity advocates the presence of numerous endemic species, many of them poorly studied from both ethnobotanical and botanical perspectives (Blancas et al. 2013; Santiago-Alvarado et al. 2016).

456

J. J. Blancas Vázquez et al.

Despite the long history of human occupation, the region remained relatively isolated from the neighboring areas until the end of the twentieth century. By that time the contact increased with the Tehuacán-Cuicatlán Valley due to the construction of the main roads communicating Coyomeapan with the towns of Coxcatlán, Ajalpan, and Tehuacán, in the state of Puebla and Córdoba, in the state of Veracruz. Such isolation for centuries enhanced local people to be self-sufficient, developing several strategies to achieving it. These included the establishing of nets of interchange and commercialization among towns within the Sierra Negra and, although at low intensity, with towns outside the region. Interchange involved diverse products from agriculture and, especially, from the forests of the region (Blancas et al. 2013; Lotero-Velásquez et al. 2022). A high diversity of products from the Sierra Negra are currently interchanged with the Tehuacán-Cuicatlán region through the important markets of Tehuacán, Ajalpan, Zinacatepec, Coxcatlán, Tecomavaca, Cuicatlán, and Córdoba (Arellanes et al. 2013). These interchanges occurred in the past but were less frequent and involving small amounts of products. The long and continuous interaction of local people and ecosystems has determined complex systems of knowledge, management, and classification of environmental units, plants, and animals (Mota 2007; Blancas et al. 2013; Zarazúa 2020, 2022; Lotero-Velásquez et al. 2022). However, there is no ethnobotanical work synthesizing and analyzing the whole regional information. This chapter aims to summarize the general panorama of the useful flora of the region recorded until present, aspiring to identifying general patterns and main topics, as well as the areas of the region that should be sampled for an integral assessment of the ethnobotany of the Sierra Negra.

Study Site This research was conducted in eight communities of the municipality of Coyomeapan, Puebla. The communities are Ahuatla, Aticpac, Caxalli, Chimalhuaca, Ixtacxochitla, Santa María Coyomeapan, Yohuajca, and Xocotla, which are representative of the environmental heterogeneity of the region (Fig. 1). These communities are located in an elevation and environmental gradient from 1045 to 2509 m. Annual mean temperatures vary from 10 to 24  C while the annual mean rainfalls vary from 1000 to 3000 mm. Climates are predominately temperate wet, with abundant rainfall during the summer; semiwarm wet with rains throughout the whole year; and warm humid, with rains the year round. Temperate subhumid climate with summer rains and semicold subhumid with summer rains and semiwarm subhumid with summer rain are all also present in the region in restricted areas (INEGI 2020). The predominant vegetation types are: (a) Pine-oak forest, in temperate areas, characterized by the dominance of Pinus ayacahuite C. Ehrenb. ex Schltdl., Pinus cembroides Zucc., Pinus leiophylla Schiede ex Schltdl. & Cham., Pinus montezumae Lamb., Pinus patula Schltdl. & Cham., Abies religiosa (Kunth) Schltdl. & Cham., Roldana candicans (Née) Villaseñor, S. Valencia, & Coombes, Alnus acuminata Kunth, and Cupressus lusitanica Mill.; (b) oak forest, also in temperate areas, with the dominance of Roldana candicans, Quercus laurina Bonpl., Quercus rugosa Née, Arbutus xalapensis Kunth and Fraxinus uhdei (Wenz.) Lingelsh.;

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

457

Fig. 1 Study area. The villages studied in the Sierra Negra, Puebla, Mexico. (Modified from Blancas et al. 2013)

(c) tropical rain forest, in warm humid areas, characterized by the presence of Cedrela odorata L., Chamaedorea tepejilote Liebm., Croton gossypiifolius Vahl, and Pouteria sapota (Jacq.) H.E. Moore & Stearn; (d) cloud forests, in semiwarm wet areas, dominated by Talauma mexicana (DC.) G. Don, Heliocarpus appendiculatus Turcz., Eugenia capuli (Schltdl. & Cham.) Hook. & Arn., Liquidambar styraciflua L., Platanus mexicana Moric., and Cecropia obtusifolia Bertol.; (e) tropical deciduous forest, in semiwarm subhumid areas, dominated by Lysiloma acapulcence (Kunth) Benth., Bursera grandifolia (Schltdl.) Engl., and Salix taxifolia Kunth; and (f) xerophytic thorn scrub forest, in semiwarm drier areas, with the presence of Agave angustifolia Haw, Dasylirion serratifolium (Karw. ex Schult. f.)

458

J. J. Blancas Vázquez et al.

Fig. 2 Vegetation types of the Sierra Negra, Puebla, Mexico: (a) Pine forest; (b) oak forest; (c) tropical evergreen forest; (d) cloud forest; (e) tropical deciduous forest; and (f) xerophytic thorn scrub forest. (Photos: José Blancas)

Zucc., Beaucarnea gracilis Lem., Pachycereus weberi (J.M. Coult.) Backeb. and Myrtillocactus geometrizans (Mart. ex Pfeiff.) Console (INEGI 2020) (Fig. 2). The communities studied are mostly inhabited by Indigenous people, speakers of five dialectal variants of Nahuatl (INALI 2022). Table 1 shows details of the environmental and socio-demographic aspects of these communities. All of them are ancient villages, with pre-Columbian occupation, and some were refunded after the Spanish invasion (Sepúlveda 2005). The oldest written document recording the colonial population of the area is from 1540, in which the territorial limits of the Encomiendas (a colonial institution that gave land, people, and labor force to the conquerors), and the tributes demanded to the subject peoples, were established. However, the territorial conformation of the current villages was clearly consigned until 1690 in the El Lienzo de San Juan Cuauhtla, a document that indicates the limits of the villages in relation to the main neighboring village of Coyomepan (Sepúlveda 2005).

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

459

Table 1 Socio-ecological aspects of the villages studied in the Sierra Negra, Puebla, Mexico Type of vegetation Pine-oak forest, oak forest

Altitude (masl) 2021

Number of inhabitants 554

Ethnic ascription Nahua

Aticpac

Tropical evergreen forest, cloud forest

1045

160

Nahua

Caxalli

Pine-oak Forest, Oak forest

2509

216

Nahua

Villages Ahuatla

Principal economic activities Seasonal agriculture, mainly self-supply (maize and peas) Cultivation of the Golden apple variety (Malus domestica) in agroforestry systems Chile canario cultivation (Capsicum pubescens) Firewood recollection Laurel harvest (Litsea glaucescens) Flor de tila recollection (Ternstroemia lineata) Sheep husbandry Seasonal agriculture, mainly self-supply (maize) cultivation of coffee (Coffea arabica), purple banana (Musa paradisiaca), and orange (Citrus  sinensis) Cultivation of quelite hoja ancha or nda tu ijné (Witheringia solanacea), and quelite huele de noche (nda dzóo, tzopelikílitl) (Cestrum nocturnum) Recollection and cultivation of tepejilote (Chamaedorea tepejilote) Tequelite recollection (Peperomia peltilimba) Recollection and cultivation of mototetl (Eugenia capuli) Orange liqueur preparation Flor de corazón recollection (Talauma mexicana) Seasonal agriculture, mainly self-supply (maize and peas) (continued)

460

J. J. Blancas Vázquez et al.

Table 1 (continued) Villages

Type of vegetation

Altitude (masl)

Number of inhabitants

Ethnic ascription

Chimalhuaca

Tropical deciduous forest, xerophytic thorn scrub forest

1820

112

Nahua, mestizos

Ixtacxochitla

Tropical evergreen forest

1316

402

Nahua, Mazatec

Principal economic activities Cultivation in agroforestry systems of capulín (Prunus serotina) and tejocote (Crataegus mexicana) Maguey pulquero cultivation (Agave atrovirens) Flor de tila recollection (Ternstroemia lineata) Sheep husbandry Seasonal agriculture, mainly self-supply (maize) Cultivation of mandarin (Citrus reticulata) and Mexican lemon (Citrus aurantifolia) Cultivation and recollection of quelites: pápalo (Porophyllum ruderale) and pipicha (Porophyllum macrocephalum) Firewood recollection Goat husbandry Seasonal agriculture, mainly self-supply (maize) Cultivation of mamey (Pouteria sapota), orange (Citrus  sinensis), coffee (Coffea arabica), purple banana (Musa paradisiaca), and cinnamon (Cinnamomum verum) Cultivation and recollection of tepejilote (Chamaedorea tepejilote) Cow husbandry Hunting of wild animals (continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

461

Table 1 (continued) Type of vegetation Pine-oak forest, oak forest

Altitude (masl) 2027

Number of inhabitants 1288

Ethnic ascription Nahua, Mestizos

Yohuajca

Oak forest, tropical deciduous forest

1980

251

Nahua

Xocotla

Pine-oak forest, oak forest

1998

466

Nahua, Mestizos

Villages Santa María Coyomeapan

Principal economic activities Seasonal agriculture, mainly self-supply (maize) Cultivation of plum (Prunus domestica), apple (Malus domestica), peach (Prunus persica), quince (Cydonia oblonga), and avocado (Persea americana) Chile canario cultivation (Capsicum pubescens) Firewood recollection Trade of various products Sheep husbandry Seasonal agriculture, mainly self-supply (maize) Cultivation of ornamental flowers Avocado cultivation (Persea americana) Flor de tila recollection (Ternstroemia lineata) Firewood recollection Edible plants recollection (flower buds of Dasylirion serratifolium) Seasonal agriculture, mainly self-supply (maize) Apple cultivation (Malus domestica) in agroforestry systems Chile canario cultivation (Capsicum pubescens) Cultivation of plum (Prunus domestica), apple (Malus domestica), and peach (Prunus persica) Firewood recollection Sheep husbandry

462

J. J. Blancas Vázquez et al.

Fig. 3 Main crops in the Sierra Negra, Puebla, Mexico: (a) Coffee (Coffea arabica); (b) milpa (maize, squash, and bean); (c) apple (Malus domestica); and (d) chile canario (Capsicum pubescens). (Photos: (a), (c), and (d) José Blancas; (b) Mariana Vallejo)

Currently, the communities of the region carry out diverse economic activities, among them the principal agricultural system which is the rainfed cultivation of milpa (the traditional Mesoamerican polyculture of maize, beans, and squash, commonly including other crops like chile pepper). The milpa production is mainly destined to self-consumption (Blancas et al. 2013). Other commercial crops like banana, orange, and coffee (in lowlands) and apple, chile canario (Capsicum pubescens), peaches, plums, and peas (in highlands) are also cultivated (Fig. 3). Agriculture is complemented by raising of sheep and goats, as well as gathering of a great number of wild species products, especially mushrooms, medicinal and edible plants, and fuelwood (Blancas et al. 2013; Lotero-Velásquez et al. 2022).

Interviews, Vegetation, and Ethnobotanical Sampling Information was obtained through semistructured interviews, participant observation, ethnobotanical walks, and vegetation sampling (Aguilar et al. 1994; Martin 1995;

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

463

Bernard 2011). This information forms part of a research project on the ethnobotany of Coyomeapan, which included studies in the eight villages referred to. The information was complemented with records of plants from the Sierra Negra in the regional markets of the Tehuacán-Cuicatlán Valley by Arellanes et al. (2013); a study on management practices on edible wild and weedy plants by Blancas et al. (2013); the role of homegardens in conserving local flora, their composition, and management by Larios et al. (2013); the diversity of plants in agroforestry systems of the zone (Vallejo et al. 2015); the role of wild and weedy plants (quelites) in local diet and food patterns by Olvera (2016); a study on the ecological complementarity and interchange of forest and agricultural products or eco-symbiotic relations (Lotero-Velásquez et al. 2022); and a study in progress on the relative cultural and economic importance of nontimber forest products in the community of Ixtacxochitla.

Ethnobotany of the Sierra Negra Until present, our studies have recorded 496 species of plants used by local people, grouped in 118 plant families and 351 genera (Appendix 1). The plant families richest in useful species are Asteraceae, Fabaceae, Asparagaceae, Solanaceae, Rosaceae, Araceae, Cactaceae, and Lamiaceae (Table 2). Most useful plant species (90%) are herbs, trees, and shrubs. The remaining 10% of species includes sclerophyllous, vines, and epiphytes (Table 3). Our records have identified 320 species (64.2%) native to México, 131 of them endemic; 173 species (34.88%) are introduced, 2 are of uncertain origin, and 1 is naturalized. Table 4 shows that most plant species recorded are wild (39.3%), domesticated (34%), ruderal (23.9), and weedy (13.7%). We have recorded 16 use categories; 112 species (22.5%) have 12 different uses while the rest (77.5%) are of exclusive use. On average, each species has 3.48 uses. Ornamental plants are the most numerous, constituting 42.34% of the useful flora, over the medicinal and edible plants, which commonly are recorded as the most abundant in ethnobotanical studies in Mexico. In the region, medicinal and edible species are 31.65% and 28.43%, respectively, while species used as fuel are 8.67%. The general panorama of the number and proportion of these and other use categories are shown in Table 5. Most useful plants are in situ managed, mainly gathered in forests (33.27%), tolerated or let standing during vegetation clearing (33.06%), and enhanced or promoted (19.75%). Other management forms and the number of species recorded are shown in Table 6.

Importance of Ornamental Plants As mentioned, ornamental plants conform the most numerous use category in the Sierra Negra. Of the 210 species recorded with this use, 51.9% (109 species) are native to Mexico and 46.2% (97 species) are introduced from other parts of the

Family Asteraceae Fabaceae Asparagaceae Solanaceae Rosaceae Araceae Cactaceae Lamiaceae Acanthaceae Apocynaceae Brassicaceae Crassulaceae Cucurbitaceae Euphorbiaceae Commelinaceae Orchidaceae Poaceae Amaryllidaceae Fagaceae

Species 70 26 21 21 19 15 14 14 8 8 8 8 8 8 7 7 7 6 6

Family Bignoniaceae Burseraceae Ericaceae Lythraceae Oleaceae Passifloraceae Pinaceae Polypodaceae Rhamnaceae Sapindaceae Scrophulariaceae Verbenaceae Aizoaceae Anacardiaceae Annonaceae Araliaceae Asclepiadaceae Asphodelaceae Begoniaceae

Species 3 3 3 3 3 3 3 3 3 3 3 3 2 2 2 2 2 2 2

Family Alliaceae Alstroemeriaceae Altinginaceae Aquifoliaceae Asreraceae Betulaceae Bixaceae Calophyllaceae Cannabaceae Chrysobalanaceae Cyatheaceae Dennstaedtiaceae Dioscoreaceae Ebenaceae Equisetaceae Gesneriaceae Heliconiaceae Hydrangeaceae Hypericaceae

Table 2 Number of useful species per botanical family recorded in the Sierra Negra, Puebla, Mexico Species 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

Family Selaginellaceae Siparunaceae Theaceae Tropaeolaceae Ulmeaceae Violaceae Xanthorrhoeaceae

Species 1 1 1 1 1 1 1

464 J. J. Blancas Vázquez et al.

Iridaceae Malvaceae Piperaceae Rubiaceae Rutaceae Amaranthaceae Apiaceae Bromeliaceae Cupressaceae Arecaceae Balsaminaceae Lauraceae Moraceae Myrtaceae Papaveraceae Plantaginaceae Sapotaceae Zingiberaceae

6 6 6 6 6 5 5 5 5 4 4 4 4 4 4 4 4 4

Cannaceae Capparidaceae Caricaceae Caryophyllaceae Convolvulaceae Geraniaceae Juglandaceae Liliaceae Magnoliaceae Nyctaginaceae Onagraceae Pentaphylacaceae Primulaceae Pteridaceae Ranunculaceae Salicaceae Urticaceae Adoxaceae

2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1

Labiatae Loranthaceae Lycopodiaceae Malphigiaceae Melastomataceae Musaceae Myricaceae Oronbachaceae Oxalidaceae Paeoniaceae Phytolaccaceae Picramniaceae Platanaceae Polygonaceae Portulacaceae Santalaceae Sapindaceae Sapotaceae

1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

Ethnobotany of the Nahua People: Plant Use and Management in the. . . 465

466

J. J. Blancas Vázquez et al.

Table 3 Growth habit of the useful flora in the Sierra Negra, Puebla, Mexico

Growth habit Herbs Trees Shrubs Rossettes Epiphytes Climbs

Table 4 Ecological status of the useful plants of the Sierra Negra, Puebla, Mexico

Ecological status Wild Domesticated Ruderal Weedy

Table 5 Flora uses in the Sierra Negra, Puebla, Mexico

Uses Ornamental Medicinal Edible Firewood Construction Fodder Shade Ritual/ceremonial Living fences Utensils Alcoholic beverages Saponiferous Resine Dyer Poison Glue Soil retention Fibers

Number of species 255 113 74 20 18 16

Number of species 195 169 119 68

Number of species 210 157 141 43 18 16 15 13 7 6 3 2 2 1 1 1 1 1

% 51.41 22.78 14.92 4.03 3.63 3.23

Percentage (%) 39.31 34.07 23.99 13.71

Percentage (%) 42.34 31.65 28.43 8.67 3.63 3.23 3.02 2.62 1.41 1.21 0.60 0.40 0.40 0.20 0.20 0.20 0.20 0.20

world. Nearly 45.71% (96 species) are domesticated and cultivated through seeds and vegetative parts; 38% (80 species) are wild, gathered in forests or planted in homegardens through vegetative propagation; and 14.76% (31 species) are weedy or ruderal plants, which may also be cultivated sexually or vegetatively. Ornamental plants are significant elements in local people’s life, forming part of their homegardens with the purpose of embellishing these spaces, or collected from forests to decorate the house, the table, the altars, and even the milpas (Blancas et al. 2013; Larios et al. 2013). Some outstanding native species are the camedor palm (Chamaedorea elegans Mart.), oloxóchitl (Justicia carnea Lindl.), cacaloxóchitl

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

467

Table 6 Documented forms of management among the useful flora of the Sierra Negra, Puebla, Mexico Management plant Gathering Tolerance Enhancement Seed sowing Transplanting of whole plants Protection Propagation of vegetative parts

Number of species 165 164 98 94 28 22 22

Percentage (%) 33.27 33.06 19.76 18.95 5.65 4.44 4.44

Fig. 4 Native ornamental plants of the Sierra Negra, Puebla, Mexico: (a) Oloxóchitl (Justicia carnea); (b) tehuanxochitl (Stanhopea tigrina); (c) iztacxóchitl (Spathiphyllum cochlearispathum); (d) tlalteztli (Tigridia pavonia); and (e) camedor palm (Chamaedorea elegans). (Photos: (a); (b); (c); and (d) José Blancas; (e) Carolina Larios)

(Plumeria rubra L.), velijmolli (Renealmia mexicana Klotzsch ex Petersen), iztacxóchitl (Spathiphyllum cochlearispathum (Liebm.) Engl.), tehuanxochitl (Stanhopea tigrina Bateman ex Lindl.), and tlalteztli (Tigridia pavonia (L. f.) DC.) (Fig. 4). Among the most important introduced ornamental plants we should mention tzapa or floripondio (Brugmansia candida Pers.), bebecho rojo (Caladium bicolor (Aiton) Vent.), panispatl (Canna indica L.), listón (Chlorophytum comosum (Thunb.) Jacques), campechana (Chrysanthemum morifolium Ramat.), española

468

J. J. Blancas Vázquez et al.

Fig. 5 Exotic ornamental plants registered in the Sierra Negra, Puebla, Mexico: (a) Campechana (Chrysanthemum morifolium); (b) cartucho or alcatráz (Zantedeschia aethiopica); (c) tzapa or floripondio (Brugmansia candida); (d) Mexican petunia (Ruellia sp.); and (e) panispatl (Canna indica). (Photos: (a); (b); (c); and (e) José Blancas; (d) Carolina Larios)

(Kniphofia uvaria (L.) Oken), amapola (Papaver rhoeas L.), azalea (Rhododendron kaempferi Planch.), and cartucho or alcatráz (Zantedeschia aethiopica (L.) Spreng.) (Fig. 5). Some of these species are maintained and cultivated in small spaces inside homegardens, and women oversee their care and commercialization in the local markets (Fig. 5). These plants are part of the inheritance that mothers leave to daughters, since some species have a high value, as it is the case of cacaloxóchitl (Plumeria rubra), nopalxóchitl (Disocactus phyllanthoides (DC.) Barthlott), and numerous orchids (e.g., Laelia superbiens Lindl.; Prosthechea cochleata (L.) W.E. Higgins; Stanhopea oculata (G. Lodd.) Lindl) and some species of bromeliads (e.g., Tillandsia rubra Ruiz & Pav.; Tillandsia imperialis E. Morren ex Mez and Tillandsia lucida E. Morren ex Baker).

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

469

Medicinal Plants A total of 157 plant species were recorded to have medicinal use, 81.5% of which are native and 18.5% introduced. Plant families providing more medicinal species are Asteraceae (30%), Lamiaceae (8%), Fabaceae (4%), Solanaceae (4%), and Rosaceae (3.3%). Most medicinal plants are herbs (58.3%), followed by trees (19.86%) and shrubs (19.5%). The rest are climber plants and epiphytes. Nearly 41% of the medicinal plants are obtained through gathering in primary forests, and others (40%) are collected in anthropogenic disturbed areas like secondary forest, roads, and fallow agricultural fields. Nearly 17.8% of the medicinal species are obtained from agroecosystems like homegardens and milpas, and 15.23% are domesticated species cultivated in homegardens, or bought in the market. Among the most important species, the xometl (Sambucus mexicana C. Presl ex DC.) can be mentioned. This is a shrubby plant tolerated and promoted in agroforestry systems, which is used in the traditional steam baths (called temazcal) with therapeutic purposes, especially for women recently given birth (Vázquez-Medina 2010; Blancas et al. 2013). It is also used in infusions to treat respiratory affections and to alleviate constipation. This plant grows in temperate environments, especially oak and pine-oak forests. Another important species from temperate ecosystems is a linden blossom or hachévitl (Ternstroemia lineata DC.), a shrubby or small tree collected and promoted in oak and pine-oak forests. Flowers of this species are used to treat affections of the nervous system, as a relaxant against stress, or to help falling asleep. This species is economically important since it is commercialized within the communities and in the regional markets (Arellanes et al. 2013). The chimalacate (Viguiera dentata (Cav.) Spreng.) is also a relevant species. It is a perennial herb from oak and tropical dry forests used as antiseptic in wounds, prepared as infusion with their leaves, stems, and flowers. The papalogüitl (Platanus mexicana Moric.) is a big, beautiful tree growing in the riparian vegetation in neighboring pine-oak and cloud forests. It is an important therapeutic plant whose leaves and cortex are widely used for treating skin affections like dermatitis or insect bites. Among the medicinal plants of the tropical rain forest, the mototetl is preeminent (Eugenia capuli (Schltdl. & Cham.) Hook. & Arn.), a tree growing wild and commonly introduced to cultivation in homegardens, grasslands, and agroforestry systems. Its inflorescences and fruits are appreciated as medicine prepared in infusions, but it is also a condiment (Blancas et al. 2013); it is common to find these products in regional markets. Another outstanding species of the rain forest is the cocacabatzi (Hamelia patens Jacq.), a shrub promoted and cultivated in homegardens, which is used to heal wounds and as hemostatic. Its leaves are used in poultices to alleviate bumps and sprains, and it is appreciated for its analgesic and febrifuge properties. The cloud forest provides several medicinal species; one of the principals is the yoloxóchitl (Talauma mexicana (DC.) G. Don), a tree whose fruits are used to prepare infusions against depression and other emotional ailments. This species is

470

J. J. Blancas Vázquez et al.

Fig. 6 Native medicinal plants of the Sierra Negra, Puebla, Mexico: (a) Flor de tila or hachévitl (Ternstroemia lineata); (b) yoloxóchitl (Talauma mexicana); (c) papalogüitl (Platanus mexicana); (d) mototetl (Eugenia capuli); and (e) xometl (Sambucus mexicana). (Photos: José Blancas)

widely valued, and trees are protected and propagated in homegardens and other agroecosystems. It is common to see the commercialization of its fruits in local and regional markets (Arellanes et al. 2013). Another important species is the epiphyte cactus called mazazalitzi (Rhipsalis baccifera (Sol.) Stearn), whose fruits are used to regulate the levels of glucose in blood by people with diabetes (Fig. 6).

Edible Plants and Their Importance in Local Food Patterns Edible plants are the third richest use category in the Sierra Negra region; 141 species are used with this purpose, 61.7% (87 species) being native plant species and 37.5% (53 species) introduced from other regions of the world. Nearly 44.68% (63 species) are herbaceous plants, 30.46% (43 species) are trees, 10.63% (15 species) are shrubs, and the rest are epiphytes, climber, and rosetophyllous plants. A total of 64 species (45.39% of all edible plants) are domesticated, propagated through seeds and/or vegetative parts, while 32.62% (46 species) are wild plants obtained through gathering in forest areas, and 19.14% (27 species) are weedy plants, which are tolerated, enhanced, and protected in plots of agroforestry systems. The remaining 16.31% (23 species) are ruderal plants, gathered and tolerated along the roads and other anthropogenic sites. Some examples of species used as food, obtained from the oak and pine-oak forests, are the mexcalli cacaya (Agave obscura Schiede) whose flower buds are consumed and very much appreciated by local people. These buds are boiled, and the water poured and changed after boiling, with the purpose of removing secondary

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

471

compounds conferring bitter flavor, then the buds are cooked in different ways, fried with garlic, onion, and tomato, with sauces, and in other forms. This is a wild plant, tolerated and enhanced in homegardens and milpas, as live fences and agroforestry systems (Blancas et al. 2013). The baquilitl (Amaranthus hybridus L.) is also important. It is an annual herbaceous plant consumed as greens or quelite and promoted by spreading its seeds in the milpas when maize and squashes are sown (Fig. 7). Another relevant edible plant is the chile canario (Capsicum pubescens Ruiz & Pav.), which is introduced from the Andean region in South America, but that has been strongly incorporated to the local culture. It is cultivated in monocultures or combined with other species as maize, squash, and peas. Its cultivation is an important source of income. Apple (Malus domestica (Suckow) Borkh.) is another exotic plant highly rooted in local culture. We have identified four varieties, but there are probably more. The most widely distributed variety in the region is the Golden, but there are others called “criollas” (creole), some of which appear to be descendants from ancient varieties

Fig. 7 Some examples of native edible plants in Pine-Oak Forest of the Sierra Negra, Puebla, Mexico: (a) Mexcalli cacaya (Agave obscura); (b) baquilitl (Amaranthus hybridus); (c) capulín (Prunus serotina); and (d) guaje (Leucaena esculenta). (Photos: José Blancas)

472

J. J. Blancas Vázquez et al.

introduced during the early colonial period. This has been a key species in the process of molding the landscape of the highlands of the Sierra Negra, since they form part of diverse agroforestry systems (Lotero-Velásquez et al. 2022), that commonly include exotic components like peaches (Prunus persica (L.) Batsch), quinces (Cydonia oblonga Mill.), and plums (Prunus domestica L.). But, in addition, these species coexist with native plants, some of them managed wild species, like the tetzmolli (Vaccinium leucanthum Schltdl.), capulín (Prunus serotina Ehrh.), tejocote (Crataegus mexicana DC.), and the xoxonte or zarzamora (Rubus adenotrichos Schltdl.) (Fig. 8). In these ecosystems, people also use to collect the sogogotl or Mexican laurel (Litsea glaucescens Kunth), a wild tree whose leaves are widely appreciated as condiment in the regional and national food. Leaves of this tree are mainly collected from the wild, but people promote its abundance in the forests and young plants may be transplanted to homegardens. The cloud forest and tropical rain forests provide important edible products. Among the principal species, we should mention the tepejilote (Chamaedorea tepejilote Liebm. ex Mart), whose immature male inflorescences are collected in the wild. But in addition, plants of this species are promoted, protected, and cultivated in coffee plantations. The male inflorescences of this dioecious species are locally consumed and commercialized in regional markets, which strongly enhances its

Fig. 8 Exotic edible species with highly regional importance in pine-oak forest of the Sierra Negra, Puebla, Mexico: (a) Apple (Malus domestica); (b) plums (Prunus domestica); (c) field mustard or colesh (Brassica rapa); (d) peas or alverjón (Pisum sativum); and (e) chile canario (Capsicum pubescens). (Photos: José Blancas)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

473

propagation and care. Local people recognize at least three varieties (called corpus, de cafetal, and metlapil), with different attributes but whose differential phenology motivate people to have all tepejilote products for a long period throughout the year. The zopelilquilitl (Cestrum nocturnum L.) is an edible product also considered to be a quelite. The tender young leaves are collected and prepared, boiled, or roasted. It is vegetatively propagated in coffee plantations and homegardens. Another interesting species is the velijmolli (Renealmia alpinia), whose leaves are employed to wrap maize dough to prepare tamales. This species is appreciated as wrapping material, but also because the leaves confer to the tamales a special flavor. Fruits of this species are also consumed, and these are used to prepare several types of sauces and moles. This plant is cultivated through vegetative propagation in homegardens, and it is also enhanced or promoted in coffee plantations. Leaves of this plant species are commercialized in the market of Coyomeapan, where people sell them by dozen units. The tequilitl (Peperomia peltilimba C. DC.) is also a quelite, gathered from wild vegetation. It is consumed raw, with a flavor like coriander. It is widely appreciated, and it is commonly found in the markets of Coyomeapan and Zoquitlán, in the Sierra Negra, as well as in the regional markets of the Tehuacán-Cuicatlán Valley. This fact has generated an increasing pressure on the plant populations, and now people say that they have to go further to collect it (Fig. 9).

Fig. 9 Native edible species of the Cloud Forest in the Sierra Negra, Puebla, Mexico: (a) Mexican laurel (Litsea glaucescens); (b, c) velijmolli (Renealmia alpinia); (d) tequilitl (Peperomia peltilimba); (e) zopelilquilitl (Cestrum nocturnum); and (f) tepejilote (Chamaedorea tepejilote). (Photos: (a); (b); (c) and (d) José Blancas; (e) and (f) Carolina Larios)

474

J. J. Blancas Vázquez et al.

Plants Used as Fuel People inhabiting the Sierra Negra strongly depend on fuelwood for their daily life, especially for cooking. A total of 43 species have been identified to be used for this purpose, 79% of them being native (34 species) and 21% introduced (9 species). Native species are part of the forests neighboring the villages while the introduced species commonly occur in agroforestry systems and homegardens. Trees are the main source of fuel (35 species; 81.39%), but some shrubs (6 species; 13.95%) and herbs (2 species; 4.65%) are also used. Most plant species used as fuel (28 species; 65%) are wild and therefore gathered from forests, but some of them are let standing or tolerated, enhanced, or promoted in forests and agroforestry systems. A total of 11 species are domesticated (25.5%), which are cultivated through vegetative propagation or seeds in agroforestry systems and homegardens; six species are ruderal or weedy plants (13.95%), which are gathered from anthropogenic areas. In the highlands, use of fuelwood is markedly focused on oak species since, according to local people, their wood has the best quality in terms of caloric emission, durability, and the flavor that confer to food prepared with it. For instance, the tamalabatl (Roldana candicans) is a species that can be easily propagated since after being chopped down its trunk, its stumps commonly sprout. The copazole (Quercus laurina) is widely distributed in the region and is a species tolerant to intense management, having the capacity of colonizing open areas and relatively fast growth. Although its wood is considered a good quality fuel, big-size individuals are scarce and, therefore, its use is limited to cooking maize tortillas. The tlilibicawatl (Quercus rugosa Née) and tepoztenawatl (Quercus urbanii Trel.) are species commonly used by people living near temperate forests. They have good fuelwood and are more abundant than copazole. Where oaks are absent, people use as fuel other species like madroño (Arbutus xalapensis), tetzmolli (Vaccinium leucanthum), tepellili (Cercocarpus fothergilloides Kunthy), and elite (Alnus acuminata). Villages close to the cloud forest commonly use the papalouwit (Liquidambar styraciflua), ilite (Ulmus mexicana (Liebm.) Planch.), and the chichistlahui (Fraxinus uhdei). The communities settled in the lowlands, near the tropical rain forest, commonly use low density wood, as it is the case of the topetli (Inga flexuosa Schltdl.), as well as old trees or dry branches of the tliltlzapotl (Diospyros digyna Jacq.). Coffee plants (Coffea arabica L.) are also used as fuel for cooking in this zone. People living in villages close to tropical dry forest, in elevations about 900 to 1800 m, use several species as fuel, outstandingly the huizache (Acacia farnesiana (L.) Willd), tepehuaje (Lysiloma sp.), guaje (Leucaena esculenta (Moc. & Sessé ex DC.) Benth.), abeshotl (Salix taxifolia Kunth), and cuetzalcahuitl (Cupania dentata DC.). Despite the broad spectrum of species used as fuel, it is clear that people have preferences over some species, because they are obtained more easily or because of their quality. Species of the genus Quercus are the most preferred; however, the population growth and the increasing demand of fuelwood have determined pressures on oak forests which have decreased in area in the last decades. People say they have to go further to collect fuelwood. In addition, the size of trees used is progressively smaller,

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

475

Fig. 10 Some species that are used as firewood in the Sierra Negra, Puebla, Mexico: (a) Trunks of Pinus teocote; (b) trunks of chichistlahui (Fraxinus uhdei); (c) branches of copazole (Quercus laurina); (d) madroño (Arbutus xalapensis); (e) abeshotl (Salix taxifolia); (f) tamalabatl (Roldana candicans); and (g) woman collecting firewood in Coyomeapan. (Photos: (a); (b); (c); (d); (e); and (f) José Blancas; (g) Carolina Larios)

which determines risks on oak populations and increases pressure on other species that were considered of secondary preference in the past. Further research is needed to evaluate the vulnerability of populations and ecosystems to fuelwood extraction. In addition, studies are necessary about the historical role of women in this activity, since women were and currently continue to be the members of the household with the main responsibility of collecting and making decisions about the type and quality of fuelwood. However, recent changes associated to the introduction of new practices and technology have changed that role. For instance, the introduction of chainsaw, which turned the extraction of fuelwood to men and allowed intensifying the felling of trees and the commercialization of both wood and firewood (Fig. 10).

Ecological Complementarity and Multiple Use of Ecosystems in Plant Use As referred to above, the Sierra Negra is a mosaic of vegetation types where people carry out different activities according to the differences in climate, relief, and soil, among other factors. Use of plant products is based on principles of multiple use of both plant resources and ecosystems; it is common that people cultivate, manage, and gather plants and plant products in different ecological zones. Such strategy is a response to different conditions of uncertainty and are part of the aim to manage ecosystems and complement the benefits obtained from them through activities directed to mitigate such uncertainty (Blancas et al. 2013). For instance, the territory of the community of Yohuajca covers zones where species of the highlands (oak and pine-oak forests) and

476

J. J. Blancas Vázquez et al.

lowlands (tropical dry forest and xerophilous thorn-scrub forest) occur. It is common that people sow maize of a 9-month life cycle in the highlands, whereas they plant maize of 90 days varieties in the lowlands. Clearly, the fast growing varieties produce small plants and cobs, but this production is crucial to have available some maize during the most critical period of food availability of the year. The diversifying strategy of using ecosystems, plant species, and plant products allows complementing subsistence of local people through wild plant products from forests in different ways. For instance, some households go to collect hachevitl (Ternstroemia lineata) from April to June, when the maize reserves are close to running out. The commercialization of hachevitl allows income that make possible buying maize and other food. This community directly use and commercialize edible products; one of the most outstanding is the matzitzi (Dasylirion serratifolium), whose edible flower buds are highly valued, and its availability also occurs during the dry season, when agricultural products are scarce. The availability of this product in markets allows income for household gathering them, as well as edible products to communities where this plant is absent. A similar situation occurs with different products, and it is an expression of the reciprocity that the mechanisms of interchange allow to maintain the regional patterns of people’s subsistence. Such strategy is also practiced by communities settled in other ecological contexts. For instance, in Aticpac, whose territory is predominantly tropical rain forest and cloud forest, people produce coffee, bananas, and velijmolli (Renealmia alpina) in agroecosystems. These are main products for commercialization and interchange. In addition, they collect numerous products accompanying the activities of commercialization, outstandingly tequelite (Peperomia peltilimba), tepejilote (Chamaedorea tepejilote), laurel or sogogotl (Litsea glaucescens), and mototetl (Eugenia capuli). In the highlands, people of the villages of Caxalli, Ahuatla, and Coyomepan produce apples, pears, plums, peaches, chile canario, and peas, which are main products for commercialization. But in addition, people interchange products gathered from forests and agroforestry systems, mainly capulín (Prunus serotina), mexcall (Agave atrovirens Karw. ex Salm-Dyck), mexcalcacaya (Agave obscura), izotl (Yucca guatemalensis Baker), tetzmolli (Vaccinium leucanthum), iquimite (Erythrina americana Mill.), ilamatzin (Phaseolus coccineus L.), tamalabatl (Roldana candicans), aguacatl (Persea americana Mill.), matetecomatl (Passiflora subpeltata Ortega), and texocotl (Crataegus mexicana). All these resources are available throughout the year in the market of Coyomepan, granting food, medicinal, and ornamental plant products diversity and incomes to regional people. Surplus is brought to the regional markets of the Tehuacán-Cuicatlán Valley (Tehuacán, Ajalpan, Zinacatepec, Puebla, and Teotitlán, Tecomavaca and Cuicatlán, and Oaxaca; see Arellanes et al. 2013). In this way, it appears that each region contributes with particular resources growing in specific types of ecosystems and that on the whole contribute to the regional metabolism of ecological complementarity through interchange (Lotero-Velásquez et al. 2022). Peoples of the Sierra Negra have developed mechanisms of multiple use of resources and ecosystems at local level, achieving complementarity at regional scale. Such complementarity is based on strong historical, cultural, and economic

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

477

ties which have constructed nets of reciprocity that substantially contribute to construct a food sovereignty system and maintenance of biocultural diversity in the region.

Ecological and Sociocultural Strategies Influencing Plant Management Management of plant species in the complex mountainous environments of the Sierra Negra has represented an option to face the regional challenges for subsistence, which have been solved based on a great resilience capacity developed throughout history. These forms of management are part of the traditional ecological knowledge as defined by Berkes (1999), which has allowed developing strategies of interaction with plants and ecosystems adapted to the ecological and sociocultural contexts. Such interactions look for ensuring the availability of resources, increase their quality, and manage the sources of uncertainty (Blancas et al. 2013). Among the main strategies we can mention the following: (a) Use of wild and cultivated plant resources. This strategy looks for widening in time and space the availability of useful plant species and their products, not only those that people can cultivate, but also others obtained from the neighboring forests (Fig. 11). It is common to find that the regional communities destine part of their agricultural products to self-consumption and part to commercialization, and these are main activities; therefore, some people gather products from forests and agroforestry systems with these purposes, thus complementing their

Fig. 11 Some examples of plant species harvested from the forests of the Sierra Negra, Puebla, Mexico: (a) Coyol (Acrocomia mexicana); (b) hachevitl or flor de tila (Ternstroemia lineata); and (c) bromeliads and palms for ritual purposes. (Photos: (a) and (c) José Blancas; (b) Ignacio TorresGarcía)

478

J. J. Blancas Vázquez et al.

economy. This is, for instance, the case of people dedicated to collect mushrooms and fuelwood and other nontimber forest products referred to above. (b) Privileging vegetative over sexual propagation for reproducing plants. This management technique looks for ensuring success of propagation and survival of plant species considered particularly valuable and to shorten the time of harvesting their products. An example of this fact is the maintenance and care of stumps of copazole (Quercus laurina), which are expected to sprout within the next year. It is also the case of the izote (Yucca guatemalensis), which can be propagated through vegetative cuttings and placed around plots of agricultural fields and homegardens as live fences. These plants provide benefits as fences, edible flowers, and fiber obtained from their leaves (Fig. 12). The long-term consequence of this management type and the strategies that ensure maintaining genetic diversity through this management are yet to be evaluated. (c) Rotation of gathering areas. This practice is a common response to the perception of scarcity of one or a group of resources in an area, especially those with higher cultural value. This is, for instance, the case of the sogogotl or Mexican Fig. 12 Vegetative propagated species: (a) Izote (Yucca guatemalensis); and (b) mexcalli (Agave atrovirens). (Photos: José Blancas)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

479

laurel (Litsea glaucescens), whose leaves are mainly gathered in forests for selfconsumption and commercialization. Populations of this species have decreased progressively as the gathering intensity increases. It is common that people migrate to collect in other areas to leave a population to recover. According to people, by doing this, after 2 or 3 years they can come back to continue using a population where they used to practice gathering. In principle, gathering leaves from this plants should not have high impact, but until recently some people used to cut the tree to make the leaf collection easier. This practice has been documented in different regions of Mexico. Propagation of this plant together with regulations in the form of collection is necessary to conservation of the sogogotl. (d) Rotation of crops and agricultural areas. This is a strategy analogous to the rotation of extraction areas, but in this case with domesticated plants and spaces. Continual cultivation in an area has restrictions associated to impoverishment of soils and increasing risk of pests. It is common that people sow maize in a plot for 2–3 years, then they change to cultivate peas or potatoes, and then they reuse fallow plots to cultivate maize. Such actions, according to people, mitigate the negative effects on their plots, but an experimental evaluation in this respect is still needed. (e) Seed storing. Storing of staple grains is a strategy for ensuring availability of agricultural products throughout the year, as well as the availability of seeds for the following agricultural cycle. Selection of seeds is the expression of the continual process of domestication operating in crops. People use to choose seeds with the best attributes (size, vigor, color, texture, and resistance against pests are all indicators; some other attributes are related with the performance of particular plants in the plots). People use different techniques to store their seeds. And these principles are used not only for different crops, but also for the cultivation of wild and weedy plants. For instance, several species of quelites like the pipicha (Porophyllum linaria (Cav.) DC.), Brassica rapa, Amaranthus hybridus, and others (Blancas et al. 2013) are cultivated and promoted in the milpas and homegardens. People use to collect seeds in agricultural fields (commonly practicing some kind of selection), then seeds are stored and eventually spread in plots to enhance their abundance (Fig. 13). (f) Use of different environmental units. As discussed above, this strategy looks for using plant resources restricted to specific environments, thus ensuring spatial and temporal availability of plant products and their ecosymbiotic complementarity (Lotero-Velásquez et al. 2022). (g) Use of emergency food. This strategy looks for mitigating the negative consequence of food scarcity associated to catastrophic events (climatic or social) limiting agricultural production in a year. Emergency food is commonly identified in rural communities throughout Mexico, and the Sierra Negra is not an exception. Use of this type of food is recurrent, especially during atypical drought, excess of rain, incidence of frosts, hail, fires, and pests (Zinyama et al. 1990; Guinand and Lemessa 2000). It is also referred to periods of social conflicts that limited agricultural production, but these in general have been

480

J. J. Blancas Vázquez et al.

Fig. 13 Storage of seeds of different species: (a–c) malagalquilitl or pipicha (Porophyllum linaria); (d) maize (Zea mays). (Photos: José Blancas)

poorly studied (Redžić 2010; Vorstenbosch et al. 2017; Sulaiman et al. 2022). In the Sierra Negra, people commonly evocate the Mexican Revolution period. In such circumstances, local people refer to some plants that have been used to avoid famine situations. The references vary from village to village, for instance, in a village people mentioned the consumption of the rhizome of the tlalteztli (Tigridia pavonia), and in communities of the highlands people refer the use of tender leaves (quelite) of molquilitl (Phytolacca icosandra L.), fruits of tetzmolli (Vaccinium leucanthum), which are used to prepare atole (a local beverage), flowers of the xochiquilitl (Phaseolus coccineus), and the inflorescence of the elotlquilitl (Spathiphyllum cochlearispathum), which is mostly considered an ornamental plant (Fig. 14). People of the Sierra Negra remember the famine of the local inhabitants during the Mexican Revolution between 1910 and 1920. Old people refer that in the lowlands people interchanged small amounts of maize and the dough prepared was mixed with banana to prepare a kind of sweet tortilla. In other places, people mixed and cooked the dough maize with early floral escapes of agave, and oak acorns flour. Emergency food is currently rarely consumed, or when consumed, people do not recognize the fact, which appears to be related with a sort of shame, since the consumption of this food is associated to poverty (Olvera 2016). Further studies on this group of food are important, their actual role in diet, the recurrency of difficult periods, and the regional mechanisms of resilience.

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

481

Fig. 14 Some examples of species used as emergency food: (a) Field mustard or colesh (Brassica rapa); (b) tlalteztli (Tigridia pavonia); (c) zopelilquilitl (Cestrum nocturnum); (d) xochiquilitl (Phaseolus coccineus); and (e) tetzmolli (Vaccinium leucanthum). (Photos: José Blancas)

(h) Commerce and interchange. As discussed above, numerous plant products are interchanged to directly satisfy different needs and to obtain income destined to satisfy other needs established by the increasing market of external products (materials for the school, clothes, materials for construction, medicines, and agrochemicals, among others). Interchange of products may involve cultivated products or those obtained from the forest (Fig. 15). Hitherto, we have identified in the region 121 species used for barter or commercialization, but more research is needed for a more precise evaluation of forest products in the local people’s economy. (i) Migration and local employment. It is common at least one member of every family of the Sierra Negra has migrated, most commonly young men, who are mainly employed in the construction industry in big cities like Puebla and Mexico City. Such migration is usually temporary, generally during the pauses of the agricultural labors. However, there are also cases of permanent migration, mainly to Mexico City, states of northern Mexico, and the USA. The permanent migrants periodically send remittances to their families, which are important

482

J. J. Blancas Vázquez et al.

Fig. 15 The regional market of Coyomeapan fulfills provision and cultural functions, where diverse species of plants are commercialized and exchanged: (a) Tepejilote, mamey, coffee, and bananas; (b) exchange of maize for oranges; (c) sale of custard apples, chayote, sweet potato, papalo, and guavas; (d) sale of velijmolli leaves, beans, coffee, tequelite, and bananas; (e) sale of medicinal plants, black zapote, tempesquixtle fruits, velijmolli leaves, and citlaltomatl; and (f) general appearance of a Sunday in the Coyomeapan market. (Photos: José Blancas)

incomes to attend the needs of products obtained in the market, as well as salaries to people helping in agricultural labors. Ethnobotanical literature has emphasized the negative consequences of migration, especially in relation to the maintenance and transmission of traditional ecological knowledge (Nesheim et al. 2006; Brandt et al. 2013). Certainly, migration has generated changes in food patterns, health, materials for constructing houses, ornamental plants, and the role of homegardens, among numerous other cultural patterns. However, it has been poorly evaluated the role of migration in the maintenance of the way of life among people that remain in the communities, who have to attend payments in the communitarian feasts, labors to construct public places, and importantly, to maintain their traditional ways of life. All these activities are important to maintain ties and social relations in the community, may favor exchange of knowledge, and, in general, may contribute to maintaining the local culture and ways of life. Local employment allows maintaining and

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

483

complementing household subsistence, and this employment is possible because of the remittances received by some families in the community. (j) Interchange. Interchange at the different scales described above may involve plants, as well as knowledge, and this is an important process allowing printing the dynamic nature of traditional knowledge. Particularly important is the exchange of knowledge about using and managing plants and agroforestry systems. It is common that young people ask for advice to older people about how to deal with pests, how to increase agricultural production, how to cultivate new species and varieties, and how to manage soils. For instance, in the community of Ahuatla old people used to tolerate and enhance the trees of elite (Alnus acuminata) in the outline and inside the plots. This agroforestry practice contributes to maintaining soil fertility, since leaves are an excellent organic fertilizer. This knowledge has recently been added to the knowledge of younger agriculturalists (Fig. 16). And, importantly, this type of knowledge and techniques is shared and reinforced in informal spaces such as markets, fairs, and even the religious or civilian celebrations. Through these ways, a complex knowledge exchange network is formed that goes beyond the local.

Fig. 16 The traditional management of the elite (Alnus acuminata), an example of knowledge exchange: (a) Elite trees placed as living fence in the milpas; (b) saplings of elite trees in a milpa with 2 years without cultivation; and (c) elite tree tolerated in the milpa whose leaves are used as organic fertilizer. (Photos: José Blancas)

484

J. J. Blancas Vázquez et al.

(k) Construction and application of communitarian agreements, rules, and institutions to access plant resources. In Mexico, there is a complex system of norms regulating land tenure, and in a village the following two or three types of tenure may coexist: private property, ejidal property, and communal property. The two latter are collective property regimes. Ejido is a type of social property resulting from the Agrarian Reform after the Mexican Revolution. This type of property may be part of the territory of a community, and people that may make use of this land are called ejidatarios. The ejido areas are designated through national decrees and generally include land with attributes for using in agriculture. The communal property includes forests and cultivation areas, and people of the community recorded as part of this property are called comuneros. Both ejido and communal land are regulated by ad hoc authorities, the Comisariado de Bienes Ejidales, and the Comisariado de Bienes Comunales, which are periodically elected and make decisions based on general assemblies of the owners. Ejidatarios and comuneros therefore have their authorities and rules are constructed through complex processes of agreements combining national norms and laws and local agreements. Therefore, use of plant resources either for satisfying a direct need or commercialization requires permits of the communitarian authorities. For instance, the areas for extracting fuelwood or for extracting wood for construction are specifically agreed by the general assemblies and monitored and guarded by ejidal and communal authorities. It is also established the amount and type of fuelwood that is permitted, and even the persons that are authorized to perform the extraction. This type of regulation is also used to define the way of gathering some edible wild plants, especially those considered endangered, like sogogotl (Litsea glaucescens) or the tequelite (Peperomia peltilimba); some medicinal plants like the linden flower or flor de tila (Ternstroemia lineata) and the yoloxóchitl (Talauma mexicana); fuelwood like the ocote (Pinus patula) and the tlaxca (Cupressus lusitanica Mill.); ornamental plants like the camedor palm (Chamaedorea elagans); and species of ritual value like coyol (Acracomia mexicana Karw. Ex. Mart) and the xolochi (Tillandsia lucida E. Morren ex Baker), which are all species under special local regulations. In addition, the ejidal and communal authorities take care of designing and guarding norms to access the forest benefits. For instance, it is definitely prohibited using timber for commercial purposes. People caught cutting trees without authorization receive sanctions, which can be warning or payment of fines. (l) Cooperation. This strategy is very well rooted in the communities of the Sierra Negra. People use to collaborate with others in agricultural activities and, in turn, receive collaboration. This practice is called mano-vuelta, which can be translated as today “I help you and then you help me when I need your help.” Commonly, not only it is carried out among relatives and friends, but also the communitarian authorities design persons who should help others that need help. There are in addition communitarian activities that require the participation of all members of the community. Cooperation is a mechanism to lightening the work, maintaining communitarian ties, and helping vulnerable people of the community.

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

485

Conclusion The Sierra Negra is a mosaic of biocultural diversity, characterized by diverse forest ecosystem and agroecosystem types, all occurring in a context of also diverse forms of understanding, conceiving, using, conserving, and recovering the components of nature. The geomorphological particularities of the Sierra Negra have been used by the cultures inhabiting the region, determining strategies of multiple use and management of resources and landscapes. The relative historical isolation of the region promoted different forms of selfsufficiency, risk management associated to uncertainties in the availability of the local flora and ecosystems, and strategies enhancing the ecosystem complementarity. However, the opening to external influences has promoted changes in the forms of using plant resources and forests. Further studies should focus on evaluating the impact of these influences on conservation, enrichment, or erosion of traditional knowledge. Hitherto, we have identified 496 useful plant species, and this information reveals the strong relation of local people and the surrounding ecosystems to satisfy their needs of ornament, medicine, food, and fodder. However, this inventory is far to be complete. It is necessary to continue the ethnobotanical exploration, ecological and ethnographic studies, not only to increase the number of records, but also, above all, to understand the relations and importance of the interactions between local cultures and ecosystems, their mutual influence, and the conditions to their conservation and recovering as biocultural landscape. The ecological complementarity is a topic of great importance, and its study should be enhanced. This issue as well as the interaction with cultures existing in the Sierra Negra and the neighboring Tehuacán-Cuicatlán Valley is an ideal setting to analyze the dynamics of the construction of traditional ecological knowledge systems. Designing forms of sustainable use of plant species and ecosystems of the region should be based on the local experiences to be viable. The Sierra Negra is a biocultural refuge, with a vast reservoir of species, ecosystems, knowledge, and practices. The particularities of the regional biocultural diversity may provide important lessons for other areas of the region and for other regions of Mexico. Studying and understanding the mountain regions of Mexico is undoubtedly a priority for scholars and governmental sectors interested in the immense biocultural heritage of this mountainous country. Acknowledgments The authors thank the communities who supported our work, especially to Mrs. Guillermina Copas, Bertha Sánchez, Yolanda Orozco, Genoveva, Porfirio Pérez, Margarito Perea, and Gerardo Garzón. Also, we thank financial support from CONACYT, Mexico (project A1S-14306), the GEF Project ID 9380 CONABIO-GEF-FAO/ RG023 “Manejo y domesticación de agrobiodiversidad en Mesoamérica: Bases para la soberanía alimentaria sustentable,” and PAPIT, UNAM (project IN206520, IA207721, and IN224023).

486

J. J. Blancas Vázquez et al.

Appendix 1 Plants used by local people in Sierra Negra, Puebla, Mexico Scientific name

Common name

Growth type

Origin

Acanthaceae Acanthus mollis L.

Acanto

Herb

Exotic

Aphelandra squarrosa Nees



Herb

Exotic

Beloperone guttata Brandegee

Illegaypig, camarón

Shrub

Native

Herb

Exotic

Hypoestes phyllostachya Baker

Usesa

Ecological status

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Tolerated

O

Domesticated

O

Domesticated

O

Domesticated

O

Domesticated

O

Domesticated

M, O

Domesticated

O

Domesticated

O

Ruderal

Management status

Justicia aurea Schltdl.

Oloxochitl

Shrub

Native

Justicia carnea Lindl.

Oloxochitl

Herb

Native

Pachystachys lutea Nees

Cuatlinenepil

Shrub

Native

Thumbergia alata Bojer ex Sims Adoxaceae Sambucus mexicana C. Presl ex DC.

Ixtololomazatl

Climb

Exotic

Xometl, sauco

Shrub

Native

Gathering, enhancement, tolerated

E, M

Wild

Cebolla

Herb

Exotic

E

Domesticated

Aptenia cordifolia (L. f.) Schwantes



Herb

Exotic

O

Domesticated

Carpobrotus edulis (L.) N.E. Br.



Herb

Exotic

Cultivated by seed Cultivated by vegetative propagation Cultivated by vegetative propagation

O

Domesticated

Comegatl

Climb

Exotic

Tolerated

F, LF

Weedy

Papalouwit cimarro

Tree

Native

Enhancement, tolerated

FW, M Wild

Baquilitl, quintonil

Herb

Native

Cultivated by seed, enhancement, and tolerated

E

Aizoaceae Allium cepa L.

Alstroemeriaceae Bomarea hirta Schenk Altinginaceae Liquidambar styraciflua L. Amaranthaceae Amaranthus hybridus L.

Weedy

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . . Common name

Growth type

Origin

Flor de terciopelo Amole, quelite Blanco

Herb

Exotic

Herb

Native

Chenopodium nuttalliae Saff.

Huauzontle

Herb

Native

Dysphania ambrosioides (L.) Mosyakin and Clemants Amaryllidaceae Agapanthus campanulatus F.M. Leight. Amaryllis belladonna L.

Epazotl

Herb

Native

Ciento uno, agapando

Herb

Exotic



Herb

Exotic

Amaryllis sp.



Herb

Exotic

Clivia miniata Regel



Herb

Exotic

Hippeastrum elegans (Spreng.) H. E. Moore

Palma, lirio

Herb

Exotic

Narcissus sp.

Narciso

Herb

Exotic

Anacardiaceae Cyrtocarpa procera Kunth Mangifera indica L.

Mashaxocotl, chupandilla Mango

Tree

Native

Tree

Exotic

Chirimoya

Tree

Exotic

Guanábana

Tree

Exotic

Cilantro

Herb

Exotic

Zanahoria

Herb

Exotic

Eryngium foetidum L.

Cilantro cimarrón Herb

Native

Foeniculum vulgare Mill. Petroselinum crispum (Mill.) Fuss

Hinojo

Herb

Exotic

Perejil

Herb

Exotic

Scientific name Celosia cristata L. Chenopodium berlandieri Moq.

Annonaceae Annona cherimola Mill. Annona muricata L. Apiaceae Coriandrum sativum L. Daucus carota L.

487

Usesa

Ecological status

Cultivated by seed Cultivated by seed

O

Domesticated

E

Cultivated by seed, enhancement Cultivated by seed

E

Weedy, domesticated, ruderal Weedy, ruderal

E, M

Weedy, domesticated, ruderal

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by seed

O

Domesticated

O

Domesticated

O

Domesticated

O

Domesticated

O

Domesticated

O

Domesticated

Enhancement, tolerated Cultivated by seed

E

Wild

E

Domesticated

Cultivated by seed Cultivated by seed

E, FW

Domesticated

E

Domesticated

Cultivated by seed Cultivated by seed Gathering, tolerated Cultivated by seed Cultivated by seed

E

Domesticated

E

Domesticated

E

Wild

M

Domesticated

E

Domesticated

Management status

(continued)

488

J. J. Blancas Vázquez et al.

Common name

Growth type

Origin

Copa de oro

Shrub

Exotic

Asclepias curassavica L. Asclepias linaria Cav. Catharanthus roseus (L.) G. Don

Chilillo

Herb

Native

Roromero Coronaxibitl

Shrub Herb

Native Exotic

Nerium oleander L.

Adelfa

Shrub

Exotic

Plumeria rubra L.

Cacaloxochitl

Tree

Native

Tree

Native

Scientific name Apocynaceae Allamanda cathartica L.

Stemmadenia sp.

Usesa

Ecological status

Cultivated by vegetative propagation Tolerated

O

Domesticated

M

Ruderal

Tolerated Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by seed

M O

Ruderal Domesticated

O, S

Domesticated

O, M

Wild

G

Wild

C

Domesticated

Management status

Vinca major L.

Candelaria

Herb

Exotic

Aquifoliaceae Ilex sp.



Tree

Exotic

Cultivated by vegetative propagation

O

Domesticated



Herb

Exotic

O

Domesticated



Herb

Exotic

O

Domesticated

Anthurium crassinervium (Jacq.) Schott Anthurium cubense Engl.

Anturio

Herb

Exotic

O

Domesticated

#

Climb

Exotic

O

Domesticated

Caladium bicolor Vent.

Bebetcho rojo

Herb

Exotic

O

Domesticated

Dieffenbachia maculata (Lodd.) G. Don Dieffenbachia seguine (Jacq.) Schott

Bebetcho pinto

Herb

Exotic

O

Domesticated



Herb

Native

O

Domesticated

Monstera deliciosa Liebm.

Piñanona, calavera

Herb

Native

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation, transplantation of whole individuals

O

Domesticated

Araceae Alocasia macrorrhizos (L.) G. Don. Anthurium andreanum Linden

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

Scientific name

Common name

Growth type

Origin

Monstera sp.

Piñanona

Herb

Native

Philodendron hederaceum (Jacq.) Schott Spathiphyllum cochlearispathum (Liebm.) Engl.

Teléfono

Herb

Native

Ixtaxochitl, elotlquilitl, eloxóchitl

Herb

Native

Spathiphyllum wallisii Regel

Ixtaxochitl (cuna Herb de Moisés)

Native

Xanthosoma robustum Schott.

Bebetcho, malangar

Herb

Native

Xanthosoma sagittifolium (L.) Schott Zantedeschia aethiopica (L.) Spreng. Araliaceae Acracomia mexicana Karw. ex. Mart.

Bebetcho pequeño

Herb

Native

Cartucho, alcatráz

Herb

Exotic

Coyol

Tree

Native

Oreopanax xalapensis (Kunth) Decne. & Planch. Schefflera arboricola (Hayata) Merr.



Tree

Native



Tree

Exotic

Chamaedora elegans Mart

Camedora

Herb

Native

Chamaedorea tepejilote Liebm. ex Mart

Tepexilotl, tepejilote

Herb

Native

489

Usesa

Ecological status

Cultivated by vegetative propagation Cultivated by vegetative propagation Transplantation of whole individuals, enhancement, and tolerated Cultivated by vegetative propagation Cultivated by vegetative propagation, enhancement, and transplantation of whole individuals Cultivated by vegetative propagation Cultivated by vegetative propagation

O

Domesticated

O

Domesticated

O, E

Wild

O

Wild

O, E

Wild

O

Wild

O

Domesticated

Gathering, tolerated, transplantation of whole individuals Gathering

C, E, O Wild

Management status

O

O Cultivated by vegetative propagation O Cultivated by seed, enhancement, protected, gathering, tolerated O, E Cultivated by seed, cultivated by vegetative propagation, enhancement, protected, gathering, tolerated

Wild

Domesticated

Wild

Wild

(continued)

490

J. J. Blancas Vázquez et al.

Usesa

Ecological status

Cultivated by vegetative propagation

O

Domesticated

Native

Gathering

M

Wild

Climb

Native

Enhancement

E

Wild

Rosette

Native

CO, FI Domesticated

Agave americana L.

Mexcaltenechtli Rosette

Native

Agave americana var. marginata Trel.

Maguey

Rosette

Native

Agave angustifolia Haw.

Népulli

Rosette

Native

Agave atrovirens Karw. ex Salm-Dyck

Mexcalli

Rosette

Native

Agave celsii Hook

Magueyito

Rosette

Native

Agave ellemetiana Jacobi



Rosette

Native

Agave pendula Schnittsp.

Mexcalli

Rosette

Native

Agave obscura Schiede

Mezcalcacaya

Rosette

Native

Agave salmiana Otto ex Salm-Dyck.

Mexcalli

Rosette

Native

Agave salmiana var. ferox (K. Koch) Gentry Asparagus officinalis L.

Maguey de pulque

Rosette

Native

Espárrago

Herb

Exotic

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation, enhancement, and transplantation of whole individuals Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by seed, enhancement, and tolerated Cultivated by vegetative propagation Cultivated by vegetative propagation

Scientific name Phoenix canariensis Wildpret Asclepiadaceae Asclepias curassavica L. Gonolobus uniflorus Kunth. Asparagaceae Agave aff. angustifolia

Common name

Growth type

Origin

Palma

Tree

Exotic



Herb

Pochotl



Management status

O, E, F, Domesticated CO, LF O

Domesticated

E, AB

Wild

AB

Domesticated

O

Domesticated

O

Wild

E

Wild

C

Wild

E

Domesticated

E

Domesticated

E

Domesticated

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

Scientific name

Common name

Growth type

Origin

Chlorophytum comosum (Thunb.) Jacques Cordyline fruticosa (L.) A. Chev.

Listón, listoncillo Herb

Exotic

Pabellón

Exotic

Dasylirion serratifolium (Karw. ex Schult. f.) Zucc.

Tehuizotl, Rosette tebixtli, cucharita

Native

Dasylirion sp.

Matzitzi

Rosette

Native

Nolina longifolia (Karw. ex Schult. f.) Hemsl. Polianthes tuberosa L.

Sotolli

Rosette

Native

Nardo

Herb

Native

Sansevieria trifasciata Prain



Herb

Exotic

Yucca guatemalensis Baker

Izotl, izote

Rosette

Native

Asphodelaceae Aloe vera (L.) Burm. f.

Sábila

Herb

Exotic

Gasteria sp.



Herb

Exotic

Kniphofia uvaria (L.) Oken

Española

Herb

Exotic

Micashibitl, alcanfoshibitl Teogopultzin, teogopaltzi

Herb

Cogopatzi Tegopultzin, teoyopultzin Chichilchibitl

Asteraceae Achillea millefolium L. Ageratina sp.

Ageratina sp. Ageratina sp. Ageratina sp.

Herb

491

Usesa

Ecological status

Cultivated by vegetative propagation Cultivated by vegetative propagation Gathering, enhancement, protected, and tolerated Gathering, enhancement, protected, tolerated Enhancement, tolerated

O

Domesticated

O

Domesticated

E

Wild

E

Wild

O

Wild

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation

O

Domesticated

O

Domesticated

Management status

E, LF, C Domesticated

Transplantation O, M of whole individuals, cultivated by vegetative propagation O Cultivated by vegetative propagation O Cultivated by vegetative propagation, enhancement, and tolerated

Ruderal

Native

Gathering

M

Ruderal

Herb

Native

M

Weedy

Herb Herb

Native Native

Cultivated by vegetative propagation Gathering Gathering

M M

Ruderal Wild

Herb

Native

Gathering

M

Wild

Domesticated

Weedy, ruderal

(continued)

492

Scientific name Ageratina espinosarum (A. Gray) R.M. King and H. Rob. Ageratina tomentella (Schrad.) R. M. King and H. Rob. Ambrosia sp. Artemisia absinthium L. Artemisia absinthium var. insipida Stechmann Artemisia ludoviciana subsp. mexicana (Willd. ex Spreng.) D.D. Keck Artemisia mexicana Willd. ex Spreng. Baccharis salicifolia (Ruiz & Pav.) Pers. Baccharis conferta Kunth. Barkleyanthus salicifolius (Kunth) H.E. Robins. And Brett. Bidens sp. Brickellia veronicifolia (Kunth) A. Gray Calea zacatechichi Schltdl. Calendula officinalis L. Calyptocarpus vialis Less. Chrysanthemum morifolium Ramat. Cirsium sp. Cosmos bipinnatus Cav. Erigeron karvinskianus DC. Erigeron sp. Galinsoga parviflora Cav. Gnaphalium americanum P. Mill.

J. J. Blancas Vázquez et al.

Common name

Growth type

Origin

Management status

Usesa

Ecological status



Herb

Native

Gathering

M

Wild



Herb

Native

Gathering

M

Wild

– Hierba maestra

Herb Herb

Native Exotic

Gathering Gathering

M M

Wild Wild

Hierba maestra

Herb

Exotic

Gathering

M

Ruderal

Iztauhyatl, iztagafiah, estafiate

Shrub

Native

Enhancement

M

Ruderal

Hierba maestra

Herb

Native

M

Ruderal



Shrub

Native

Cultivated by seed Gathering

M

Ruderal

Tlalpabashtli, tlachpabashtli, escobilla Yoyotli, azomiate

Shrub

Native

Gathering

U

Ruderal

Herb

Native

Gathering

M

Ruderal

Amozo –

Herb Shrub

Native Native

Gathering Gathering

M M

Ruderal Wild

Zacatechichi

Herb

Native

Gathering

M

Ruderal

Caléndula

Herb

Exotic

M

Domesticated

Iloshibitl

Herb

Native

Cultivated by seed Gathering

M

Ruderal

Campechana, crisantemo Pitzobiztli Mirasol

Herb

Exotic

O

Domesticated

Herb Herb

Native Native

Cultivated by seed Gathering Gathering

M O

Ruderal Wild

Ictebashibitl, iktigabashibitl Teshibitl –

Herb

Native

Gathering

M

Ruderal

Herb Herb

Native Native

M F

Ruderal Weedy, ruderal

Gordolobo

Herb

Native

Gathering Enhancement, gathering, and tolerated Enhancement, gathering, and tolerated

M

Ruderal, weedy

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . . Common name

Growth type

Origin

Gymnosperma glutinosum (Spreng.) Less. Jaegeria hirta (Lag.) Less. Jefea sp. Helianthus sp. Heterotheca inuloides Cass. Lactuca sativa L.

Escobilla

Shrub

Native



Herb

Native

– Shaltacto Arnica

Shrub Herb Herb

Lechuga

Herb

Leucanthemum vulgare Tourn. ex Lam. Matricaria chamomilla L. Melampodium divaricatum (Rich) DC.

Margarita

Herb

Manzanilla

Montanoa grandiflora DC. Montanoa sp. Parthenium tomentosum DC Piqueria trinervia Cav.

Scientific name

493 Ecological status

Management status

Usesa

Gathering, tolerated

AB, M, Ruderal U

Gathering, tolerated Native Gathering Exotic Gathering Native Cultivated by seed Exotic Cultivated by seed Naturalized Cultivated by seed

F, M

Weedy, ruderal

M O M

Wild Weedy, ruderal Weedy, ruderal

E

Domesticated

O

Weedy, ruderal

Herb

Exotic

Enhancement

M

Domesticated

Ojo de conejo

Herb

Native

E, F, M Ruderal

Tzatzaztli

Herb

Native

Cultivated by seed, enhancement, and tolerated Gathering

O

Wild

Tlatlatzca Sibashibitl

Herb Shrub

Native Native

Gathering Gathering

O M

Wild Wild

Sanigulas, hierba de San Nicolás Ahuabitl

Herb

Native

Tolerated

M

Weedy

Shrub

Native

M

Ruderal

Malagalquilitl, pipicha, pipitza

Herb

Native

E, M

Weedy

Porophyllum ruderale (Jacq.) Cass.

Papaloquilitl, papaloquelite

Herb

Native

E

Weedy

Pseudognaphalium conoideum Anderb. Telanthophora sp.

Aztoquilitl

Herb

Native

E

Weedy

Ecashibitl

Herb

Native

E

Wild

Sanvitalia procumbens Lam. Sigesbeckia jorullensis Kunth. Sonchus oleraceus L. Stevia sp.



Herb

Native

M, O

Weedy, ruderal



Herb

Native

M

Weedy

Memella Xogobashi, hierba del borracho Micaxochitl

Herb Herb

Exotic Exotic

Gathering, tolerated Cultivated by seed, enhancement, protected, and tolerated Cultivated by seed, enhancement Enhancement, tolerated Gathering, tolerated Gathering, tolerated Gathering, tolerated Tolerated Gathering

E E, M

Weedy, ruderal Ruderal

Herb

Native

Gathering

C

Ruderal

Pluchea carolinensis (Jacq.) G. Don Porophyllum linaria (Cav.) DC.

Stevia sp.

(continued)

494

J. J. Blancas Vázquez et al.

Usesa

Ecological status

Gathering, tolerated

M

Ruderal

Gathering, tolerated Cultivated by seed Enhancement, tolerated Gathering, enhancement, and tolerated Gathering, tolerated Gathering, tolerated

O

Domesticated

O

Weedy

M

Ruderal

M

Ruderal

O

Wild

M

Ruderal

Exotic

Tolerated

M

Weedy, ruderal

Herb

Native

O

Weedy, ruderal

Acahuale

Shrub

Native

Gathering, enhancement, and tolerated Tolerated

F

Weedy, ruderal

Vitzcobitl, capitaneja Zazastli

Herb

Native

Shrub

Native

FW, M Ruderal, weedy O Ruderal

Zazagapali (teresita) Tronadora

Herb

Native

Herb

Native

Viguiera dentata (Cav.) Spreng.

Chimalacate

Herb

Native

Viguiera sp. Xanthium strumarium L. Zaluzania montagnifolia (Sch. Bip.) Sch. Bip. Zinnia elegans Jacq.

Zojiejshibitl –

Herb Herb

Native Native

Gathering, tolerated Gathering, tolerated Gathering, tolerated Gathering, tolerated Gathering, enhancement, and tolerated Gathering Tolerated

Pepextli, pepebeshtli

Shrub

Native

Cabecitas

Herb

Native

Zinnia peruviana L.



Herb

Native

Balsaminaceae Impatiens balsamina L.

Chino, gachupín Herb

Exotic



Exotic

Common name

Growth type

Origin

Symphyotrichum subulatum (Michx.) G.L. Nesom Symphyotrichum sp.

Chichishibitl, cotchua

Herb

Exotic



Herb

Exotic

Tagetes erecta L.

Herb

Native

Tagetes filifolia Lag.

Xempaxóchitl, flor de muerto Anís

Herb

Native

Tagetes lucida Cav.

Pericón

Herb

Native

Tagetes lunulata Ort.

Flor de muerto silvestre Santa María shibitl

Herb

Native

Herb

Exotic

Diente de león

Herb

Acahual

Scientific name

Tanacetum parthenium (L.) Sch. Bip. Taraxacum officinale F. H. Wigg. Tithonia rotundifolia (Mill.) S. E. Blake Tithonia diversifolia (Hemsl.) A. Gray Verbesina crocata (Cav.) Less. Verbesina sp. Verbesina sp. Verbesina sp.

Impatiens zombensis Baker f.

Herb

Management status

C

Wild

O

Wild

M

Wild

M O

Wild Ruderal

Gathering, tolerated

M, F

Wild

Cultivated by seed Gathering, enhancement, and tolerated

O

Domesticated

O

Ruderal

Cultivated by vegetative propagation

O

Domesticated

O

Domesticated

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

Scientific name

Common name

Growth type

Origin

Management status Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation

495

Usesa

Ecological status

O

Domesticated

O

Domesticated

O

Domesticated

O

Domesticated

Impatiens sodenii Engel and Warb.



Shrub

Exotic

Impatiens walleriana Hook f.

Belén, chino

Herb

Exotic

Cecigxochitl, begonia

Herb

Exotic

Ala de ángel

Herb

Exotic

Elite comegaxóchitl, Bejuco de cera, eliguitl, elite, abedul

Tree

Native

Wild Transplantation M, FW, C, of whole SH individuals, enhancement, and tolerated

Chichictegomatl, Tree jícara de Monte Jacaranda Tree

Native

Wild

Cuaxilotl, cuajilote

Tree

Native

Cultivated by U seed Cultivated by O seed Transplantation E of whole individuals, enhancement, and tolerated

Axiote

Shrub

Exotic

Cultivated by seed

E

Wild

Mostaza

Herb

Exotic

Tolerated

E

Ruderal

Colesh, colechtenechtli, colechtenso

Herb

Exotic

E

Weedy, ruderal

Brassica sp.

Colesh cuagio

Herb

Exotic

E

Weedy

Lepidium virginicum L. Matthiola incana (L.) W.T. Aiton Raphanus raphanistrum L. Rorippa nasturtiumaquaticum (L.) Hayek

Chilchibitl

Herb

Native

M

Weedy, ruderal

Alhelí

Herb

Exotic

M

Domesticated

Rabanoquilitl

Herb

Exotic

Cultivated by seed, enhancement, and tolerated Gathering, tolerated Gathering, tolerated Cultivated by seed Enhancement, tolerated

E

Weedy

Atlanquilitl, berro

Herb

Exotic

E, M

Ruderal, wild

Begoniaceae Begonia semperflorens Link and Otto Begonia tuberosa Lam. Betulaceae Alnus acuminata Kunth

Bignoniaceae Crescentia alata Kunth Jacaranda mimosifolia D. Don. Parmentiera aculeata (kunth) Seem.

Bixaceae Bixa orellana L. Brassicaceae Brassica juncea (L.) Czern. Brassica rapa L.

Exotic

Cultivated by vegetative propagation Cultivated by vegetative propagation

Wild Wild

(continued)

496

Scientific name

J. J. Blancas Vázquez et al.

Common name

Growth type

Origin

Management status

Usesa

Ecological status

Cultivated by vegetative propagation Bromeliaceae Tillandsia deppeana Steud. Tillandsia gymnobotrya Baker Tillandsia imperialis E. Morren ex Mez Tillandsia lucida E. Morren ex Baker Tillandsia usneoides (L.) L. Burseraceae Bursera aspleniifolia Brandegee Bursera biflora (Rose) Standl. Bursera grandifolia (Schltdl.) Engl. Cactaceae Disocactus ackermannii (Haw.) Ralf Bauer Disocactus phyllanthoides (DC.) Barthlott Echinopsis sp.



Epiphyte Native

Gathering

O, R

Wild



Epiphyte Native

Gathering

O

Wild



Epiphyte Native

Gathering

O, R

Wild

Xolochi

Epiphyte Native

Gathering

O, R

Wild

Heno, pashtle

Epiphyte Native

Gathering

O

Wild

Copalhuabitl

Tree

Native

Gathering

RE

Wild

Copalhuabitl

Tree

Native

Gathering

RE

Wild

Chichiltihuabitl, Tree tepebashibitl

Native

Gathering

M

Wild

Cultivated by vegetative propagation Gathering

O

Wild

O

Wild

O

Wild

O

Wild

O, E

Wild

O

Wild

O

Wild

Nopalxochitl

Epiphyte Native



Epiphyte Native



Rosette

Epiphyllum sp.



Epiphyte Native

Hylocereus undatus (Haw.) Britton and Rose Mammillaria carnea Zucc. Ex Pfeiff. Mammillaria elegans DC. Mammillaria flavicentra Backeb. ex Mottram Mammillaria polyedra Mart. Opuntia ficus-indica (L.) Mill.

Pitahaya

Epiphyte Native

Cactus

Rosette

Native

Cactus

Rosette

Native

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Gathering, tolerated Gathering

Cactus

Rosette

Native

Gathering

O

Wild

Cactus

Rosette

Native

Gathering

O

Wild

Nopalli, nopal

Shrub

Native

E, M

Domesticated

Opuntia microdasys (Lehm.) Pfeiff.



Shrub

Native

Cultivated by vegetative propagation Cultivated by vegetative propagation

O

Wild

Exotic

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

497

Management status

Usesa

Ecological status

Epiphyte Native

Gathering

M

Wild



Epiphyte Exotic

Domesticated

Pitaya

Columnar Native cacti

O Cultivated by vegetative propagation E, FW Tolerated, enhancement, protected, transplantation of whole individuals, and cultivated by vegetative propagation

Cuatzapotl

Tree

Native

Enhancement, tolerated

E, O

Wild



Tree

Native

Gathering

FW

Wild

Panispatl

Herb

Exotic

O, E

Wild, ruderal

Canna x generalis L.H. Bailey and E.Z. Bailey Capparidaceae Cleome sp.



Herb

Native

Cultivated by seed Cultivated by seed

O

Domesticated



Herb

Native

Cleoserrata speciosa (Raf.) Iltis

Mabilquilitl

Herb

Native

Caricaceae Carica cauliflora Jacq.

Papaya cimarrona

Tree

Native

Carica papaya L.

Papaya

Tree

Native

Caryophyllaceae Dianthus caryophyllus L.

Clavel

Herb

Exotic

Nube

Herb

Exotic

Zapote de niño

Tree

Native

Scientific name Rhipsalis baccifera (Sol.) Stearn Schlumbergera truncata (Haw.) Moran Stenocereus pruinosus (Otto ex Pfeiff.) Buxb.

Calophyllaceae Mammea americana L. Cannabaceae Celtis caudata Planch. Cannaceae Canna indica L.

Gypsophila paniculata L. Chrysobalanaceae Couepia polyandra (Kunth) Rose

Common name

Growth type

Mazazalitzi

Origin

Transplantation O of whole individuals E Cultivated by seed, enhancement, tolerated

Domesticated

Weedy

Weedy

Cultivated by vegetative propagation Cultivated by vegetative propagation

E

Domesticated

E

Domesticated

Cultivated by vegetative propagation Cultivated by seed

O

Domesticated

O

Wild

Gathering, enhancement, tolerated

E

Wild

(continued)

498

J. J. Blancas Vázquez et al.

Origin

Management status

Usesa

Ecological status

Hierba del pollo Herb

Native

Gathering

M

Ruderal

Matlali, hierba del pollo Matlali

Herb

Native

Enhancement

M

Ruderal

Herb

Native

M

Wild

Tradescantia pallida (Rose) D.R. Hunt

Matlali

Herb

Native

M

Domesticated

Tradescantia sillamontana Matuda

Oreja de burro

Herb

Native

O

Domesticated

Tradescantia spathacea Sw.

Matlali

Herb

Native

O

Domesticated

Tradescantia zebrina Heynh. ex Bosse

Matlali

Herb

Native

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation

O

Domesticated



Herb

Native

O

Domesticated

Camote

Climb

Native

Cultivated by seed Cultivated by vegetative propagation

E

Weedy



Herb

Exotic

Cultivated by vegetative propagation

O

Ruderal

Santuario

Herb

Exotic

O

Ruderal

Sedum allantoides Rose

Deditos

Herb

Native

O

Domesticated

Sedum morganianum E. Walther

Cola de borrego Herb

Native

O

Wild

Sedum pachyphyllum Rose

Deditos del niño Herb Jesús

Native

O

Domesticated

Sedum palmeri S. Watson



Herb

Native

O

Domesticated

Sedum rubrotinctum R.T. Clausen

Deditos del niño Herb Jesús

Native

O

Domesticated

Sedum sp.

Deditos del Niño Herb Jesús

Native

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation

O

Domesticated

Scientific name Commelinaceae Commelina tuberosa Willd. Tinantia erecta (Jacq.) Schltdl. Tradescantia sp.

Convolvulaceae Dichondra sp. Ipomoea batatas (L.) Lam. Crassulaceae Kalanchoe daigremontiana Raym.-Hamet and H. Perrier Kalanchoe pinnata (Lam.) Pers.

Common name

Growth type

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

Usesa

Ecological status

Cultivated by seed Cultivated by seed

E

Domesticated

E

Weedy, domesticated

Cultivated by seed Cultivated by seed Cultivated by seed Cultivated by seed, enhancement Cultivated by seed Tolerated

E

Domesticated

E

Domesticated

E

Domesticated

E, M

Domesticated

E

Domesticated

S

Wild, ruderal

Cultivated by seed Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation, enhancement, tolerated Transplantation of whole individuals, tolerated

M, FW, Domesticated CO O Domesticated

Native

Gathering

O

Wild

Herb

Native

Tolerated

M

Wild

Barbasco

Herb

Native

Gathering, tolerated

M

Ruderal

Tliltlzapotl, zapote negro

Tree

Native

Cultivated by seed

O, E, Domesticated FW, M

Cola de caballo

Herb

Native

Gathering

O, M

Scientific name

Common name

Growth type

Origin

Cucurbitaceae Cucumis sativus L.

Pepino

Climb

Exotic Native

Cucurbita moschata Duchesne Cucurbita pepo L.

Climb Ziligayotl, cilagayotl, chilacayote Abayotli, Herb calabaza amarilla Calabaza Herb

Cucurbita sp.

Calabaza melón Herb

Native

Lagenaria siceraria (Molina) Standl.

Cuatecomatl

Climb

Native

Sechium edule (Jacq.) Sw. Sicyos parviflorus Willd. Cupressaceae Cupressus lusitanica Mill. Cupressus macrocarpa Hartw. ex. Gordon Cupressus sempervirens L.

Chayoctli, chayote Camachichio

Climb

Native

Climb

Native

Cedro

Tree

Exotic

Tlátzca, ciprés

Tree

Exotic

Tlátzca, ciprés

Tree

Exotic

Cupressus sp.

Teocobitl, Cedro Tree

Native

Juniperus deppeana Steud.

Tlátzca, ciprés

Tree

Native



Tree

Ogopetatl

Cucurbita ficifolia Bouché

Cyatheaceae Cyathea sp. Dennstaedtiaceae Pteridium sp. Dioscoreaceae Dioscorea floribunda M. Martens and Galeotti Ebenaceae Diospyros digyna Jacq. Equisetaceae Equisetum hyemale L.

499

Native Native

Management status

O

Domesticated

M, CO Wild

FW, LF Wild

Wild

(continued)

500

J. J. Blancas Vázquez et al.

Ecological status

Common name

Growth type

Origin

Management status

Usesa

Madroño

Tree

Native

Gathering

M, FW Wild

Azalea

Shrub

Exotic

O

Domesticated

Tetzmolli, tetzmole

Shrub

Native

Cultivated by seed Enhancement, tolerated

E, FW

Wild



Herb

Native

M

Wild, ruderal



Tree

Native

M

Wild

Corona de Jesús Herb

Exotic

O

Ruderal

Shrub Nexóchitl, nezpatlahuac, neztlixochtl, flor de nochebuena – Shrub

Native

Gathering, tolerated Enhancement, tolerated Cultivated by vegetative propagation Cultivated by vegetative propagation

O

Domesticated

O

Domesticated

Jatropha curcas L.

Piñón

Shrub

Native

E, M

Domesticated

Omphalea sp. Ricinus communis L.

Jonote Cuyabatzi, higuerilla

Tree Shrub

Native Exotic

Cultivated by vegetative propagation Cultivated by vegetative propagation Gathering Enhancement, tolerated

M, U M

Wild Ruderal

Huizache

Shrub

Native

Tolerated

FW

Ruderal

Cedro rosado

Tree

Exotic

FW, SH Wild

Calia secundiflora (Ortega) Yakovlev Canavalia villosa Benth.



Shrub

Native

Comegatl

Climb

Native

Conzattia multiflora (B.L. Rob.) Standl. Crotalaria pumila Ortega Desmodium aparines (Link) DC. Diphysa americana (Miller) M. Sousa Erythrina americana Mill.



Tree

Native

Cultivated by vegetative propagation, cultivated by seed Gathering, tolerated Cultivated by vegetative propagation Tolerated

Xayacatl

Herb

Native



Herb

Native

Enhancement, tolerated Tolerated

M, FW, Weedy, ruderal CO M Weedy, ruderal

Chiquiliche

Tree

Native

Tolerated

CO

Wild

Iquimite, Tree nigimitl, colorín

Native

Cultivated by vegetative propagation

E, LF

Wild

Scientific name Ericaceae Arbutus xalapensis Kunth Rhododendron indicum (L.) Sweet Vaccinium leucanthum Schltdl. Euphorbiaceae Acalypha alopecuroidea Jacq. Croton draco Schltdl. and Cham. Euphorbia milii Des Moul. Euphorbia pulcherrima Willd. ex Klotzsch Euphorbia trigona Haw.

Fabaceae Acacia farnesiana (L.) Willd Acrocarpus fraxinifolius Arn.

Exotic

M, O

Wild

O

Wild

SH

Wild

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

Scientific name

Common name

Growth type

Origin

Inga flexuosa Schltdl.

Topetli

Tree

Native

Leucaena diversifolia (Schltdl.) Benth. Leucaena esculenta (Moc. & Sessé ex DC.) Benth. Leucaena leucocephala (Lam.) de Wit Lupinus elegans Kunth Lupinus montanus HBK

Guaje

Tree

Native

Guaje guajilitl

Tree

Native

Guaje cimarrón

Tree



Management status

Usesa

501 Ecological status

Transplantation of whole individuals Enhancement, tolerated Cultivated by seed

E, FW, Wild SH

Native

Herb

Yoloczixihuitl

Lysiloma sp.

E

Wild

E, FW

Domesticated

Tolerated

E

Wild

Native

Gathering

O

Wild

Herb

Native

O, M

Weedy, ruderal

Tepebashitli

Tree

Native

FW

Wild

Mimosa aculeaticarpa Ortega Phaseolus coccineus L.



Shrub

Native

FW

Ruderal

Ilamatzin, xochiquilitl, bepa, frijol ayocote

Shrub

Native

E

Weedy, ruderal

Phaseolus sp.



Herb

Native

E

Domesticated

Phaseolus vulgaris L.

Pitza, frijol negro Chícharo

Herb

Native

E

Domesticated

Herb

Exotic

E

Domesticated

Nesneshtamaltzin

Tree

Native

O

Wild



Shrub

Native

Cultivated by seed, gathering, and tolerated Gathering, tolerated Gathering, tolerated Cultivated by seed, enhancement, protected, and tolerated Cultivated by seed Cultivated by seed Cultivated by seed Transplantation of whole individuals Transplantation of whole individuals

O

Ruderal



Shrub

Native

Tolerated

M

Ruderal



Shrub

Native

Tolerated

M

Ruderal

Haba

Herb

Exotic

Cultivated by seed

E

Domesticated

Copazole, apipitzactli

Tree

Native

M, FW Wild

Tlilibicawatl, nahuacocomole

Tree

Native

Gathering, enhancement, protected, and tolerated Gathering, tolerated

Pisum sativum L. Senna holwayana (Rose) Irwin et Barneby Senna multiglandulosa (Jacq.) H.S. Irwin and Barneby Senna obtusifolia (L) H.S. Irwin and Bareby Senna septemtrionalis (Viv.) H. S. Irwin and Barneby Vicia faba L. Fagaceae Quercus laurina Humb. and Bonpl.

Quercus rugosa Née

FW

Wild

(continued)

502

J. J. Blancas Vázquez et al.

Common name

Growth type

Origin

Management status

Usesa

Ecological status

Tlilibiabatl Abatl cuchara Tepoztenawatl, tepozabatl Tamalabatl

Tree Tree Tree

Native Native Native

Gathering Gathering Gathering

FW FW FW

Wild Wild Wild

Tree

Native

Gathering, enhancement, protected, and tolerated

E, FW, Wild M



Herb

Exotic

Domesticated

Geranio

Herb

Exotic

O Cultivated by seed, cultivated by vegetative propagation O Cultivated by seed, cultivated by vegetative propagation

Violeta

Herb

Exotic

Cultivated by seed

O

Domesticated

Ebespatl

Herb

Native

Cultivated by vegetative propagation

O

Wild

Hortensia

Herb

Exotic

Cultivated by vegetative propagation

O

Domesticated

Pulcobitl, camparaguey

Shrub

Native

Gathering, tolerated

O

Wild

Palma cimarrón Herb

Exotic

O

Domesticated

Palma

Herb

Exotic

O

Domesticated

Gladiolus grandiflorus Andrews

Palma

Herb

Exotic

O

Domesticated

Gladiolus  hortulanus L.H. Bailey Iris germanica L.

Palma, lirio

Herb

Exotic

O

Domesticated

Palma, lirio

Herb

Exotic

O

Wild

Tigridia pavonia (L. f.) DC.

Tlalteztli, flor de trueno

Herb

Native

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation, enhancement, protected, gathering, and tolerated

O, E

Ruderal, weedy

Scientific name Quercus sp. Quercus sp. Quercus urbanii Trel. Roldana candicans (Née) Villaseñor, S. Valencia and Coombes Geraniaceae Pelargonium betulinum (L.) L’Hér. ex Aiton Pelargonium zonale (L.) L’Hér.

Gesneriaceae Saintpaulia ionantha H. Wendl. Heliconiaceae Heliconia bihai (L.) L. Hydrangeaceae Hydrangea macrophylla (Thunb) Ser. Hypericaceae Vismia camparaguey Sprague and L. Riley Iridaceae Crocosmia aurea (Pappe ex Hook.) Planch. Gladiolus communis L.

Domesticated

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

Usesa

Ecological status

Cultivated by seed Cultivated by seed

E

Domesticated

Gathering, tolerated Gathering

M

Ruderal

M

Wild

Cultivated by vegetative propagation Cultivated by seed, tolerated Cultivated by vegetative propagation Gathering, tolerated Cultivated by seed Cultivated by vegetative propagation Cultivated by vegetative propagation Gathering

O

Domesticated

M

Weedy, ruderal

E, M

Domesticated

M

Wild

E

Domesticated

M, O

Domesticated

E, M

Domesticated

M

Wild

Gathering, tolerated Gathering, tolerated Gathering

M, O

Wild

M

Ruderal

Native Exotic

Gathering Cultivated by seed

M E

Wild Domesticated

Sogogotl, laurel Tree

Native

E, R

Wild

Tree Tree

Native Native

Gathering, enhancement, protected, tolerated, and transplantation of whole individuals Gathering Cultivated by seed

Scientific name

Common name

Growth type

Origin

Juglandaceae Juglans nigra L.

Nuez

Tree

Exotic

Juglans regia L.

Nuez, nogal

Tree

Exotic

Hierba del borracho Payaniltzin, Payamile

Shrub

Native

Shrub

Native

Siempre me verás Shrub así

Exotic

Marrubium vulgare L. Mentha spicata L.

Marrubio

Herb

Exotic

Hierbabuena

Herb

Exotic

Ocimum micranthum Willd. Origanum majorana L. Plectranthus tomentosus Benth.

Xoshashibitl, clavo Mejorana

Herb

Native

Shrub

Exotic

Planta del vaporub

Herb

Exotic

Rosmarinus officinalis L.

Romero

Shrub

Exotic

Salvia boegei Ramamoorthy Salvia elegans Vahl

Tetenhuitzi

Herb

Native

Mirto rojo

Herb

Native

Salvia leucantha Cav.

Cordón de Jesús Herb

Native

Salvia polystachya Cav. Salvia sp. Thymus vulgaris L.

Totochía

Shrub

Native

– Tomillo

Herb Shrub

Lamiaceae Clinopodium laevigatum Standl. Clinopodium mexicanum (Benth.) Govaerts Coleus scutellarioides (L.) Benth.

Lauraceae Litsea glaucescens Kunth

Nectandra sp. Persea americana Mill.

503

Ahuacatl, Aguacate

Management status

E, FW, Wild CO

M, O, R Wild

CO Wild, ruderal E, FW, Domesticated SH

(continued)

504

J. J. Blancas Vázquez et al.

Growth type

Origin

Chinene

Tree

Native

Enhancement, E, CO, transplantation SH, U of whole individuals, and tolerated

Liliaceae Lilium candidum L.

Lirio, azucena

Herb

Exotic

O

Domesticated

Lilium sp.

Azucena

Herb

Exotic

Cultivated by vegetative propagation Cultivated by vegetative propagation

O, M

Domesticated

Salvarón

Shrub

Native

Cultivated by vegetative propagation

O

Wild

Risco

Herb

Native

Gathering

O

Wild

Cola de ratón

Herb

Native

Gathering

O

Wild

Ehegashibitl Herb Axochitl, flor de Herb agua Granada roja Shrub

Native Native

Gathering Gathering

M O

Ruderal Wild

Exotic

Cultivated by vegetative propagation

E

Domesticated

Magnoliaceae Magnolia grandiflora L.

Magnolia

Tree

Native

O

Wild

Talauma mexicana (DC.) G. Don

Yoloxóchitl, flor Tree de corazón, magnolia

Native

Cultivated by vegetative propagation Gathering, enhancement, protected, and tolerated

O, M

Wild

Totolzapotl

Tree

Native

Enhancement, tolerated

SH

Wild

Jonote Blanco

Tree

Native

Gathering

M

Wild

Xonocobitl, xonotl, chachamacobitl jonote

Tree

Native

Tulipán, rosa china

Shrub

Exotic

Enhancement, SH transplantation of whole individuals, and tolerated O Cultivated by vegetative propagation

Persea schiedeana Nees

Loranthaceae Psittacanthus calyculatus (DC.) G. Don Lycopodiaceae Lycopodium clavatum L. Lycopodium sp. Lythraceae Cuphea cyanea DC. Lythrum sp. Punica granatum L.

Malphigiaceae Bunchosia lindeniana A. Juss. Malvaceae Hampea integerrima Schltdl. Heliocarpus appendiculatus Turcz.

Hibiscus rosasinensis L.

Management status

Ecological status

Common name

Scientific name

Usesa

Wild

Wild

Domesticated

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

505

Usesa

Ecological status

Cultivated by seed Enhancement, tolerated Enhancement, tolerated

O

Ruderal

M

Weedy, ruderal

M, O

Weedy, ruderal

Native

Gathering

O

Wild

Tree

Exotic

O

Ruderal

Higo, breva

Tree

Exotic

M

Ruderal

Ficus elastica Roxb. Ex Hornem

Hule

Tree

Exotic

O

Ruderal

Morus celtidifolia Kunth Musaceae Musa paradisiaca L.

Moro

Tree

Native

Cultivated by seed Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by seed, tolerated

O

Wild

Plátano, xochicualli

Herb

Exotic

Cultivated by vegetative propagation

E

Domesticated

Itllahaldtenzo

Tree

Exotic

Cultivated by vegetative propagation

O

Wild

Eucalipto

Tree

Exotic

M, SH

Ruderal

Kowiawak

Tree

Native

Cultivated by seed Gathering

M

Wild

Mototetl, pimientón

Tree

Native

Psidium guajava L.

Guayaba

Tree

Native

Nyctaginaceae Bougainvillea spectabilis Willd.

Bugambilia

Shrub

Exotic

Scientific name

Common name

Growth type

Origin

Hibiscus trionum L.



Herb

Exotic

Malva parviflora L.

Orashibitl

Herb

Native

Sida rhombifolia L.

Tlaxpabaxtli, escobilla

Shrub

Native



Shrub



Ficus carica L.

Melastomataceae Chaetogastra scabriuscula (Schltdl.) P.J.F. Guim. and Michelang. Moraceae Ficus benjamina L.

Myricaceae Morella cerifera (L.) Small Myrtaceae Eucalyptus globulus Labill. Eugenia acapulcensis Steud. Eugenia capuli (Schltdl. & Cham.) Hook. and Arn.

Management status

Transplantation E, SH of whole individuals, enhancement, protected, gathering, and tolerated E, M, Cultivated by FW vegetative propagation, cultivated by seed Cultivated by vegetative propagation

O, M

Wild

Domesticated, ruderal

Domesticated

(continued)

506

Scientific name Pisoniella arborescens (lag & Rdr.) Standl. Oleaceae Fraxinus uhdei (Wenz.) Lingelsh. Jasminum fruticans L.

Jasminum mesnyi Hance Onagraceae Fuchsia magellanica Lam. Lopezia racemosa Cav. Orchidaceae Laelia superbiens Lindl. Prosthechea citrina (Lex.) W.E. Higgins Prosthechea cochleata (L.) W.E. Higgins Prosthechea rhynchophora (A. Rich. & Galeotti) W.E. Higgins Prosthechea vitellina (Lindl.) W.E. Higgins Stanhopea oculata (G. Lodd.) Lindl. Stanhopea tigrina Bateman ex Lindl. Oronbachaceae Castilleja tenuiflora Benth. Oxalidaceae Oxalis corniculata L. Paeoniaceae Paeonia lactiflora Pall. Papaveraceae Argemone mexicana L.

Argemone platyceras Link and Otto

J. J. Blancas Vázquez et al.

Management status

Ecological status

Common name

Growth type

Origin



Shrub

Native

Cultivated by vegetative propagation

FW, M, Wild SR

Chichistlahui, chichiscuitl, Fresno Jazmín

Tree

Native

Enhancement, tolerated

M, FW, Wild SH

Shrub

Exotic

O, LF

Domesticated

Jazmín

Shrub

Exotic

Cultivated by vegetative propagation Cultivated by vegetative propagation

F, O

Domesticated

Aretillo

Shrub

Exotic

O

Domesticated



Shrub

Native

Cultivated by vegetative propagation Enhancement, tolerated

F

Weedy, ruderal



Epiphyte Native

Gathering

O

Wild



Epiphyte Native

Gathering

O

Wild



Epiphyte Native

Gathering

O

Wild



Epiphyte Native

Gathering

O

Wild



Epiphyte Native

Gathering

O

Wild



Epiphyte Native

Gathering

O

Wild

Tehuanxochitl

Epiphyte Native

Gathering

O

Wild



Herb

Native

Tolerated

M, O

Wild

Limonadas, trébol

Herb

Exotic

Gathering

E

Ruderal

Rosa de Monte

Herb

Exotic

Cultivated by seed

O

Domesticated

Papalotito

Herb

Native

M

Weedy, ruderal

Chicalote

Herb

Native

Gathering, enhancement, and tolerated Tolerated

O

Weedy, ruderal

Usesa

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

Scientific name

Common name

Growth type

Origin

Bocconia frutescens L.

Gordolobo

Herb

Native

Papaver rhoeas L.

Amapola

Herb

Exotic

Maracuyá, granadilla

Climb

Exotic

Passiflora ligularis L.

Granadilla

Climb

Native

Passiflora subpeltata Ortega Pentaphylacaceae Ternstroemia sp.

Matetecomatl

Climb

Native

Flor de tila

Shrub

Native

Hachevitl, tila

Tree

Native

Molquilitl, molli, yamolli

Herb

Soplador

Passifloraceae Passiflora edulis Sims

507

Usesa

Ecological status

Gathering, tolerated Cultivated by seed

M

Ruderal

M, O

Weedy, ruderal

Cultivated by vegetative propagation Cultivated by vegetative propagation Gathering, tolerated

O, E, M Domesticated

Management status

E

Wild

M

Ruderal

Gathering, enhancement, protected, and tolerated Enhancement, tolerated

M

Wild

M

Wild

Native

Gathering, enhancement, and tolerated

E

Weedy, ruderal

Herb

Native

Gathering

CO

Wild

Pino

Tree

Native

Gathering

CO

Wild

Ocotl, ocote, pino

Tree

Native

O, FW, Wild CO

Pinus teocote Schltdl. and Cham.

Ocotl, ocote, pino

Tree

Native

Cultivated by seed, enhancement, and tolerated Cultivated by seed, enhancement, and tolerated

FW

Wild

Piperaceae Peperomia caperata Yunck.



Herb

Exotic

O

Domesticated

Peperomia nigropunctata Miq.

Tequilitl, tequelite

Herb

Native

E

Wild

Peperomia obtusifolia (L.) A. Dietr.



Herb

Native

O

Wild

Peperomia peltilimba C. DC. ex Trel. Peperomia sp.

Tehuantequilitl

Herb

Native

Cultivated by vegetative propagation Gathering, enhancement, tolerated Cultivated by vegetative propagation Gathering

E

Wild

Huetshibitl

Herb

Native

Gathering

M

Wild

Ternstroemia lineata DC. Phytolaccaceae Phytolacca icosandra L. Picramniaceae Picramnia antidesma Sw. Pinaceae Pinus montezumae Lamb. Pinus patula Schltdl. and Cham.

(continued)

508

J. J. Blancas Vázquez et al.

Usesa

Ecological status

Cultivated by vegetative propagation, enhancement

E, M

Domesticated

O

Domesticated

O

Domesticated

F

Weedy, ruderal

Native

Cultivated by vegetative propagation Cultivated by vegetative propagation Enhancement, tolerated Tolerated

M

Wild

Tree

Native

Tolerated

M, FW Wild

Agatl, Carrizo

Herb

Exotic

CO

Ruderal

Bambusa vulgaris Schrad. ex J.C. Wendl. Bothriochloa laguroides (DC.) Herter Cymbopogon citratus (DC.) Stapf

Agatl, Carrizo

Herb

Exotic

CO

Domesticated

Calzácatl

Herb

Native

Enhancement, tolerated Cultivated by vegetative propagation Enhancement, tolerated

CO

Weedy, ruderal

Zacate limón

Herb

Exotic

M

Domesticated

Pennisetun purpureum Schumach

Molotzin, zacatillo

Herb

Exotic

F

Ruderal and weedy

Saccharum officinarum L.

Caña de azúcar

Herb

Exotic

E

Domesticated

Zea mays L.

Maíz

Herb

Native

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by seed

Xoghilitl, Herb Lengua de vaca

Exotic

Enhancement, tolerated

E

Weedy, ruderal



Herb

Native

Gathering

O

Wild

Nogopetatl, asta Herb de ciervo

Exotic

Cultivated by vegetative propagation

O

Wild

Common name

Growth type

Origin

Tlanepaquilitl, tlanilpayilitl, hierba santa

Shrub

Native

Plantaginaceae Antirrhinum majus L.

Perrito

Herb

Exotic

Digitalis purpurea L.



Herb

Exotic

Plantago alismatifolia Pilg. Russelia obtusata S. F. Blake Platanaceae Platanus mexicana Moric.

Lengua de vaca, Herb Toro Lengua Aretillo Herb

Exotic

Papaloguitl, papalohuitl, álamo

Scientific name Piper auritum Kunth.

Poaceae Arundo donax L.

Polygonaceae Rumex crispus L. Polypodaceae Phlebodium pseudoaureum (Cav.) Lellinger Platycerium alcicorne Desv.

Management status

E, F, C, Domesticated M

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . . Ecological status

Scientific name

Common name

Growth type

Portulacaceae Portulaca oleracea L.

Verdolaga

Herb

Uncertain Cultivated by seed, enhancement, protected, and tolerated

O, E, M Weedy, ruderal

Primulaceae Anagallis arvensis L.



Herb

Exotic

F, M

Weedy, ruderal

Lysimachia nummularia L. Pteridaceae Adiantum capillusveneris L.



Herb

Exotic

O

Ruderal



Herb

Exotic

Ruderal

Llavea cordifolia Lag.



Herb

Native

Transplantation O of whole individuals O Cultivated by vegetative propagation

Chichibitl Botón de oro, tepostrompo

Herb Herb

Native Native

Gathering Gathering

M M

Wild Weedy, ruderal

Espire –

Shrub Tree

Native Native

M O

Wild Wild

Sageretia elegans (Kunth) Brongn. Rosaceae Amelanchier denticulata (Kunth) K. Koch Cercocarpus fothergilloides Kunth Cercocarpus macrophyllus C.K. Schneid. Crataegus mexicana DC.



Tree

Native

Gathering Gathering, tolerated Gathering

E

Wild

Tlaxistle

Shrub

Native

Gathering

M

Ruderal

Tepellili

Tree

Native

Tepoztenabatl

Tree

Native

Enhancement, tolerated Tolerated

CO, Wild FW, SH M, FW Wild

Texocotl, tejocote

Tree

Native

E, LF, M

Wild

Crataegus pubescens (C. Presl) C. Presl

Texocotl

Tree

Native

Transplantation of whole individuals, enhancement, protected, and tolerated Transplantation of whole individuals, enhancement, protected, and tolerated

E, M, FW

Wild

Ranunculaceae Ranunculus sp. Ranunculus petiolaris Humb., Bonpl. and Kunth ex DC. Rhamnaceae Ceanothus sp. Rhamnus sp.

Origin

Management status

509

Enhancement, tolerated Enhancement, tolerated

Usesa

Wild

(continued)

510

J. J. Blancas Vázquez et al.

Usesa

Ecological status

E

Domesticated

E

Domesticated

E

Domesticated

E, FW

Domesticated

E

Domesticated

E

Domesticated

E, FW

Domesticated

Native

Cultivated by seed, cultivated by vegetative propagation Cultivated by seed Cultivated by seed Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Transplantation of whole individuals, enhancement, protected, and tolerated Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by vegetative propagation Enhancement, tolerated Tolerated

Shrub

Native



Shrub



Café

Common name

Growth type

Origin

Cydonia oblonga Mill.

Membrillo

Tree

Exotic

Eriobotrya japonica (Thunb.) Lindl. Fragaria mexicana Schltdl. Malus domestica Borkh.

Xoxogotl, níspero Fresa

Tree

Exotic

Herb

Exotic

Manzana

Tree

Exotic

Prunus armeniaca L.

Chabacano

Tree

Exotic

Prunus domestica L.

Ciruela

Tree

Exotic

Prunus persica (L.) Batsch

Durazno

Tree

Exotic

Prunus serotina Ehrh.

Capolcuahuitl, capulín

Tree

Native

Pyrus communis L.

Pera

Tree

Exotic

Pyrus malus L.

Perón

Tree

Exotic

Rosa sp.

Rosa

Shrub

Native

Rosa gallica L.

Rosa de castilla Shrub

Exotic

Rubus adenotrichos Schltdl. Rubus eriocarpus Liebm. Rubiaceae Bouvardia ternifolia (Cav.) Schltdl. Chiococca alba (L.) Hitchc. Coccocypselum cordifolium Nees and Mart. Coffea arabica L.

Xoxonte, zarzamora Zarzamora

Shrub

Native

Shrub

Cagastli

Scientific name

Management status

E, FW, Wild M

E

Domesticated

E

Domesticated

O

Domesticated

O

Domesticated

E, LF

Ruderal

E

Wild

Tolerated

P

Ruderal

Native

Tolerated

O

Wild

Herb

Native

Gathering

M

Wild

Shrub

Exotic

Cultivated by vegetative propagation

E, FW

Domesticated

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

511

Usesa

Ecological status

Cultivated by vegetative propagation Enhancement, tolerated

O

Domesticated

M

Weedy

Cultivated by vegetative propagation Cultivated by seed Cultivated by seed Cultivated by seed Cultivated by vegetative propagation Enhancement, tolerated

E

Domesticated

E, FW

Domesticated

E

Domesticated

E, FW

Domesticated

O

Wild

M

Ruderal

Native

Tolerated

FW, SH Wild

Herb

Native

Gathering

O

Acezintle

Tree

Native

CO, SH Wild



Tree

Exotic

Cuetzalcahuitl

Tree

Native

Tolerated, enhancement, protected, and transplantation of whole individuals Enhancement, tolerated Tolerated

Urvillea ulmacea Kunth Sapotaceae Manilkara zapota (L.) P. Royen



Tree

Native

Tolerated

Zapote Blanco

Tree

Native

Pouteria campechiana (Kunth) Baehni

Zapote de niño

Tree

Native

Pouteria sapota (Jacq.) H.E. Moore and Stearn

Mamey

Tree

Native

Cultivated by vegetative propagation Cultivated by seed, enhancement, protected, and tolerated Cultivated by seed, propagación vegetativa

Common name

Growth type

Origin

Gardenia jasminoides J. Ellis

Gardenia

Shrub

Exotic

Hamelia patens Jacq.

Cocacabatzi

Shrub

Native

Rutaceae Citrus aurantifolia Swingle

Limón

Tree

Exotic

Citrus limetta Risso

Lima

Tree

Exotic

Citrus reticulata Blanco Citrus sinensis (L.) Osbeck Ptelea sp.

Mandarina

Tree

Exotic

Naranja

Tree

Exotic



Tree

Native

Ruta chalepensis L.

Ruda

Shrub

Exotic

Abeshotl

Tree

Tempala

Scientific name

Salicaceae Salix taxifolia Kunth Santalaceae Phoradendron sp. Sapindaceae Acer negundo L.

Allophylus cominia (L.) Sw. Cupania dentata DC.

Management status

F

Wild

Wild

CO, Wild FW, SH F Wild

E

Wild

E, M

Wild

E, M

Wild

(continued)

512

J. J. Blancas Vázquez et al.

Growth type

Origin

Tempesquistle

Tree

Native

Transplantation E, CO, SH, U of whole individuals, enhancement, protected, and tolerated

Wild

Pepeshtli Globitos

Shrub Herb

Native Native

Gathering Gathering

M M

Wild Weedy

Flor de cochi

Herb

Native

Tolerated

F

Weedy, ruderal

Tlanpatzitzi

Herb

Native

Gathering, tolerated

F, O

Weedy, ruderal



Shrub

Native

Gathering

CO, Wild FW, SH

Ogopetatl, doradilla

Herb

Native

Gathering

O, M

Wild

Sojacobitl, soquicobitl, limoncillo

Tree

Native

Protected

M

Wild

Tree Tzapa, xóchitltecomatl, floripondio, florifundio Tree Tzapa, xóchitltecomatl, floripondio, florifundio Chile poblano Herb

Exotic

Cultivated by vegetative propagation

O, M

Domesticated

Exotic

Cultivated by vegetative propagation

O, M

Domesticated

Native

E

Domesticated

Capsicum annuum var. glabriusculum (Dunal) Heiser and Pickersgill Capsicum pubescens Ruiz and Pav. Cestrum nocturnum L.

Chiltepin

Herb

Native

E

Wild

Chile canario

Herb

Exotic

E

Domesticated

Zopeliquilitl, huele de noche

Herb

Native

E

Weedy

Datura stramonium L.

Toloatzin, toloache Xaltocto, aztoquilitl, quelite de zorra

Herb

Native

M

Weedy, ruderal

Herb

Exotic

Cultivated by seed Cultivated by seed, enhancement, and tolerated Cultivated by seed Transplantation of whole individuals, cultivated by vegetative propagation Enhancement, tolerated Cultivated by seed, enhancement, and tolerated

E

Weedy, ruderal

Sideroxylon palmeri (Rose) T.D. Penn.

Scrophulariaceae Buddleja sp. Calceolaria tripartita Ruiz and Pavón Castilleja arvensis Schltdl. and Cham. Lamourouxia multifida Kunth Salicaceae Xylosma flexuosa (Kunth) Hemsl. Selaginellaceae Selaginella rupestris (L.) Spring Siparunaceae Siparuna thecaphora (Poepp. & Endl.) A. DC. Solanaceae Brugmansia  insignis (Barb. Rodr.) Lockwood ex R.E. Schult. Brugmansia candida Pers.

Capsicum annuum L.

Jaltomata procumbens (Cav.) J. L. Gentry

Management status

Ecological status

Common name

Scientific name

Usesa

(continued)

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

Scientific name Lycopersicon esculentum Mill.

Nicotiana glauca Graham Petunia hybrida Vilm. Physalis philadelphica Lam. Solandra grandiflora Sw. Solanum americanum Mill. Solanum glaucescens Zucc.

Solanum lanceolatum Cav. Solanum laxum Spreng.

Common name

Growth type

Origin

Management status

513

Usesa

Ecological status

Herb Citlaltomatl, tomate de lucero, tomate guajillo, tomate de ratón Tepebeshotl, Tree tebebeshotl Betunia Herb

Exotic

Cultivated by seed, enhancement, and tolerated

E

Weedy, ruderal

Native

M

Ruderal

O

Domesticated

Miltomatl, tomate de cáscara Copa de oro

Herb

Native

Gathering, tolerated Cultivated by seed Enhancement, tolerated

E

Weedy, ruderal

Shrub

Native

O

Wild

Tomaquilitl, hierba mora Cuatomatl, tomate de corazón o de Monte Valeriana

Herb

Native

E

Weedy, ruderal

Herb

Native

E, M

Weedy

Herb

Native

Cultivated by vegetative propagation Enhancement, tolerated Cultivated by seed, enhancement, and tolerated Gathering, tolerated Cultivated by seed

E, M

Ruderal

O

Wild

Exotic

Comecaxochitl, Climb lágrima, ramo de novia, velo de novia Xaltocto, Herb coccoyo

Exotic

Native

Enhancement, tolerated

E

Weedy, ruderal

Tlahuatetl, tlacotetl, papa

Herb

Exotic

E

Domesticated

Solanum wendlandii Hook. f.



Climb

Exotic

O

Wild

Witheringia solanacea L’hér. Theaceae Camellia japonica L.

Hierbamora cimarrona

Shrub

Native

Cultivated by vegetative propagation Cultivated by vegetative propagation Cultivated by seed

E

Weedy

Camelia

Shrub

Exotic

Cultivated by vegetative propagation

O

Domesticated



Herb

Exotic

Cultivated by vegetative propagation

O

Ruderal

Ilite

Tree

Native

Enhancement, tolerated

CO, Wild FW, SH

Guarumbo hormiguero

Tree

Native

Tolerated

M

Solanum nigrescens M. Martens and Galeotti Solanum tuberosum L.

Tropaeolaceae Tropaeolum majus L.

Ulmeaceae Ulmus mexicana (Liebm.) Planch. Urticaceae Cecropia obtusifolia Bertol.

Wild

(continued)

514

J. J. Blancas Vázquez et al.

Usesa

Ecological status

Cultivated by vegetative propagation, enhancement, and tolerated

O

Weedy, ruderal

O, P

Ruderal

Native

Cultivated by vegetative propagation, tolerated Gathering

E, M

Wild

Herb

Native

Gathering

M

Ruderal

Violeta cimarrona

Herb

Exotic

Cultivated by vegetative propagation

M

Weedy

Palma cimarrona

Herb

Exotic

Cultivated by vegetative propagation

O

Wild

Zingiberaceae Alpinia purpurata (Vieill.) K. Schum.



Herb

Exotic

O, C

Wild

Renealmia alpinia (Rottb.) Maas

Huazmole, veligmolli

Herb

Native

E

Wild

Renealmia mexicana Klotzsch ex Petersen

Palma velijmolli

Herb

Native

O

Wild

Zingiber officinale Roscoe

Gengibre

Herb

Exotic

Cultivated by vegetative propagation Transplantation of whole individuals, cultivated by vegetative propagation, enhancement, and protected Cultivated by vegetative propagation, enhancement Cultivated by vegetative propagation

E

Domesticated

Scientific name Pilea microphylla (L.) Liebm.

Verbenaceae Lantana camara L.

Lippia graveolens Kunth Phyla scaberrima (Juss. ex Pers.) Moldenke Violaceae Viola odorata L.

Xanthorrhoeaceae Phormium sp.

Common name

Growth type

Origin

Lenteja

Herb

Native

Mala suegra

Shrub

Native

Orégano

Herb

Egegachibitl

Management status

a Uses (AB alcoholic beverages, C ceremonial, CO construction, E edible, F fodder, FI fibers, FW firewood, LF living fence, M medicinal, O ornamental, P poison, R ritual, RE resine, S saponiferous, SH shade, SR soil retention, Us utensils)

References Aguilar A, Camacho JR, Chino S, Jacques P, López M. Herbario medicinal del Instituto Mexicano del Seguro Social (IMSS). Información etnobotánica. Ciudad de México: IMSS; 1994. Arellanes Y, Casas A, Arellanes A, Vega E, Blancas J, Vallejo M, Torres I, Rangel-Landa S, MorenoCalles AI, Solís L, Pérez-Negrón E. Influence of traditional markets on plant management in the Tehuacan Valley. J Ethnobiol Ethnomed. 2013;9(1) https://doi.org/10.1186/1746-4269-9-38.

Ethnobotany of the Nahua People: Plant Use and Management in the. . .

515

Berkes F. Sacred ecology. Traditional ecological knowledge and resource management. Philadelphia: Taylor & Francis; 1999. Bernard HR. Research methods in anthropology. Rowman: Altamira Press; 2011. Blancas J, Casas A, Pérez-Salicrup D, Caballero J, Vega E. Ecological and socio-cultural factors influencing plant management in Náhuatl communities of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013:9(1). https://doi.org/10.1186/1746-4269-9-39. Brandt R, Mathez-Stiefel SL, Lachmuth S, Hensen I, Rist S. Knowledge and valuation of Andean agroforestry species: the role of sex, age, and migration among members of a rural community in Bolivia. J Ethnobiol Ethnomed. 2013;9:83. https://doi.org/10.1186/1746-4269-9-83. Guinand Y, Lemessa D. Wild food plants in southern Ethiopia: reflections on the role of “famine foods” at a time of drought. Addis Ababa: UN-EUE; 2000. https://reliefweb.int/report/ethiopia/ wild-food-plants-southern-ethiopia-reflections-role-famine-foods-time-drought INEGI. Puebla. Aspectos geográficos. 2020. https://www.inegi.org.mx/contenidos/app/areasgeo graficas/resumen/resumen_21.pdf Instituto Nacional de Lenguas Indígenas (INALI). Agrupación lingüística: náhuatl. Familia lingüística: Yuto-nahua. 2022. https://www.inali.gob.mx/clin-inali/html/l_nahuatl.html Larios C, Casas A, Vallejo M, Moreno-Calles AI, Blancas J. Plant management and biodiversity conservation in Náhuatl homegardens of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013;9:74. https://doi.org/10.1186/1746-4269-9-74. Lotero-Velásquez E, García-Frapolli E, Blancas J, Casas A, Martínez-Ballesté A. Eco-symbiotic complementarity and trading networks of natural resources in Nahua communities in mountain regions of Mexico. Hum Ecol. 2022;50:307–19. https://doi.org/10.1007/s10745-022-00311-x. Martin GJ. Ethnobotany. London: Chapman & Hall; 1995. Morrone JJ. Biogeographic regionalization of the Sierra Madre del Sur province, Mexico. Rev Mex Biodivers. 2017;88(3):710–4. Mota C. Plantas comestibles de la Sierra Negra de Puebla, México. MSc Thesis. Texcoco: Universidad Autónoma Chapingo; 2007. Nesheim I, Dhillion SS, Anne SK. What happens to traditional knowledge and use of natural resources when people migrate? Hum Ecol. 2006;34:99–131. https://doi.org/10.1007/s10745005-9004-y. Olvera AA. ¡Mayolito in kilitl! / ¡Que viva el quelite! Recetario de quelites de Coyomeapan, Puebla. Bachelor Thesis. Morelia: Universidad Nacional Autónoma de México; 2016. Redžić S. Use of wild and semi-wild edible plants in nutrition and survival of people in 1430 days of siege of Sarajevo during the war in Bosnia and Herzegovina (1992–1995). Coll Antropol. 2010;34:551–70. Santiago-Alvarado M, Montaño-Arias G, Espinosa D. Áreas de endemismo de la Sierra Madre del Sur. In: Luna-Vega I, Espinosa D, Contreras-Medina R, editors. Biodiversidad de la Sierra Madre del Sur. Ciudad de México: UNAM; 2016. p. 431–48. Sepúlveda MT. Los lienzos de San Juan Cuauhtla, Puebla. Ciudad de México: CONACULTAINAH; 2005. Sulaiman N, Pieroni A, Sõukand R, Polesny Z. Food behavior in emergency time: wild plant use for human nutrition during the conflict in Syria. Foods. 2022;11(2):177. https://doi.org/10.3390/ foods11020177. Vallejo M, Casas A, Pérez-Negrón E, Moreno-Calles AI, Hernández-Ordoñez O, Téllez O, Dávila P. Agroforestry systems of the lowland alluvial valleys of the Tehuacán-Cuicatlán Biosphere Reserve: an evaluation of their biocultural capacity. J Ethnobiol Ethnomed. 2015;11(1):8. https://doi.org/10.1186/1746-4269-11-8. Vázquez-Medina B. Diferencias de género en el conocimiento y uso de plantas medicinales en Ahuatla y Xocotla, municipio de Coyomeapan, Puebla. MSc Thesis. Texcoco: Colegio de Postgraduados; 2010.

516

J. J. Blancas Vázquez et al.

Vorstenbosch T, de Zwarte I, Duistermaat L, van Andel T. Famine food of vegetal origin consumed in the Netherlands during World War II. J Ethnobiol Ethnomed. 2017;13:63. Zarazúa-Carbajal M, Chávez-Gutiérrez M, Romero-Bautista Y, Rangel-Landa S, Moreno-Calles A, Alvarado Ramos L, Smith S, Blancas J, del Val E, Arizmendi MC, Casas A. Use and management of wild fauna by people of the Tehuacán-Cuicatlán Valley and surrounding areas, Mexico. J Ethnobiol Ethnomed. 2020;16(4). https://doi.org/10.1186/s13002-020-0354-8 Zarazúa-Carbajal M, Chávez-Gutiérrez M, Peña-Mondragón JL, Casas A. Ecological Knowledge and Management of Fauna Among the Mexicatl of the Sierra Negra, México: An Interpretive Approach. Front. Ecol. Evol. 2022;10:760805. https://doi.org/10.3389/fevo.2022.760805 Zinyama LM, Matiza T, Campbell DJ. The use of wild foods during periods of food shortage in rural Zimbabwe. Ecol Food Nutr. 1990;24:251–65.

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca Leonor Solís and Alejandro Casas

Abstract

An ethnoecological approach was conducted to identify the role of plants and animals in subsistence of the Cuicatec people. The study was carried out in the indigenous community of San Lorenzo Pápalo, Oaxaca, where we documented and analyzed local people’s knowledge about plants and animals and their local uses and management. This chapter focuses on the ethnobotanical research, while the study of interactions with fauna was published elsewhere. We interviewed 30 persons who are the heads of nearly 20% of the households of the community. All of them are Cuicatec speakers, and some are bilingual (Cuicatec and Spanish). An inventory of plants, their uses, frequency, and quantities extracted and consumed to satisfy different needs (mainly food, medicine, firewood, construction, and ornamentation) was documented, as well as the species preferred for those uses; then, based on the local nomenclature, we analyzed the Cuicatec folk classification of plants which was corroborated in the field. We explored the local perception of the territory and landscape, identifying 12 environmental units classified based on vegetation types and different anthropogenic areas. In those units we carried out ecological vegetation sampling, analyzing the distribution and abundance of useful plants as main indicators of their spatial availability, the species richness and diversity, and the relative ecological importance index of each species, according to their density, frequency and biomass in the sampling units. In addition, we documented the local knowledge about seasonal availability of useful products. With this information we examined possible risks over some species and potentialities of using different plant resources. A total of 520 plant species were recorded, 367 having one or more uses, 176 are fodder, 84 are edible plants, 73 medicines, and 47 are appreciated as ornamental plants, among other L. Solís · A. Casas (*) Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected]; [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_55

517

518

L. Solís and A. Casas

use types. We determined the human cultural importance of plant species through free listing techniques and then evaluated their extraction rates. Nearly 98.4% of plant species showed a restricted distribution through different vegetation types. Every type of vegetation provides differential diversity and products biomass. For instance, tropical deciduous forests and riparian vegetation are diverse, supplying different types of food, medicinal, ornamental plants, and fodder, whereas the coniferous and oak forests are main sources of fuelwood and materials for construction. It is remarkable that anthropogenic areas like crop fields, fallow agricultural areas, homegardens, and even the secondary vegetation are outstanding sources of some of the most important resources like edible and medicinal plants. In general, extraction rates of plant resources are low in relation to their spatial availability, except for the most important fuelwood species (Quercus conzatii and Quercus magnolifolia, among others), which are extracted at high rates and can be identified as species in risk that deserve regulations and particular studies to recommend sustainable forms of extraction. Biocultural, ecological, and ethnobiological studies may substantially contribute social and ecological information for constructing regulation systems by local decision makers.

Introduction It is widely recognized the important role of non-timber forest products in traditional people’s life (Casas et al. 1994; Cunningham 2001; Tuxill and Nabhan 2001; Blancas et al. 2013; Rangel-Landa et al. 2016; Solís and Casas 2019; ZarazúaCarbajal et al. 2020), since rural households commonly practice a subsistence pattern based on a broad spectrum of activities and resources like agriculture, livestock, hunting, and gathering of forest products (Toledo 1990; Cavendish 2001; Casas et al. 1994; Lotero et al. 2022). In Mexico, forest products are mostly goods of common access that provide multiple satisfiers to people’s life (Epstein et al. 2021). These are collected in wild ecosystems, but some of them are also silvicultural managed and/or cultivated in agricultural plots, homegardens, and secondary vegetation (Blancas et al. 2010; Casas et al. 1996, 1997, 2017; Ford and Nigh 2015; Clement et al. 2021). Gathering and management practices are expressions of ancient interactions between humans and plants, which have generated knowledge and techniques throughout time (Cavendish 2001, Toledo and Barrera Bassols 2008; Casas et al. 2016). Traditional gathering is generally a low-impact practice not causing drastic changes on ecosystems. But in some cases, the products extraction overpasses certain thresholds determining significant changes and drastic degradation of systems (Torres-García et al. 2015, 2020). Such processes are commonly associated to economic and sociocultural changes related to the transitions to market economy, which abruptly affect traditional rural life (Salisbury 1970; Casas et al. 1994; Gómez-Baggethun et al. 2010). Identifying such thresholds and causes of pressures on ecosystems is a main challenge of both academic and management sectors (Torres-García et al. 2013). Even more, the effects of disturbance over population of a species commonly have consequences on populations of other species (ValienteBanuet et al. 2015). Therefore, sustainable perspectives of using ecosystems should

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

519

consider how use and management affect whole systems, not only particular components. But in such context, the diversified patterns of using resources and ecosystems, a common traditional practice among indigenous rural communities, provide good principles that appear to be effective to buffer the degradation of both ecosystems and their components. We documented in this chapter the subsistence patterns of the Cuicatec people in a community of the Tehuacán-Cuicatlán Valley to analyze the role of plant species on their life, how the traditional use and management processes involve the diversity of components and ecosystems, the effectiveness or not of their management for sustainability, and how these management patterns are affected by external factors. The Tehuacán-Cuicatlán Valley is one of the main reservoirs of biodiversity of the arid and semiarid environments of Mexico (Dávila et al. 2002), harboring more than 3,000 species of vascular plants, distributed in 29 vegetation types (Valiente-Banuet et al. 2009). Casas et al. (2001, 2017), Lira et al. (2009), and Blancas et al. (2010) documented in the region more than 2,000 species of plants used by local peoples, which makes the Tehuacan-Cuicatlán Valley the richest ethnobotanical inventoried region in Mexico. The Nahua, Popoloca, Chocho, Mazatec, Chinantec, Mixtec, Ixcatec, and Cuicatec live in this region, being nearly 30% of the total population. The Ixcatec and Cuicatec are human cultures exclusive to the region (Casas et al. 2001). All of them have deep ethnobotanical knowledge resulting from a cultural history of more than 10,000 years old (MacNeish 1967, 1992). San Lorenzo Pápalo is a Cuicatec rural community with a history of nearly 750 years (Doesburg 2001a) and clear features of a subsistence economy based on agriculture of maize, beans, and squashes, free raising of livestock (the last five centuries), and multiple use of forest resources. Such long history allows supposing that the traditional practices of using local resources have strong bases of sustainability. This study explores this supposition by documenting the Cuicatec knowledge of plant resources and their role in people’s subsistence. Previous ecological studies in the region showed high α and β diversity among plant communities (Osorio et al. 1996; Dávila et al. 2002; Valiente-Banuet et al. 2009); we therefore expected that the distribution of most plant resources would be rather restricted, a fact that would enhance the ecological complementarity of resources provision. We looked for analyzing sustainable management strategies by examining the general landscape management rather than sustainable harvest of particular resources. Such an approach should be based on information on distribution, abundance, richness, and diversity of plant resources in the different types of forests and comparing such information with their use rates. This would allow us to identify the risks for the permanence of some species under current management patterns, as well as to analyze the potential use of other resources and the opportunities to manage species richness under principles of ecological complementarity. We therefore emphasized the importance of joining ethnobotanical and community ecology approaches to understand the bases of local sustainable management. According to previous information: (1) San Lorenzo Pápalo has been using a broad spectrum of forest resources to complement agriculture probably for thousands of years, and livestock for five centuries (Solís 2006; Solís and Casas 2019). However, ethnobotanical studies in the region have found that, more commonly, a

520

L. Solís and A. Casas

reduced number of species is the most used to satisfy the households’ needs (Casas et al. 2008, Blancas et al. 2010; Rangel-Landa et al. 2016). We consequently expected to find this pattern in the community studied and, therefore, high pressures on the group of more used species, particularly those that are scarce or with narrow distribution. (2) The diversity of resources allows buffering pressures on the most used plant species, and this pattern supports the premise that the multiple use of resources (Toledo et al. 1976) is the base of the rationality of sustainable use patterns. For the contrary, excessive practices directed to few resources would be less sustainable. (3) The high β diversity characterizing plant communities of the region makes feasible to expect high specificity in the availability of plant resources and, therefore, the complementarity of environmental units strongly supports sustainability of plant resources management at landscape level. This chapter aims to document the Cuicatec knowledge of plants, their nomenclature and classification, and use and management by people of San Lorenzo Pápalo. We looked for identifying the resources with higher cultural importance, analyzing their spatial and temporal availability compared with their use and extraction rates. We particularly directed our effort to: (1) complete an inventory of the plants locally used, their Cuicatec names and classification, their use and management forms; (2) determine the cultural importance of plant species to satisfy different human needs, compared with those products obtained from agriculture and livestock raising, and the rates of use and extraction from the forest and other anthropogenic areas; (3) describe the distribution, abundance and temporal availability of plant resources in different ecosystems of the territory of San Lorenzo Pápalo; and (4) analyze risk and potential sustainable use forms of the most important species.

The Cuicatec Ethnographic and ethnobiological studies of the Cuicatec (Starr 1902; Belmar 1902; Elfego 1922; Basauri 1940; De la Cerda 1942; Weitlaner 1969; Bazúa 1982; Hunt 1972; Geist 1997; Solís 2006; Solís and Casas 2019) indicate that these people live in a restricted area of the Sierra of the Pápalos, at north and northeast of the state of Oaxaca, including the villages of Concepción Pápalo, Santos Reyes Pápalo, Santa María Pápalo, San Lorenzo Pápalo, and Teutila (Fig. 1). It is a territory 8,400 km2 extent, characterized by a highly heterogeneous orography with elevations between 600 to 3,300 m, and numerous rivers among them the Grande, Sendo, Cacahuatán, and Chiquito. The territory has three well-differentiated climates, one cold, humid, or dry, in the highlands of the Sierras of Teutila and Pápalos; another temperate in the area between the Pápalos and the Santo Domingo River, with pine-oak forests in higher altitudes and tropical dry forest in the lowlands, and a third warm, humid, or dry, in Cuicatlán and Quiotepec, where vegetation is dominated by columnar cacti and tropical wet and dry forests (Valiente-Banuet et al. 2009). The presence of the Nahua people is relatively recent in the region, but the older native people, descendant from prehistoric human groups of the Tehuacán-Cuicatlán

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

521

Fig. 1 Study area. (a) Location of the community of San Lorenzo Pápalo, Oaxaca in the southeastern area of the Tehuacán-Cuicatlán Valley, central Mexico, (b) a general view of the village

Valley, belong to two linguistic families of the Otomanguean group: the Popolocana (from which the Popoloca, Chocho, Ixcatec, and Mazateco derived), and the Mixtecana (from which the Triqui, Mixtec, and Cuicatec derived) (De Ávila 2004, 2010). According to Kauffman (1990), the Proto-Otomanguean language divided into the two main branches approximately 6,000 years ago. The family Mixtecana, according to De Ávila (2004, 2010) in turn divided into the Proto-Triqui of the subfamily Mixteco-Cuicateca 3,700 years ago, and the Proto-Cuicatec became differentiated from the Mixtec 2,500 years ago. Eberhard et al. (2022) distinguish two Cuicatec languages, one from Tepeuxila with approximately 8,600 speakers, including San Lorenzo Pápalo, and the other from Teutila with nearly 3,000 speakers, and a total of about 13,000 Cuicatec speakers. During the Classic period (2,100 to 700 years ago) settlements of the region became villages and their ceremonial centers were established in the tops of hills (Byers 1967), when the population of the Tehuacán Valley was approximately 20,000 to 30,000 people (Doesburg 2001a). By that time the Cuicatlán region was under the power of the great cities of Monte Albán and Huajuapan, while southern Puebla, including areas of the Tehuacán Valley, was dominated by Teotihuacán and Cholula (Doesburg 2001a). The Classic period ended with the partial abandonment of Monte Albán and Teotihuacán in the eighth century. Then, the Postclassic period lasted until the arrival of the Spaniards in 1519 (Byers 1967). By the end of the Classic period (about the year 1000 AC), all irrigable land of the area was already in use. Remains of irrigation channels show the intensive agriculture practiced in the area. During the Postclassic, the Cañada of Cuicatlán region had a high population growth and intensive use of land and water, and the chiefdoms became small states. About 1200 to 1520 AC, numerous villages settled on small hills in the Sierra, which persisted until the Spanish conquest (Doesburg 2001a). The main Cuicatec manors were Cuicatlán, Alpitazagua, and Quiotepec (Geist 1997).

522

L. Solís and A. Casas

The Cuicatec were independent until 1460, when they became dominated by the Aztec (Doesburg 2001b). The Geographic Relations of the sixteenth century refer to that these areas paid tribute to the Mexica (Geist 1997), while in the Mendocino Codex, Cuicatlán appears as tributary to the Aztec (Geist 1997). Hunt (1972) described the Cuicatec societies of the fifteenth and sixteenth centuries, demonstrating that the geographic distribution of the villages was strongly determined by the presence of water bodies, small streams tributaries of the Grande River. According to this author, the Cuicatec villages of the lowlands were small territories highly populated depending on irrigation, while settlements in the Sierra were spread, low-populated, depending on subsistence agriculture mostly seasonal and small irrigated areas (Hunt 1972; Doesburg 2001a), a pattern that is similar to what can be seen at present in San Lorenzo Pápalo. According to Doesburg (2001a), after the conquest, the indigenous chiefdoms were the base of the Spanish administration and, since the sixteenth century, San Lorenzo Pápalo appears as a village subject to the Spanish Crown. During the Colonial period and until the nineteenth century, the ancient Cuicatec chiefdoms maintained certain legitimacy owning lands, based on documents of the colonial period (Doesburg 2001a). But these lands were then opened to the market and the owners forced to sell them to Mexican and stranger great investors, causing the disintegration of the old Cuicatec chiefdoms about 1870, when vast extensions of land were used for establishing estates, mines, and coffee plantations (Doesburg 2001a). During the twentieth century several processes intensified the integration to the Mexican schemes of development. It is pertinent to mention two facts, the violent, discriminatory assimilation to educational programs directed to eliminate indigenous languages, and the establishing of a net of roads connecting the villages with the railways and highways in the 1960s. Doesburg (2001a) described that the first fact, by the mid- twentieth century, limited the teaching of Cuicatec to the children in schools and homes, while Geist (1997) referred to that the second fact favored the connection of the region with companies and markets that took advantage of the integration and changed the economic, social, and cultural relations with the country (Geist 1997). Bazúa (1982) documented a third important fact; by the 1970s the exploitation of forests was a dynamic economic activity in the region, but it was concessioned to the company “Papelera Tuxtepec S. A.” The incomes of local people derived from this concession were very low, nearly 70% of them were destined to communitarian constructions, and 30% to households’ incomes. People decided interrupting these concessions by the 1990s.

The Community Studied: San Lorenzo Pa´palo San Lorenzo Pápalo is at the northeast of the state of Oaxaca, in the Sierra of the Pápalos, at an elevation of 1,800 m (Fig. 1). The territory of the community is 3,900 ha, belonging to the municipality of Concepción Pápalo. It is a mountainous territory with elevations between 1,500 and 3,000 m, with heterogeneous

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

523

physiography and diversity of climates and vegetation types. Climate is predominantly temperate, but part of the territory are lowlands with tropical dry forest, while the highlands are colder than the area where the village is settled. The main rivers are the Grande and Sendo, tributaries of the Papaloapan River. The main vegetation types are a variety of pine-oak forests, Alnus forests, tropical deciduous and riparian vegetation. The Cuicatec classify their territory into “warm land” (yo ino) and “cold land” (ji quió), which generally correspond to tropical dry and temperate forests, respectively. The natural forests are called the “monte” (bo’cheno in Cuicatec), a term used to classify the vegetation types. For instance, the Alnus forest is called “cheno ya’a ni,” pine forest “cheno ya’a ca,” and oak forest “cheno ya’a ja’a.” People recognize geomorphological units like hills “d’u tu,” rocky areas “y’a ba,” hills “ti cu,” and mountains “ti clun.” The piedmont is called “coó jiquió,” and valleys “y’u du,” the rivers “jicu,” and gullies “ya’a.” People from San Lorenzo Pápalo migrate, our survey allowed registering that nearly 50% of people interviewed had at least one relative working outside, mostly (61.11%) in Mexico City, 27.7% in the city of Puebla, and 11.1% in the city of Oaxaca. Approximately 73.3% of migrants send money to their families, but it happens irregularly. Therefore, the main economic activity continues being agriculture, with products mainly destined to consumption by households and partly to interchange. The main production system is the milpa, called ñango’o, which is classified into different types using the prefix dat. The Cuicatec name the irrigated milpa as dat ió, the seasonal milpa as dat cubi and the seasonal milpa in the cold highlands as dat iquió. The main crops are maize, beans, and squashes, mainly destined to the households’ subsistence. For land tilling they use plough, or a stick called coa in Nahuatl and ya nda’a in Cuicatec, and most households apply chemical fertilizers. They recognize two general varieties of maize, one for warm land and the other for cold land (nin jiquió and nin yo ino, respectively), each general variety with white, yellow, or pinto more particular varieties. Local people use to cultivate “mosquito” beans (Phaseolus vulgaris) in the irrigated area, the “milpa beans” in the seasonal agricultural area and the “mayeso” beans (Phaseolus coccineus) in the cold land. The squashes cultivated are varieties from the species chompo (Cucurbita mixta), támala (Cucurbita moschata), chilacayota (Cucurbita ficifolia), and nahuayota (Cucurbita pepo). The commercial crops are granada (Passiflora ligularis), avocado (Persea americana), chirimoya (Annona cherimola), peach (Prunus persica), and chile canario (Capsicum pubescens).

Research Methods Subsistence patterns. We conducted semi-structured, qualitative interviews, and questionnaires for a quantitative survey, to 20% of households (30 in total). We asked about cultivation techniques, amount of annual production and the relative importance in their economy. Also, we explored the role of homegardens and domestic animals in subsistence.

524

L. Solís and A. Casas

Inventory of useful plants. We collected and herborized plant specimens monthly throughout one year, obtaining ethnobotanical information showing the specimens to the local authorities of the Agencia Municipal and people of the community, who provided information about names, forms of use, and preparation. The information was systematized in the database “Ethnoflora of the Tehuacan Valley” at the CIEco, UNAM, designed following the format of the “Banco de Información Etnobotánica sobre Plantas Mexicanas” (BADEPLAM), of the Botanical Garden at UNAM, including geographic and ecological information, elevation, vegetation type, life form, taxonomic, and ethnobotanical information. The specimens collected were deposited at the National Herbarium MEXU. Cuicatec nomenclature and classification. Our study documented the local nomenclature and classification of plants, as an approach to the local knowledge and perception of nature (Berlin 1992). We based our methods on generating lists of Cuicatec names, then exploring classification patterns, as suggested by De Ávila (2004). We firstly obtained the names of the collected specimens, then we analyzed nomenclatural patterns, which were finally corroborated in the field. We complemented the nomenclatural information obtained in the field with a dictionary Cuicatec-Spanish elaborated by Anderson and Concepción (1983), the names obtained from this source were also corroborated and, in some cases, corrected in the field. The information was compared with nomenclatural records documented among the Mixtec by Casas et al. (1994) and De Ávila (2004, 2010). Cultural importance of species. We evaluated the cultural importance, or the value of plant species based on their role in the Cuicatec culture (Turner 1988; Stoffle et al. 1999; Pieroni 2001). Following Turner (1988), the more frequently and intensively used is a plant the higher is its cultural importance. Such importance may vary according to the quality of plant products, their use intensity, exclusiveness, among other attributes that in turn may change through time. Several approaches have been developed to identify the cultural importance of plants. Turner (1988), Stoffle et al. (1999), Parra et al. (2021), and other research groups have proposed indices to evaluate the cultural importance of species or varieties. In this study we followed the approach based of free listing (Frei et al. 1998; Turner 1988). Based on 30 free lists, we identified the species with the highest cultural importance of the following use categories: (1) medicinal for gastrointestinal illnesses, (2) medicinal for respiratory ailments, (3) medicinal for muscular pain, (4) medicinal for cultural ailing like “limpias” (cleaning cure), “susto and espanto” (scare), and “mal de ojo” (evil eye), (5) fuelwood, (6) edible as green vegetables or “quelites,” (7) edible fruits, (8) edible seeds, (9) edible roots, tubers, and bulbs, (10) edibles flowers, (11) ornamental, (12) construction, (13) fodder, (14) ritual use. Amounts of plant products extracted. Through semi-structured interviews and a survey, we evaluated the amounts of plant products extracted from the most culturally important species identified. We obtained information from different measure units locally used, and then we transformed these units in kg or number of plant individuals; whenever possible, we directly weighted the products, we also documented the seasonal availability of products and frequency of extraction. For estimating the extraction and use of fuelwood, we weighted the amount consumed

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

525

per day, we analyzed the composition of samples of fuelwood available in the homes studied, identifying most of the species, based on local names and features of wood and cortex, as well as the proportion of each species in the samples analyzed. Distribution and abundance of useful plant products. Vegetation sampling was carried out in different environmental units to estimate density, frequency, and biomass of the species composing each unit type. We selected areas with different vegetation and anthropogenic unit types in: (1) tropical deciduous forest, (2) riparian vegetation, (3) Alnus firmifolia forest, (4) Pinus michoacana–Quercus conzatii forest, (5) Pinus lawsonii–Quercus crassifolia forest, (6) Quercus rugosa forest, (7) Quercus laurina forest, (8) Quercus magnolifolia forest, (9) granada (Passiflora ligularis) gardens, (10), homegardens, (11) irrigated milpas, and (12) seasonal cultivation milpas. In each unit we sampled at least two 500 m2 squares, 50 m long per 10 m wide, divided in 10 m2 subunits. In these squares we recorded all individuals of each species of trees and shrubs, thus estimating their density. Herbs were sampled through five 1 m2 squares randomly placed within each 10 m2 square estimating the percentage of area covered by each species. The frequency was calculated as the percentage of 10 m2 squares in which a species was present. The biomass was calculated by measuring height, two perpendicular diameters and, in trees, the breast height diameter of trunks, then we used these measures to calculate the volume of individual plants approaching their forms to ellipsoids (shrubs) and to inverted cones (trees), and finally estimating the total volume per species in the sampling area. The relative value of ecological importance (REI) of each species was estimated, following Valiente-Banuet et al. (2000), as: REI ¼ Frequency (%) X Density (ind/m2) X Biomass (m3) Frequency values were also used as estimators of distribution. Species appearing less than 33% were considered to have restricted distribution, those occurring in 45–75% sampling units were considered of intermediate distribution, and those occurring in more than 75% sampling units were considered to have wide distribution.

Results Agricultural systems. Milpa is the main agricultural system; it is the traditional polyculture of maize, beans, and squashes. The irrigated systems (dat quió), which are close to the Sendo River, benefited with a system of channels or apantles constructed in pre-Hispanic times. The seasonal systems are in the lowlands areas (data cubi) and in the highlands (dat iquió). According to the survey, plots of irrigated milpas may vary from 0.5 to 3 ha, on average 1.06 (SD 0.62) ha per household. Most households (86%) sow in March and April, and harvest in July and September. Others (14%) sow in February. All households use plough for tilling the ground, except in rocky or pronounced slope terrains, where they use the stick called coa. All households sow the native varieties (white, yellow, pinto, and black) of the creole maize of “de tierra caliente” (“nin jiquió”). Approximately one-half of producers combine white and yellow varieties in a plot, and the rest mix all the varieties recorded. Only 3% of people interviewed

526

L. Solís and A. Casas

said to use herbicides, most people weed their plots manually once or twice per production season. Most people (90%) sow the “mosquito” beans (Phaseolus vulgaris) and 84% some of the squash species referred to in the irrigated milpa. Presence and maintenance of quelites in milpas were reported by 20% of people surveyed, mostly mentioning the yerbamora (Solanum nigrescens) and quintonil (Amaranthus hybridus). Maize production per household in this system was 743.4 kg/ha (EE  160.83). The seasonal milpas’ plots are on average 1.8 ha (DE  1.42) per household. All households use creole varieties of maize. The average production is 639 kg/ha (EE  122.40). Sowing is carried out with the first rains in May, June, or July, while the harvest occurs in November, December, or January. Milpa is also a polyculture with the maize varieties, the “milpa” beans or the “mayeso” beans (P. coccineous). The squash chilacayota (Cucurbita ficifolia) is the most used but it is commonly combined with the other species. The most common quelites are quintonil and yerbamora, and this pattern is due to people’s preference and the consequent procurement of these plants (Fig. 2). The main productive systems complementing the milpa in local subsistence of households are the fruit production system called “huertas,” which generate monetary incomes and may include the ganada (Passiflora ligularis) garden, the chirimoya (Annona cherimola) avocado (Persea americana), peach (Prunus persica), and chile canario (Capsicum pubescens), cultivated in homegardens. The survey allowed estimating that the average income from commercialization of fruit is $3,789 ( DE 1149) pesos per year per household.

Fig. 2 Frequency of presence of quelites in the milpa system sampled. Quintonil ¼ Amaranthus hybridus, yerbamora ¼ Solanum nigrescens, mostaza ¼ Brassica campestris, husk tomato ¼ Physalis philadelphica (weedy green tomato), red tomato ¼ Solanum lycopersicum (weedy red tomato)

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

527

Homegardens complement the subsistence activities of households, mainly managed by women. These systems have a high diversity of plants and a complex structure with herbaceous, shrubby and arboreal strata, including native and introduced species mainly edible and some medicinal and ornamental. The plot size is on average 1,167 m2; the dominant species being orange, lemon, banana, chirimoya, prickly pears, coffee, avocado, peaches, chile, granada, guava, epazote, and camomile, all products destined to households’ use (Fig. 3).

Fig. 3 The main production systems for local people’s subsistence in San Lorenzo Pápalo (a) the milpa, (b) homegardens, and (c) the practice of tilling the land in San Lorenzo landscape. (Photos: Leonor Solís)

528

L. Solís and A. Casas

Production and consumption of agricultural products. According to our surveys, maize production per household is on average 959.31 kg/ha, 95% is directly consumed and the rest is commercialized. On average, each household consumes 22  3.38 kg of maize per week, and 1182.55  176.12 per year. Each household has to buy on average 106.3 kg of maize. For producing maize, people must acquire fertilizers. Therefore, the production is deficient, and households need money to produce it. Production of beans is on average 63.92 kg ( DE 63.25) per household per year. All this product is directly consumed by households, but these have to buy on average 35.68 kg per year. Households commonly produce squashes and mainly consume their seeds, while pulp is mainly used to feed pigs. On average they produce 29 squashes per household, all production directly consumed. Diet. Food of the Cuicatec households is mainly composed by maize tortillas and beans. These elements are commonly complemented with pasta soup or rice. Stoves of the main meal at midday may vary including seasonal forest products (quelites, guajes, plants gathered, animals hunted, or mushrooms collected in forests; Fig. 4). Wild food is particularly abundant during the first months of the rainy season, but some products are available throughout the year. Consumption of chicken meat has increased, with the establishment of the communitarian store; now, households consume this meat once per week on average. Consumption of pork or cattle meat is occasional and mainly associated to parties, twice per year on average; according to the survey 8.06 kg  26.06 of meat per household per year. Other products that are common are cheese, pasta, bread and soda, as well as agricultural products that are not produced in the community (mainly rice, tomato, onion, garlic, jalapeño chiles, and potatoes). Table 1 shows a summary of the amounts of these products consumed per household per week. Cuicatec nomenclature and classification of plants. The Cuicatec nomenclature of plants is generally binomial, with a generic term commonly grouping life forms, characteristics of the environments where they live, among other aspects, and a specific term alluding some characteristics (phenological, morphological, ecological, or other aspects to their use). The classification is based on life forms (mainly herbs, trees, and vines, aspects of use (for instance medicinal plants, quelites, or species producing flowers) and/or morphological characteristics or related to their habitat. In addition, some names are accompanied with varietal terms associated to colors, particular forms, and other attributes. All trees are called ño’o, which means stick, and ya’an, which refers to trees and shrubs (Table 2), but it is unclear why in some cases it is used a term and in other cases the other. Herbaceous plants are called yata. Shrubs may be classified as trees or herbs. Fruit producing trees are called ño’o (stick) ndut (seed). For instance, the hawthorn is named ño’o dut iñu, while its fruit is called ndut iñu, peach trees are called ño’o n’deyi while their fruit is nd’e yi. The guajes (Leucaena spp.) and legume trees and shrubs are grouped with the term nin (Tables 2 and 11). For figs they use the term n’ga, for oaks ya ‘a nde ‘e, ya’a or ye (Table 2). Pines are grouped with the term ya’a ca; for instance, several pines are called ya’a ca cuó jiquió, Pinus michoacana is ya’ac toó, P. teocote, P lawsonii, P. pringlei, and P. herrerae are

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

529

Fig. 4 Some groups of wild and weedy food obtained from forests and anthropogenic areas. (a) Mushrooms collected from temperate forests, (b) Leucaena or guaje pods, (c) larvae of “cuetlas,” a nocturnal butterfly, collected from trunks of Heliocarpus appendiculatus in the tropical dry forest area, and (d) some agricultural products consumed as greens. (Photos: Leonor Solís)

530 Table 1 Average weekly consumption of food obtained in markets per household

L. Solís and A. Casas

Food Rice Tomato Onion Garlic Chile Pasta soup Eggs Bread Soda Chicken meat Pig meat Cheese

Weekly consumption/household 1.0  (0.55) kg 1.8  (0.87) kg 0.9  (0.60) kg 2.0  (2.20) units 0.3  (0.25) kg 2.7  (1.49) bags 1.27  (1.20) kg 7.0  (7.00) pieces 4.5  (8.50) l 1.38  (1.00) kg 0.12  (0.20) kg 1.6  (1.33) units

called ya’ac yudi. Trees of the genus Cupressus are grouped with the term ya’a cú, while grasses are grouped with the term yuni. These among other examples. Columnar cacti are named with the generic term n’un, followed by a term distinguishing the species. The spherical cacti are called i ndin yava and Opuntia plants are grouped with the term ditu (Table 3). The herbaceous plants are grouped with the term yata, meaning leaf, herb, or plant. The term is used as yat for naming different herbs (Table 4). A group of herbs is called nanda, the Cuicatec term for flower, and herbs with beautiful flowers are named through the term nanda or nan (Table 4). The orchids belong to this group and are grouped with the term nanda ’tca. Vines are grouped with the term chivi, and some of them are considered herbs and called yat chivi. Some examples can be seen in Table 4. Plants producing tubers and bulbs are grouped with the term m’in. For instance, they recognize the Dioscorea spp. bitter camote (appreciated for food) distinguishing one from the cold land (m’in go’ yó) and other from the warm land (min goó). The sweet potato (Ipomoea batatas) is called m’in ya dí and another edible tuber called stinky camote, also edible, is called m’in yatim yacú. According to information from interviews, the term ji is commonly used as a prefix to name herbaceous plants used as food or medicine. For instance, the quelites are grouped with the term ji v, and the medicinal plants ji quiud (Tables 5 and 6, respectively). Ferns are grouped with the term ya cua (Table 7).

Inventory of Useful Plants In this study we identified 520 plant species belonging to 110 botanical families, mainly Asteraceae (83 species), Fabaceae (42), Euphorbiaceae (20), Solanaceae (19), Poaceae (13), Cactaceae (13), Lamiaceae (12), Adianthaceae (12), Fagaceae (11), Malvaceae (11), Commelinaceae (10), and Rubiaceae (10). We identified that 367 (70%) species of 87 families that are used with 17 use categories, and 23% of the plant species recorded have more than one use. Most of them (33.84%) are used as

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

531

Table 2 Cuicatec nomenclature of trees and shrubs with the prefix ño’o y ya’an Species Rapanea jurgensenii Calliandria eryophylla Juliana adstringens Gyrocarpus mocinoii Myrica mexicana* Barkleyanthus salicifolius* Bursera simaruba Bursera bipinnata Arctostaphylos pungens* FRUIT TREES Crataegus mexicana (tejocote) Annona reticulata (anona) Musa paradisiaca (plátano) Bunchosia palmeri (nanche) GUAJES Leucaena esculenta (Guaje colorado) Leucaena esculenta (Guaje zopilote) Desmanthus virgatus (Guaje de ratón) Especie no identificada (unidentified species) OAKS Quercus rugosa Quercus obtusata Quercus glaucoides Quercus peduncularis Quercus magnolifolia Quercus conzatii Quercus salicifolia Quercus crassipes Quercus crassifolia Quercus acutifolia

Life form Tree Tree Tree Tree Shrub Shrub Tree Tree Shrub

Cuicatec life form ño’o ño’o ño’o ya’an ya’an ya’an ya’an ya’an ya’an

Cuicatec generic name tu’u ya yo chama de go guo gu dut

Cuicatec specific name de gada ino cuá cheno ido ó ú inn’i

Tree Tree Tree Tree

ño’o ño’o ño’o ño’o

ndut ndut ya’a nun

iñu mé tiaca güe’e

Tree

ño’o

nin

guo’ó

Tree

ño’o

nin

jaca

Shrub

ño’o

nin

du’o

Shrub

ño’o

nin

güi

Tree Tree Tree Tree Tree Tree Tree Tree Tree Tree

ño’o ño’o ño’o ño’o ño’o ño’o ño’o ño’o ño’o ño’o

ya’a nde ya’a nde ya’a nde ya’a ja ya’a ya’a ye ye i gño’o

cuá cua tu’u tu’u jaba jaá cú co yoo cuó

fodder, 16.15% as food, 14.03% as medicine, and 9.03% as ornamental. Table 8 summarizes the general information of this inventory, while Table 9 shows information on the types of use provided by plant families identified, and Tables 10 and 11 summarize information on the useful plants provided by the different environmental units of the territory. From the 84 medicinal plant species recorded, people mentioned 36 used for gastrointestinal illnesses, the most important mentioned were Matricaria recutita,

532

L. Solís and A. Casas

Table 3 Cuicatec nomenclature of cacti Species Pachycereus grandis Pilosocereus chrysacanthus Eschontria chiotilla Stenocereus pruinosus Stenocereus stellatus OPUNTIA Opuntia tomentosa (Nopal Amarillo) Nopalea auberi (Nopal de perrito) Opuntia ficus-indica (Nopal de castilla)

Cuicatec generic name n’un n’un n’un n’un n’un

Cuicatec specific name no’o chicú ya’a ti na’a cuo’o

ditu ditu ditu

coó ya’a na da’a

Table 4 Cuicatec nomenclature of herbaceous plants with showy flowers “nanda” Species Eryngeron longipes Perymenium mendezii Geranium sp. Crotalaria rotundifolia Bacopa monieri Mirabilis jalapa

Cuicatec generic name nan nan nan nan nan nan

Cuicatec specific name cuá jiquió cuó yoino dio diyu guó caya tin bocheno

Table 5 Examples of the Cuicatec names of quelites “ji uv” Species Portulaca oleracea (verdolaga) Phytolacca icosandra (lengua de vaca) Solanum nigrescens (yerbamora) Brassica rapa (mostaza) Amaranthus hybridus (quintonil) Chapoquelite

Cuicatec generic name jiv jiv jiv jiv duc jiv jiv

Cuicatec specific name di Tú duv inó du n’e iya do’o co’o

Table 6 Examples of Cuicatec names of medicinal plants “ji quiud” Species Acalypha sp. Oenothera rosea Loeselia caerulea Iresine celosia Plantago australis

Cuicatec life form ji quiud ji quiud ji quiud ji quiud ji quiud

Cuicatec generic name van lun yande du cu

Cuicatec specific name yudi chi bocheno atá cho

Artemisia mexicana, Ruta chalepensis, Foeniculum vulgare, and Mentha viridis. For attending respiratory problems, people mentioned 15 species, the most important were Bouganvillea glabra, Eucalyptus globulus, Allium sativum, Psidium guajava,

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

533

Table 7 Examples of Cuicatec names of ferns “ya cua” Species Adiantum sp. Cheilantes sp. Pallaea sp. Phlebodium aureum Pleopeltis sp. Polypodium polypoidioides Pteridium aquilinium Schizaeocese sp. Table 8 General panorama of useful plant species recorded in San Lorenzo Pápalo

Cuicatec life form Ya cua Yac Ya cua Ya cua Ya cua Yac eno

Cuicatec generic name nuni nuni nuni nuni nuni yaba

Ya cua Ya cua

caya

Use Fodder Edible Medicinal Ornamental Construction Fuel Ceremonial Utensils Toys Shade Beverages Insecticide

Cuicatec specific name caya

jiquió jiquió

Species 176 84 73 47 37 33 6 5 4 3 3 2

Table 9 Number of useful plant species per family. Use categories: 1 ¼ medicinal, 2 ¼ fodder, 3 ¼ edible, 4 ¼ ornamental, 5 ¼ ceremonial, 6 ¼ fuel, 7 ¼ utensils, 8 ¼ construction, 9 ¼ live fences, 10 ¼ shade, 11 ¼ handcraft, 12 ¼ poison, 13 ¼ beverage, 14 ¼ glue, 15 ¼ cosmetic, 16 ¼ toy Family Asteraceae Fabaceae Solanaceae Euphorbiaceae Poaceae Cactaceae Fagaceae Adiantaceae Malvaceae

Species recorded 83 42 19 20 13 13 11 12 11

Useful species 60 38 16 12 12 12 11 7 7

Use type 1,2,3,4,5,8 1,2,3,4,6,7,8,10,16 1,2,3,4 1,2,4,8,12 1,2,3 2,3,4,5 2,6,8 2,4 1,2, 12

534

L. Solís and A. Casas

Table 10 Number of useful plants per vegetation type Vegetation type Tropical dry forest Milpas Riparian vegetation Pinus michoacana–Q. conzatii forest Pinus lawsonii–Q. crassifolia forest Homegardens Secondary vegetation Alnus forest Ruderal areas close to the village Oak forest (Quercus rugosa) Oak forest (Quercus laurina) Granada gardens (Passiflora ligularis) Oak forest (Quercus magnolifolia)

Species number 144 90 90 71 71 66 57 44 41 36 28 16 10

Useful species number 97 70 68 52 42 62 48 19 38 22 16 16 10

Table 11 Type of useful species per vegetation type. Uses: 1 ¼ medicinal, 2 ¼ fodder, 3 ¼ Edible, 4 ¼ Ornamental, 5 ¼ Ceremonial, 6 ¼ Utensils, 7 ¼ Construction, 8 ¼ Shade, 9 ¼ Handcrafts, 10 ¼ insecticide, 11 ¼ Poison, 12 ¼ Beverages Vegetation type Tropical dry forest Riparian vegetation Pinus michoacana– Quercus conzatii forest Pinus lawsonii– Q. crassifolia forest Quercus magnolifolia forest Quercus laurina forest Quercus rugosa forest Alnus forest Secondary vegetation Milpa Granada gardens Homegardens

1 18 16 7

2 49 34 29

3 29 9 3

4 14 8 13

5 1 1 0

Use 6 6 4 11

7 0 3 1

8 9 5 12

9 1 1 0

10 0 0 1

11 1 1 0

12 0 2 0

7

20

3

9

1

9

0

10

0

0

0

0

1

7

0

0

0

4

0

4

0

1

0

0

0 2 3 7 19 1 11

8 7 9 32 47 7 7

2 2 3 6 18 8 41

0 3 1 2 2 1 5

0 0 0 1 1 0 2

6 10 5 6 1 0 0

1 1 0 2 1 0 1

8 11 6 8 1 3 3

0 0 0 0 0 0 1

0 0 0 0 0 1 3

0 0 0 0 0 0 0

0 0 0 0 1 0 1

and Matricaria recutita, while for fever they mentioned Sida acuta, Gnaphallium spp., and Fraxinus purpusii. For alleviating muscular pain people mentioned 17 species, the most important Amphipteringium adstringens, Pinus spp., Equisetum hyemale, Ruta chalepensis, Selaginella lepodophyla. For cultural illnesses people mentioned 12 species, the most important Sambucus mexicana, Schinus mole, Ruta chalepensis, Bursera bipinnata, and Ocimum basilicum.

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

535

Table 12 Quelites with the highest cultural importance in San Lorenzo Pápalo Common name Yerbamora Quintonil Berro Mostaza Chapoquelite Verdolaga Pepicha Papaloquelite Lengua de vaca Col blanca

Species Solanum nigrescens Amaranthus hybridus Roripa nasturtium–officinale Brassica rapa Not determined Portulaca oleracea Porophyllum tagetoides Porophyllum ruderale subsp. macrocephalum Rumex crispus Brassica rapa

Frequency of mention 30 27 17 14 14 13 6 5 4 4

Table 13 Wild edible fruit with the highest cultural values in San Lorenzo Pápalo (species with * are cultivated domesticated species escaped and adapted into wild environments Common name Pitaya Ciruela amarilla Chupandía Mango Tuna Guayaba Anona Zarzamora Tempesquisle

Species Stenocereus pruinosus Spondias mombin Cyrtocarpa procera Mangifera indica* Opuntia sp. Psidium guajava* Annona reticulata Rubus liebmanii Sideroxylon palmeri

Frequency of mention 30 18 16 11 10 10 6 6 6

Table 14 Wild edible seeds with the highest cultural importance in San Lorenzo Pápalo Common name Guaje colorado Guaje zopilote Guaje de agua Guaje verde Guaje blanco Guaje de ratón Pochote Bonete Acazle

Species Leucaena esculenta Leucaena esculenta Leucaena leucocephala var. glabrata Leucaena leucocephala Leucaena leucocephala Desmanthus virgatus Ceiba aesculifolia Jacarita mexicana No identificada

Frequency of mention 7 3 2 2 2 2 1 1 1

From a total of 84 species recorded as edible plants, local people referred to 18 species as the most important quelites (Table 12), 17 species are the most important wild fruit (Table 13), the edible seeds, mainly from Leucaena spp., are indicated in Table 14; whereas the main plants providing subterranean edible parts are of the genera Dioscorea (camote amargo, camote de agua, and camote hediondo)

536

L. Solís and A. Casas

Table 15 Plants used as fuelwood with the highest cultural importance in San Lorenzo Pápalo Common name Encino cucharo Ocote Ya’a jaba Madroño Encino negro Encino hoja ancha Tepeguaje Ya’a já yoó Ilite Chamizo

Species Quercus conzatii Pinus michoacana Quercus magnolifolia Arbutus xalapensis Quercus glaucoides Quercus peduncularis Lysiloma acapulcensis Quercus crassifolia Alnus firmifolia Barkleyanthus salicifolius

Frequency of mention 26 20 12 11 11 6 4 3 3 3

and Ipomoea. The edible flowers most mentioned are those from Pilosocereus chrysacanthus, Agave spp., and Erythrina americana. Among the plants used as fuelwood, people mentioned numerous species, but those most frequently used and mentioned by people are indicated in Table 15, while those most frequently mentioned as used for fabricating tools are Pinus spp. and Lysiloma acapulcensis.

Extraction Rates of Plant Resources Our evaluations were directed to those exceptional important resources mentioned by local people. Among the wild fruit, for instance, the free lists indicated 17 species but the group with 100% of mention was the columnar cacti fruit (Stenocereus pruinosus), the chupandía (Cyrtocarpa procera), Spondias mombin, and Diospyros digyna. We will focus our attention on these species in this chapter. For people of San Lorenzo Pápalo, the populations of columnar cacti are relatively far from the village, approximately 2 h by walking to arrive to the area of tropical dry forest. But gathering of this fruit is carried out by 90% of people interviewed, from April to early June. Stenocereus pruinosus is called n’un na’a; households go to collect these fruits once per week during the production season, seven to eight occasions in a year. On average, they extract 7.36 kg of fruits per household per week, a total of 65.5 kg per household per year (nearly 315 fruits). Since 90% of the families go to collect these fruits, we estimate that the whole village extract annually 10,142.49 kg or 48,762 fruits. Other columnar cacti are also collected (Stenocereus stellatus, Escontria chiotilla, and Pilosocereus chrysacanthus) but these events are occasional. In the case of Pilosocereus chrysacanthus, only 3% of people interviewed collect its fruit, but 80% collect its edible flowers. Prickly pears (Opuntia spp.) and pitahaya Hylocereus undatus are also collected, but they cultivate O. ficus-indica in homegardens and therefore they do not have to

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

537

go to collect other species in the wild. There is a wild Hylocereus species growing in oak forests, but they become established and tolerated in homegardens, where people mostly extract its fruits. The Sapotaceae tree called tempesquisle (Sideroxylon palmeri) was not especially mentioned in the free lists (only 20% of people interviewed mentioned it). However, it is a wild species actively cultivated in homegardens, with a high seasonal consumption (cooked or as fresh fruit) and commercialization in April. The survey revealed that 76% of local people consume tempesquisle fruits in March and April. They said to collect fruits once per week during the production season, and, on average, each household consumes 4.96 kg annually, and the whole community 854.26 kg. Fruits of chupandía (Cyrtocarpa procera) are collected in tropical dry forest areas during September and October. These are collected by 36.6% of the local population, once or twice per year, on average 1.2 kg of fruits per household, or 217.8 kg per year by the whole community. The ciruelas (Spondias mombin) are collected in April and May in areas with tropical dry forest by 20% of households, which consume on average 0.58 kg of these fruit per household per year, and the whole village consume 100.33 kg per year. The fruits of tempesquisle amarillo (Bumelia laetevirens) is collected by 10% of people interviewed in May. Households consume approximately 0.18 kg per year, on average, and the whole village 31.53 kg per year. Fruits of the vine “chivi yucu” (Gonolobus grandiflorus) are consumed after being roasted. It is collected in the tropical dry forest in October by 10% of the households. Other native fruits collected and consumed are guava (Psidium guajava), custard apple (Annona reticulata), hawthorn (Crataegus mexicana), and the black cherry or capulín de monte (Prunus serotina subsp. capuli), but because these species are cultivated in homegardens, their gathering in the wild is rather occasional. Seeds from wild plants are obtained mainly from the “guajes”, basically two species of the genus Leucaena. These seeds are consumed raw or prepared in several stews and sauces. The green or white guaje (Leucaena leucocephala) is consumed by 60% of local people, in March–April and September–October; nearly 20% said to buy these seeds in Tehuacán or Quiotepec the rest collect pods in the tropical dry forest and homegardens. Households collect guaje pods three times per year, nearly 0.25 to 0.5 kg, and 0.66 kg of pods per year, and 113.95 kg by the whole community. The red guaje (Leucaena esculenta) is consumed by 43% of households in October and November. It is collected in tropical dry forest areas three times per year. Each household collects 1.3 kg of pods per year, nearly 216.43 kg per year in the whole community. The “camote amargo” (Dioscorea sp.) is consumed by 76.6% of local people in March and April. It is collected in the tropical dry forest, on average 4.6 kg per household per year 804.1 kg per year in the village. Among the main edible flowers, the local people mentioned the pipe or colorín flowers (Erythrina americana), the agave or maguey flower buds, also called “cacayas” (Agave spp.), and the flowers of the “nanabuela” Pilosocereus chrysacanthus.

538

L. Solís and A. Casas

Flowers of Erythrina americana are found in all homegardens and granada gardens since their trunks are support of the vines of Passiflora ligularis and chayote (Sechium edule), and provide shade to coffee. These flowers are consumed by 90% of households between October and December when they collect once or twice per week, on average 0.5 kg. According to the survey, each household collects on average 4.2 kg of flowers per year, and the whole community 715.95 kg of flowers per year. The “nanabuela” flowers (Pilosocereus chrysacanthus) or “nun chicú” in Cuicatec are very much appreciated and consumed by 80% of households. These flowers are collected in areas with tropical dry forest from March to May. People collect these flowers twice per year and trap edible lizards (“yati” in Cuicatec), to prepare a soup with lizards and flowers. On average, each household consume 1.6 kg of nanabuela flowers per year, and the whole community 288.81 kg of flowers per year. Flower buds of maguey or agave maguey (Agave peacockii and A. potatorum) are called “cacayas” and consumed by 70% of the households. Local people collect cacayas in two periods, from March to May (A. peacockii), and from September to November (A. potatorum). The flower buds of Nopalea auberi (or Opuntia auberi) are called “cocoches” and are consumed throughout the year. The latter are cultivated in homegardens, from where people obtain the buds. These are prepared boiled, with tomato, onion and garlic, or in soup. Through the free lists we identified 18 species of plants called quelites. Most of them are available in the rainy season, growing in milpas or fallow land. The rest grow near or inside streams. The yerbamora (Solanum nigrescens) received 100% of mention by people interviewed, while the survey about consumption reported that 93.3% of households consume it. It is collected in milpas, on average 13 days per year; people interviewed said they go to collect this quelite twice per week during the time it is available, at the beginning of the rainy season. Each household consumes, on average 8.2 kg of yerbamora per year. Therefore, the community consumes 772 kg per year. The quintonil (Amaranthus hybridus) is consumed by 93.3% of households, mainly collected in milpas and areas near streams. It is collected when the plant is young and tender, on average 13 times per year, twice or three times during the season when it is available. On average, each household consumes 10.8 kg per year, and 1223.54 kg is consumed by the whole village. The berro (Roripa nasturtium-offincinale) is consumed by 90% of the local households. It is available the whole year, growing in the streams around the village. On average, each household collect 16.96 kg of this quelite per year, which means a total of 2918.26 kg per year is collected by the whole community. The papaloquelite (Porophyllum ruderale subsp. macrocephalum) is consumed by 73.3% of households. It is available in milpas, homegardens, and secondary vegetation during the rainy season. On average, a household collect this plant in small amounts 31 times per year, in total 2.74 kg per household per year, 490.96 kg by the whole community. The pepicha (Porophyllum tagetoides) is consumed by 63.3% of households during the rainy season, in the warm land area. Households collect this plant twice

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

539

per year, on average 0.41 kg, and the whole community consumes approximately 73.39 kg of this plant per year. The verdolaga (Portulaca oleracea) is consumed by 43.3% of households. It is collected in milpas of the warm land area, on average five times per year (1.84 kg) per household, nearly 329.36 kg by the whole community. The mostaza (“mustard”) (Brassica rapa) is consumed by 36.6% of households. It is collected in milpas of the cold land area. On average, each household collect three or four times per year and consume 0.93 kg per year, which means that the whole community consumes approximately 166.47 kg of this plant per year. The chapoquelite is consumed by 30% of households. It is collected in pine-oak forests; it is very much appreciated but because of the distance of the places where it grows, it is collected once per year. On average, each household consumes 0.4 kg per year, approximately 71.6 kg consumed by the whole community. The quelite lengua de vaca (cow tong quelite) (Rumex crispus) is consumed by 26.6% of households. It is collected on average twice per year per household in areas near streams. Each household consumes 0.165 kg per year, which means that the whole community consumes approximately 28.38 kg per year (Table 16). Extraction of medicinal plants. Local people are progressively switching the traditional remedies for commercial medicines, and, therefore, they extract few amounts of medicinal plants occasionally. Approximately 55% of the medicinal plant species used are cultivated in homegardens, 20% are weedy or ruderal plants, and 25% are wild. All this information indicates that the impact associated to this activity is insignificant. Fuelwood extraction. Collecting of fuelwood is highly relevant in life of people from San Lorenzo Pápalo (Fig. 5). Although 10% of households have gas stoves, the provision of gas is irregular and, therefore fuelwood is primordial for all households. In addition, it is pertinent to comment that local people said to prefer fuelwood for cooking since flavor and texture of food is better than that prepared with gas stove. Fuelwood extraction is mainly carried out in oak forest areas, mainly those where Quercus conzatii and Q. magnolifolia, the preferred species, dominate in the vegetation composition. More recently, some trucks have been acquired by people of the village and this condition has favored extracting fuelwood from areas of higher elevations. Nearly 40% of households buy fuelwood to extractors, intensively dedicated to this activity, who have chainsaw and trucks. Fuelwood extractors cut on average 3 trees per charge contracted, mainly extracting Quercus conzatii and Q. magnolifolia; this charge provides, on average, fuelwood for 4.5 months to a household. This means that the 40% of local households that buy fuelwood consume on average 558.48 trees of these species per year. Our survey found that samples of fuelwood were composed nearly 60% by Q. conzatii and 40% of Q. magnolifolia, which indicates that households using this source of fuelwood consume 335.08 trees of Q. conzatii per year and 223.39 of Q. magnolifolia. The remaining 60% of local households collect their fuelwood, which is transported by donkeys and the units of measurement of consumption are the “donkey

540

L. Solís and A. Casas

Table 16 Amounts extracted of the main edible wild and weedy plants in San Lorenzo Pápalo

Plant species Stenocereus pruinosus Roripa nasturtiumofficinale Amaranthus hybridus Sideroxilon palmeri Dioscorea sp. (Camote amargo) Solanum nigrescens Erythrina americana Porophyllum ruderale subsp. macrocephalum Portulaca oleracea Pilosocereus chrysacanthus Cyrtocarpa procera Leucaena esculenta Brassica rapa Leucaena leucocephala Spondias mombin Porophyllum tagetoides Chapoquelite Bumelia laetevirens Rumex crispus

Type of food Fruit Quelite Quelite Fruit Root Quelite Flower Quelite Quelite Flower Fruit Seed Quelite Seed Fruit Quelite Quelite Fruit Quelite

Seasonality April–June Year round

Consumption per household (kg) 65.5 16.96

Annual consumption in the community (kg) 10,142.50 2,918.2

Sept–Nov March– April March– April Sept–Nov Oct–Dec Sept–Nov

10.8 4.96

1,223.5 854.26

4.6

804.1

8.16 4.2 2.74

772 715.9 490.96

Sept–Nov March– May Sept–Oct Oct–Nov Sept–Nov Oct–Nov April–May Sept–Nov Year round April–May Year round

1.84 1.6

329.36 288.81

1.2 1.3 0.93 0.66 0.58 0.41 0.4 0.18 0.16

217.8 216.4 166.47 113.95 100.33 73.93 71.6 31.53 28.38

charges” (the amount of fuelwood that on average a donkey is able to bear). According to our survey, each household extracts approximately 2.64 donkey charges per week. These charges are on average 65.5 kg, which means that the 60% of local households consume 169 kg of fuelwood per week, which means nearly 940.316 t per year. Through measuring the daily consumption of fuelwood, we estimated that each household consumes 6.75 ton of fuelwood per year (Table 17). This figure allows estimating that the annual consumption by the whole community is nearly 1,206.08 ton per year. Based on the analysis of average composition of samples of fuelwood used in the daily life, we found that households mainly consume Q. conzatii, Q. magnolifolia, Q. glaucoides, Pinus spp., Lysiloma acapulcensis, and Q. peduncularis (Fig. 6). These are not the only species used; in fact, the survey recorded the mention of other species like Arbutus xalapensis, Acacia pennatula, and A. farnesiana. Figure 6 shows the panorama of the average composition of fuelwood samples recorded in a day per household.

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

541

Fig. 5 General aspect of extraction and use of fuelwood for subsistence in San Lorenzo Pápalo. (Photos: Leonor Solís)

Distribution and Abundance of the Main Plant Resources Table 18 summarizes the average estimation of density, biomass, and frequency of the main plant species recorded in the vegetation sampling carried out in each vegetation type analyzed. These samplings allowed identifying the distribution and

542

L. Solís and A. Casas

Table 17 Average weight and volume of fuelwood consumed per household and per the whole community per year Species Encino cucharo (Quercus conzatii) Encino blanco (Q. magnolifolia) Encino negro (Q. glaucoides) Encino (Q.peduncularis) Pino (Pinus spp.) Tepeguaje (Lysiloma acapulcensis) Total

Tons/village/ year 718.685

Volume (m3)/ household/year 120.77

Volume (m3) / village/year 21,619

420.321

73.29

13,119

91.46

8.39

1,502.7

23.52 20.907 7.84

6.3 5.84 0.69

1,143.3 1,045 124.13

1,206.068

215.28

38,553.13

Fig. 6 The average composition of fuelwood documented in households sampled. Aguacatillo is an unidentified species of the family Lauraceae (it was not identified from wood samples), while the category “others” include several unidentified species occurring in small amounts in fuelwood samples

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

543

Table 18 General plant species richness and useful plant species richness per environmental unit type Environmental unit type Tropical dry forest Milpas Riparian vegetation Pinus michoacana–Q. conzatii forest Pinus lawsonii–Q. crassifolia forest Homegardens Secondary Vegetation Alnus forest Ruderal areas near the village Quercus rugosa forest Quercus laurina forest Granada gardens Quercus magnolifolia forest

Species number 144 90 90 71

Number of useful species 97 70 68 52

% of useful species 67.36 77.78 75.56 73.24

71

42

59.15

66 57 44 41 36 28 16 10

62 48 19 38 22 16 16 10

93.94 84.21 43.18 92.68 61.11 57.14 100 100

abundance of the most used plants and compare this information with that obtained from surveys of extraction of products. In this chapter we only show the information considered more relevant for such an analysis. Other details will be published elsewhere and can be consulted in Solis (2006). A first aspect to highlight is in relation to the differential richness of species recorded on each vegetation type and the proportion of species that supply a benefit to people. As indicated in Table 18, the tropical dry forest has the highest plant species richness, most of them contributing to satisfy different human needs. The riparian vegetation is also outstanding in terms of diversity, while temperate forests (oak forest and pine forests) are important because of their extent and the arboreal biomass they harbor. On average, more than 60% of the species recorded in the vegetation sampling are plants used by local people. But it is also relevant to highlight that the anthropized vegetation types sampled (milpas, homegardens, and secondary vegetation) are all areas with exceptionally high proportion of plant species used by people. To summarize the information, Figs. 7, 8, 9, 10, 11, and 12 provide graphic illustrations of the different types of resources that are more abundant in the environmental units sampled. Each environmental unit provides different types of products. For instance, tropical dry forest, riparian vegetation, milpas, and secondary vegetation are main sources of fodder for livestock (Fig. 7); milpas, homegardens, and tropical dry forests are by far the main sources of food, milpas, and homegardens providing cultivated and weedy food, and the tropical dry forest supplying wild food (Fig. 8). Milpas, homegardens, riparian vegetation, and tropical dry forest are the main sources of medicinal plants, which, as mentioned, are mainly cultivated (Fig. 9). Oak and pine forest types and the tropical dry forests are the main sources of fuelwood and materials for construction, while ornamental plants are obtained

544

L. Solís and A. Casas

Fig. 7 Main environmental sources providing fodder species

from some pine forest, tropical dry forest, and riparian vegetation (Figs. 10, 11, 12). The analysis of distribution of plant species indicates that 61.5% of the useful plant species occur in only one vegetation type. The figures referred to and this information illustrate that the different ecosystem types are providing also different types of plant resources, and this is an important aspect to analyze the complementarity of them to people’s life.

Discussion and Conclusions The Cuicatec are an ancient group of people inhabiting the mountains of Mexico, closely related with the Mixtec, but relatively poorly known. Ours is one of the few, if not the only ethnobotanical study with the Cuicatec. Such scarcity of information is a condition that still predominates for most native cultures of Mexico and indicates the importance of enhancing priority research among these groups (Camou et al. 2016). Nomenclature and classification of plants have great similarities to those reported for the Mixtec, for whom life forms like trees (yutu), herbs (yuku), vines (yo’ó), the shrubs either herbs or trees; the cultural categories analogous to life forms, for instance quelites (yiwa), flowers (ita), and grasses (icha), are the principles for the general classification of plants, then followed by terms indicating

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

545

Fig. 8 Main environmental sources providing edible species

generic, specific, and varietal categories (Casas et al. 1994; De Ávila 2004). Deeper linguistic studies among the Mixtec were conducted by De Ávila (2010), and it is widely recommended to the readers to follow that interesting piece of work, which is extraordinarily helpful to understand the Cuicatec classification of plants. That study helped very much to organize the main units of classification that we documented among the Cuicatec. As reviewed in this chapter, the subsistence of the Cuicatec strongly depends on agriculture, for direct consumption of products that are staple food (maize and beans principally), but also to obtain monetary income from fruit production in granada (Passiflora ligularis) plantations and homegardens, which are commercialized and allow obtaining the staple products that their agricultural systems are not able to satisfy. Livestock is important for obtaining monetary income (mainly from sheep and goats), as labor animals (cattle, donkeys, and horses), while backyard animals (mainly hens, turkeys, and pigs) are destined to occasional consumption of their products. Monetary incomes are complemented by the remittances derived from jobs out of the village although these are irregular and difficult to evaluate. And all these sources of money have gained importance in the last decades, through a process in which commercial products associated to the food system (chemical fertilizers for agriculture, chicken meat in the communitarian store, buying staple food in markets, and buying fuelwood by nearly half of local households).

546

L. Solís and A. Casas

Fig. 9 Main sources of medicinal plants

Maybe, the most relevant finding of this study is the ongoing use and strong human culture associated to using wild and weedy plants to satisfy different needs, mainly food, fuelwood, and construction. Our study identified that the satisfaction of these needs could be also the most meaningful in terms of possible impact on natural ecosystems and biodiversity occurring in the territory of San Lorenzo Pápalo. We have seen that procuring of food impacts on local ecosystems mainly through changes in land use, the transformation of forest into agricultural plots. Gathering of food has low impact on natural ecosystems since the highest diversity and abundance of edible biomass occur in the production systems, as weedy plants (quelites), while other significant wild food consists of fruits and seeds, whose gathering is in important amounts (as it is the case of Stenocereus pruinosus), but the impact of collecting these parts is generally low (Blancas et al. 2010; TorresGarcía et al. 2015, 2020; Arellanes et al. 2018). The impact could be more relevant through the gathering of edible tubers, roots and bulbs, since the extraction of these parts may cause severe damage to plants and may affect their survival (Blancas et al. 2010); however, the gathering of these edible parts is relatively low. Therefore, the main challenge for the community to maintain sustainable ways of procuring food is mainly related to strategies to maintain the food production systems functioning in the long term, in order to decrease the rate of clearing land for new agricultural plots. The impact of livestock has not been evaluated; for the moment we identified the main areas used for grassing the animals through the free raising regime. But it is still

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

547

Fig. 10 Main sources of fuelwood

necessary to analyze the number of animals raised, and the real impact on these areas to develop proposals of optimal management of grassing areas and size of herds with ecological criteria. Hunting and gathering of animals and their products are mainly directed to obtain food (Solís and Casas 2019). Our study analyzed the potential impact of these activities on particular species, those more frequently hunted (squirrels and armadillo), and those affected by ecosystem transformation (deer), but these studies require more specific monitoring of the effects of these practices on particular populations. These studies would be important for contributing ecological criteria for processes of local regulations construction. Apart from transformation of forests to production systems, wood extraction is apparently the most important practice impacting on forests. The extraction of wood for construction may cause a significant impact during certain periods. For instance, because of the occurrence of severe landslides and floods part of the town had to move from one area to another and the construction of new houses represented a sudden impact associated to this activity, but these events are apparently rather rare, and the construction of new houses or the renovation of part of their structure may be occasional. What is more significant, due to the frequency and systematic occurrence, is fuelwood extraction. There are several options developed in areas of Mexico that could be locally adopted. One of the most viable, because of cultural and technical reasons, is the promotion of programs for constructing efficient fuelwood stoves,

548

Fig. 11 Main environmental sources of materials for construction

Fig. 12 Main environmental sources of ornamental plants

L. Solís and A. Casas

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

549

which are being designed by several organizations and have demonstrated to significantly reduce the amounts of fuelwood used (Berrueta et al. 2017; Pine et al. 2011). According to our current diagnostic, this activity is the most impacting on natural ecosystem, threatening two oak species. Using efficient stoves not only would help to reduce the amount of biofuel but would favor a diversification of species suitable to be used, thus buffering the impact on those identified oak species. Undoubtedly, the use of a broad spectrum of plants for satisfying each important human need is a way to guarantee buffering the effect of impact on particular species. As in the case of fuelwood, looking for the ways different species may be similarly valued for the different activities could be a way to contribute to strengthen the buffering effect. This is a process that continually happens in the real situations. An appreciated edible plant like the chapoquelite is not frequently consumed since its availability is restricted by distance from the village and its low abundance. In part this is a situation that has motivated people to put in practice management techniques such as cultivation of some species. But not only cultivation is an option, in fact, commonly it is not feasible (Casas et al. 1997; Blancas et al. 2010); therefore, documenting the local strategies that people practice to substituting valuable products and developing innovations in this direction is probably as important as to experiment cultivation. Using diversity has been part of the successful experience of humans to maintain ecosystems they interact with. In part because they have several options to satisfy a need and this fact allows increasing their resilience capacity. Similarly, strategies of using different environmental units represent extraordinary important ways to complement the acquisitions of products and buffering impact on particular ecosystems. Our study illustrates how the Cuicatec of San Lorenzo Pápalo makes use of different species and different environmental units, complementing the products that can be obtained from each unit. The location of the village is strategic, it is settled on the intermediate elevation of the territory of the community, which favors the access to products from the warm and cold lands, and therefore favors the complementarity of products and ecosystems in their life. The local diversity of plant resources documented in this study represents nearly 15% of the whole diversity of plants utilized by peoples in the whole region. And, as mentioned in the Introduction section, the ethnofloristic inventory of the Tehuacan Valley is one of the richest documented in Mexico. This figure helps to dimension the relevance of diversity in the Cuicatec life, which is similar to what has been documented among other human cultures of the region, like the Ixcatec, the Nahua, the Mixtec, the Popolocan, and even the mestizo rural people. Using the diversity favors the conservation of diversity, and this is an important premise for designing management programs in different contexts, and especially in the Biosphere Reserve Tehuacán-Cuicatlán, where these people live, and where they have developed experiences and techniques to do it. Acknowledgments The authors thank the Consejo Nacional de Ciencia y Tecnología (CONACYT research project A1-S-14306), Mexico, the Dirección General de Asuntos del Personal Académico (DGAPA, UNAM, research project IN206520 and IN224023), and the Comisión Nacional para el Conocimiento y Uso de la Biodiversidad (CONABIO/GEF/FAO project ID 9389 770, research project RG023) for financial support.

550

L. Solís and A. Casas

References Anderson ER, Concepción H. Diccionario cuicateco. Serie de Vocabularios y Diccionarios Indígenas “Mariano Silva y Aceves” núm. 26. Mexico City: Instituto Lingüístico de Verano; 1983. Arellanes Y, Romero-Sosa MA, Vega E, Maza-Villalobos S, Casas A. Ecological bases for sustainable management of Ceiba aesculifolia (H.B. & K.) Britten & Baker f. subsp. parvifolia (Rose) P.E. Gibbs & Semir through demographic analysis. Econ Bot. 2018;72(1):20–37. Basauri C. La población indígena de México. Oficina editora Popular, Secretaría de Educación Pública, Mexico City: Etnografía; 1940. Bazúa S. Los cuicatecos. Mexico City: Instituto Nacional Indigenista; 1982. Belmar F. El cuicateco. Oaxaca: Imprenta del Comercio; 1902. Berlin B. Ethnobiological classification: principles of categorization of plants and animals in traditional societies. Princeton: Princeton University Press; 1992. Berrueta VM, Serrano-Medrano M, García-Bustamante C, Astier M, Masera OR. Promoting sustainable local development of rural communities and mitigating climate change: the case of Mexico’s Patsari improved cookstove project. Clim Chang. 2017;140:63–77. Blancas J, Casas A, Rangel-Landa S, Torres I, Pérez-Negrón E, Solís L, Moreno-Calles AI, Delgado A, Parra F, Arellanes Y, Caballero J, Cortés L, Lira R, Dávila P. Plant management in the Tehuacán-Cuicatlán Valley. Econ Bot. 2010;64(4):287–302. Blancas J, Casas A, Pérez-Salicrup D, Caballero J, Vega E. Ecological and sociocultural factors influencing plant management in Nahuatl communities of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013;9:39. Byers DS, editor. The prehistory of the Tehuacan Valley, vol. I y II. Austin: Robert Peabody Foundation, University of Texas Press; 1967. Camou-Guerrero A, Casas A, Moreno-Calles AI, Aguilera-Lara J, Garrido-Rojas D, Rangel-LandaS, Torres-García I, Pérez-Negrón E, Solís L, Blancas JJ, Guillén S, Parra F, Rivera E. Chapter 2. Ethnobotany in Mexico: history, development and perspectives. In: Lira R, Casas A, Blancas JJ, editors. Ethnobotany of Mexico. Interactions of peoples and plants in Mesoamerica. Utrecht: Springer; 2016. p. 21–39. Casas A, Viveros JL, Caballero J. Etnobotánica mixteca. Sociedad, cultura y recursos naturales en La Montaña de Guerrero. Mexico City: Instituto Nacional Indigenista/Consejo Nacional para la Cultura y las Artes; 1994. Casas A, Vázquez MC, Viveros JL, Caballero J. Plant management among the Nahua and the Mixtec of the Balsas River basin: an ethnobotanical approach to the study of plant domestication. Hum Ecol. 1996;24(4):455–78. Casas A, Caballero J, Mapes C, Zárate S. Manejo de la vegetación y domesticación de plantas y origen de la agricultura en Mesoamerica. Bol Soc Bot Méx. 1997;61:31–47. Casas A, Valiente-Banuet A, Viveros JL, Caballero J, Cortés L, Dávila P, Lira R, Rodríguez I. Plant resources of the Tehuacán- Cuicatlán Valley, Mexico. Econ Bot. 2001;55:129–66. Casas A, Rangel-Landa S, Torres-García I, Pérez-Negrón E, Solís L, Parra F, Delgado A, Blancas JJ, Farfán B, Moreno-Calles AI. In situ management and conservation of plant resources in the Tehuacán-Cuicatlán Valley, Mexico: an ethnobotanical and ecological approach. In: Albuquerque UP, Alves-Ramos M, editors. Current topics in ethnobotany. Kerala: Research Signpost; 2008. p. 1–25. Casas A, Lira R, Torres I, Delgado-Lemus A, Moreno-Calles AI, Rangel-Landa S, Blancas J, Solís L, Pérez-Negrón E, Vallejo M, Parra F, Farfán-Heredia B, Arellanes Y. Ethnobotany for sustainable ecosystem management: A regional perspective in the Tehuacán Valley. In: Lira R, Casas A, Blancas J, editors. Chapter 8 Ethnobotany of Mexico. Interactions of peoples and plants in Mesoamerica. Utrecht: Springer; 2016. p. 179–206. Casas A, Parra F, Aguirre-Dugua X, Rangel-Landa S, Blancas J, Vallejo M, Moreno-Calles AI, Guillén S, Torres I, Delgado-Lemus A, Pérez-Negrón E, Figueredo CJ, Cruse-Sanders JM, Farfán-Heredia B, Solís L, Otero-Arnaiz A, Alvarado-Sizzo H, Camou-Guerrero A. Manejo y

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

551

domesticación de plantas en Mesoamérica. Una estrategia de investigación. Capítulo 3. In: Casas A, Torres-Guevara J, Parra F, editors. Domesticación en el Continente Americano. Volumen 2. Perspectivas de investigación y manejo sustentable de recursos genéticos en el Nuevo Mundo. Chapter 3. Morelia: Universidad Nacional Autónoma de México /Universidad Nacional Agraria La Molina / CONACYT; 2017. p. 69–102. Cavendish W. Rural livelihoods and non-timber forest products. In: de Jong W, Campbell B, editors. The role of non-timber forest products in socio-economic development. Wallingford: CABI Publishing; 2001. Clement CR, Casas A, Parra-Rondinel F, Levis C, Peroni N, Hanazaki N, Cortés-Zárraga L, RangelLanda S, Alves RP, Ferreira MJ, Cassino MF, Coelho SD, Cruz-Soriano A, Pancorbo-Olivera M, Blancas J, Martínez-Ballesté A, Lemes G, Lotero-Velásquez E, Bertin VM, Mazzochini GG. Disentangling domestication from food production systems in the Neotropics. Quaternary 2021;4(1):4. Cunningham AB. Applied ethnobotany. People, wild plant use and conservation. People and plants conservation manual. London: Earthscan, WWF, UNESCO, Royal Botanic Gardens Kew; 2001. Dávila P, Arizmendi MC, Valiente-Banuet A, Villaseñor JL, Casas A, Lira R. Biological diversity in the Tehuacán-Cuicatlán Valley, Mexico. Biodivers Conserv. 2002;11:421–42. De Ávila A. Clasificación de la vida en las lenguas de Oaxaca. In: García-Mendoza AJ, Ordóñez MJ, Briones-Salas M, editors. Biodiversidad de Oaxaca. México: Instituto de Biología UNAM, Fondo Oaxaqueño para la conservación de la Naturaleza, WWF; 2004. p. 481–539 De Ávila A. Mixtec plant nomenclature and classification. Dissertation of Doctor in Philosophy in Anthropology. Graduate Division of the University of California, Berkeley. 2010. De la Cerda SR. Los cuicatecos. Rev Mex Sociol. 1942;IV:4. Doesburg S. Códices cuicatecos: Porfirio Díaz y Fernández Leal. Edición facsimilar, contexto histórico e interpretación. El Gobierno Constitucional del Estado de Oaxaca and Editorial Porrúa. Oaxaca. 2001a. Doesburg S. Época prehispánica, los antiguos pobladores del valle. In: El Valle de TehuacánCuicatlán patrimonio natural y cultural. ICA, Banamex y Fundación Cuicatlán; 2001b. p. 137–52. Eberhard DM, Simons GF, Fennig CD, editors. Ethnologue: languages of the world. Twenty-fifth edition. Dallas, Texas: SIL International; 2022. Online version: http://www.ethnologue.com. Consultation date: 15 March 2022 Elfego A. Los cuicatecos actuales en: Anales del Museo Nacional de Arqueología, Historia y Etnografía. 1ra. Mexico City: Época, Imprenta del Museo Nacional de Antropología e Historia; 1922. Epstein G, Gurney G, Chawla S, Anderies JM, Baggio J, Unnikrishnan H, Villamayor-Tomas S, Cumming GS. Drivers of compliance monitoring in forest commons. Nat Sustain. 2021;4: 450–6. Ford A, Nigh R. The Maya forest garden. Eight millennia of sustainable cultivation of the tropical woodlands. New York: Routledge; 2015. Frei B, Baltisberger M, Sticher O, Heinrich M. Medical ethnobotany of the Zapotecs of the IsthmusSierra (Oaxaca, México): documentation and assessment of indigenous uses. J Ethnopharmacol. 1998;62:149–65. Geist I. Comunión y disensión: prácticas rituales en una aldea cuicateca. Instituto oaxaqueño de las culturas. Instituto Nacional de Antropología e Historia. Fondo Estatal para la Cultura y las Artes. Mexico City. 1997. Gómez-Baggethun E, Mingorría S, Reyes-García V, Calvet L, Montes C. Traditional ecological knowledge trends in the transition to a market economy: empirical study in the Doñana natural areas. Conserv Biol. 2010;24(3):721–9. Hunt E. Irrigation and the socio-political Organization of Cuicatec cazicazgo. In: Johnson F, editor. The Prehistory of the Tehuacán Valley, vol. IV. Austin: University of Texas Press; 1972. p. 162–253.

552

L. Solís and A. Casas

Kauffman T. Early Otomanguean homelands and cultures; some premature hypotheses. University of Pittburgh Working Papers in Linguistics. 1990;1:91–136. Lira R, Casas A, Rosas R, Paredes M, Pérez-Negrón E, Rangel-Landa S, Solís L, Torres I, Dávila P. Traditional knowledge and useful plant richness in the Tehuacán-Cuicatlán Valley, Mexico. Econ Bot. 2009;63:271–87. Lotero-Velásquez E, Garcia-Frapolli E, Blancas-Vázquez JJ, Casas A, Martinez-Ballesté A. Eco-symbiotic complementarity analysis and trading networks of natural resources in Nahua communities in mountain regions of Mexico. Hum Ecol. 2022; https://doi.org/10.1007/ s10745-022-00311-x. MacNeish RS. A summary of the subsistence. In: Byers D, editor. The prehistory of the Tehuacán Valley. Volume One: environment and subsistence. Austin: University of Texas Press; 1967. p. 290–310. MacNeish RS. The origins of agriculture and settled life. Norman: University of Oklahoma Press; 1992. Osorio O, Valiente-Banuet A, Dávila P, Medina R. Tipos de Vegetación y diversidad β en el Valle de Zapotitlán Salinas, Puebla, México. Bol Soc Bot Méx. 1996;59:35–58. Parra F, Casas A, Begazo D, Paco A, Márquez E, Cruz A, Segovia J, Torres-García I, Zarazúa M, Lizárraga L, Torres-Guevara J. Natural and cultural processed influencing gene flow among wild (atoq papa), weedy (araq papa and K’ipa papa), and crop potatoes in the Andean region of Southern Peru. Frontiers in Ecology and Evolution 2021;9:617969 Pieroni A. Evaluation of the cultural significance of wild food botanicals traditionally consumed in Northwestern Tuscany Italy. J Ethnobiol. 2001;21(1):89–104. Pine K, Edwards R, Masera O, Schilmanne A, Marrón-Mares A, Riojas-Rodríguez H. Adoption and use of improved biomass stoves in rural Mexico. Energy Sustain Dev. 2011;15(2):176–83. Rangel-Landa S, Casas A, Rivera-Lozoya E, Torres I, Vallejo M. Ixcatec ethnoecology: biocultural principles of plant management in Oaxaca, Mexico. J Ethnobiol Ethnomed. 2016;12:30. Salisbury RF. Vunamami. Economic transformation in a traditional society. Berkeley: University of California Berkeley; 1970. Solís L. Etnoecología cuicateca en San Lorenzo Pápalo, Oaxaca. Master in Sciences Thesis. Posgrado en Ciencias Biológicas. Morelia: Universidad Nacional Autónoma de México;. 2006. Solís L, Casas A. Cuicatec ethnozoology: traditional knowledge, use and management of fauna by people of San Lorenzo Pápalo, Oaxaca, Mexico. J Ethnobiol Ethnomed. 2019;15:58. Starr F. The physical characters of the Indians of Southern Mexico. Chicago: The University of Chicago Press; 1902. Stoffle RW, Halmo DB, Evans MJ. Puchuxwavaats uapi (to know about plants). Traditional knowledge and the cultural significance of southern Paiute plants. Hum Organ. 1999;58(4): 416–29. Toledo VM. The ecological rationality of peasant production. In: Hecht SB, editor. Agroecology and small farm development Altieri MA. Boca Raton: CRC Press; 1990. p. 53–60. Toledo VM, Barrera BN. La memoria biocultural. La importancia ecológica de las sabidurías tradicionales. Barcelona: Icaria Editorial; 2008. Toledo VM, Argueta A, Rojas P, Mapes C, Caballero J. Uso múltiple del ecosistema: Estrategia del desarrollo. Ciencia y Desarrollo. 1976;2:33–9. Torres-García I, Casas A, Delgado-Lemus A, Rangel-Landa S. Aprovechamiento, demografía y establecimiento de Agave potatorum en el Valle de Tehuacán, México: Aportes etnobiológicos y ecológicos para su manejo sustentable. Zonas Áridas. 2013;15(1):92–109. Torres-García I, Casas A, Vega E, Martínez-Ramos M, Delgado-Lemus A. Population dynamics and sustainable management of mescal agaves in central Mexico: Agave potatorum in the Tehuacán-Cuicatlán Valley. Econ Bot. 2015;69(1):26–41. Torres-García I, León A, Vega E, Moreno-Calles AI, Casas A. Integral projection models and sustainable forest management of Agave inaequidens in western Mexico. Front Plant Sci. 2020;11:1224.

Cuicatec Ethnobotany: Plants and Subsistence in San Lorenzo Pa´palo, Oaxaca

553

Turner NJ. The importance of a rose. Evaluating the cultural significance of plants in Thompson and Lilooet Interior Salixh. Am Anthropol. 1988;90(2):272–90. Tuxill J, Nabhan GP. People, plants and protected areas: a guide to in situ management. London: Earhscan; 2001. Valiente-Banuet A, Casas A, Alcántara A, Dávila P, Flores N, Arizmendi MC, Villaseñor JL, Ortega J, Soriano JA. La vegetación del Valle de Tehuacán-Cuicatlán. Bol Soc Bot Méx. 2000;67:25–74. Valiente-Banuet A, Casas A, Dávila P, Arizmendi MC, Solís L, Treviño J, Rangel S. La vegetación del Valle de Tehuacán-Cuicatlán. México City: UNAM, CONABIO, Fundación Cuicatlán; 2009. Valiente-Banuet A, Aizen MA, Alcántara JM, Arroyo J, Cocucci A, Galetti M, García MB, García D, Gómez JM, Jordano P, Medel R, Navarro L, Obeso JR, Oviedo R, Ramírez N, Rey PJ, Traveset A, Verdú M, Zamora R. Beyond species loss: the extinction of ecological interactions in a changing world. Funct Ecol. 2015;29(3):299–307. Weitlaner RJ. The Cuicatec. In: Wauchope R, editor. Handbook of Middle American Indians, vol. 7. Austin: University of Texas Press; 1969. Zarazúa-Carbajal M, Chávez-Gutiérrez M, Romero-Bautista Y, Rangel-Landa S, Moreno-Calles AI, Alvarado-Ramos F, Blancas-Vázquez JJ, del Val-de Gortari E, Arizmendi MC, Casas A. Use and management of wild fauna by peoples of the Tehuacán-Cuicatlán Valley and the surrounding areas, Mexico. J Ethnobiol Ethnomed. 2020;16(4):1–23.

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding the Tehuacán Valley Selene Rangel-Landa and Alejandro Casas

Abstract

Peoples’ knowledge about their natural surroundings is tightly linked with practices to appropriate natural components, ecological processes occurring in the ecosystems, and to their cosmovision. Peoples and their surroundings are mutually transformed through management practices and the influence of landscapes on peoples’ culture. A portion of the peoples’ knowledge is codified by languages as names, classification systems, word meanings, and the contexts in which these words are pronounced. The loss of languages, therefore, leads to the risk of losing an essential part of peoples’ memories. Santa María Ixcatlán is a town located in the state of Oaxaca, Mexico, with a population of nearly 500 inhabitants, with an alarming risk of disappearance of the Ixcatec language (only eight Ixcatec speakers) and the loss of knowledge about the environment that it implies. Ethnobotanical documentation may contribute to safeguarding the invaluable memory of the Ixcatec people. In this chapter, we provide a summary of our records about the people-plant interactions in this town. Our study attempts to understand how a rich biocultural legacy has been built in one of the most biodiverse zones in the semiarid mountains of Mexico. Also, to analyzing the challenges the Ixcatec people face to maintain this legacy and provide an account of the ways in which they confront them. We reviewed the literature with S. Rangel-Landa (*) Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Consejo Nacional de Ciencia y Tecnología (CONACYT) – Escuela Nacional de Antropología e Historia (ENAH), Mexico City, Mexico e-mail: [email protected] A. Casas Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_20

555

556

S. Rangel-Landa and A. Casas

information about the Ixcatec people and their interactions with plants, and analyzed the published and unpublished results of ethnobotanical and ethnoecological work conducted by our research team from 1999 to the present. In Santa María Ixcatlán, agriculture is the axial activity around which all other productive activities and the community’s ceremonial life are organized. Local farmers grow the basic staple foodstuffs consumed throughout the year like maize, beans, and supplementary crops like wheat and squash. They also manage edible weedy plant species that are essential for the local gastronomy. They use a total of 627 plant species to satisfy food, medicine, construction, handcrafts, fodder, and other needs. We analyze these aspects in this chapter. In addition, they manage 401 species through practices for ensuring their availability including tolerance (206 spp.), protection (251 spp.), transplanting (139), enhancement (34 ssp.) and ex situ propagation (155 spp.). In addition, they practice the gathering of 299 spp., and the raising of livestock that forage 243 plant species in different ecosystems of their territory. People mentioned 94 plant names corresponding to 129 species that are considered essential for living, and we centered our attention on these species. The Ixcatec ethnobotanical knowledge is possibly the most deeply documented in the Tehuacán-Cuicatlán region. However, we identified a process of loss of local knowledge associated to the substitution of local products for others, the loss of knowledge codified in the Ixcatec language, and the high rate of migration, especially the young people. Improving the conditions of interchange through organizational processes as well as the innovation of productive practices to generate profitable products are viable ways identified by local people to enhance local inhabitants to remain linked to the community. We refer to the initiatives of Xula Palma Artesanal collective and the Ixcateco collective brand, which have provided new opportunities to improve the lives of some households and demonstrated to be effective to maintain appropriate ways to manage plants and vegetation and to value the Ixcatec culture. For the members of Xula and Ixateco, their experience may be the base to strengthening customs and cultural aspects that give identity to their products. This experience illustrates the importance of documenting and understanding the local productive processes through ethnobotanical studies. In this way, ethnobotany may contribute to support the local efforts to improve life and maintain the valuable biocultural heritage of the Ixcatec people.

Introduction Through their interactions with the environment, peoples construct a body of knowledge, perceptions, beliefs, emotions, worldviews, attitudes, and practices on which they base their coexistence with the natural surroundings. These factors strongly influence processes of using and adapting components and functions of ecosystems to satisfy different human needs (Casas et al. 2014). As a result, peoples and their surroundings are mutually transformed. Management practices act on individual organisms modifying their populations, and landscapes are transformed by changes in their biophysical conditions as a consequence of their utilization.

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

557

Conversely, the available resources in managed landscape contribute to modulate the peoples’ lifestyles, costumes, social interactions, and culture (Boege 2008; Toledo and Barrera-Bassols 2008; Casas et al. 2015). According to Berkes (1999) and Toledo (2002), peoples’ knowledge about their natural surroundings (the corpus) is tightly linked both to their practices to appropriate natural components, ecological processes, and ecosystems (the praxis), and to their beliefs, values, and cosmovision (the kosmos). The kosmos modulates how human beings conceive themselves relative to the natural environment and determine the ways in which they interact with it. Corpus, praxis, and kosmos conform the systems of traditional knowledge that are maintained and transmitted through generations by management practices, use forms, tales, stories, and chronicles (Berkes et al. 2000; Gadgil et al. 1993; Toledo 2002). Therefore, all people possess a particular history and unique legacy, from which humanity is and has been benefited. A portion of the peoples’ knowledge is codified by languages as names, classification systems, word meanings, and the contexts in which these words are pronounced (Wehi et al. 2009). The loss of languages, therefore, leads to the risk of losing an essential part of peoples’ memories. When a language is abandoned, a part of the original knowledge is transmitted or translated, but there is a partial loss of this knowledge (Zent 2001; Harrison 2007; Si 2016). Additionally, the loss of knowledge is increased by the gradual loss of peoples’ original interactions with natural components because of changes in their activities and lifestyles. These are for instance the loss of food habits and healthcare practices, the transformation of ecosystems, or the modification of other socioecological elements. Santa María Ixcatlán is a town located in the state of Oaxaca, Mexico, with a population of nearly 500 inhabitants, which is currently undergoing the continual emigration of youngsters. The town is the only place where the Ixcatec people exist as a community linked to its ancestral territory (INEGI 2020). The Ixcatec is one of the 48 endangered languages in Mexico, with less than 10 people recognized as speakers of the native language (Eberhard et al. 2022; Embriz and Zamora 2012). The undeniable and alarming risk of disappearance of the Ixcatec language and the loss of their knowledge about the environment make ethnobotanical documentation a way to contribute to safeguarding the invaluable memory of the Ixcatec people. It is crucial to identify and understand the processes that drive the loss, the way of transmission, and the innovation of the environmental knowledge of the Ixcatec people. Such aspects would contribute criteria and tools to face the challenges involved in managing the ecosystems and plant species they considered essential for their subsistence strategies. Also, these aspects would aid their initiatives to vindicate their identity, language, and culture. In this chapter, we provide a summary of our records about the people-plant interactions in Santa María Ixcatlán. Our study attempts to understand how a rich biocultural legacy has been built in one of the most biodiverse zones in the semiarid mountains of Mexico. Also, to documenting the challenges the Ixcatec people face to maintain this legacy and provide an account of the ways in which they confront them. For these purposes, we reviewed the literature with information about the Ixcatec people and their interactions with plants, and analyzed the published and unpublished results of ethnobotanical and ethnoecological work conducted by our

558

S. Rangel-Landa and A. Casas

research team from 1999 to the present. Also, we show a panorama of our collaboration with community members in events, actions, and projects directed to support the recognition of the value and maintenance of the Ixcatec biocultural heritage (Rangel-Landa et al. 2014, 2016a, b, 2017; Smith-Aguilar et al. 2016; Swanton 2022). The methods we used in this research were: (i) open and semi-structured interviews and free listing of plants used for different purposes; (ii) participant observation in the different activities of local households; (iii) sampling of the different vegetation types in order to evaluate distribution, abundance, and ecological importance of plant species in the local environmental contexts; (iv) documentation of Ixcatec botanical knowledge through sessions with local experts and Ixcatec speakers, by using fresh and herborized plant vouchers as prompts; (v) field trips and sampling to document the floristic composition of home gardens, agricultural fields, and relevant anthropized areas; (vi) exchanges of local experiences of plant management with other peoples; and (vii) follow-up of experiments for the innovation of management practices. Further details about the methodology and analyses can be consulted in Rangel-Landa et al. (2016a, 2017), and in Smith-Aguilar et al. (2016).

The Ixcatec and Their Territory The Ixcatec are settled in what is presently known as the town and municipality of Santa María Ixcatlán in the state of Oaxaca. Santa María Ixcatlán is located in the mountains surrounding the Southeastern Tehuacán-Cuicatlán Valley, a semiarid zone with an exceptional biocultural diversity. For this reason, the region was decreed in 1998 as the Tehuacán-Cuicatlán Biosphere Reserve (CONANP 2013). In this region, nearly 36 plant associations harbor over 3000 vascular plant species, more than 565 vertebrate species, and a high diversity of little-studied invertebrates. Human populations have found refuge and sustenance in the Tehuacan-Cuicatlán Valley since at least 12,000 years ago, when hunter-gatherer groups left remains that have allowed archaeologists documenting their presence in the area (MacNeish 1967; CONANP 2013; Dávila et al. 2002). At present, the region is inhabited by Popolocan, Chocholtec, Chinantec, Mazatec, Cuicatec, Ixcatec, Mixtec, Nahua, and mestizo people and it harbors a large biocultural diversity with more than 2000 useful plant species. Until now, over 600 plant species have been recorded whose permanence and abundance is secured by agricultural and silvicultural management practices incorporated into ecological and productive systems covering a wide range of agroforestry practices (Casas et al. 2007; Blancas et al. 2010, 2013; Larios et al. 2013; Lira et al. 2009; Moreno-Calles et al. 2010; Vallejo-Ramos et al. 2016). The origin and history of the Ixcatec people is little known, but based on Fernández et al. (1959) estimation of the period presumably elapsed since the Ixcatec language differentiated from the Otomanguean languages most related to the Ixcatec (the Chocholtec, Popolocan, and Mazatec), whose speakers coexist in the region together with the Cuicatec and Mixtec speakers. We may assume it has had a long interaction with its present territory during at least the past 1300 years (Fig. 1). The Ixcatec people

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

559

Fig. 1 Location of the community of Santa María Ixcatlán, Oaxaca, and neighboring towns

are recognized as an indigenous community governed by the “uses and costumes” regime, through which the community assembly designates its own civic and agrarian authorities by choosing them among the community members. The assembly also makes most of the decisions regarding the community’s management of the territory, civic life, and even part of its religious organization (Nava and Romero 2007). According to Cook (1958) and Hoppe and Weitlaner (1969), the Ixcatec population was approximately 10,000 inhabitants at the time of the arrival of the Spaniards (Hironymous (2007). Since that time, due to diseases, the bad conditions in mines owned by Spanish encomenderos, among other factors, by 1803, the population decreased to 361 inhabitants. In 1945, the population reached its maximal recovery with 1113 inhabitants, and has been decreasing since then (Cook 1958; Hironymous 2007). INEGI (2020) reported the present population to be 461 inhabitants, grouped in 163 households. The continued population decrease since the mid-twentieth century has been due to a high emigration rate caused by conflicts with the neighboring community of San Miguel Huautla, which had violent expressions between the 1940s and 1960s (Cook 1958; Hironymous 2007). After that period, emigration has been mostly driven by the search for employment and providing economic support to family members that remain in the community (parents, younger siblings, and grandparents). In addition, the birth rate decreased mostly due to the Mexican government’s birth control campaigns (Hironymous 2007; Nava and Romero 2007).

560

S. Rangel-Landa and A. Casas

Anthropologists and linguists considered that the Ixcatec population became bilingual in 1956, when some inhabitants began to abandon the Ixcatec language (Swanton 2008). In early 2022, eight inhabitants older than 50 years, and five of them older than 80, were fluent speakers of Ixcatec. In addition, several adults older than 50 years remembered some words and short phrases they listened to from their grandparents, but they declared not being able to speak the language (Swanton 2022). Some youngsters and children had learned some basic words and phrases in the Ixcatec language recovery courses and workshops carried out by several institutions since 1993 (Molina 2010). According to official documents, the Ixcatec territory has an area of 41,530 ha (DOF 1948). The whole territory is mountainous, with elevations ranging from 800–2600 m. The three zones in the Ixcatec territory characterized by altitudinal range, climate, and vegetation are: (i) temperate zone, located between 1600 and 2000 m, where the town of Santa María Ixcatlán is settled, having temperate climate, oak forests dominated by Quercus liebmani Oersted and Q. laeta Liebm., microphyllous scrub (locally called mexical), palm scrubland dominated by Brahea dulcis (Kunth) Mart., and gallery forests with Taxodium mucronatum Ten.; (ii) cold zone, located in the highlands at elevations between 2000 and 2600 m, with a cooler temperate climate with presence of fog during the rainy season and winter, mixed oak forests with Q. urbanii Trel., Q. castanea Née, Q. conspersa Benth., and a diverse community of epiphytic plants; and (iii) warm zone, located in the elevations from below 800 to about 1600 m, dry and warm climate, dominated by a great variety of small-sized trees and shrubs such as Bursera spp., Pistacia mexicana Kunth, Sarcomphalus amole (Sessé & Moc.) Hauenschild, Cephalocereus columnatrajani (Karw. ex Pfeiff.) K.Schum., Yucca periculosa Baker, C. fulviceps (F.A.C. Weber ex K.Schum.) H.E.Moore, Pseudalcantarea grandis (Schltdl.) Pinzón and Barfuss, Agave potatorum Zucc., Lindleya mespiloides Kunth, and Lippia origanoides Kunth in tropical deciduous forests, thorn scrub, and cacti forests (Rangel-Landa et al. 2016a; Fig. 3). The limits and settlements of Santa María Ixcatlán have been dynamic throughout history, mainly in response to environmental factors like droughts, and, as mentioned above, to population decrease. The Ixcatec people came to constitute a seigniory with seven dependent towns, but since the sixteenth century only Santa María Ixcatlán has remained as the only permanent settlement (Hironymous 2007). The pre-Hispanic settlements are represented by several vestigial constructions and irrigation and soil management infrastructure (Hironymous 2007) (Fig. 2). Elder members of the community reported that, until nearly 50 years ago, some families from the community inhabited these sites for some months or years to take advantage of the more fertile and resource-rich land. These land occupations by small families were called ranchos and have preserved the memory of landscape domestication processes (Fig. 2). Examples of these occupations are places with the presence of scrubland dominated by Brahea dulcis (palma criolla) – a type of secondary vegetation favored by management (Illsley et al. 2001; Rangel-Landa et al. 2014; Rzedowski 1978). Also, small patches of plants like Leucaena esculenta (DC) Benth. (guaje), Lophocereus marginatus (DC.) Backeb. (órgano), Opuntia huajuapensis Bravo (nopal), Agave salmiana

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

561

Fig. 2 Landscapes in pre-Hispanic settlements Santa María Ixcatlán. (a) Vestiges of constructions used until the middle of the mid-twentieth century for the production of mescal in the Santiago River near the pre-Hispanic settlement of La Iglesia; (b) cross-channel dam in El Bartolo, near the settlement of Santa María Ixcatlán; (c) and (d) pre-Hispanic settlement of San Juan Viejo where the low-intensity harvest of leaves maintains the palm grove Palmonar of Brahea dulcis, other plants associated with settlements are present, such as Leucaena esculenta, Lophocereus marginatus, Opuntia huajuapensis, and Agave salmiana subsp. tehuacanensis. (Photos: Selene Rangel-Landa)

subsp. tehuacanensis (Karw. ex Salm-Dyck) García-Mend. (maguey), Morus celtidifolia Kunth (moral), and Celtis caudata Planch. (moralillo) were mentioned by people to have been important resources for a long time. These species are still abundant in homegardens (locally known in Spanish as solares) and agricultural fields that conform agroforestry systems (Figs. 2 and 3). The lifestyles and livelihoods of rural people are defined by two environmental factors: precipitation and the predominant soil type. In Santa María Ixcatlán, precipitation is low (646 mm per year on average) with a marked canicule, and frequent interannual drought periods (SMN-CONAGUA n.d.). The predominant soil type at intermediate elevations is of calcareous origin, which is known to be of low suitability for agriculture. In the memory of the inhabitants of Santa María Ixcatlán, “rainfall is lessening.” However, the aridity of this region was one of the first observations reported in the sixteenth century by Velazquez de Lara (1579), who described the settlement as lacking permanent water sources except for seasonal streams. Likewise, Cook’s (1958) description in the 1940s mentions low and irregularly distributed precipitation. According to meteorological records of the SMN-CONAGUA (n.d.) from 1955–2016 and considering the 33 year period with

562

S. Rangel-Landa and A. Casas

Fig. 3 Agroforestal systems and management practices. (a) and (b) Agricultural fields around the town, where several wild species are kept, constituting the agroforestry complex Milpa-palmonar of Brahea dulcis; (c) milpa cultivation on solares (yards and backyards); (d) space in the backyard of the family garden (corral) dedicated to the care of plants; (e–g) sites where mescal is prepared (palenque), where plants associated with springs such as Taxodium mucronatum are maintained,

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

563

complete data, the average annual precipitation in the region is 646 mm (SD ¼ 186 mm), with a maximum of 1069 mm and a minimum of 291 mm, becoming as low as under 200 mm, as recorded in 1938 (Cook 1958). At present, two springs supply drinking water for the community’s household members, and their livestock drink from five small artificial reservoirs.

A Subsistence Strategy Based on Diversity In Santa María Ixcatlán, agricultural production is the axial activity around which all other productive activities and the community’s ceremonial life are organized. Local farmers grow the basic staple foodstuffs consumed throughout the year like maize – known as maíz criollo, pole and bush black beans –frijol negro enrredador and frijol negro de mata or de tierra, and supplementary crops like wheat and squash. They also manage edible weedy plant species that are essential for the local gastronomy like Physalis philadelphica Lam. (miltomate), Amaranthus hybridus L. (quelite tintonil), and Dysphania ambrosioides (L.) Mosyakin and Clemants. Crops are grown in plots near the town, where several wild species are managed within home gardens, or in yards transformed into agricultural plots during the rainy season (Fig. 3). Although agricultural production is essential for the subsistence strategy of the inhabitants of Santa María Ixcatlán, it is insufficient. For example, in the year 2000, the local production of maize was only enough to cover 37% of the local consumption of this staple food, and it was 20% in 2012, while the production of bean covered only 46% and 49% of the community’s needs for the same years, respectively (Rangel-Landa et al. 2016a). People’s concern regarding uncertainty in the agricultural cycle is expressed in their worries about whether rainfall will be sufficient, if it will occur before August so that crops can be harvested before mid-October when temperature decreases. Also, if crops seedlings will survive after seed germination, or if once plants have bloomed, there will be enough precipitation for the development of fruits and seeds (Hironymous 2007; Rangel-Landa et al. 2016a). Uncertainty and an unfavorable balance between costs and benefits have led to the loss of some crops. The former cultivation of the maíz de cajete – a slow-growth maize variety sown early in the year in canyons – is registered in the elders’ memory and records of the mid-twentieth century (Hoppe and Weitlaner 1969). However, the crop was abandoned due to the lack of winter rainfall and conflicts with the neighboring community of San Miguel Huautla for land where the crop was grown. Some households cultivated potato, barley, and peas in the high-elevation ä Fig. 3 (continued) and plants whose maintenance is only possible in these sites such as avocado; (h) sites in the oak forests called paddocks (potreros) conditioned for the maintenance of cattle; (i) “cold land” dominated by oak forest; and (j) Cephalocereus columna-trajani shrubland and tropical deciduous forests in the “warm land”. (Photos: Selene Rangel-Landa)

564

S. Rangel-Landa and A. Casas

temperate zones (Fernández 1950), until these crops were abandoned in the early 1990s. Their abandonment was caused by repeated droughts and because the effort of traveling for several hours or days to work in these upper lands surpassed the benefits of the harvest. In addition, these crops became available in local stores at accessible prices. Many elders expressed the hardships they experienced for subsisting in this uncertain scenario. But local people have managed to survive using sociocultural strategies including: (i) maintaining strong family and community links to ensure mutual support in case of emergencies; (ii) diversifying household’s members activities; and (iii) using and managing diverse plant species, productive systems, and territorial units (Blancas et al. 2013, 2014; Rangel-Landa et al. 2017). The diversification of productive activities within households is one of the main socioeconomic strategies for ensuring enough income to satisfy their members’ needs (Rangel-Landa et al. 2016a). Besides providing the households with maize, beans, and other crops, agricultural production includes management of weedy plant species to supplement the basic diet of people, and crop wastes that are a significant input of fodder. The use of 627 plant species and the management of 401 species through practices for ensuring their availability provide the base for maintaining the local population’s productive activities. Plant management includes tolerance (206 spp.), protection (251 spp.), transplanting (139), enhancement (34 ssp.) and ex situ propagation (155 spp.), as well as the gathering of 299 spp., and the foraging of 243 spp. by livestock. These practices allow satisfying the basic needs of foodstuffs, healthcare, firewood, construction, tools and utensils manufacturing, ornament, shade, living fences, as well as ritual and ceremonial inputs (Table 1; RangelLanda et al. 2016a). The products of some plant species supply currency to community members, such as firewood from Quercus spp., leaves from the palm B. dulcis, and resin from Bursera spp. trees (copal). There is an agreement among people of the community to restrict the sale of unprocessed natural products with other members of the community. The only exception are leaves of the palm B. dulcis that are allowed to be bartered for maize, fruits, vegetables, and other products offered by a small number of traders visiting the community. The basic productive activities of the community are agriculture, manufacture, and trading of hats and artisanal products made with Brahea dulcis leaves, the elaboration of artisanal mescal from stems of Agave potatorum, and small-scale extensive livestock production of cattle, goats, and sheep. Hats are traded or bartered in local stores and mescal is mostly retailed to local inhabitants and visitors, but in a growing amount. Palm-woven handicraft products are sold outside the community and some mescal producers are distributing their product through wholesale traders who sell it in regional, domestic, and foreign markets. For most households in the community, livestock grazing in common use areas represents a source of saving money for engagements or emergencies, and only less than 3% of households depend on it for income to satisfy basic needs. Other activities unrelated with the use and management of natural resources, but that are essential for the community’s subsistence strategy, are the occupation in communitarian services like commerce, masonry, music, day labor, and others.

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

565

Table 1 Useful plant species of Santa María Ixcatlán. (Based on Rangel-Landa et al. (2016a)) Use Fodder Ornamental Medicinal Edible Ceremonial Firewood Utensils Living fences Timber products and construction Shade Food additive (flavor) Handcrafts Insects repellent Soil control Animals medicine Facilitator Toys Alcoholic beverages Cosmetic Soap Paint Weather predictors Aromatizing Tannin source Water attracter Glues Poisons Useful species

Native 238 160 166 72 73 44 29 24 27

Introduced 30 110 53 66 55 2 4 6 2

Total 268 270 219 138 128 46 33 30 29

12 9 10 8 6 5 3 5 2 2 2 3 2 1 1 1 1 1 439 (Ixc)

23 15 11 8 8 6 5 5 3 3 3 3 2 1 1 1 1 1 627

Unknown use TOTAL

150 589

11 6 1 0 2 1 2 0 1 1 1 0 0 0 0 0 0 0 154 from other regions 31 from other areas of the Tehuacan-Cuicatlán Valley 3 of unknown origin 3 191

153 780

These activities, for the majority of the community members, are temporal or occasional. Remittances are especially important for emergency expenses and communitarian engagements, and the support from governmental programs has gained importance for the community members’ subsistence strategy. The latter source of income provides the local inhabitants resources for covering their daily needs, as we observed during the coronavirus pandemic in 2020 and 2021. In this period, the regional trade of palm hats was completely stalled, and many families depended on it for acquiring maize and other basic foodstuffs. Then, the remittances were crucial for the maintenance of local people’s life.

566

S. Rangel-Landa and A. Casas

The Essential Plants for Livelihood Although there is a perception that all plants and everything that is part of the mountains are important, when we asked people about the plants that were essential for living, they mentioned 94 plant names corresponding to 129 species (Fig. 4). Oaks Quercus spp. (yange in Ixcatec) were among the 30 most valued species. They provide the highest quality firewood (essential for cooking, mescal production, and bread baking), acorns to feed livestock and wild animals, lumber for building houses, and because they provide shade and fresh air – these are expressed as: “... they give life to the land . . .” and “... they attract and hold water.” Q. acutifolia Née and Q. conspersa Benth are particularly valued for women’s healthcare, Juniperus flaccida and Morus celtidifolia Kunth are especially appreciated, the former tree for providing shade in agricultural fields and the latter in homegardens, besides for other benefits (Rangel-Landa et al. 2016a). The members of households that we surveyed included within the most essential plants for livelihood basic crops like Zea mays L., Phaseolus vulgaris L., and Triticum aestivum L., and other edible plants that form part of their basic diet like Opuntia spp., Leucaena spp., A. hybridus, and Dysphania ambrosioides. Also, they mentioned the economically important plant species Brahea dulcis and Agave potatorum (Table 1; Appendix 1; Fig. 4). People also included in the list plants used in healthcare like Grindelia inuloides Willd., Mentha x piperita L., Aloe vera (L.) Burm.f., Casimiroa edulis La Llave and Lex., and Lippia oaxacana B.L.Rob. and Greenm. (Table 1; Fig. 4). People of Ixcatlán value grass species as source of fodder and because they cover the ground and give life to the mountains. They also mentioned plants used in rituals, as it is the case of Litsea glaucescens Kunth whose branches are used to make bouquets with which they start processions and adorn the church facade. Of the 35 Ixcatlán inhabitants we interviewed, 20% included water in the list. Oak trees in the forests surrounding the settlement are highly valued for their relationship with water, which suggests that local people associate forests with the water cycle. This connection appears to have an influence on their decision-making about the management of plants and vegetation.

Nourishment and Edible Plants Of the plant species we recorded in the community, 138 were edible and nearly 50 were basic for the local people’s diet (Rangel-Landa et al. 2016a; Table 1; Appendix 1). The inhabitants of Santa María Ixcatlán eat their meals in the morning and in the afternoon, and, in some households, their members also eat a light meal at night. Like centuries ago, maize and beans are part of the everyday meals (Velazquez de Lara 1579; Appendix 1). Breakfast includes tortillas, cooked beans flavored with Dysphania ambrosioides (epazote), eggs, sauces prepared with Capsicum annuum L. (chili) and miltomate (Physalis philadelphica), coffee or atole, and bread. The afternoon meal includes a soup, stew or broth with tortillas, and beans. The nocturnal light meal consists of coffee or atole accompanied by bread, or a small serving of beans with tortillas.

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

567

Fig. 4 The essential plants for livelihood. Cognitive prominence values expressed as Sutrop relative prominence index (S), where S ¼ F/(N mP), where F represents the frequency of the species, N the total number of interviewed people per use category, and mP is the medium position in which the term or species was named (Sutrop 2001)

568

S. Rangel-Landa and A. Casas

Tortillas, beans, and sauces are consumed daily in all households, and are always present in festive meals, for which they can be considered as the core of the diet and gastronomic culture in Santa María Ixcatlán. The stews and broths vary and may include meat or vegetables acquired in the local stores, produced in home gardens, or gathered in agricultural fields or wild vegetation. Other foodstuffs available the yearround like eggs, rice, pasta soup, and chicken soup are consumed in most households at least once a week. Because of their nutritional contribution to the local people’s diet, these fodds can be considered as part of the basic local diet (Appendix 1). The most frequent dishes served in traditional festivities include barbacoa (beef cooked in earth ovens) and tesmoles (chili-based sauces generally served with chicken). For the Todos los Santos (Day of the Dead) festivity, totopos – a special kind of dry tortilla – are always present in the altars set for the symbolic consumption by the deceased people, together with the pan de muerto (a kind of bread) prepared by community members (Fig. 5). Festive foodstuffs have a cultural significance as part of local tradition but are also important because they are shared during celebrations to acknowledge the community members that collaborated in their preparation. Participation in religious festivities and giving support to mayordomos (community members in charge of organizing activities of religious celebrations) allow members of less prosperous households, unable to purchase meat, to have access to a significant source of animal protein. The religious ceremonies take place at least once a month, and every one or two weeks, the mayordomos of the community’s patron saint festivity must organize and provide supplies for celebration events called calendas (Hironymous 2007). Game animals are another supplement of the community members’ diet. The consumption of deer (Odocoileus virginianus Zimmermann, 1780) and collared peccaries (Dicotyles tajacu Linnaeus, 1789) has been part of the diet of people in the region since about 10,000 years ago and it continues to be practiced. However, since the designation of the Biosphere Reserve in 2012, the conservation policies promoted by the Reserve’s administrative authorities have diminished the consumption of these species and other wild mammals and insects (Cook 1958; Flannery 1986; Solís and Casas 2019; Velazquez de Lara 1579; Zarazúa-Carbajal et al. 2020). Other wild animals and their products continue to be consumed by community members, or at least they acknowledge their consumption. These are the cases of D. tajacu, small wild mammals – mainly squirrels and rabbits (Sylvilagus spp.; Lepus spp.; Spermophilus variegatus Erxleben, 1777; Sciurus aureogaster F. Cuvier, 1929), pigeons (Zenaida spp.), and insects like Eucheira socialis Westwood, 1834 (madrone worms), honey and bee larvae, especially of Apis mellifera Linnaeus, 1758, and of Brachygastra spp. (panal de tierra; Appendix 1). The consumption of wild animals is occasional and depends on when activities conducted in the field provide the opportunity to hunt or gather them. The exception is gathering of honeycombs of A. mellifera, since it is carried out during expeditions to the warm zones for obtaining copal, which together with honey forms part of the Day of the Death offerings. Other exception is the hunting of deer and collared peccaries. These activities require specialized knowledge about the environment and about the animal

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

569

species, and are carried out exclusively by males, except from some households that have recently allowed the participation of young women. Among the gathered wild plants, the most outstanding are quelites, called nyiya in Ixcatec, which are edible plants whose tender leaves are consumed boiled or fried (Costaouec and Swanton 2015; Rangel-Landa et al. 2016b) including Amaranthus hybridus (quelite tintonil, nyiya xacújù), Chenopodium berlandieri Moq. (quelite de manteca, nyiyaxije), and Anoda cristata (L.) Schltdl. (violeta, nyiya ñundu). Of these three species of quelite, A. hybridus was consumed by members of all the households interviewed in 2000, and in 95% of the households enquired in 2012. Local people said to consume it from one to five times a week during the season when it is available (Appendix 1). The consumption of C. berlandieri has decreased drastically. In 2000, it was eaten by members of 30% of the surveyed households, and, in 2012, the members of only 15% of the surveyed households said to consume it. The use and knowledge of A. cristata is being lost. In 2012, only the members of one household said to have consumed it in the past 10 years. At present, only the elders that consumed A. cristata 20 years ago could relate its consistency with the Ixcatec name, whose epithet means slimy, referring to the consistency of broths and stews prepared with it. Like other plant species consumed as quelites having nutraceutical, nutritional, and antimicrobial properties, A. cristata is a present and future valuable resource (Gomez-Chang et al. 2018; Mateos-Maces et al. 2020). However, the memory of the interaction of local people with A. cristata is at risk of being lost because of changes in diet and the forgetting of the name of this plant in Ixcatec, which codifies the knowledge associated with plant use. The availability of edible plants is closely associated with practices carried out to ensure their availability for use and improve their quality. Native plant species managed in home gardens and cropland include several species of Opuntia spp. (nopales or ñunda in Ixcatec), A. hybridus, Porophyllum linaria (Cav.) DC. (pepitza or xaxcunyà), P. ruderale var. macrocephalum (DC.) Cronquistand (paploquelite or myeni), and Leucaena spp. (guaje or nyatsje), which are important components of the household members’ diet (Fig. 5). Although the local availability of these species lasts for only one to four months, these are consumed once per month to three times per week because people procure them from markets or stores outside Ixcatlán (Appendix 1). Other plants having short production seasons like Sideroxylon palmeri (Rose) T.D.Pennm.(tempesquisle or chixu), whose edible fruits are valued for their use in religious ceremonies and require a laborious preparation, are commonly obtained through interchange with sellers coming from outside the community. The inhabitants of Ixcatlán use 15 plants to flavor food or beverages (Table 1), some of which provide a substantial supply of nutrients. These condiments include some used for dishes prepared for celebrations, like Piper auritum Kunth used for tamales and Origanum majorana L. that confers mole a distinctive taste. A. hybridus must be cooked with P. linaria (Cav.) DC (pepitsa or xaxcunyà). The leaves of A. salmiana subsp. tehuacanensis (maguey cimarron) confer to the barbacoa a particular flavor. In addition, the cuticle of these leaves function as a utensil for wrapping the meat, thus preventing it from burning, dehydrating, or coming into contact with the hot stones of the oven (Fig. 5).

570

S. Rangel-Landa and A. Casas

Fig. 5 Uses of plants for daily life and celebrations in Santa María Ixcatlán. (a) Altar during the celebration of the Day of the Dead in which Laelia anceps, Tagetes erecta, T. lunulata Ortega, and copal resin (Bursera spp.) are offered to deceased relatives; (b) Tenates de flor with candles that will be offered to be lit on deceased family graves or given to relatives and friends for their deceased; (c) lit trunk in front of the church on the nights of the Day of Death celebration; (d) leaves of A. salmiana subsp. tehuacanensis in the oven for barbacoa cooking; (e) bunch of palm leaves Brahea dulcis adorned with flowers Prosthechea karwinskii (Mart.) J.M.H. Shaw

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

571

The relationship between people’s cultural values and the perception of risk in the availability of edible plants motivate management practices. These practices are directed to ensure the availability of products to satisfy their needs, as was documented by Rangel-Landa et al. (2017) for the cases of P. philadelphica, A. hybridus, and D. ambrosioides. Having useful plants nearby “tenerlas a mano” and ensuring their quality are factors that drive people’s management practices like cultivation. For example, the Ixcatec inhabitants perceive that P. linaria is abundant, but since it is necessary to collect the plant from crop fields far from the house, they cultivate them in their homegardens to ease their availability (Fig. 5). They also cultivate Solanum lycopersicum L. ( jitomate), Opuntia spp., and Coriandrum sativum L. because they consider that their cultivated plants have better flavor and are healthier than plants purchased in stores. In the period between 2000 and 2012, we identified a trend of change in the local people’s food patterns expressed as the decrease of the average weekly consumption per household of maize (from 14.7  8.3 to 13.5  6.3 kg) and bean (3  1.7 to 2.2  1.9 kg), and an associated increase in the consumption of eggs and meat. Our observations suggest that households are changing their food patterns, although that trend occurs at different rates among households. Emergency foodstuffs formerly used in times when maize became scarce remain in the collective memory of the current inhabitants of Santa María Ixcatlán. The most important plants used in the past as emergency foodstuffs are agaves. The cooked scapes, flower buds, and stems of agave were mixed with cooked maize to prepare tortillas. The boiled and cooked flower buds of A. potatorum and A. kerchovei Lem, the stems of A. potatorum prepared with leaves of Oxalis aff. latifolia Kunth, and agave stems cooked in underground ovens were appreciated energy-rich foodstuffs in the past. The consumption of agave was recorded in the archaeological studies in the region from at least 9000 years ago (MacNeish 1967), and for Ixcatlán, it was described in written records of the sixteenth century (Velazquez de Lara 1579; Cook 1958). However, the use of agave as food has decreased during the last 30 years, partly because of the use of A. potatorum for producing mescal, and partly because harvesting flower buds is recently seen as a practice affecting the conservation of agave populations. Social factors also appear to have played a role in the decreased consumption of agave products, as we observed in the Ixcatlán inhabitants associating that habit with poverty. Something similar could have happened with the consumption of the flowers of Opuntia spp., which was recorded in the sixteenth century (Velazquez de Lara 1579), but that we failed to observe in our study.

ä Fig. 5 (continued) (monjita amarilla), which will be blessed in the celebration of Palm Sunday; (f) Leucaena esculenta pods and Porophyllum linaria herbs; (g) construction of a kitchen roof with B. calcarea leaves; (h) P. ruderale bouquet to accompany the food during the celebration of the Señor de las Tres Caídas; and (i) transport of Tillandsia spp. (Soluche) collected in oak forests to feed pack animals in the dry season. (Photos: Selene Rangel-Landa)

572

S. Rangel-Landa and A. Casas

Other plants like Peperomia quadrifolia (L.) Kunth – an epiphyte distributed in oak forests in the highest elevations of the territory – and the tender inflorescences of Dasylirion serratifolium (Karw. ex Schult. & Schult.f.) Zucc. Hook. are highly valued by the inhabitants of Santa María Ixcatlán for their flavor, and because they recognize them as healthy foodstuffs. Both species are consumed by most households in Santa María Ixcatlán at least once or a few times per year, even by household members that cannot gather them. In the latter cases, these products are shared by relatives, neighbors, or friends, which play a role in maintaining reciprocal relationships for strengthening the bonds among community members (RangelLanda et al. 2017).

Plants and Healthcare Before the establishment in Ixcatlán of a public health clinic of the Mexican Institute for Social Security (IMSS for its acronym in Spanish) in 1990, the community members attended their health issues mainly through traditional medicine practices (Hironymous 2007). Since then, a sort of hybrid healthcare system has operated in Ixcatlán. Currently, practically all household members make use of public health services within or outside the community. But, depending on the type of illness, the household’s economic resources, and the household members’ trust in each type of treatment, they also use traditional medicine to complement physician’s prescribed treatments, or when they consider the latter to be ineffective (Rangel-Landa et al. 2017). The most common health issues reported by the inhabitants of Santa María Ixcatlán are cold symptoms, stomachache, fever, headache, earache, traumatisms, and culture-specific illnesses like empachos (stomach inflammation, loss of appetite, and constipation), sustos and aires (general discomfort caused by strong impressions, envies, or having had contact with ill or deceased persons), and alferecía affecting children (irritability, loss of appetite, and weakness). Cardiovascular diseases, obesity, and diabetes have been increasingly manifested among the current inhabitants of Santa María Ixcatlán, which could be due in part to changes in their exercise and dietary habits during recent years. For example, the increasing consumption of sweet industrial beverages (refrescos) – one to seven times a week or twice a month – was practiced by 29% of the household members in 2000, and by 81% in 2012. Within households, the older women are the most experienced and knowledgeable of traditional healthcare treatments, and – both in cases of culture-specific illnesses or when supplementing physician’s prescriptions – they decide when they are able to treat the ill household members, or when to consult a healer. These experts are in charge of the traditional medical practice, make cleanings, prescribe specialized remedies, therapeutic massages (sobas), and steam house baths (baño de temazcal). In Ixcatlán, we recorded 219 species of medicinal plants (Table 1, Fig. 4). Among the plants recognized as basic for life the Ixcatec indicated the red oak species Quercus acutifolia and Q. conspersa, Casimiroa edulis, Ageratina mairetiana (DC.)

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

573

R.M.King and H.Rob, and Clinopodium mexicanum (Benth.) Govaerts. All these plants are used in temascal baths for recovering women after childbirth. Except for C. edulis, that is cultivated, all the medicinal plant species we recorded were wild and are gathered from forests when needed (Rangel-Landa et al. 2016a, 2017). The inhabitants of Santa María Ixcatlán maintain some of the plants they consider basic in home gardens (65 species) and crop fields (79 species). The forms of management of medicinal plants can be cultivation, like in the cases of Matricaria chamomilla L., Mentha  piperita L., Aloe vera (L.) Burm., and A. potatorum. Also, tolerance, enhancing protection, or removal for controlling their abundance, as in the cases of Marrubium vulgare L., Malva parviflora L., and Ricinus communis A. Gray. Other wild plants like Grindelia inuloides Willd., Lippia oaxacana B.L.Rob. and Greenm., Tagetes lucida Cav., Turnera diffusa Willd. ex Schult., Artemisia ludoviciana Nutt., and Chrysactinia mexicana A. Gray are gathered, dehydrated, and stored to use them when needed. Although some people have tried to cultivate the abovementioned medicinal plants in homegardens, gathering is the main practice to obtain them, and some plants like T. lucida are difficult to cultivate. In general, people consider that when these wild plants are cultivated could be less effective than plants gathered from the wild because they lose their scent when they are removed from the mountain (“pierden su aroma al sacarse del monte”).

Ceremonial Plants Very little is known about the religious beliefs and practices of the Ixcatec people before the arrival of the Spanish colonizers, but the available information underscores the interaction between plants and their ritual life. Velazquez de Lara (1579) mentioned that during one of the four main celebrations, called Malinaltzi, which was dedicated to fire, the most important offering was composed of plants. In this celebration a fire was lit with wood of several trees and incensed with copal. The main deities were two, Ocelotl and Acatl (meaning ocelot and reed in Nahuatl, respectively), whose representations were displayed in the temples to be seen by people who gave offerings consisting in flowers. Santa María Ixcatlán is currently considered a Catholic community, but the use of offerings consisting of flowers and copal incense, and the lighting of ritual fires using tree trunks, continues to form part of local religious ceremonies and rites. Currently, during the Day of Dead celebration, a ritual is carried out in which young men go to the forest surrounding the town to cut oak logs (Quercus spp.), drag them to town, carry them to the church through the main streets, and gather all the trunks together to light a big fire to warm the dead (“calentar a los muertos”; Fig. 5). As in the sixteenth century, people who die are shod with ceremonial sandals elaborated with leaves of the palm Brahea dulcis, which are blessed during the Palm Sunday. Blessing of palms is associated with the catholic rite representing the entering of Jesus to Jerusalem, but also with the pre-Hispanic rite of putting shoes to the deceased persons for their walk away from life. In addition, the ritual is related

574

S. Rangel-Landa and A. Casas

to a procession during the dry season asking God to continue making available the palm and other products obtained from the mountain. Offering flowers has been an ongoing practice in all religious ceremonies and funerals in Santa María Ixcatlán, as a way to demonstrate devotion and thanking the saints whose images people place in their home altars and in the church. We recorded 128 plant species used in ceremonies and rites in Santa María Ixcatlán, 43 of them exotic and cultivated in home gardens (Table 1). Among the native species, Santa María Ixcatlán inhabitants also cultivate the orchids Laelia albida Bateman ex Lindl. and L. anceps Lindl. because of their beauty and to be offered in home and church altars during the celebration of the Day of the Dead (Rangel-Landa et al. 2016a, 2017). Another plant cultivated for the latter celebration is Tagetes erecta L. (cempasuchil or tsjucájà), which, together with copal, is essential (Fig. 5). Maintaining home altars is a main motivation for cultivating many of the plants found in home gardens. The mayordomos make a commitment of continuously providing fresh flowers to maintain the ornamentation of church altars, a commitment they achieve involving a complex network of social relations and organization. Also, through the use of other local resources to generate the required income for having fresh flowers for the saints of the community. Bursera spp. resin (copal or ska in Ixcatec) is highly appreciated as an offering for the saints and dead people, and to clean the environment and the body of negative feelings and culture-specific illnesses like aires or sustos. Management and care of the species of the genus Bursera from which this resin is obtained, mainly Bursera biflora (Rose) Standl. and B. fagaroides (Kunth) Engl., have the purpose to protect these plants and ensure the quality of the aroma and amount of resin they produce. The common practice in other copal producing regions is to make incisions in the trunk and branches of trees to promote resin exudation. But this practice is uncommon in Santa María Ixcatlán, since people consider it as bad because it damages the trees. Cultivation of copal trees is not practiced in Santa María Ixcatlán since people consider that the only good resin is that which is naturally produced by trees after being perforated by lepidopteran larvae (Rangel-Landa et al. 2017; Blancas et al. 2022). Other plants highly valued for their use in ceremonies and whose management involves carefully avoiding overharvesting are Litsea glaucescens and Beaucarnea stricta Lem., from which gatherers extract only some branches, and Dasylirion serratifolium (Karw. ex Schult. & Schult.f.) Zucc. and P. grandis, from whose populations the gatherers extract a few individual plants to obtain their leaves that are used for ornamenting the celebrations.

Ornamental and Luxury Plants We identified 270 plant species considered by Ixcatlán inhabitants as plants that have the purpose of enhancing the homes’ and home gardens’ appearance and feeling of comfort. These plants give satisfaction and pride to household members, and their role is described by people as “for adding luxury” (para dar lujo). Some of these plants are in the forest and beautify the sites where they grow, while some grow in

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

575

the village and people appreciate them because of their beauty (Rangel-Landa et al. 2016a, 2017). Of these plant species, 190 are considered as house luxury plants (lujo de las casas), 99 of them are native, like Agave potatorum and Plumeria rubra L, and some of them are cultivated since a long time ago, like the orchids Laelia albida and L. anceps (Fig. 5). Other 139 plant species are perceived to adorn the landscape. People of Santa María Ixcatlán describe them as elements of “forest luxury” (lujo de monte), as are animals such as deer, or sites like springs. These plants include dominant forest trees, plants with beautiful flowers, species of grass and herbs covering the ground, and some like Quercus spp. and Tillandsia spp. are occasionally transplanted to home gardens. The concept of luxury varies among households’ members, but those people having more ornamental plants in their homegardens are recognized as a motive of pride. The Ixcatlán inhabitants’ love for plants and their beauty is linked with their ceremonial life, so it is not surprising that they use 58 plant species considered to be beautiful both for adorning their houses and to be offered to the saints.

The Feeding and Care of Domestic Animals The inhabitants of Santa María Ixcatlán refer to livestock as animals (animales) and they recognize and name to use over 260 plant species for feeding them. Nearly 25% of the community households raise cattle, goats, and sheep, and 85% own backyard livestock including donkeys, mules, horses, poultry, and pigs (Table 1). Cattle mostly obtain their feed from herbs and the young branches, pods, and acorns of wild plant species in zones dominated by oak forest and the transition of this vegetation type with tropical dry forest. Sheep and goats are herded by shepherds in areas close to the town, especially in scrubland dominated by B. dulcis, thorn scrub, and oak forests. Shepherds sometimes drive large goat herds away from the settlement for several days or weeks. Donkeys, mules, and horses consume a broad spectrum of plants, but their main sources of fodder are maize and the leaves and stalks from maize, wheat, beans, oat, and barley crops, which is supplemented by herbs managed in backyards, crop fields, and homegardens. This fodder stock is only enough for 6 months of the year. People’s need of ensuring fodder for their animals helps to understand why it is important to maintain a high diversity of herbs, shrubs, and trees in agroforestry systems, and why maize is cultivated even though the need for human food is not met (Moreno-Calles et al. 2012, 2013; Vallejo et al. 2014). In the homegardens and crop fields of Ixcatlán, we recorded 114 plant species used as fodder, nearly 20 of which are agrestal plant species that the cultivators remove from inside the crop field but allow their growth in its edges. Some plant species recognized for its good quality as fodder are tolerated, such as Tithonia spp., Viguiera spp., Bidens spp., Perymenium spp., Mirabilis jalapa, Malva parvifolia, A. hybridus, and A. salmina subsp. tehuacanensis. Livestock owners in Ixcatlán consider some wild species as good sources of fodder; these are for instance the cases of the epiphytes Tillandsia gymnobotrya

576

S. Rangel-Landa and A. Casas

Baker, T. recurvata (L.) L., and T. usneoides (L.) L., and Hechtia oaxacana BurtUtley, Utley and García-Mend., and P. grandis (Fig. 5). During fodder shortage periods, when leaves and stalks of crops are scarce (from March to June), backyard livestock owners gather the abovementioned plants and acorns from the forests, which is the only economically viable source of fodder for 25% of the households. The need for forest fodder plant species contributes to explain why Ixcatlán inhabitants assign high value to oaks, epiphytes, and the forests where they grow. Livestock owners in Ixcatlán value six plants used for the healthcare of their animals (Rangel-Landa et al. 2016a). The most frequently cause for applying treatment to livestock (and dogs) is snake bite.

Firewood People of Santa María Ixcatlán use 48 plant species as fuelwood. Although gas stoves are increasingly being used for cooking, their use is limited because gas is difficult to in this remote rural area and expensive. Therefore, most households in Santa María Ixcatlán use firewood for making tortillas and cooking their meals. The most appreciated firewood is produced by the 11 species of oak growing in the community because it burns for a longer time and produces less smoke. The stems and branches of shrubs (called varejón in Spanish) are used for igniting the fire. We identified a decrease in the consumption of firewood for cooking from 143.4  11.3 in 2000 to 18.8  12 kg per week in 2012, which people attribute to the recent use of firewood-saving stoves. The communal statute of Santa María Ixcatlán allows community members to extract firewood for cooking and other productive activities, and to trade it within the community. Commercialization of firewood outside the community or at regional level is prohibited. The statute contains the practices that community members must follow for firewood extraction, which include only cutting dead trees or branches. When green firewood or firewood from live trees are required, only cutting the branches is allowed (C. Santa María Ixcatlán 2009). Despite the sale of firewood outside the community remains under control, the number of community members having access to chainsaws and trucks has recently increased. These tools make harvesting easier, and allow their users attending the growing demand for oak firewood for the production of mescal, which increased from 16.2 t in 2000 to 63.36 t per year in 2012. Although not recently evaluated, it is possible to say that the current increase of mescal production has determined an even more drastic increase of firewood extraction. Some Ixcatlán community members involved in mescal production are experimenting with cutting techniques allowing regeneration of oak trees after harvesting firewood with chainsaws and promote the follow-up of practices to ensure tree survival and regeneration (Fig. 6). However, some mescal producers and other community members do not follow the firewood extraction practices prescribed in the community’s statute, and the assembly has not developed yet appropriate control mechanisms (Rangel-Landa et al. 2016a).

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

577

Fig. 6 Innovation for facing the challenges involved in the palm and maguey trades in Santa María Ixcatlán. (a) Agave potatorum stand generated by transplanting some individuals and seed spreading; (b) production of Agave potatorum seedlings in backyards by mescal producers; (c) green firewood cutting technique that favors the regeneration of branches; and (d) palm handicrafts with contemporary design made by members of Xula Palma Artesanal. (Photos: Selene Rangel-Landa)

Housing In the last 30 years, the houses in Ixcatlán have changed substantially, which contrasts with the few changes in house construction observed in studies made between the mid-twentieth century and the early twentyfirst century (Cook 1958; Hironymous 2007). At present, the houses in the community range between those preserving all the architectural elements and materials of traditional housing and recently built houses following modern architectural styles. The latter include brick walls, concrete roofs, the kitchen, rooms, and bathroom constructed within the same building. Most houses combine both extremes of techniques, in which the household members aim at preserving the functionality of traditional houses but incorporating the resistance and convenience of modern materials and construction practices.

578

S. Rangel-Landa and A. Casas

In traditional houses, the kitchen, rooms, latrine, storerooms, granaries, corrals, and other facilities are in separate buildings (Fig. 5). Most houses preserve the traditional kitchen built with a wood structure (Quercus sp. and Juniperus flaccida), palm (Brahea calcarea Liebm. or B. dulcis) or clay tile roof, walls made of trunks (B. dulcis) or scapes (Agave scaposa Gentry or A. salmiana subsp. tehuacanensis), and earthen floor (Fig. 5). In some houses built using modern architecture and materials, the traditional kitchens are maintained because they provide better ventilation for cooking with firewood, or to cook large amounts of food. In traditional houses, rooms are in a building separated from the kitchen, usually in numbers going from one to three. One of the rooms has a space for the home altar where saints are venerated. These rooms are used to store valuable goods such as seed to be sown in the next agricultural cycle and is where guests are received and entertained. In most houses, room buildings have tin or clay tile roofs and walls are built with limestone blocks extracted within the community’s territory. A decreasing number of houses maintain room buildings built using trunks and Agave scapes, and in many of these households their members wish they could rebuild them with longlasting materials. The other buildings of traditional houses, all within the household’s premises (solar), are the latrine, a dugout cave or cellar for storing harvested palm leaves and for weaving them, a granary (troja) for maize, a small storeroom, and corrals for livestock. There is a trend in Santa María Ixcatlán to build the latrines closer to the rooms. Wood from tree trunks and branches, flowering stalks (scapes), and fiber from 29 plant species provide materials to the inhabitants of Santa María Ixcatlán for constructing the abovementioned house buildings and spaces, and for making doors (Rangel-Landa et al. 2016a). Live fences on the limits of household’s premises include 30 species of trees and shrubs, including Lophocereus marginatus, Condalia mexicana Schltdl., Opuntia spp., and Agave spp. (Rangel-Landa et al. 2016a). A zone close to the kitchen serves as a space to do laundry, wash dishes, and grow plants to take advantage of gray water produced in domestic activities. In the openair spaces used by household members for domestic activities and livestock management, we recorded 20 tree and shrub species that provide shade, which improves the environment and allows growing shade plants. Celtis caudata Planch. and Morus celtidifolia Kunth. are the most common trees in household premises in Santa María Ixcatlán. The open places in the household premises are versatile, during celebrations serving as gathering areas, and during favorable rainy seasons as crop fields. In the mid-twentieth century, Cook (1958) observed that nearly half of the households in Ixcatlán sowed maize in homegardens (Fig. 3). At present, we recorded that milpa (the multicrop system of maize, bean, and squash) is present in the homegardens of nearly 25% of the households in the community. Mostly in households whose members are elders or women who are unable of cultivating a field outside of the town because of the time, labor, or physical effort that this practice would require.

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

579

Utensils and Other Uses We recorded 34 plant species used as domestic utensils or tools (Rangel-Landa et al. 2016a). Oak (Quercus spp.) wood is appreciated by Ixcatlán inhabitants for its hardness and durability, in particular for making agricultural tools and the mallets used in mescal production (Fig. 4). The leaves of Malacomeles denticulata (Kunth) Decne. Forestiera rotundifolia (Brandegee) Standl., locally called tlasisle, are used for lining steam-cooked foodstuffs like tamales to avoid water coming into contact with them. Some plants have infrequent uses associated with specialized activities like the treatment of bovine, ovine, and caprine skins, and for repelling insects and poisoning wild animals that are harmful to livestock (Table 1). For example, Aralia humilis Cav. is appreciated by honey gatherers for producing a smoke that tranquilizes and repels bees. Among the useful plants we recorded in Ixcatlán, 14 species have uses (ludic, cosmetic, soap, dyes, or glue) that are becoming less frequent or have been abandoned. Mostly because currently people have access to products to replace them. For example, the resin of Bursera spp. was used as a glue for repairing pots, the fruits of Comarostaphylis polifolia (Kunth) Zucc. ex Klotzsch was used for dyeing, and Lysimachia arvensis (L.) U.Manns and Anderb. and Cucurbita pedatifolia L.H. Bailey were used for personal care and as soap for washing clothes.

The Palm: From Mats, Baskets, and Hats to Artisanal Contemporaneous Designs Of the 10 plant species used for manufacturing objects that the Ixcatec consider as handicrafts, palms B. dulcis, B. calcarea, and its hybrid B. dulcis x B. calcarea (palma or tjen;) are the most important (Table 1; Rangel-Landa et al. 2014). In the region, there is archaeological evidence of human use of B. dulcis dating to 11,959 years ago (Smith 1967). During the sixteenth century, weaving of palm mats (petates) and palm baskets (tenates) were among the main economic activities in Ixcatlán, and palm mats were paid as tribute and exchanged for gold, which was also a part of the tribute paid by the community (Velazquez de Lara 1579). No record exists about the time when hat weaving started in Ixcatlán, but based on eighteenth century documents, Cook (1958) estimated the activity to have been practiced since approximately 250 years ago. Hat weaving is an essential activity within the survival strategy of the households in Santa María Ixcatlán. Although the income obtained per woven hat is low (US$ 0.17 in 2000, US$0.23 in 2012, and US$0.20 in 2022), the activity has represented a reliable and constant income for households in Santa María Ixcatlán. Except for during extraordinary eventualities like the COVID-19 pandemic, this activity allows bartering palm products for maize and other groceries in local stores for satisfying the daily needs or sell them for money (Cook 1958; Rangel-Landa et al. 2014).

580

S. Rangel-Landa and A. Casas

Compared with palm hats, palm basket weaving provides very little income to the community of Ixcatlán but is culturally important for its presence in everyday life as a container of tortillas, bread, bean, maize, and other things. Decorated palm baskets, called flower palm baskets (tenates de flor), express a colorful representation of the local nature and costumes of Ixcatlán (Fig. 5). These are made by a few palm weavers recognized by the community as artisans because of their skill and ingenuity to create elaborate designs going beyond simple hat weaving. The tenates de flor are culturally important; there is a tradition of weaving and using them for the first time to carry the candles shared by the inhabitants of Ixcatlán during their visits to cemeteries in the Day of the Dead (Fig. 5). Also considered as handicrafts, some palm weavers create purses, small boxes, covers for mescal bottles, and animal figures traditionally used in the community of Ixcatlán as toys or pendants. Palm baskets and the abovementioned handicrafts are normally sold among community members, relatives, and visitors. These products have a higher economic value than hats, and provide a higher profit for the weavers relative to hat weaving because they use less palm, and obtain a higher profit with less invested time. However, when comparing the price of these handicrafts in markets outside Ixcatlán, the local prices continue to be low and do not compensate for the work invested in their manufacture, nor the impact of the extraction of the raw material. The palm hat weavers in Ixcatlán are not worried about having enough supply of palm leaves because they have a reserve of over 100 ha where they obtain the raw matter. However, they are concerned about events of palm leaf scarcity with repeated years of drought, as they remember it happened in the 1960s. Instead, artisans require longer and stronger fibers than those used by palm hat weavers, the latter being scarce in palmonares near the town. Palm weavers in Santa María Ixcatlán attributed the scarcity of long palm leaves near the town to the continued harvest by the community members, because of which artisans solve the problem by going to farther away areas where the rate of leaf extraction is lower. The main concerns of palm hat weavers in Ixcatlán are focused on their low income and the possibility of palm hat wholesalers losing their interest in acquiring their production. During the COVID-19 pandemic, the large hat stores in Tehuacán, Puebla, stopped buying the palm hats woven in the region, because the distribution centers where palm hats are processed – boiled, molded, and adorned – and sold are located in tourism facilities that were closed. Since 2014, community’s palm hat weavers in association with the nonprofit Weaving Alliances Civil Association (Tejiendo Alianzas A.C.) began a project to organize, innovate, and promote the local palm weaving trade, which gave rise to the Xula Palma Artesanal collective. The Xula collective is currently integrated by five families of palm weavers and traders of palm handicrafts with contemporary design who have found an opportunity to be recognized as artisans, both within and beyond the community (Fig. 6). This fact provides them a higher and more reliable income from their work, in comparison with palm hat weaving and other palm handicrafts (Hanson 2022).

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

581

The abundance of B. dulcis leaves, on which palm weavers in Ixcatlán depend, is the product of landscape management. It goes from carefully harvesting the leaves during the full moon, avoiding damaging the apical meristem (cogollo), tolerating palms in the edges of houses or maize fields, and promoting and protecting the palms (Rangel-Landa et al. 2014). Wild palm stand management has a long history that led to the formation of a scrubland dominated by palms (Rzedowski 1978; ValienteBanuet et al. 2009). This vegetation type is known by the Ixcatec as palmonar, which covers the landscapes surrounding the town of Santa María Ixcatlán, and the areas of abandoned settlements like San Juan Viejo, where palm leaves harvest continues albeit at a low intensity (Figs. 2 and 3). Palmonares are landscape units in which people’s management involves the interaction of the harvest of palms and other 104 species and where agricultural fields and small livestock herding are also included (see chapter ▶ “Agroforestry Complexes in the Mountain Regions of Mexico”). The management of palm-dominated scrubland allows the continuity between anthropic landscapes and forests and maintains several essential ecosystem services and benefits for the landscape managers. However, it also causes a loss of biodiversity and processes like soil erosion which limit agricultural productivity (Cook 1958; Rangel-Landa et al. 2014; Vallejo et al. 2014; Moreno-Calles et al. chapter ▶ “Agroforestry Complexes in the Mountain Regions of Mexico” Fig. 3).

Maguey as Sources of Sacred and Spirituous Drinks Agave species have been sources of food and fiber for the inhabitants of the Tehuacán-Cuicatlán Valley region since at least 12,000–14,000 years ago (MacNeish 1967). In the sixteenth century, the aguamiel producing magueys were abundant, providing a drink rich in sugar, minerals, and probiotics during drought periods, and a commodity sold or bartered to pay tribute (Velazquez de Lara 1579). Currently, we identified 11 species of the genus Agave (maguey or tsu), of which A. americana L. provides aguamiel. Some inhabitants of Ixcatlán remembered that, over 50 years ago, aguamiel was also extracted from A. applanata K.Koch and A. salmiana subsp. tehuacanensis, and that the cultivation of A. americana had been decreasing in the past two decades. A. americana was recorded in ten of the 21 homegardens sampled in this study; in eight of them the owners hoped to obtain pulque and only in 1 we recorded more than one variety of A. americana and individuals of different size. Despite the current scarcity of A. americana in Santa María Ixcatlán, its inhabitants prepare a traditional drink called tepache during the festivities in December. This beverage is prepared with unrefined sugar (panela), and pulque acquired from neighboring communities. The decrease in pulque consumption by Santa María Ixcatlán inhabitants might be due to the present availability of drinking water and beer, the consumption of which they associate with a higher economic and social status relative to drinking pulque. The availability of beer in Santa María Ixcatlán could have started in the 1980s, when local store owners acquired their first motor vehicles, and it intensified with the inauguration of the Cuacnopalán-Oaxaca

582

S. Rangel-Landa and A. Casas

highway in 1994, as it happened in other regions of the country (Ramírez-Rodríguez 2018). Another agave-derived drink with biocultural importance for the Ixcatec people is mescal (ndamasalíyè), and according to the collective memory of Ixcatlán, its production dates back to at least generations. In the mid-twentieth century, Cook (1958) described that the inhabitants of Santa María Ixcatlán elaborated and consumed mescal in ceremonies. In Santa María Ixcatlán, mescal has been produced following a tradition of materials and practices – which gives it a distinguishing quality. It is mostly prepared with A. potatorum but occasionally also supplemented with wild individuals of A. angustifolia Haw. (Rangel-Landa et al. 2016b; AlvaradoÁlvarez and Toolan n.d.). The growth of the international trade of agave distilled beverages in recent years has boosted the economic role of mescal elaboration in Santa María Ixcatlán, its value incrementing by nearly tenfold in the past 20 years (US$ 2.5 in 2000, US$ 6 in 2011, US$ 9 in 2015, and US$ 22.5 in 2022; RangelLanda et al. 2016a). Most of the local mescal production is retailed among households and visitors within Ixcatlán and sold to local and neighboring towns’ stores. However, since 2015, most of the mescal produced in Ixcatlán has been traded in regional and national markets. Since 2018, a Santa María Ixcatlán mescal producer gained access to the international markets through registration of the Ixcateco trade name. This was achieved in collaboration with wholesale distributors of artisanal spirits that promote the environmental responsibility of producers and consumers, shortening the distribution chains, and increasing the income of producers. The producers currently export their product to the USA, the UK, and France (Alvarado-Álvarez and Toolan n.d.). The increase in price of mescal has led to a larger demand of the inputs needed for its production in Santa María Ixcatlán, in particular of firewood as mentioned above, and of maguey stems. For 2011, we estimated that the whole community produced 192 mescal batches using 91.14  9.78 agaves per batch, which means a total of nearly 17,500 agave plants per year, while for the year 2000 we estimated the use of 4900 individuals, that is, nearly a fourfold increment in 9 years (Rangel-Landa et al. 2016a). By 2000, mescal producers in Santa María Ixcatlán and the community assembly expressed their concern about the scarcity of the agave plants necessary for their trade, and made the first attempts to find ways of management that improved the availability of the input, but without encouraging results. It was until 2012 when several mescal producers in Ixcatlán began to exchange experiences with other communities that propagated the agave plants in nurseries. As a consequence, they began propagating the plants in their homegardens and also increased their participation in institutional programs aimed at the management of agave plants (Fig. 6). However, unlike the Xula collective, the organization of mescal production in Ixcatlán is still incipient because facing the challenges implied in ordering the activity requires the participation of the whole community. Compared to palm weaving, the artisanal spirit trade is subjected to stronger constraints from market interests and foreign institution regulations (Álvarez and Laird 2021). The lack of appropriate legislation and institutions directed to real protection of products such as the Ixcatec mescal allowed that in 2019 wholesale distributors usurped the name of

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

583

the community to give value to their products in the American market. These events violate the rights of the Ixcatec people but have gone unpunished and the marketers covered up by the institutions that regulate the commercialization of mezcal. The conflict has enhanced a process of organization of several producers who, in addition to taking advantage of the collective brand, began to take an interest in the collective management of the agave (MILPA 2020). For the members of Xula and Ixateco, the change in the way they relate to their buyers, the organization, the care of the base resources of their productive activity, and the strengthening of customs and cultural aspects that give identity to their products have been key to have better pay. This indicates the importance of documenting, understanding, and encouraging these processes in ethnobotanical studies, so that the work of ethnobotany may contribute support actions looking for the well-being of communities.

Conclusion The long history of interaction of the Ixcatec people with their environment has conformed a deep biocultural heritage. Nearly 80% of the plant species reported to occur in the area have one or more uses, while 51% of them is managed in one or more ways to take care of them or to maintain or increase their availability and/or quality. Such management has made possible the continuity of the domestication of species that are crucial for the local subsistence, among them the staple crops maize, beans, and squash, as well as Physalis philadelphica, Dysphania ambrosioides, Cosmos bipinnatus, and Tagetes erecta., among other species. In addition, it has favored a complex matrix of landscape units including agroforestry systems like homegardens (solares), crop land immersed in palmar vegetation (Milpa-palmonar of Brahea dulcis), sites where mescal is prepared (palenque), and sites of abandoned settlements, which coexist with the diverse forest types described above (Casas and Parra 2016; Rangel-Landa et al. 2016a; Fig. 3). The Ixcatec ethnobotanical knowledge is possibly the most deeply documented in the Tehuacán-Cuicatlán region (Lira et al. 2009; Blancas et al. 2010, 2013; Casas et al. 2017). This is because all households carry out primary productive activities. However, we identified a process of loss of local knowledge associated to the substitution of local products for others, the loss of knowledge codified in the Ixcatec language, and the high rate of migration, especially the young people. Improving the conditions of interchange though organizational processes, as well as the innovation of productive practices to generate profitable products is probably a viable way to enhance people to remain linked to the community. Initiatives like those of Xula Palma Artesanal collective and the Ixcateco collective brand have provided new opportunities to improve the lives of some households, and have demonstrated to be effective to maintain appropriate ways to manage plants and vegetation and to value the Ixcatec culture. These organizational experiences have promoted the propagation of A. potatorum, in home gardens, reforestation of areas of communal land, and good practices of gathering palm and agave, and initiated

584

S. Rangel-Landa and A. Casas

actions for the vindication of culture and contributing to revitalize the Ixcatec language (Swanton 2022). In both cases mentioned, the maintenance of sustainable productive practices and fair commercialization are great challenges. In all these processes the Ixcatec community has gained the collaboration of external institutions, including civil organizations and scholars, which have contributed to support the local efforts to improve life and maintain the valuable biocultural heritage of the Ixcatec people. For the members of Xula and Ixateco, the change in the way they relate to their buyers, the organization, the care of the base resources of their productive activity, and the strengthening of customs and cultural aspects that give identity to their products have been key to get better pay. This indicates the importance of documenting, as well as understanding these processes in ethnobotanical studies, so that the work of ethnobotany may contribute to support programs and actions directed to the well-being of communities. Acknowledgments The authors thank the community members of Santa María Ixcatlán for their hospitality, allowing our work, and collaborating with us in different projects. We thank Ricardo Lemus Fernández, Erandi Rivera Lozoya, Sandra E. Smith Aguilar, Emmanuel E. González Arevalo, Michael W. Swanton, and Amando Alvarado Álvarez for their collaboration in the projects, field work, and activities that we carried out in Ixcatlán. We acknowledge the institutional support we received from the IIES-UNAM and the Escuela Nacional de Antropología e Historia, and CONACYT through the program Investigadoras e Investigadores por México (project 3856) and the project A1S-14306 (financial support and a postdoctoral fellowship for the first author). We also thank the support given to us by the Alfredo Harp Helú Oaxaca Foundation, the GEF project ID 9380 770 CONABIO-GEF-FAO/RG023 “Manejo y domesticación de agrobiodiversidad en Mesoamérica: Bases para la soberanía alimentaria sustentable,” and PAPIIT, UNAM (project IN206520 and IN224023).

Appendix 1 Basic diet of the Ixcatec people: past and present. Data obtained from surveys conducted in 2000 and 2012 and bibliographical review. * Local availability  Coming from outside the community. Species or food Zea mays L.

Main preparations Tortillas

Temporality Jan–Mar Apr–Jun Jul–Sep Oct–Dec Consumption * * * * 2000: 100% daily, 14.7  8.3 kg/week 2012: 100% daily, 13.5  6.3 kg/week In sixteenth century, maize is mentioned as part of the main foods (Velazquez de Lara 1579)

(continued)

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

Species or food Phaseolus vulgaris L.

Main preparations Soup

Capsicum annuum L.

Fresh or dry chili sauces

Dysphania ambrosioides (L.) Mosyakin and Clemants Zea mays L., Triticum aestivum, oats, and rice Egg

Bean soup condiment

Temporality Jan–Mar Apr–Jun Jul–Sep Oct–Dec Consumption * * * * 2000: 100% daily, 3  1.7 kg/week 2012: 100% daily, 2.2  1.9 kg/week In sixteenth century, beans are mentioned as part of the main foods (Velazquez de Lara 1579) * * * * 2000: 100% daily, green 1.8  1.4 kg/month and dry 0.63  0.51 kg/month 2012: 90% daily, 10% 1–4/week, green 1.6  1.3 kg/month and dry 0.44  0.38 kg/month * * * * 2000, 2012: 100% daily

Atoles

*

*

*

*

Diverse preparations

*

*

*

*

Rice

Dry soup with tomato sauce

*

*

*

*

Pasta

Pasta soup

*

*

*

*

*

*

*

*

Bread

585

2000: ND 2012: 100% 1–7 (3.3  2.2) times/week 2000: 15% daily, 70% 3–5 times/week, 10% 1–2 times/week, 5% 1 time/month 2012: 15% daily, 30% 3–5 times/week, 50% 1–2 times/week, 5% 1 time/month 2000: 100% 1–3 (1.8  0.9) times/week, 1  0.5 kg/week 2012: 100% 1–4 (2.1  1.0) times/week 2000: 100% 1–5 times (2.2  1.1)/week 2012: 95% 1–4 times (2.1  1.1)/week, 5% without consumption 2000: 100% 1–7 (4.6  2.4) times/week 2012: 100% 1–7 (4.8  2.5) times/week

(continued)

586

S. Rangel-Landa and A. Casas

Temporality Jan–Mar Apr–Jun Jul–Sep Oct–Dec Consumption * * * * 2000: 95% 1–7 (5.3  2.5) times/week, 5% without consumption 2012: 90% 1–7 (3.4  2.6) times/week, 10% without consumption Milk * * * * 2000: 30% 0.25–3 (1  1.2) times/week, 70% without consumption 2012: 90% 0.25–7 (2.8  3.3) times/week, 10% without consumption Poultry Soup, mole, * * * * 2000: 5% 3 times/week, and stews 20% 2 times/week, 40% 1 time/week, 35% 1–2 times/month 2012: 10% 3 times/ week, 20% 2 times/ week, 25% 1 time/week, 45% 1–2 times/month Beef, goat, turkey, Baked in * * * * 2000: 55% consume pork earth oven, them as part of the carnitas common diet (1–4 times (fried), mole, to month) and and stews celebrations, 45% only consume them at celebrations 2012: 30% consume them as part of the common diet (1–8 times to month) and celebrations, 70% only consume them at celebrations Soft drink * * * * 2012: 20% daily, 40% 2–5 times/week, 25% once a week or a month, 15% no usual consumption Opuntia With bean  * *  2000: 10% daily, 60% lasiacantha Pfeiff., soup, stews 3–4 times/week, 30% Opuntia tomentosa 1–2 times/week Salm-Dyck, 2012: 30% daily, 20% Opuntia ficus3 times/week, 45% 1–2 indica (L.) Mill. times/week, 5% without consumption Species or food Coffee

Main preparations

(continued)

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

Species or food Leucaena esculenta (Moc. and Sessé ex DC.) Benth., Leucaena leucocephala (Lam.) de Wit Amaranthus hybridus L.

Porophyllum ruderale (Jacq.) Cass.

Roots

Main preparations

Temporality Jan–Mar Apr–Jun Jul–Sep Oct–Dec Consumption   *  2000, 2012: 100% consume them, in the L. esculenta season (Aug–Oct) at least 1 time/week

Boiled and flavored with Porophyllum linaria, with sauce and fried Fresh

Unknown

Wild animals: Roasted, Odocoileus mole virginianus Zimmermann, Dicotyles tajacu Herrera, Sylvilagus spp., Lepus spp., Spermophilus variegatus Erxleben, Sciurus aureogaster F. Cuvier, and Zenaida spp. (Zarazúa-Carbajal et al. 2020)

587

*

*

*

*

*

*

*

*

*

2000: 20% 3–5 times/ week, 10% 2 times/ week, 70% 1 time/week 2012: 15% 3–5 times/ week, 35% 2 times/ week, 50% 1 time/week 2000: 10% 3–7 times/ week, 45% 1–2 times/ week, 10% 1–2 times/ month, 20% 1–2 times/ year, 15% without consumption 2012: 30% 3–7 times/ week, 20% 1–2 times/ week, 10% 1–2 times/ month, 30% 1–2 times/ year, 10% without consumption In the sixteenth century, roots were mentioned as staple food (Velazquez de Lara 1579). But currently, the only tuber or root that is habitually consumed is the potato, which was cultivated until 30 years ago. Roots of wild plants are currently consumed and even known by few people (Rangel-Landa et al. 2016a) 2012: 20% consumed bush meet. Consumption is occasional, from five to less times per year

(continued)

588

Main Species or food preparations Larvae and honey: Roasted Hymenoptera spp., madrone worms Eucheira socialis Westwood, Brachygastra spp., Apis mellifera Linnaeus, meliponini spp., Comadia redtenbacheri Hammerschmidt, Lepidoptera spp. Rhynchosphorus sp., Sphenarium sp. and other unknown species oak worms (Zarazúa-Carbajal et al. 2020)

S. Rangel-Landa and A. Casas

Temporality Jan–Mar Apr–Jun Jul–Sep Oct–Dec Consumption * * * * 2012: 45% hymenoptera larvae and/or honey, 10% Sphenarium sp., 5% Lepidoptera and other larvae. Consumption is occasional, from five to less times per year

References Alvarado-Álvarez A, Toolan M. Ixcateco [Internet]; n.d.. https://www.ixcateco.com/. Accessed May 2022. Álvarez, Laird. Certification of mezcal production: challenges for small-scale producers. Traditional foodways and, voices for biojustice [Internet]; 2021. https://www.peopleandplants.org/mezcal. Accessed May 2022. Berkes F. Sacred ecology: traditional ecological knowledge and resource management. Philadelphia/New York: Taylor & Francis; 1999. Berkes F, Colding J, Folke C. Rediscovery of traditional ecological knowledge as adaptive management. Ecol Appl. 2000;10(5):1251–62. Blancas J, Casas A, Rangel-Landa S, Moreno-Calles A, Torres I, Pérez-Negrón E, et al. Plant management in the Tehuacan-Cuicatlan Valley, Mexico. Econ Bot Springer New York. 2010;64(4):287–302. Blancas J, Casas A, Pérez-Salicrup D, Caballero J, Vega E. Ecological and socio-cultural factors influencing plant management in Náhuatl communities of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013;9:39. Blancas J, Pérez-Salicrup D, Casas A. Evaluando la incertidumbre en la disponibilidad de recursos vegetales. Gaia Sci. 2014;8(2):137–60. Blancas J, Abad-Fitz I, Beltrán-Rodríguez L, Cristians S, Rangel-Landa S, Casas A, et al. Chemistry, biological activities, and uses of copal resin (Bursera spp.) in Mexico. In: Murthy HN, editor. Gums, resins and latexes of plant origin. Cham: Springer; 2022. p. 14. Boege E. El patrimonio biocultural de los pueblos indígenas de México. Instituto Nacional de Antropología e Historia & Comisión Nacional para el Desarrollo de los Pueblos Indígenas: México; 2008. Casas A, Parra F. La domesticación como proceso evolutivo. In: Casas A, Torres-Guevara J, Parra F, editors. Domesticación en el Continente Americano, vol. 1. Molina: Lima Universidad Nacional

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

589

Agraria La Molina del Perú (UNALM) & Universidad Nacional Autónoma de México (UNAM); 2016. p. 133–58. Casas A, Otero-Arnaiz A, Pérez-Negrón E, Valiente-Banuet A. In situ management and domestication of plants in Mesoamerica. Ann Bot. 2007;100(5):1101–15. Casas A, Blancas J, Pérez-Negrón E, Torres I, Vallejo M, Rangel-Landa S, et al. Manejo sustentable de recursos naturales: naturaleza y cultura. In: Arias G, Farfán-Hereida B, Corral JC, Rendón H, Herrera ML, Servin H, et al., editors. Sustentabilidad e Interculturalidad: Paradigmas entre la relación Cultura y Naturaleza. Pátzcuaro: Universidad Intercultural Indígena de Michoacán; 2014. p. 10–9. Casas A, Parra F, Blancas J. Evolution of humans and by humans. In: Albuquerque UP, Medeiros PM, Casas A, editors. Evolutionary ethnobiology. Cham: Springer; 2015. p. 21–36. Casas A, Parra-Rondinel F, Aguirre-Dugua X, Rangel-Landa S, Blancas J, Vallejo M, et al. Manejo y domesticación de plantas en Mesoamérica: una estrategia de investigación y estado del conocimiento sobre los recursos genéticos. In: Casas A, Parra F, Torres-Guevara J, editors. Domesticación en el Continente Americano, vol. 2. Morelia: Universidad Nacional Autónoma de México (UNAM), Universidad Nacional Agraria La Molina del Perú (UNALM); 2017. p. 69–102. CONANP. Programa de manejo de la Reserva de la Biosfera Tehuacán-Cuicatlán. Mexico City: Secretaría de Medio Ambiente y Recursos Naturales SEMARNAT & Comisión Nacional de Áreas Naturales Protegidas CONANP; 2013. Cook SF. In: Ibero-Amer, Sauer CO, Woodrow B, Cook SF, Rowe JH, editors. Santa María Ixcatlán: habitat, population, subsistence. Berkeley: University of California Press; 1958. Costaouec D, Swanton M. Classification nominale en ixcatèque. La Linguist. 2015;51(2):201–38. Dávila P, Arizmendi M d C, Valiente-Banuet A, Villaseñor JL, Casas A, Lira R. Biological diversity in the Tehuacán-Cuicatlán Valley, Mexico. Biodivers Conserv. 2002;11:421–42. DOF. Resolución sobre conflicto por límites de bienes comunales al poblado de Santa María Ixcatlán, municipio del mismo nombre, Estado de Oaxaca. Diario Oficial de la Federación (DOF) May 10th, 1948. México; 1948. http://www.dof.gob.mx/. Accessed 25 May 2015. Eberhard DM, Simons GF, Fennig CD, editors. Ethnologue: languages of the world. 25th edition [Internet]. Dallas: SIL International; 2022. https://www.ethnologue.com Embriz A, Zamora Ó. México lenguas indígenas nacionales en riesgo de desaparición. México: Instituto Nacional de Lenguas Indígenas; 2012. Fernández MT. Fonémica del ixcateco. Bachelor thesis. Mexico City: Escuela Nacional de Antropología e Historia; 1950. Fernández MT, Swadesh M, Weitlaner RW. Some findings on Oaxaca language classification and culture terms. Int J Am Linguist. 1959;25(1):54–8. Flannery KV. In: Flannery KV, editor. Guilá Naquitz. New York: Academic; 1986. Gadgil M, Berkes F, Folke C. Indigenous knowledge for biodiversity conservation. Ambio. 1993;22 (2–3):151–6. Gomez-Chang E, Uribe-Estanislao GV, Martinez-Martinez M, Gálvez-Mariscal A, Romero I. Antihelicobacter pylori potential of three edible plants known as quelites in Mexico. J Med Food. 2018;21(11):1150–7. Hanson R. The World of Xula Palm Craft [Internet]. 2022. https://www.youtube.com/watch? v¼pT8ymAJ1UA8. Accessed 2022 May. Harrison KD. When languages die: the extinction of the world’s languages and the erosion of human knowledge. New York: Oxford University Press; 2007. Hironymous MO. Santa María Ixcatlan, Oaxaca: from colonial cacicazgo to modern municipio. Phd thesis, University of Texas at Austin; 2007. Hoppe WA, Weitlaner RJ. The Ichcatec. In: Wauchope R, Vogt E, editors. Handbook of Middle American Indians, vol. 7. Austin: University of Texas Press; 1969. p. 499–505. Illsley C, Gómez T, Edouard F, Marshall E. Brahea dulcis (Arecaceae). Trenzado simultáneo de las hojas: Producción familiar de sombreros y artesanías. In: Marshall E, Schreckenberg K, Newton A, editors. Comercialización de Productos Forestales no Maderables. Factores que

590

S. Rangel-Landa and A. Casas

Influyen en el Éxito. Cambridge, UK: Centro Mundial de Vigilancia de la conservación del PNUMA (UNEP-WCMC); 2001. p. 43–6. INEGI. Carta topográfica San Pedro Atzumba E14B85 Puebla y Oaxaca 1:50 000. 2nd ed. México: Instituto Nacional de Estadística Geografía e Informática (INEGI); 2001. INEGI. Continuo de Elevaciones Mexicano 3.0 (CEM 3.0). Raster, 15 m. Aguascalientes: Instituto Nacional de Estadística y Geografía (INEGI); 2013. INEGI. México en cifras, Santa María Ixcatlán, Censo 2020 [Internet]. Instituto Nacional de Estadística y Geografía (INEGI); 2020. https://www.inegi.org.mx/app/indicadores/?t¼255& ag¼20416#D255. Accessed Jan 2022. Larios C, Casas A, Vallejo M, Moreno-Calles AI, Blancas J. Plant management and biodiversity conservation in Náhuatl homegardens of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013;9:74. Lira R, Casas A, Rosas-López R, Paredes-Flores M, Pérez-Negrón E, Rangel-Landa S, et al. Traditional knowledge and useful plant richness in the Tehuacán–Cuicatlán Valley, Mexico. Econ Bot. 2009;63(3):271–87. MacNeish RS. A summary of the subsistence. In: Byers DS, editor. Prehistory of the Tehuacán Valley, Vol.1: Environment and subsistence. Austin: University of Texas Press; 1967. Mateos-Maces L, Chávez-Servia JL, Vera-Guzmán AM, Aquino-Bolaños EN, Alba-Jiménez JE, Villagómez-González BB. Edible leafy plants from Mexico as sources of antioxidant compounds, and their nutritional, nutraceutical and antimicrobial potential: a review. Antioxidants. 2020;9(6):541. MILPA AC. Patrones de intensificación en la producción de destilados de Agave de México [Internet]. 2020. https://www.youtube.com/watch?v¼kSh80VwQmcw. Accessed May 2022. Molina M. La recuperación de la lengua xuani-ixcateca de Oaxaca a través del video. Oaxaca: CEDELIO, FAHHO; 2010. Moreno-Calles A, Casas A, Blancas J, Torres I, Masera O, Caballero J, et al. Agroforestry systems and biodiversity conservation in arid zones: the case of the Tehuacán Valley, Central México. Agrofor Syst. 2010;80(3):315–31. Moreno-Calles AI, Casas A, García-Frapolli E, Torres-García I. Traditional agroforestry systems of multi-crop “milpa” and “chichipera” cactus forest in the arid Tehuacán Valley, Mexico: their management and role in people’s subsistence. Agrofor Syst. 2012;84(2):207–26. Moreno-Calles AI, Toledo VM, Casas A. Los sistemas agroforestales tradicionales de México: una aproximación biocultural. Bot Sci. 2013;91(4):375–98. Nava C, Romero M. Ixcatecos, pueblos indígenas del México contemporáneo. México: Comisión Nacional para el Desarrollo de los Pueblos Indígenas; 2007. Ramírez-Rodríguez R. La querella por el pulque: Auge y ocaso de una industria mexicana, 1890–1930. Zamora: El Colegio de Michoacán; 2018. Rangel-Landa S, Rivera-Lozoya E, Casas A. Uso y manejo de las palmas Brahea spp. (Arecaceae) por el pueblo ixcateco de Santa María Ixcatlán Oaxaca, México. Gaia Sci. 2014;8(2):62–78. Rangel-Landa S, Casas A, Rivera-Lozoya E, Torres-García I, Vallejo-Ramos M. Ixcatec ethnoecology: plant management and biocultural heritage in Oaxaca, Mexico. J Ethnobiol Ethnomed. 2016a;12(1):30. Rangel-Landa S, Smith-Aguilar S, Rivera-Lozoya E, Swanton MW, Casas A, Solís L, et al. Patrimonio biocultural ixcateco de la comunidad de Santa María Ixcatlán Oaxaca México. Morelia: Universidad Nacional Autónoma de México (UNAM); 2016b. Rangel-Landa S, Casas A, García-Frapolli E, Lira R. Socio-cultural and ecological factors influencing management of edible and non-edible plants: the case of Ixcatlán, Mexico. J Ethnobiol Ethnomed. 2017;13(1):59. Rzedowski J. Vegetación de México. México: Limusa; 1978. Santa María Ixcatlán C. Estatuto comunal de Santa María Ixcatlán, municipio del mismo nombre, distrito de Teotitlán, Oaxaca. Santa María Ixcatlán, Oaxaca: Desarrollo Rural Alternativo A.C. (facilitator); 2009. p. 25.

Ixcatec Ethnobotany: Plant Knowledge in the Mountains Surrounding. . .

591

Si A. The traditional ecological knowledge of the Solega, a linguistic perspective. Cham: Springer; 2016. p. 57–96. Smith-Aguilar SE, Rangel-Landa S, Casas A, Rivera-Lozoya E, Swanton MW. Patrimonio biocultural ixcateco: investigación y colaboración para su documentación, valoración y difusión. Diálogos Campo Morelia. 2016;II(1):168–200. SMN-CONAGUA. Lluvia total mensual: Estación 20129 Santa María Ixcatlán, Oaxaca [Internet]. n.d. Sistema Meteorológico Nacional (SMN) & Comisión Nacional del Agua (CONAGUA). https://smn.conagua.gob.mx/tools/RESOURCES/Mensuales/oax/00020129.TXT. Accessed Jan 2022. Solís L, Casas A. Cuicatec ethnozoology: traditional knowledge, use, and management of fauna by people of San Lorenzo Pápalo, Oaxaca, Mexico. J Ethnobiol Ethnomed. 2019;15:58. Sutrop U. List task and a cognitive salience index. Field Methods. 2001;13(3):263–76. Swanton M. La escritura indígena como “material lingüístico”. Una carta en lengua ixcateca al presidente Lázaro Cárdenas. In: van Doesburg S, editor. Pictografía y Escritura alfabética en Oaxaca. Oaxaca: Instituto Estatal de Educación Pública de Oaxaca; 2008. p. 353–87. Swanton M. Una celebración del Día Internacional de la Lengua Materna en Santa María Ixcatlán. Boletín Digital de la Fundación. Alfredo Harp Helú Oaxaca [Internet]. Oaxaca; 2022. (14): 9–10. https://fahho.mx/una-celebracion-del-dia-internacional-de-la-lengua-materna-en-santamaria-ixcatlan/ Toledo VM. Ethnoecology: a conceptual framework for the study of indigenous knowledge of nature. In: Steep JR, editor. Ethnobiology and biocultural diversity. Athens: International Society of Ethnobiology; 2002. p. 511–22. Toledo VM, Barrera-Bassols N. La Memoria biocultural: la importancia ecológica de las sabidurías tradicionales. Barcelona: Icaria Editorial; 2008. Valiente-Banuet A, Solís L, Dávila P, Arizmendi M del C, Silva C, Ortega-Ramírez J, et al. Guía de la vegetación del Valle de Tehuacán-Cuicatlán. México D.F.: Universidad Nacional Autónoma de México, Comisión Nacional para el Conocimiento y Uso de la Biodiversidad, Instituto Nacional de Antropología e Historia, Universidad Autónoma de Tamaulipas, Fundación para la Reserva de la Biosfera Tehuacán-Cuicatlán; 2009. Vallejo M, Casas A, Blancas J, Moreno-Calles AI, Solís L, Rangel-Landa S, et al. Agroforestry systems in the highlands of the Tehuacán Valley, Mexico: indigenous cultures and biodiversity conservation. Agrofor Syst. 2014;88(1):125–40. Vallejo-Ramos M, Moreno-Calles AI, Casas A. TEK and biodiversity management in agroforestry systems of different socio-ecological contexts of the Tehuacán Valley. J Ethnobiol Ethnomed. 2016;12(1):31. Velazquez de Lara G. Relación de Ixcatlán, Quiotepeque y Tecomahuaca [Internet]; 1579. p. 28. https://fromthepage.lib.utexas.edu/llilasbenson/relaciones-geograficas-of-mexico-and-guate mala/ixcatlan-quiotepeque-tecomavaca-oaxaca-1579/display/2054 Wehi PM, Whaanga H, Roa T. Missing in translation: Maori language and oral tradition in scientific analyses of traditional ecological knowledge (TEK). J R Soc N Z. 2009;39(4):201–4. Zarazúa-Carbajal M, Chávez-Gutiérrez M, Romero-Bautista Y, Rangel-Landa S, Moreno-Calles AI, Ramos LFA, et al. Use and management of wild fauna by people of the Tehuacán-Cuicatlán Valley and surrounding areas, Mexico. J Ethnobiol Ethnomed. 2020;16(1):4. Zent S. Acculturation and ethnobotanical knowledge loss among the Piaroa of Venezuela: demonstration of a quantitative method for the empirical study of traditional environmental knowledge change. In: Maffi L, editor. On biocultural diversity, linkage language, knowledge, and the environment. Washington, DC/London: Smithsonian Institution Press; 2001. p. 190–211.

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of Craft Traditions and Biocultural Landscapes Citlalli Lo´pez-Binnqu¨ist, Belinda Contreras-Jaimes, Fortunata Panzo-Panzo, and Edward A. Ellis

Abstract

The case of wool textiles and the Nahua artisans of the Sierra de Zongolica in the state of Veracruz, Mexico, is addressed from an ethnobotanical study interwoven with a biocultural landscape approach. In this framework, the life stories of artisans of the Sierra de Zongolica reveal their history around weaving wool and their interactions with the mountainous environment, considering its biophysical, symbolic, cultural, economic, and political dimensions. We show the results of our analysis of changes in the landscape that frame the management of flora and fauna species used for the provision of wool and dyes for making wool textiles. The scope was enriched through a participatory exercise and dialogue which materialized in a storybook and an awareness raising infographic on the artisanal work. We conclude that efforts framed in theoretical and methodological procedures that promote creativity and participation support analytical experiences that can foster shared learning. In this interaction, the biocultural landscape, used as a comprehensive approach, allowed us to recognize the work of the artisans, integrating aspects of the Indigenous worldview, gender, use of spaces, and local resources, as well as the future of the craft studied.

C. López-Binnqüist (*) Centro de Investigaciones Tropicales, Universidad Veracruzana, Xalapa, Veracruz, Mexico e-mail: [email protected] B. Contreras-Jaimes Doctorado en Ciencias de la Sustentabilidad, Universidad Nacional Autónoma de México, Ciudad de México, Mexico F. Panzo-Panzo Campus Atlahuilco, Sierra de Zongolica and People and Plants International, Universidades para el Bienestar “Benito Juárez García”, Atlahuilco, Veracruz, Mexico E. A. Ellis Centro de Investigaciones Tropicales, Universidad Veracruzana, Veracruz, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_14

593

594

C. Lo´pez-Binnqu¨ist et al.

Introduction In the Sierra de Zongolica, located in the center of the state of Veracruz, natural and cultural richness combine to form one of the most bioculturally important regions in Mexico (Alatorre-Frenk et al. 2014; Boege 2008). Historically, wool textiles have been produced in this region for making traditional clothing, and today the artisans of the area have recreated their pieces for sale as handicrafts, integrating a growing range of new designs created with plant dyes. In this chapter, we recognize the importance of wool textiles as part of the regional biocultural landscapes. Since pre-Hispanic times, the Sierra de Zongolica has been inhabited by Indigenous peoples, the current Nahua, and their predecessors, who have developed a rich knowledge about the management of multiple use of spaces such as the traditional crop field called milpa, and the forests or monte1 and homegardens. Wool textiles are elaborated as part of the regional production and cultural symbolic systems; they are transformed and recreated, as part of the biocultural landscapes of the Sierra de Zongolica. Our interest in this topic and approach is driven by the following questions, (a) what have been the techniques and natural resources (fauna and flora) used for the production of wool textiles in the Sierra de Zongolica, (b) how have they been modified throughout the history of their use, (c) what factors have influenced these changes, as part of the diversity of uses and management of the natural resources of the region, and (d) how to contribute to the social relevance of the ethnobotanical work through a transdisciplinary, participatory, and dialogical proceeding. This chapter is divided into five sections. In the first one, we establish the theoretical underpinnings of the biocultural approach on the analysis of handicrafts and regional landscapes, and we describe the methodology used. In the second section, we describe the Sierra de Zongolica’s biophysical conditions and its historical evolution, in particular the changes in vegetation cover and land use. In the third section, we provide ethnobiological information about the natural dyes and sheep providing wool; in addition, we show the collaborative work process that expanded this research to produce teaching and dissemination materials. In the fourth section, we discuss how the changes in the elaboration of handicrafts and in multiple-use management of the Sierra de Zongolica are intertwined, through the biocultural landscape approach. Finally, in the fifth section, we discuss the perspectives and a methodological reflection on this transdisciplinary experience and its contributions to the ethnobotanical work. This chapter is part of a participatory research program developed during the past ten years in the Sierra de Zongolica, along with students and researchers of the Universidad Veracruzana, the organization People and Plants International, artisans,

1 In Zongolica, monte (kohyo in Nahuatl) refers to all the areas where trees are growing; they may contain a variety of combinations and densities of species, both native and introduced, and encompass the different land use type from forest to agriculture in different stages; the term is associated with forests and mountains.

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

595

students, and Nahua families. Collectively, and based on a transdisciplinary approach, we worked on valuing local knowledge, addressing issues about territory, handicrafts, and diversified production systems such as milpa and agroforestry systems, including the production of firewood and charcoal.

Theoretical and Methodological Bases Textiles Representing Nature and Human Linkages Handicrafts, including textiles, have their origin in the human process of appropriation of nature, in which the environment supplies raw materials, and the tradition and worldview of the artisan gives them form and function (Turok 1988; Zimmerer et al. 2017). Therefore, handicrafts are considered biocultural resources. In Mexico, the artisan tradition has taken place for at least 4,000 years (Sarmiento 2000; Cornejo-Rodríguez 2009) in contexts of great biological and cultural richness (Caro-Bueno et al. 2009). Textiles are, only after painting, the means that has most served humanity to record its dreams, its religiosity, and its ideas, allowing to depict the world of today and that of its ancestors (Beauregard-García et al. 2008). They emerge to satisfy the physical need to cover the body, and this involved the search for appropriate materials among the natural elements. The raw materials that the flora and fauna of each region have provided, as well as the progress in techniques, have influenced the evolution of the ways of dressing (Velasco-Rodríguez 2002). Once the primary need was satisfied, decorative elements were included, which reveals the keys to understanding the society that created them, since patterns, colors, and materials reflect a worldview (Anawalt 2005). The textile in Mexico has pre-Hispanic roots, with a permanence of traditional techniques, such as the spinning with spindle and the waist loom, still in use (Velasco-Rodríguez 2002). Approaching the study of handicrafts requires a gender perspective. Women artisans have kept alive a creative work that involves ancestral empirical knowledge and are “the ones who, with their knowledge, work and interests, have preserved the craft” (De la Borbolla 2010). The weavers hold a central position as decision-makers regarding the family unit and the use and management of the natural resources, and as multiplying agents of knowledge, playing a highly important role in the life of their communities. In this chapter, the notion of handicrafts as biocultural resources is underpinned and their relationship with the environment is highlighted. The biocultural perspective enables the explanation of handicrafts as an expression in which beliefs, knowledges, and practices around the linkage of natural resources management, production processes, uses, and symbolic representations converge under a constant negotiation between tradition and innovation.

596

C. Lo´pez-Binnqu¨ist et al.

Biocultural Landscapes in Mountainous Regions In Mexico, mountainous regions are often sites of settlement of Indigenous peoples and are commonly referred to as refuge areas (Aguirre-Beltrán 1967), where different groups settled to flee from the strong yoke of the Mesoamerican empires and later from the Spanish conquerors. They are sites with high biodiversity and have been favorable areas for species conservation, standing out for their importance for agrobiodiversity, developed from interaction and experimentation in different environmental niches and spaces of adaptation and local isolation (Brush 1998; Zimmerer et al. 2017). However, as Wallace et al. (2015) point out, mountain regions are also the most vulnerable regions, especially in the face of climate change, and the most marginalized in socioeconomic and health care terms. Various factors impact both, the practical relationship of use and management of resources and the symbolic (intangible) aspects related to worldview and the mythical history. The latter can be seen in institutional development programs that incorporate new techniques and supplies, often not adequate to the high environmental variability of the mountains. In addition, it can be seen in pressures on resources due to local population growth, and the lack or short supply of economic activities for a growing young population. In this study, the biocultural landscape approach is adopted to elucidate the multiscale processes that impact the production of wool textiles. Biocultural landscapes are socio-environmental spaces managed and configured by human populations, who throughout history have assigned them specific symbolic and cultural values (Berkes and Davidson-Hunt 2006). This approach provides a basis for interpreting the relationships between cultural and biological diversity, considering environmental and cultural heterogeneity, and including elements of diversification and multifunctionality (Agnoletti and Rotherham 2015; Zimmerer et al. 2017), considering domestication as biocultural processes (Casas et al. 2014). Some studies about crafts from mountain regions frame their analysis based on territoriality and landscapes (Silva et al. 2017; Terry 2020). In the case analyzed, we consider that the biocultural landscape approach allows us to observe and study the elaboration of wool textiles as products in which knowledge about the natural environment, manufacturing techniques, and symbolic representations converge, constantly adapting to environmental conditions as well as to the social, cultural, and economic transformations that occur at different scales.

Methodological Underpinning for a Participatory and Dialogical Ethnobotany In the experience with the textiles of Zongolica, operationalizing the theoreticalconceptual frameworks required a methodology supported in constructivism, with a qualitative approach and based on participatory methods, which are approaches of growing interest and relevance in the field of ethnobotany (Medley and Kalibo 2005; Alexiades 1996).

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

597

Techniques from ethnobiology and ethnography were used, including field trips, collection of botanical specimens and identification of dye species, in-depth interviews, focus groups, audiovisual and work material production, life stories, participant observation, field journal, and relationship mapping (Albuquerque et al. 2014; Martin 1995). From these methods, a list of the species was generated, in which their use and management during textile production was recorded; the routes by which raw materials are obtained, consumed, and marketed were identified. The mapping of the organizational and commercial relations was carried out, making it possible to explain the impact on gender relations, derived from female work, as well as to identify the changes in the management of the territory. Conducting a participatory and dialogical research (Encina et al. 2009) responds to the need to bring ethnobotanical work closer to its societal relevance and to acknowledge the interest of the artisans to revitalize their knowledge through their mother language and disseminate the value of their textile production. It was necessary to strengthen the transdisciplinary nature of the project, incorporating elements of analysis from communication studies (Encina and Ávila 2010) and Indigenous pedagogy (Gómez 2009), that allowed a deeper understanding of the ways in which knowledge flows and is appropriated (Argueta-Villamar 2011; Toledo 2011). This dialogue materialized in two audiovisual products, the book and CD Los textiles de la Tlasesekya: cuentos sobre hilos, telares y vellones (The textile of the Tlasesekya: tails about threads, looms and fleeces) (Contreras and Sosme 2014) and an infography on the textile production process. All these materials were created from an interpersonal and interdisciplinary collaboration exercise, with community participation to build, test, and validate both the content and the narrative, as well as the graphics and the audio production that captured the voices of the artisans themselves, always prioritizing the Nahuatl language and their oral tradition.

Landscape Dynamics in the Sierra de Zongolica Main Historical and Socioeconomic Conditions The Sierra de Zongolica has been occupied by populations of pre-Hispanic origin, and within Mexico it currently comprises the second largest Nahua territory. According to the interpretation of pre-Hispanic sources (Kirchhoff et al. 1976), the original population came from regions located in northern Mexico. In the twelfth century, the Nonoalcas, a group from which the current Nahuas descend, settled in the Sierra de Zongolica, taking military and cultural power and imposing their lordship over the inhabitants of the region (Rodríguez and Hasler 2000). The mountains have always been home and pillar for the lives of the Nahuas of Zongolica. In the Nahua conception of the world, which is based on duality (everything that exists has its opposite) (López-Austín 1994), the universe appears divided from the surface of the earth: Tlaltikpatli is the land surface, nutritious mother of the farmers, and Tlalokan is the wild paradise located in the subsoil, place of agricultural fertility, cradle of wild animals and vegetables, and receptacle of

598

C. Lo´pez-Binnqu¨ist et al.

springs and water sources that flow into the caves of the mountains (Aguirre-Beltrán 1986; Rodríguez 2003). The intimate relationship of the Nahuas with their environment is expressed in various ritual practices that remain current and dynamic throughout the Sierra (Sevilla et al. 2020). Physiographically, the Sierra de Zongolica constitutes a mountain range which is part of the Sierra Madre, located in the center of the state of Veracruz (Fig. 1); composed of sedimentary limestone rocks and made up of eastern ridges composed mainly of Jurassic-Cretaceous limestone. Its morphology is modified by karst processes that have carved sinkholes, sinks, and natural bridges. Its topography is rugged, with steep slopes, and narrow valleys, some of them dislocated by normal and reverse faults. It presents an altitude variation ranging from 100 masl to more than 2,500 masl. Due to its climatic and altitudinal characteristics, it is divided locally into two main zones: the warm and cold zones (López-Binnqüist et al. 2014). In the cold zone, the climate is temperate-extreme to cold, with an annual average temperature of 12  C, with some frosts in winter. The elaboration of wool textiles is carried out in the municipalities of the cold zone, mainly in Soledad Atzompa, Tlaquilpa, Atlahuilco, and to a lesser extent in Mixtla, Los Reyes, Tequila, and Texhuacan, which make up the transition between the cold zone and the warm zone (Fig. 1). According to CONAFOR (2017), the forest cover in municipalities of the cold zone – with elevations up to 2,800 masl – includes pine forests (29%) and pine-oak or oak forests (21%). In the transition zone – with elevations below 1700 masl – montane cloud forests (13%) are found, while in the lower areas a portion of tropical rain forest (3%) is found mainly in a secondary vegetation status. In both altitudinal zones, each family generally owns one or up to three plots of 0.5 to 1 hectare. Sometimes, these portions of land are distributed in different areas, encompassing some variability in climates, soils, and altitudes. Land tenure is private, and over time families have subdivided their plots to pass on to their children, atomizing their size. Currently, as part of the forest-agroforestry-agricultural management, patches of forest or vegetation are observed in different states of succession that include areas that are part of agricultural production (milpa and acahuales2) accompanied with fruit trees, and areas covered with trees such as Quercus species for charcoal production or planted individually for timber or as part of forest plantations and/or reforestation (Fig. 2). Among the different production systems, the milpa has been the base of subsistence of the Nahua families of Zongolica. Corn is grown in the milpa in combination with beans, squash, and quelites, among other species. It also provides other resources such as animal fodder, medicinal and seasoning products, and natural fertilizers. The varieties of corn in the cold zone are the following: blue (poxahuac), yellow (kostik), red (chichiltik), and white (iztak).

2

Acahual refers to the secondary vegetation derived from the agricultural system involving fellingslashing-burning-cropping-fallow; its composition is characterized by a diversity of trees, shrubs, grasses, creepers, and herbaceous plants.

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

599

Fig. 1 Location of the Sierra de Zongolica in the center of the state of Veracruz

Temporary migration within the region has occurred since pre-Hispanic times. In colonial times, the seasonal migration from the cold to the warm zone to work in coffee, tobacco, and cane crops was documented, and in the last two decades, migration from the warm to the cold zone to work in wood production activities, especially in sawmills, has been reported (El Hidalgo-Ledesma 2016). Unlike other regions of Mexico, international migration began only recently, at the beginning of the nineteenth century, and according to Martínez Canales (Martínez-Canales 2010), it is not considered to be massive and is mainly represented by migration of men.

Tree Cover Changes as Part of Regional Land Use Changes At the beginning of the Spanish Colony, the Sierra de Zongolica remained mostly isolated. Later, in the seventeenth century, the Jesuits introduced herds of goats and sheep that they kept in the pastures of the mountains. The isolated conditions changed due to the introduction of crops of commercial importance, such as tobacco, sugar cane, and coffee, especially in the warm zone of the Sierra. During the eighteenth century, the Spanish Crown established a monopoly on tobacco cultivation for its exclusive sale in Spain. Later, at the end of the nineteenth century, coffee was introduced, and the Sierra de Zongolica became the second highest coffee-

600

C. Lo´pez-Binnqu¨ist et al.

Fig. 2 Fragmented landscape of multiple use with acahuales, milpa, and Pinus patula plantations of different ages, grass and pine-oak forest in Tlaquilpa municipality, 2018. (Photo by the authors)

producing region nationwide. This boom ended in the 1990s, when price and credit subsidies for agriculture in general, including coffee, were canceled (MestriesBenquet 2003; Reyes-Sánchez 2016) as part of imposed neoliberal policies. One of the most recent changes in the cold and transition zones has been the introduction of plantations and/or individual trees, especially of Pinus patula. This trend is part of the regional forestry history, nested in local and external power and production relations and more recently as part of development programs aiming at recovering vegetation cover. As mentioned before, the Nahua maintain strong cultural linkages with their forests while management practices have changed from a subsistence based to a more commercial oriented. During the 1970s and 1980s, the forests of the region were severely exploited by loggers from cities and regions bordering the Sierra de Zongolica, who, through repression or very low pay, extracted wood from the forests. Given this situation, the planting of trees in deforested areas and the establishment of nurseries and sawmills were part of the strategies that the Nahua families carried out to regain control of their forests and their territory. A new commercial activity started, consisting in the management of pine trees and their use for the construction of basic furniture pieces, which have been sold via a wide network of Nahua intermediaries in different states of Mexico. The introduction of timber trees intensified with the most recent forest policies, which began in the 1980s

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

601

and whose objective has been the introduction of forest plantations for recovery, conservation, and/or commercialization, following international guidelines (El Hidalgo-Ledesma 2016; López-Binnqüist et al. 2014). In order to identify the land use changes in relation to the dynamics of vegetation cover, a spatial analysis was carried out for the cold and transition zones of the Sierra de Zongolica, by using Landsat satellite images and the application of Geographic Information Systems (GIS). The vegetation cover and land use for 1997 and 2019 was determined and reclassified into four forest cover categories: (1) conserved, (2) secondary, (3) acahual or shrubby (agricultural), and (4) deforested (agricultural). Subsequently, changes in forest cover areas and rates of change between 1997 and 2019 were calculated; transitions in forest cover during the analysis period were characterized into the following four categories: (1) deforestation (total loss of conserved or secondary tree cover), (2) degradation (deterioration in tree cover without total loss), (3) areas of tree cover without change, and (4) recovery (gain in tree cover). The analysis indicates that the increase in deforested areas between 1997 and 2019 was of 8022 ha, with an annual average rate of 2.04%. The changes in forest cover are characterized mainly by deforestation of conserved and secondary forests (6,173 ha), but mostly (97%) of conserved forests. The annual loss of conserved and secondary forest areas during the same period was 2.04% and 0.04%, respectively. Shrubby or acahual vegetation remains relatively stable during the period with a change rate of 0.25% per year. There are marked differences in the transitions of forest cover change, with forest degradation being the dominant change (695 ha year 1). However, the recovery (498 ha year 1) of tree cover is also predominant. On the other hand, deforestation represents the second largest change (638 ha year 1). Deforestation is mostly attributed to the conversion of forest areas into agricultural use, although fires in 2017 and 2019 were also found to be a major cause of forest cover loss. Degradation is characterized by the deterioration of the well-conserved vegetation, which results from activities such as firewood collection, charcoal production, and grazing. This analysis demonstrates that forest degradation processes are severe in the montane cloud forests and secondary tropical vegetation, while in the pine and pine-oak forests it shows that loss and fragmentation are the main changing processes. Despite the current recovery of vegetation due to the increase of secondary vegetation and tree plantations, this process does not exceed those of degradation and deforestation. The GIS analysis also shows that higher rates of deforestation and degradation are mainly occurring in the higher elevations of the two altitudinal zones – the cold and the transitional one. On the other hand, tree cover is increasing in lower elevations, where most agricultural-agroforestry and grazing plots are located. Due to the general reduction of forests and the biodiversity they harbor, women and men identify scarcity of certain game species and edible fungi, among others, as well as the reduction of pasture areas, formerly used for sheep grazing; some have been converted into Pinus patula plantations. In some cases, the milpa has also been displaced by these plantations, with severe consequences for family subsistence and

602

C. Lo´pez-Binnqu¨ist et al.

germplasm conservation. Management practices, as well the collective fabric, identity, and local worldview around the milpa production, are also affected. In summary, the changes in forest cover and land uses, at the plot level, entail various impacts on the lives of the Nahua families who face decisions that involve incorporating diverse uses in reduced spaces, considering the number of family members and the diversity of economic activities increasingly affected by external dynamics. Changes in land use and vegetation cover determine decisions and options regarding the management of certain local resources, including those used for textile production.

Weaving as Part of the Biocultural Landscape Women Artisans: Weavers of Life According to Toledo and Alarcón (2012), the women who inhabit and manage their natural environment have a great capacity for decision and action on the landscape, in addition to their key role in the transmission of beliefs, knowledge, and practices to younger generations. In the Sierra de Zongolica, the life of the artisans reveals their history of weaving and their relationship with the environment. They begin to shepherd while cleaning the wool between the ages of 8 and 16; the beginning of their apprenticeship and mastery of the complexity of weaving takes place between the ages of 20 and 30. They finally consummate the learning by being able to shear the sheep and dye the wool with natural dyes. In the case of the Zongolica textiles, spaces and natural resources are delimited and distributed according to gender, based on myths. “They say that the first and great mother of humanity, Tonantzin, after the Virgin Mary, wanted to cover her son from the cold of the high mountains. So, she sheared the wool of her sheep, and covered her little one with it, but when he ran through the woods, the wool always stuck in the branches. This led her to think of a way to braid this thread, and she asked her husband, who was carpenter, to prepare some sticks for a spindle3 and a loom.4 With her ancient and miraculous wisdom, she created a simple fabric and made the garments to dress her son. One afternoon she left her loom to go prepare the food, but when she returned, she found that someone (the devil, they say) had made a prank: he tangled the threads and put more sticks on the loom; and so, the other types of weavings emerged. This is how the tradition that this mother passed on to all her weaving daughters, was born” (Contreras-Jaimes 2015). From generation to generation, the Indigenous women of Zongolica have inherited the gift of Tonantzin and passed on the knowledge of their ancestors. For

3

The spindle (malacatl) is the tool used for spinning (tzaba). It is a rotating artifact made of wood that, driven by the movement of women’s hands, twists the wool to turn it into thread. 4 The waist loom is a weaving technique of pre-Hispanic origin.

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

603

this reason, through the centuries, spinning and weaving has been part of the life of Nahua women, who express their worldview through the selection of the raw materials and in the designs. Regarding practices and management, spatial distribution between genders is also observed. The pine-oak areas men are dedicated mainly to wood and charcoal production, while women practice the collection of certain nontimber forest products. The plot is the family land, and it includes spaces typical of the domestic unit such as the milpa, the orchard, and the backyard, in addition to the residential area. While the cultivation of the milpa is mostly carried out by men, the backyard constitutes a space managed by women, where they raise poultry and grow medicinal, edible, dye, and ornamental plants. It should be noticed that this traditional distribution has also been transformed as men have migrated, leaving the management of all family spaces in the women’s hands.

Wool Textile Production Process Textile production in Zongolica fulfills a triple function: that for satisfying the need of protection from environmental conditions; the cultural, being a means of expression of symbolic aspects and worldview; and the economic, since artisan work is an important alternative for generating income. Originally, the craft began with the production of clothing made on a waist loom: the women’s skirt or kueitl, the men’s sleeve or jorongo, the tlalpiales worn by women to braid and decorate their hair, and blankets (Fig. 3); these continue to be used mainly by the elderly and for important celebrations. Production has diversified by adapting traditional designs to rugs, tablecloths, bedspreads, carpets, blankets, shawls, scarves, and bags (Larios-León 1991), also incorporating needle and hook techniques. Currently, mañanitas, hats, bags, vests, stuffed animals made with hook, slippers, key chains, pins, and earrings and necklaces made with needle are offered in a wide variety (Fig. 3). The textile elaboration process involves knowledge and practices from daily management and care of sheep. In Nahuatl, tlamalbis is a term used for any type of care, which implies taking them to graze, and protecting them from environmental conditions and hazards. When a sheep is about 1 year old, they shear it, in Nahuatl tlaxima, and repeat the process every year. The wool should be allowed to cool for 4–7 days before washing it. The next step is to wash the wool, an activity called tlapakii in Nahuatl. For this process, a tuber known as chicamole, previously crushed, is used as soap; water is added, and they beat the wool with a large stick until it is free of mud and dirt. After being left to dry for several days, the next phase begins, the combing, tlapochina iban tlachipaba, which consists of cleaning the wool by hand while spreading and arranging it to facilitate its later spinning. This work is carried out during free time, not intensively, and it usually takes several days. The spinning or tzaba comes later, using the spindle to twist the wool and turn it into a thread of the thickness that the woman considers appropriate.

604

C. Lo´pez-Binnqu¨ist et al.

Fig. 3 Traditional clothing from the cold zone of the Sierra de Zongolica, such as the skirt and tlalpiales for women and the male jorongo, woven using the waist loom technique (above) and examples of the diversification of textile production from traditional designs, adapted to a wide range of products for today’s consumer (below). (Photos by the authors)

The thread can be used in its natural color, or it can be dyed. At this time, the collection of dye species is carried out, in Nahuatl maniktikite xivitl ipan nitlayapas. The women collect the necessary amount of plants according to the amount of thread to be dyed; these plants are abundant in a certain season, in the surroundings of the town, the forest, or their backyards. The dyeing process or tlapalbis includes a first boil of the skeins of the thread with a mordant such as alum, to later be boiled with the fresh dye species. Women must collect the firewood for the boils and take care of the entire dyeing process. The skeins are dipped in the dyes as many times as necessary to achieve the desired color or tone, and once achieved, they must be let to dry. This is followed by the process of arranging the threads or nik yektlali, which is the most creative moment of the production, in which the artisan imagines the size and color of the garment and then arranges the thread in the correct order. Finally, the weaving, or ijkiti, is done with the pre-Hispanic waist loom technique (Fig. 4), which can be done kneeling or standing, carrying the weight of the garment on the waist, which is why women do it in 3-h sessions maximum. The finished garments are the product of months of work.

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

605

Fig. 4 Illustrations of waist loom weaving and spinning (Mendoza Codex), as well as the instruments used (Florentine Codex) in prehispanic times, compared to the techniques and instruments used today in the Sierra de Zongolica. (Photos by the authors)

Dye Plants In Mexico, the natural colorants tradition dates to pre-Hispanic times (Ortiz 1991; Velasco-Rodríguez 2002). Color was an important element in Mesoamerican social and religious life, as part of clothing, ornaments, arts, food, and drink; the pigments from animal, vegetable, and mineral origin were a valuable resource. Vegetable dyes in Mesoamerica came from roots, stems, leaves, flowers, and fruits, such as indigo (Indigofera tinctoria L.), palo de Campeche (Haematoxylon campechianum L.), achiote (Bixa orellana L.), zacatlaxcalli (Cuscuta americana L.), cascalote (Caesalpinia coriacea Pépin), cempasúchil (Tagetes erecta L.), muicle (Justicia spicigera Schltdl.), and palo de Brasil (Haematoxylon brasiletto H. Karst.), among others (Velasco-Rodríguez 2002). In addition, two dyes of animal origin reached great importance: the cochineal (Dactylopius coccus) which is an insect that feeds on the cactus plant (several species of Opuntia), and the secretion produced – as a defense mechanism – by the purple snail (Plicopurpura pansa). In Mexico, there are 541 registered plant species for artisanal use, 90 of which are used as colorants (Bravo Marentes 1999; Bravo Marentes and Neyra González 2009). Forest areas provide a large number of plants that produce dyes that are not used in an industrial way, since the extraction of the dye is expensive and not very productive, but they are used by artisans with good results. The ancestral use of natural dyes in the cold zone of the Sierra de Zongolica is limited to two commercially acquired species that are produced in the states of Oaxaca and Puebla: indigo, of plant origin, and cochineal, of animal origin. The dyeing technique has been

606

C. Lo´pez-Binnqu¨ist et al.

preserved since pre-Hispanic times, which consists of immersing the skein of thread in a hot aqueous solution of the dye, after applying the mordant or incorporating it into the dye (Mastache 2005). The use of indigo that dyes in dark blue tones requires a special preparation before its use, which consists of a controlled fermentation process that lasts between three and 6 months. In clay jugs, the powdered indigo is mixed with water, rotten maguey leaves, and the urine of women or babies, fed with a special diet in the previous days. Specialized knowledge is needed to achieve the proper levels of fermentation. Finally, the dye is obtained, and the wool thread is immersed in it. For the yarn to reach the desired tone, it can be dyed up to eight times. Dyeing with indigo is a process that only a few women in Zongolica master. As for the cochineal, which produces a wide range of red tones, the ancient dyeing technique also remains intact. The dried cochineal is ground in a metate (composed by plain and cylindrical stone for grounding); it is added to the boiling water together with the thread, previously submerged in a mordant; and it is boiled until the desired color is obtained. Currently, the number of tones has been increased by changing the acidity of the mixture, using lemon. This skill of the artisans of Zongolica on traditional dyeing techniques allowed the success of the “Project for the rescue and promotion of textile crafts in the Sierra de Zongolica,” (Proyecto de rescate y fomento artesanal de textiles de la Sierra de Zongolica), which was launched in 1992 by the Popular Cultures Office (Dirección de Culturas Populares) of the state of Veracruz, in charge of supporting the revival and development of popular culture with an emphasis on the Indigenous and rural areas of the state. The objective of this project was to train weavers in the use of local dye flora, because of the difficulties of acquiring indigo and cochineal; this allowed new uses of local plants and the expansion of the color palette. Since then, the search for new colors has fueled constant experimentation, which is reflected in the permanent innovation of dye tones. To this day, Nahua women continue to explore their local flora in search of a wider range of colors (Beauregard-García et al. 2008; Contreras-Jaimes 2015). Twenty species used in the textile process were identified: one to clean wool, one for the indigo dye ferment, 18 dye plants (Contreras-Jaimes 2015) from 13 different families, and a total of eight genera of lichens, developed together in what is locally known as paxtle. The dye plants are listed in Table 1, which shows the nomenclature and general characteristics of the species; the genera and species identified in the set of lichens used for dyeing, locally recognized under the name of paxtle, are listed in the Table 2. Figure 5 contains the photo of each species. Finally, Table 3 presents the characteristics of use and management of the dye species. Given the nonintensive collection practices of wild species and since most of these are secondary vegetation species that are favored by disturbance, their use does not represent a risk to the abundance of the species. The reported use of trees that are of primary vegetation is practically null, due to the short durability of the particular dyes. For lichen however, in-depth studies are needed (Trueba 2008). Trueba (2008) reports the intensive harvesting of lichens from the steadily reduced forest areas. Some of the plants that are scarce in the cold zone are abundant in the warmer part of

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

607

Table 1 Nomenclature and general characteristics of dye species Scientific name (Photo id in brackets)) Agave atrovirens. var. atrovirens Karw. ex SalmDyck. (a) Alnus acuminata Kunth. (b) Alnus jorullensis Kunth. Artemisia ludoviciana Nutt. (c) Baccharis conferta Kunth. (d) Barkleyanthus salicifolius (Kunth) H. Rob. and Brettell. (e) Bidens aurea (Aiton) Sherff. (f) Bidens triplinervia Kunth. (g) Bougainvillea glabra Choisy. (h) Cuphea cyanea Moc. and Sessé ex DC. (i) Cuscuta jalapensis Schltdl. (j) Justicia spicigera Schltdl. (k) Microsechium palmatum (Ser.) Cogn. (l)

Phytolacca icosandra L. (m) Prunus domestica L. Quercus crassifolia Bonpl. (n) Rubus aff. pringlei Rydb. (o)

Family ASPARAGACEAE

Common name Biological Nahuatl/Spanish form Mexcale/Maguey Herbaceous (Maguey)

Wild/ cultivated Wild and cultivated

BETULACEAE

Ilite

Tree

Wild

BETULACEAE

Ilite

Tree

Wild

ASTERACEAE

Hierba muestra

Herb

Cultivated

ASTERACEAE

Escobilla

Wild

ASTERACEAE

Yoyotl/Somiate

Erect bush or semiprostrate Shrub

ASTERACEAE

Kuawamosoh/ Herb Mozote Amosoh/Amosoh Vine cimarrón Bugambilia Tree

ASTERACEAE NYCTAGINACEAE LYTHRACEAE

Huitzitzilxochitl/ Flor de chupamirto CONVOLVULACEAE Sakapalli/ Sacapale ACANTHACEAE Muitle/Hierba azul CUCURBITACEAE Chikahmolli o Chichikahmolli/ Chicamol, chicamole o amole amargo PHYTOLACCACEAE Totolkilitl/ Mazorquilla o jabonero ROSACEAE Ciruela FAGACEAE ROSACEAE

Tlilawatl/Encino negro Mora

Wild

Wild Wild Cultivated

Herb

Wild

Parasitic herb

Wild

Herb

Cultivated

Vine Liana

Wild and cultivated

Herb

Wild

Tree

Cultivated

Tree

Wild

Semiclimbing shrub

Wild (continued)

C. Lo´pez-Binnqu¨ist et al.

608 Table 1 (continued) Scientific name (Photo id in brackets)) Salvia polystachya Cav. (p) Sambucus canadensis L. (q) Tagetes erecta L. (r) Usnea spp. (s)

Family LAMIACEAE

Common name Nahuatl/Spanish Kuawchia

Biological form Herb

Wild/ cultivated Wild

ADOXACEAE

Xoometl/Sauco

Tree

ASTERACEAE

Cempoalxochitl/ Flor de muerto Paxtle

Herb

Wild and cultivated Cultivated

Lichen

Wild

PARMELIACEAE

The nomenclature was revised and cited according to: www.theplantlist.org and www.tropicos.org, accessed on October 25, 2019.

Table 2 Genera and species identified in the set of lichens used for dyeing and locally recognized under the name of paxtle

Genus Evernia Flavopunctelia Heterodermia Parmotrema Punctelia Ramalina Teloschistes Usnea

Species

Usnea brasiliensis Usnea cirrosa Usnea dasaea Usnea erinacea Usnea flavocardia Usnea glabrata Usnea ramillosa Usnea rubicunda Usnea subfloridana

the Sierra de Zongolica like the dye plant sacapale (Cuscuta jalapensis Schltdl.) which is considered a plague.

Sheep Wool The existing, developed pre-Hispanic textile production in Mexico was for the most part enriched by elements introduced during the colonial period. Most importantly, sheep wool (Beauregard-García et al. 2008) – given its thermal qualities – was quickly incorporated into the making of garments (Martínez-Peñaloza 1986). Wool has become, along with natural dyes, a fundamental component of textiles in the cold

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

609

Fig. 5 Plant and lichen species used in the dyeing and textile process. (Photos by the authors)

zone of the Sierra de Zongolica since the introduction of sheep to the area; its use has been consolidated over half a century of management experience. Sheep (Ovis aries) is a species introduced to Mexico from Europe; it is classified into breeds and varieties. In the Sierra de Zongolica, six locally adapted varieties are

C. Lo´pez-Binnqu¨ist et al.

610 Table 3 Use and management of dye species

Species Agave atrovirens. var. atrovirens Karw. ex SalmDyck. Alnus acuminata Kunth. Alnus jorullensis Kunth.

Artemisia ludoviciana Nutt. Baccharis conferta Kunth.

Barkleyanthus salicifolius (Kunth) H. Rob. and Brettell. Bidens aurea (Aiton) Sherff.

Bidens triplinervia Kunth.

Bougainvillea glabra Choisy. Cuphea cyanea Moc. and Sessé ex DC.

Used part Leaves

Tender leaves or bark Leaves or bark, the same color results All without root Branches with leaves, the whole plant can be removed although the root is not necessary Branches with leaves and yellow flowers Grass, and sometimes the ligules (white or yellow) and flowers Ligules (yellow) and flowers, and sometimes the leaves Bracts and flowers Flowers

Color that gives Used to ferment in indigo dyeing. Does not stain Yellow or light brown

Expires when used No

Harvest season All year long

Other uses at the local level Pulque

No

All year long

Timber and firewood

Yellow or light brown

No

All year long

Timber and firewood Medicinal

Light yellow

No

Very light yellow

No/Yes when it is extracted from the root, but it is not necessary

All year long

As broom Medicinal

Very light yellow

No

April, its flowering

Medicinal. Veterinarian (to cure chickens)

Light orange

No

From October to December, its flowering

Light yellow

No

From October to December, its flowering

Edible Medicinal

Between purple and pink * It has only been used by a person who experimented, and she does not want to

No

From May to September, its flowering

Medicinal

No

Medicinal

(continued)

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

611

Table 3 (continued)

Species

Cuscuta jalapensis Schltdl. Justicia spicigera Schltdl. Microsechium palmatum (Ser.) Cogn. Phytolacca icosandra L.

Prunus domestica L.

Quercus crassifolia Bonpl. Rubus aff. pringlei Rydb. Salvia polystachya Cav.

Sambucus canadensis L. Tagetes erecta L.

Usnea spp.

Color that gives

Expires when used

reveal the result or disclose her finding * Intense yellow

Yes

All year long

Medicinal

Green, in various shades It is used as a soap to wash the wool Between pink and very light purple

No

All year long All year long

Medicinal

Soap

Leaves (purple plum, not white) Bark

Very light green

No

Between July and December, as long as it has fruits June – July

Edible fruit

Very light brown

No

All year long

Coal. Timber

Fruits

Light purple

No

June

Edible fruit

Leaves and leaves with flowers, without stem Fruits, flowers, leaves Flowers or flowering herb

Very light yellow or lemon green

No

Scourer for washing dishes

Very light purple

No

From October to December, its flowering June

Orangeyellow

No

Orange

Yes

Used part

The whole plant, mainly the stem Leaves Tuber (root) Fruits

All

Yes

No

Harvest season

From October to December, its flowering All year long

Other uses at the local level

Medicinal

Ornamental for the Día de Muertos Medicinal

612

C. Lo´pez-Binnqu¨ist et al.

recognized (Citlahua 2007): Poxavac (blue), Istak (white), Ahwatik (furry), Chokolatik (brown), Tlilivik (black), and Nehneltik (combined). However, the local sheep of the area are not recognized as breeds, due to the great variability and little durability of their characteristics compared to the so-called improved breeds, such as Merino, Suffolk or black face, and Pelibuey, introduced to the region through government programs. Local varieties, known for having a very suitable fleece for textiles, are being replaced by other breeds with much shorter, weaker wool. As raw material, the superiority of native sheep wool lies in the strength and length of the fleece, in addition to a wide range of colors. The color and type of wool obtained from local sheep are so important that they are the basis for their categorization; it includes white, black, various shades of brown, gray, and bluish gray, with different combinations and shades. Hybridization with the introduced breeds has reduced the color range of the local sheep, and now the varieties with black, gray, and bluish-gray fleece are almost extinct. Due to the characteristics of their wool – long and loose, growing about 14 cm per year – the sheep from the Sierra de Zongolica are considered to belong to the same genetic type (Citlahua and Vargas-López 2009). The fleece of local sheep is made of two layers: The outer layer is composed of a long, thick fiber that covers 10% of the fleece; and the inner layer of a short, fine fiber that covers 90% of it, in addition to the fat typical of double-layer fleeces. A further advantage of local wool is that it shrinks only once in the first wash, while that of improved breeds shrinks several times. Women artisans have a detailed knowledge of wool, its appearance, consistency, and handling. In addition to the characteristics of its wool, there are other important aspects regarding the management of sheep. First is the quality of their meat; while improved breeds look bigger and more robust, their fat level is very high compared to that of the local varieties that are practically free of fat, although their weight is low. Another important aspect is the greater adaptation to their environment; the local varieties are more tolerant to diseases and feed on scarce grasses and shrub species found on the roadsides, forest glades, and plots, and on by-products from family crops; they require little attention, endure the weather conditions, can live in rustic facilities, and are highly productive. Local sheep varieties provide wool for textile production, leather, manure that serves as a natural fertilizer, and meat for human consumption; in addition, its commercialization represents a source of income for the family unit (Citlahua 2007). In other words, local sheep are the ones that best adapt to both the climatic conditions and the management conditions to which they are subjected, offering greater benefits besides the wool used for textiles (Citlahua 2007). Sheep are an important part of family affections; feeding and caring for herds promotes family collaboration; the exchange of live sheep for their reproduction and their products generates active commercial and exchange links. For all these reasons, despite being an introduced species, its conservation is a concern of the Nahua communities in the Sierra de Zongolica.

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

613

Weaving Knowledges The history of textiles in Zongolica has – beyond natural dyes and sheep – a transversal category: knowledge. We recognize that one of the challenges, as ethnobotanical researchers studying the ways of life of Indigenous and rural communities, is to understand the generation, transmission, and meaning of knowledge and social practices that are linked to particular worldview. For this reason, this study privileged the knowledge dialogue, materializing it in two audiovisual products that incorporate the biocultural approach and reflections from the perspective of the landscape addressed in this chapter, in a format that enables their communication in various contexts and epistemic communities and makes enjoyment possible (Encina et al. 2009). The book Los textiles de la Tlasesekya: cuentos sobre hilos, telares y vellones is a storybook aimed at making visible a social group usually excluded from ethnobiological research: children. The content of this book comes from a dialogue between traditional and scientific knowledge and encourages intergenerational exchange of information between women artisans and their families. The book consists of three stories and a riddle section. It includes information on sheep and dye plants in their cultural, social, and environmental context, as well as aspects of the empowerment of artisans. Given the relevance that languages have as the main cultural instrument for the development, maintenance, and transmission of knowledge generated in practice (Boege 2008), it was decided to use both Nahuatl and Spanish in the production of the book. Due to the orality of the Nahuatl language, the book was written in Spanish and Nahuatl versions. Their content was included in audio format on a compact disc that accompanies the publication. This book was shared in the Sierra de Zongolica and in various states of Mexico in order to disseminate the biocultural value of artisan work, as well as being one of the didactic materials that supported the educational project “Life and knowledge: Nahua peoples of the Sierra de Zongolica and their biocultural resources” (Altepetlan neskayo ixtlamachilistli iwan tlenyoltok) (Contreras-Jaimes et al. 2015), a project developed with elementary school teachers and students in the region. The book and audio can be found in the following link: https://www. peopleandplants.org/knowledge-exchange-tools/community-manuals Another material produced was infographic, about the textile production process, which arose from the shared interest of the artisans in the permanence, understanding, and appreciation of their crafts. The information and images it contains are the result of collaborative work, through a research process and group meetings to review its content and images. The infographic material (Fig. 6) shows an overview of the production process, highlighting the value of local raw materials; the natural color palette and its esthetic value; the human component of the textile process; the information regarding the quantity of raw material and the time invested in each garment; and the varieties of sheep, the range of natural dyes, and the types of weaving. The myth of Tonantzin contributes the worldview of the origin of the artisanal work, illustrated with a photograph of the woodwork of the local sculptor, José Tentzohua Cervantes.

614

C. Lo´pez-Binnqu¨ist et al.

Fig. 6 This infographic describes the production process of wool textiles, linking the dimensions of the kosmos-corpus-praxis (Toledo and Alarcón 2012) in a process that starts with taking care of the sheep, the shearing, combing, washing, and spinning of the wool, the subsequent collection of dye species and dyeing, the preparation and arrangement of the thread, and the weaving on a waist loom. Months of work with versatile results in color and weave types. (Photo by the authors)

Photographs of plant species were taken on field trips with the artisans. Sheep photographs are courtesy of Eleuterio Citlahua Apale. The color palette was taken from the sample of dyed yarns used for this research. This infographic material has been presented at various events in the region, and it is used mainly as support material in artisan exhibitions and sales. In this way, artisans can inform buyers – used to haggling – about the work they do in terms of time, resources, and knowledge. The infographic contributes to the valuing of the artisans’ work.

Changes, Innovations, and Losses The case of wool textiles, from the Sierra de Zongolica, reveals the complex social and economic relationships within and outside the region, as well as the changes in the management of multiple uses in the regional landscape that have taken place. In particular, it shows the magnitude of external interventions in generating changes in the organizational processes, the flow of resources, and the resignification of the artisan work, as well as the way people relate to the environment. The use of dye plants and management forms of sheep reveal contrasting histories of knowledge and management of resources by the artisan community. On the one hand, the history of sheep, despite being a species introduced during the colonial period, is a history of management at the hands of the artisan community for five centuries. On the other hand, dye plants, naturally present in the environment of women artisans, were recognized as natural dyes just 28 years ago, and only after a governmental intervention. And these are the ways local knowledge and experience, conservation risks, and local concerns about the introduced species are connected. The dye plants occurring in the local flora are, for the most part, successional species favored by recent changes associated to the disturbance and the change of vegetation cover toward secondary vegetation. However, in the case of sheep, recent changes in

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

615

land use, reforestation, land fragmentation, and introduction of improved breeds are all factors putting in risk the permanence of local varieties and associated traditions. Citlahua (2007) points out that local sheep are threatened by economic, social, and demographic factors, in particular the fragmentation of properties, the introduction of reforestation programs, the construction of roads, and the introduction of new services. Such circumstances have reduced the spaces for breeding and grazing, forcing families to reduce the number of livestock, or to acquire small lands away from the houses to establish their pens, as well as the substitution of free grazing practices, for the ilpitinemi management system (Citlahua et al. 2004) (Fig. 7). It is also common to hear about conflicts between neighbors due to the invasion of sheep in areas where timber trees are planted or where agricultural activities are intended. Martínez-Canales (2010) reports that, in the municipality of Tehuipango, in the Sierra de Zongolica, those who remain in the community allow their animals (sheep and goats, mainly) to feed on the sprouts in plots temporarily abandoned due to migration. Upon returning, the migrant claims the act or even his own relatives are responsible for initiating the demand on his behalf. Regarding dye plants, the contribution of this chapter is to show the complexity behind the use of local flora. From the training received in 1992 by the government, in addition to establishing the use of local flora for dyeing purposes, the title of “artisans” was coined in the conception of women weavers, transforming textile work into a “job” promoting the organization of weavers’ groups and the interaction of the weavers with the world beyond their communities. Traditional forms of knowledge transmission, installed in the domestic space and in family ties, have taken new routes with the adoption of training schemes and the figures of instructor and apprentice. These practices have complemented orality, observation, and practice with exploration, experimentation, reading, and learning through audiovisual resources. Life of women is framed in a link with the natural environment, in which they self-manage their lives and activities in everyday spaces and times, carrying out tasks at both family and social levels (Howard 2003). In the Sierra de Zongolica, Fig. 7 Ilpitinemi management system (mooring in grazing areas). (Photo by the authors)

616

C. Lo´pez-Binnqu¨ist et al.

women maintain broader links with the landscape regarding land distribution, the diversification of crops, and the use of nontimber forest resources, among others. Their textile work strengthens their link as direct users of the ecosystem, since it takes them day by day to tour their territory during grazing and to be aware of the presence and abundance of dye plants and wool. However, beyond knowledge, use, and management of biodiversity, it was the social process of female empowerment through textile work (Sosme-Campos 2015) that significantly reconfigured the relationships of artisans with the environment. Once the women transcended the private space and influence the domestic and public life of their municipalities through their income, experiences, and the prestige obtained by their work, a change was detonated in the traditional role of women. The new purchasing power has fostered an internal network of trade of dye species and even of the textile production itself, which has promoted the specialization in craft production, the distribution of work, and the specialization of knowledge and work related to raw materials and textile production. This expanded view based on an ethnobotanical study allowed us to understand that, although the visible pressures in the transformation of the landscape affect the continuity of the production of textiles with the use of local wool and traditional techniques, the greatest latent risk lies in the abandonment of the activity by young women due to the low economic gains of a demanding and undervalued job, together with lifestyle changes. The current condition of the wool textiles production is explained by the dynamism between tradition and innovation, between local knowledge and women’s creativity. The tradition has been kept alive through accommodating innovation in the dyeing work to create new colors, expanding the range of garments to suit market parameters, renegotiating the conditions for obtaining raw materials under the new circumstances of land use and management, promoting the distribution and specialization of work. The ongoing tradition resignifies knowledge and practices based on ancestral and recent learning and reaffirms the artisanś personal and artistic fulfillment (Contreras et al. 2018).

Perspectives Beyond the case study, this analysis intends to provide reflections on the perspectives about ethnobotanical work. Placing ethnobotany in the third decade of the twenty-first century is a challenge in the face of the need and morality to go further, when realities far exceed their boundaries and require ethnobotanical work committed to local and global emergencies around the world (Albuquerque et al. 2019; Alexiades 2003). In the exercise of placing ethnobotany in front of the mirror, an essential aspect emerges from its origin, approaching the way in which others look at the world (Martin 1995), which adds to what continues to be a challenge today: its encounter with other epistemologies and the construction of transdisciplinary efforts with common objectives.

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

617

Various authors recognize that ethnobotany is by definition multidisciplinary (Alexiades 1996) and has a collaborative nature that involves both disciplinary integration (Sithole et al. 2002) and the relationship between scientific and popular knowledge (Cunningham 2001). Therefore, ethnobotany has faced the challenge of incorporating integrative research methods, and in this sense, participatory ethnobotany has gained strength (Alexiades 2003; Medley and Kalibo 2005). The analysis of this research experience revealed theoretical and methodological challenges that triggered creativity and reflection, promoting shared learning about ethnobotany from the biocultural landscape approach. Regarding ethnobotanical studies in mountain regions – with the great environmental heterogeneity and diversity of the peoples that inhabit them – the biocultural landscape approach allows a more comprehensive understanding, addressing the social and environmental complexity of these regions. In the Sierra de Zongolica, this approach allowed recognizing the diversity of multiple management spaces and the elements with cultural- symbolic value, which have remained a central part of the life of the Nahuas of Zongolica. The biocultural landscape approach, as suggested by several authors (Berkes and Davidson-Hunt 2006; Boege 2008; Agnoletti and Rotherham 2015), allows identifying long-term processes, as well as key events and factors of change that directly or indirectly contribute to the configuration of landscapes. In these processes, the plots and vegetation patches constitute negotiation spaces that women and men adapt, appropriate, or discard according to their family contexts, the conditions of their production units and their needs, and present and future expectations. The competition between self-consumption and market-oriented use has started to determine the uses of plots and vegetation. Examples include the way resources traditionally used for self-consumption began to circulate in markets outside the region, such as the wool textiles and wood furniture made from pine trees. All of the above reflects the encounter between different lifestyles, between generations, and between ways of conceiving the environment and its resources. It also reflects challenges that loom in the future of the Sierra de Zongolica textile production, such as the abandonment of the feminine tradition of weaving, the need to avoid the hybridization of the local sheep and to protect local varieties of sheep, and the reconstruction of artisan work that enables a higher quality of life for its makers. The methodology of this work was characterized by its flexibility, adapting to the needs of the research, the complexity of the case, and the capabilities of the work team. Reality showed a textile work determined by gender, influenced by the dimension of public policies, sustained by knowledge and by the ways in which it flows and interacts within artisanal communities and between epistemic communities, and in which forms of communicative interactions were essential. Therefore, to the initial ethnobotanical interest were added the contributions of disciplines that are not usually formally linked to ethnobotanical studies, and yet are usually carried out collaterally, such as communication (Encina and Ávila 2010) and Indigenous pedagogy (Gómez 2009). The pedagogical component was key to understanding otherness in the field of knowledge, recognizing that the various cognitive traditions (Bertely 2016) come

618

C. Lo´pez-Binnqu¨ist et al.

from the way the world is constructed from otherness (Carrillo-Medina et al. 2006). The role of communication in disciplinary and transdisciplinary relationships was recognized as enabling the construction of interfaces or bridges for exchanges of knowledge (Contreras-Jaimes 2015). This research experience leaves the team with a conclusive learning from encountering the realities of ethnobotanical work: the need to make structures – whether mental, disciplinary, epistemological, conceptual, or methodological – more flexible. From this experience, located under the peculiarities of the Sierra de Zongolica and working together with artisans, it is hoped to contribute from the narrative of a story in which the incorporation of the biocultural landscape approach, as well as the elements of analysis and action from fields of study such as communication and Indigenous pedagogy, allowed to broaden the scope of the ethnobiological project and bring it closer to a transdisciplinary, participatory, and dialogical research exercise. Acknowledgments The work presented in this chapter was carried out as part of the project “Community-based forest management and conservation in Central Mexico; building links, networks and capacities” financially supported by the Overbrook Foundation from 2010 to 2016 and conducted in collaboration with teachers-researchers and students from Universidad Veracruzana Intercultural and People and Plants International (www.peopleandplants.org). We thank Antonio Sierra Huelsz and Gabriela Alvarez Anaya for their revision and comments to this manuscript.

References Agnoletti M, Rotherham ID. Landscape and biocultural diversity. Biodivers Conserv. 2015;24: 3155–65. Aguirre-Beltrán G. Regiones de Refugio. El desarrollo de la comunidad y el proceso dominical en Mestizo-América. Ediciones especiales 46. México: Instituto Indigenista Interamericano; 1967. Aguirre-Beltrán G. Zongolica: Encuentro de dioses y santos patronos. México: Universidad Veracruzana; 1986. Alatorre-Frenk G, García-Campos H, Hidalgo-Ledesma R, López-Binnqüist C, Negreros-Castillo P, Sánchez-Hernández AI, Silva-Rivera E. Diálogos sobre cuencas, bosques y agua en Veracruz. Una propuesta metodológica intercultural. Mexico: Universidad Veracruzana; 2014. Albuquerque UP, Fernandes da Cunha LV, Farias-Paiva de Lucena R, Nóbrega-Alves RR, editors. Methods and techniques in ethnobiology and ethnoecology. New York: Springer; 2014. Albuquerque UP, Nascimento ALB, Soldati GT, Feitosa IS, Campos JLA, Hurrell JA, Hanazaki N, Medeiros P, Vasconselos R, Ludwinsky RH, Ferreira Júnior WS, Reyes-García V. Ten important questions/issues for ethnobotanical research. Acta Bot Bras. 2019;33(2):376–85. Alexiades MN. Introduction. In: Alexiades MN, editor. Selected guidelines for ethnobotanical research: a field manual. New York: New York Botanical Garden; 1996. Alexiades MN. Ethnobotany in the third millennium: expectations and unresolved issues. Delpinoa. 2003;45:15–28. Anawalt PR. Atuendos del México antigüo. Textiles del México de ayer y hoy. Edición especial de Arqueología Mexicana. 2005;10–19 Argueta-Villamar A. Introducción. In: Argueta A, Corona E, Hersch P, editors. Saberes colectivos y diálogo de saberes en México. Mexico: Universidad Nacional Autónoma de México, Instituto Nacional de Antropología e Historia, Ibero Puebla; 2011.

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

619

Beauregard-García L, Aquino-Rodríguez L, Anaya Y. La magia de los hilos. Arte y tradición en el textil de Veracruz. Mexico: Editora de Gobierno del estado de Veracruz de Ignacio de la Llave; 2008. Berkes F, Davidson-Hunt I. Biodiversity, traditional management systems, and cultural landscapes: examples from the boreal forest of Canada. Int Social Sci J. 2006;58:35–47. Bertely M. Políticas neoliberales y afectaciones territoriales en México: Algunos “para qués” de “otras” educaciones. Liminar. 2016;14(1):30–46. Boege E. El patrimonio biocultural de los pueblos indígenas de México: hacia la conservación in situ de la biodiversidad y agrodiversidad en los territorios indígenas. Mexico: Instituto Nacional de Antropología e Historia, Comisión Nacional para el Desarrollo de los Pueblos Indígenas; 2008. Bravo Marentes C. Inventario nacional de especies vegetales y animales de uso artesanal. Mexico: Asociación Mexicana de Arte y Cultura Popular A.C; 1999. Bravo Marentes C, Neyra González L. Especies vegetales y animales de uso artesanal. In: Cruz Marueta M, López Binnqüist C, Neyra González L, editores. Artesanías y medio ambiente. Mexico: Fondo Nacional para el Fomento de las Artesanías, Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2009. Brush SB. Crop diversity in mountain areas and conservation strategy. Rev Geogr Alp. 1998;86(4): 115–30. Caro-Bueno E, Cruz-Marueta M, Navarrete-Zamora N, López-Binnqüist C. Artesanías, medio ambiente y salud ocupacional. In: Cruz-Marueta M, López-Binnqüist C, Neyra-González L, editors. Artesanías y medio ambiente. Mexico: Fondo Nacional para el Fomento de las Artesanías, Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2009. Carrillo-Medina P, Jaulis-Cancho P, Núñez-Machaca M. Kururay y kawakuy. La vida, en un momento es saber ovillarse y en otro saber ser madeja. In: Calendario agrofestivo en comunidades andino-amazónicas y escuela. Peru: Proyecto Andino de Tecnologías Campesinas; 2006. Casas A, Camou A, Otero-Arnaiz A, Rangel-Landa S, Cruse-Sanders J, Solís L, et al. Manejo tradicional de biodiversidad y ecosistemas en Mesoamérica: el Valle de Tehuacán. Invest Ambient Cien Polít Públ. 2014;6:23–44. Citlahua AE. El modelo de producción ovina de las comunidades indígenas de la Sierra de Zongolica, Veracruz. Mater thesis. Mexico: Colegio de Postgraduados; 2007. Citlahua AE, Vargas-López S. La cría de borregos en la Sierra de Zongolica, Veracruz. Manual. Mexico: Fundación Produce Veracruz, Colegio de Postgraduados; 2009. Citlahua AE, Vargas S, Hernández JS. Ovinocultura en comunidades náhuatl de la Sierra de Zongolica, Veracruz (México). Memorias V Simposio Iberoamericana Sobre la Conservación y Utilización de Recursos Zoogenéticos. Peru: Universidad Nacional del Altiplano; 2004. CONAFOR. Mapa de la cobertura del suelo de México del año 2015. Mexico: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad (CONABIO), Comisión Nacional Forestal (CONAFOR) and Instituto Nacional de Estadística y Geografía (INEGI); 2017. Contreras B, Sosme MA. Los textiles de la Tlasesekya: cuentos sobre hilos, telares y vellones. Mexico: Centro de Investigaciones Tropicales, Fundación Overbrook; 2014. Contreras B, Panzo F, Álvarez L. Altepetlan neskayo ixtlamachilistli iwan tlenyoltok. Vida y saberes: Pueblos nahuas de la sierra de Zongolica y sus recursos bioculturales. [Unpublished manuscript]. Mexico: People and Plants International; 2015. Contreras B, Panzo F, López C. ArteSano: Autogestión del Kualli nikah, el “estado de estar”. In: Encina J, Urteaga E, Ezeiza A, editors. Autogestión cotidiana de la salud. Spain: Volapük Ediciones; 2018. Contreras-Jaimes B. Reconocimiento del valor biocultural de la producción artesanal a través del intercambio de saberes: El caso de los textiles de lana en Tlaquilpa, Veracruz. Master thesis. Mexico: Universidad Veracruzana; 2015.

620

C. Lo´pez-Binnqu¨ist et al.

Cornejo-Rodríguez FJ. Introducción. In: Cruz-Marueta M, López-Binnqüist C, Neyra-González L, editors. Artesanías y medio ambiente. Mexico: Fondo Nacional para el Fomento de las Artesanías, Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2009. Cunningham AB. Applied ethnobotany: people and plants conservation manual. London: Earthscan Publications; 2001. De la Borbolla SR. Las artesanías como patrimonio cultural y el uso de nuevas tecnologías en un mundo globalizado. Conferencia en el XXXII Coloquio de Antropología e Historia Regionales, Artesanías y Saberes Tradicionales; 2010 Video. Recuperado en http://etzakutarakua.colmich. edu.mx/videoteca/eventos/XXXIIColoquio/j_m3.htm El Hidalgo-Ledesma R. campesinado Zongoliqueño en la disputa territorial, saberes territoriales y manejo campesino del monte. Master Thesis. Mexico: Universidad Autónoma Metropolitana; 2016. Encina J, Ávila MA. El ilusionismo social: más allá de la última frontera metodológica. In: Encina J, Ávila MA, Begoña L, editors. Las culturas populares. Spain: Editorial Atrapasueños; 2010. Encina J, Ávila MA, Domínguez M, Alcón R. Los principios del ilusionismo social. Spain: Editorial Atrapasueños; 2009. Gómez M. Saber indígena y medio ambiente: experiencias de aprendizaje comunitario. In: La complejidad ambiental. Mexico: Siglo XXI; 2009. Howard PL. Women and the plant world: an exploration. In: Howard PL, editor. Women & plants. Gender relations in biodiversity management & conservation. London: Zed Books; 2003. Kirchhoff P, Odena-Güemes L, Reyes-García L. Historia tolteca-chichimeca. México: Centro de Investigaciones Superiores del Instituto Nacional de Antropología e Historia, Instituto Nacional de Antropología e Historia, Secretaría de Educación Pública; 1976. Larios-León S. Textiles de lana de la Sierra de Zongolica. Catálogo. Mexico: Gobierno del Estado de Veracruz; 1991. López-Austín A. Tamoachan y Tlalocan. Mexico: Fondo de Cultura Económica; 1994. López-Binnqüist C, Negreros-Castillo P, De la Hidalga-Ledesma V. Colaboración e intercambio de saberes en torno al manejo campesino de árboles maderables en la Sierra de Zongolica, Veracruz. Ponencia para el IX Congreso Internacional de Sociología Rural Asociación Latinoamericana de Sociología Rural ALASRU. Mexico; 2014. Martin GJ. Ethnobotany: a methods manual. New York: Routledge; 1995. Martínez-Canales LA. La dinámica comunitaria vista desde la migración en la sierra nahua de Zongolica, Veracruz: análisis y perspectivas de estudio. Migraciones internacionales. 2010;5(4): 7–37. Martínez-Peñaloza P. Permanencia, cambio y extinción de la artesanía en México. México: Fondo Nacional para el Fomento de las Artesanías; 1986. Mastache AG. El tejido en el México antigüo. Textiles del México de ayer y hoy. Edición especial de Arqueología Mexicana 2005: 20–31. Medley KE, Kalibo HW. An ecological framework for participatory ethnobotanical research at Mt. Kasigau, Kenya. Field Method. 2005;17(3):302–14. Mestries-Benquet F. Crisis cafetalera y migración internacional en Veracruz. Migraciones Internacionales. 2003;2(2):121–48. Ortiz V. Algunos antecedentes sobre el uso del color en el México indígena. Revista Historia de las Indias de Nueva España e Islas de Tierra Firme. 1991;48(7):105–14. Reyes-Sánchez I. Las estrategias de sobrevivencia de los cafeticultores en la zona de Zongolica, un contexto político-agrícola: 1980–2011. Master thesis. Mexico: Universidad Veracruzana; 2016. Rodríguez MT. Ritual, identidad y procesos étnicos en la sierra de Zongolica. Veracruz, Mexico: Centro de Investigaciones y Estudios Superiores en Antropología Social; 2003. Rodríguez MT, Hasler A. Nahuas de Zongolica. Mexico: Instituto Nacional Indigenista; 2000. Sarmiento FO. Breaking mountain paradigms: ecological effects on human impacts in man-aged tropandean landscapes. Ambio. 2000;29(7):423–31.

Wool Textiles of the Sierra de Zongolica, Mexico. The Reshaping of. . .

621

Sevilla RA, Licona NEA, De la Luz Sevilla-González M. The “Xochitlalli system”, an adoration ritual to mother earth among the sierra de Zongolica Nahuas in Veracruz. Adv Anthropol. 2020;10:33–47. Silva P, Cortés-Belmar M, Arce A. Public good: wheat assemblages and the revalorization of culinary and handicraft practice in Bio-Bio, Chile. In: Sherwood S, Arce A, Paredes M, editors. Food, agriculture and social change: the everyday Vitality of Latin America. Routledge Taylor & Francis Group; 2017. Sithole B, Frost P, Veeman TS. Searching for synthesis: Integrating economic perspectives with those from other disciplines. In: Campbell BM, Luckert MK, editors. Uncovering the hidden harvest: Valuation methods for woodland and forest resources. People and Plants Conservation Series. London: Earthscan Publications; 2002. Sosme-Campos MA. Tejedoras de Esperanza. Empoderamiento en los grupos artesanales de la sierra de Zongolica. Mexico: El Colegio de Michoacán; 2015. Terry C. Weaving social change (s) or changes of weaving? The ethnographic study of Andean textiles in Cusco and Bolivia. Artl@ s Bulletin. 2020;9(1):68–89. Toledo VM. Del “diálogo de fantasmas” al “diálogo de saberes”: conocimiento y sustentabilidad comunitaria. In: Saberes colectivos y diálogo de saberes en México. Mexico: Universidad Nacional Autónoma de México, Centro Regional de Investigaciones Multidisciplinarias, Universidad Iberoamericana; 2011. Toledo VM, Alarcón CP. La Etnoecología hoy: Panorama, avances, desafíos. Etnoecológica. 2012;9(1):1–16. Trueba S. Plantas tintóreas de Soledad Atzompa, Veracruz, México. Master thesis. México: Universidad Veracruzana; 2008. Turok M. Cómo acercarse a la artesanía. México: Plaza y Valdés; 1988. Velasco-Rodríguez G. Origen del textil en Mesoamérica. México: Instituto Politécnico Nacional; 2002. Wallace J, Aquilué N, Archambault C, Carpentier S, Francoeur X, Greffard MH, Laforest-LapointeI, Galicia L, Messier C. Present forest management structures and policies in temperate forests of Mexico: challenges and prospects for unique tree species assemblages. Forest Chron. 2015;91(3):306–17. Zimmerer KS, Cordova-Aguilar H, Mata-Olmo R, Jimenez-Olivencia Y, Vanek SJ. Mountain ecology, remoteness, and the rise of agrobiodiversity: tracing the geographic spaces of human–environment knowledge. Ann Am Assoc Geogr. 2017;107(2):441–55.

Ethnobotanical Knowledge and the Patterns of Plant Use and Management in the Sierra de Huautla Biosphere Reserve and the Chichinautzin Biological Corridor in Morelos, Mexico Jose´ Juan Blancas Vázquez, Araceli Tegoma-Coloreano, Itzel Abad-Fitz, Leonardo Beltra´n-Rodríguez, Belinda Maldonado-Almanza, María Idalia Villalpando-Toledo, Fabiola Mena, Ange´lica Alema´n, and Amanda Ortiz-Sa´nchez Abstract

The Sierra de Huautla Biosphere Reserve (REBIOSH by its acronym in Spanish) and the Chichinautzin Biological Corridor (COBIO) are mountain zones in the state of Morelos, Mexico, inhabited by peoples conserving a remarkable biocultural heritage. In the REBIOSH, the dominant vegetation is the tropical deciduous forest characterized by its marked seasonality. The region harbors a considerable ethnobotanical richness with 1024 useful species recorded, most of them native, belonging to the families Asteraceae, Fabaceae, Poaceae, and Lamiaceae. Contrastingly, the COBIO is covered by a heterogeneous array of vegetation types, but with 576 useful plant species of the families Asteraceae,

J. J. Blancas Vázquez (*) · I. Abad-Fitz · B. Maldonado-Almanza · M. I. Villalpando-Toledo · F. Mena · A. Alemán · A. Ortiz-Sánchez Centro de Investigación en Biodiversidad y Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico e-mail: [email protected]; [email protected]; [email protected]; [email protected]; [email protected]; [email protected]; [email protected] A. Tegoma-Coloreano Centro de Investigaciones Biológicas (CIB), Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico e-mail: [email protected] L. Beltrán-Rodríguez Laboratorio de Etnobotánica Ecológica, Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_18

623

624

J. J. Blancas Vázquez et al.

Fabaceae, Lamiaceae, and Solanaceae. In both mountain areas, nearly 1600 plant species are used, mainly for medicine, ornament, food, and fuel. Despite the differences in dominant vegetation types, both regions are similar regarding the forms of management of some species, which reveals a shared cultural heritage. The ethnobotanical knowledge in both natural protected areas is endangered by migration, land use change, and changes in the cultural patterns, which have a negative impact on the forms of use and management of plant resources. Because these regions represent havens of traditional knowledge and peoples’ cultural heritage, it is essential to systematize and analyze the pattern of plant use to establish priorities for the conservation of biocultural heritage. This chapter is a contribution in this direction.

Introduction The state of Morelos – located in central Mexico – covers only 0.3% of the total country’s surface, but its human population occupies the fifth place among the 32 states of the country. Because of its proximity to Mexico City and its touristic infrastructure including hundreds of recreation areas and aquatic parks, the state is one of the preferred destinations for thousands of visitors from the capital city. While Morelos has recently acquired a touristic vocation, its territory has staged human-nature interactions since at least 8000 years ago, when human populations dwelled in the southern part of the state (Juárez-Mondragón 2019), and the first permanent settlements date to 4000 years ago (Corona 2018). Archaeological sites in Chalcatzingo (3500 B.P.; Grove 1984), Xochicalco (1400 B.P.; Garza and Palavicini 2002), the Tepozteco (800 B.P.; INAH 2022), and Teopanzolco (700 B.P.) witness the presence in the state of the Olmec, Olmec-Xicalanca, Tlahuica, and Mexica Mesoamerican cultures (INAH 2022). During the sixteenth century, the territory of the present state of Morelos was used by the Spaniards as an experimental zone for the acclimatation of exotic crops to sustain the economy of the colony. Between 1527 and 1530, the first plantation of sugar cane (Saccharum officinarum L.) in Mexico was established near Cuauhnáhuac, the current capital city of Cuernavaca (Crespo 2018), together with the planting of mulberry trees (Morus alba L.) for rearing silkworms (Bombyx mori Linnaeus 1758; García 2018). The transformation of the landscape during the colonial period was so drastic that, by 1550, cotton – a crop of Mesoamerican origin – was replaced by mulberry trees (Maldonado 2018). The ethnobotanical importance of the present state of Morelos during the early colonial period is underscored by the exploration and research made by Francisco Hernández – court physician of Philip II, whose results were published in the Historia Natural de la Nueva España, becoming an essential ethnographic source for documenting the history of disease and medicinal plants in colonial Mexico. The work of Francisco Hernández not only listed the uses of plants of the New Spain, but it was one of the first records in the American Continent of what we now know as medical ethnobotany (Zolla 2018). Hernández documented the use in

Ethnobotanical Knowledge and the Patterns of. . .

625

treatments, forms of preparation, collection sites, vegetation types, and cultural contexts of over 519 medicinal plant species distributed in the present state of Morelos (Parodi 1991). During the colonial period, several explorers and naturalists studied the natural resources of Mexico, perhaps the most noticeable being Alexander Von Humboldt, who arrived in the country in 1803. Humboldt briefly visited the state of Morelos on his way from Acapulco to Mexico City. Perhaps the most relevant observations made during Humboldt’s transient visit to the state are embodied in his description of the contrasting climates of Morelos from the warm southern to the temperate northern portions, and his estimation of the elevation above the sea level of Cuernavaca – then known as Cuahnáhuac. Humboldt was one of the first to warn about the negative consequences of the introduction of exotic crops and livestock on soils and vegetation. The introduction of rice (Oryza sativa L.) during the first half of the nineteenth century, particularly in southern Morelos, had strong repercussions on the use of vegetation (SADER 2021), many of which persist until the present. Rice cultivation exacerbated the transformation of the landscape, above all in the tropical deciduous forest covering nearly 70% of the state’s territory. One of the largest consequences of rice cultivation was the way in which water was managed, because rice demands large quantities of irrigation and Morelos is relatively scarce in water due to one of the highest evapotranspiration rates in the country (Maderey 2002). In the twentieth century, the Mexican Revolution left a deep imprint in the way in which the vegetation was managed in Morelos due to the creation of the Ejido, a regime of social land tenure. The Ejido was a revolutionary conquest because during the dictatorship of Porfirio Díaz, private tenants accumulated land at a scandalous level. Except for a few zones, most of the state of Morelos’s territory was divided among hundreds of Ejidos having their own authorities and regulations. It was in Morelos where Emiliano Zapata was born, one of the most prominent leaders of the Mexican Revolution, he developed a political ideology summarized in the renowned expression la tierra es de quien la trabaja –translated as “the land belongs to those who work it with their own hands” – adopted by thousands of rural communities throughout Mexico, even nowadays. These historical events are closely related with the present forms of vegetation management. Currently, most of the state’s surface is covered by forests (58%) in different conservation status, followed by agriculture and livestock (34%), and human settlements (10%) (Escandón-Calderón et al. 2018). The zones covered by natural vegetation in relatively good conservation status are concentrated to the south and north of the state where the main two protected natural areas are located. These are, respectively, the Sierra de Huautla Biosphere Reserve (REBIOSH by its acronym in Spanish) and the Chichinautzin Biological Corridor (COBIO). In both of these protected areas, a diversity of traditional communities use and manage a large number of plant resources and maintain a rich biocultural heritage about the management of natural ecosystems and landscape components. Because of that, the compilation of this knowledge might become an essential contribution to the sustainable management of natural areas in the region. In this chapter, we provide

626

J. J. Blancas Vázquez et al.

a general account of the use and management of plants in these mountain zones of the state of Morelos, which are in many ways contrasting. We in addition discuss some viewpoints regarding the topics needed to be addressed by future ethnobotanical research in both regions.

Sources of Information for the Study of the Ethnobotany of the Mountain Zones of Morelos, Mexico For the present research, we made use of 42 sources of information, including scientific papers, dissertations, books, historical archives, technical reports, and databases (22 for the REBIOSH, 16 for the COBIO, and 4 for the whole state of Morelos). Three of these sources were the most important since they summarize regional views. For the REBIOSH, the study by Maldonado (1997), while for the COBIO, the book by Monroy-Ortiz and Castillo-España’s (2007), and also outstanding was the database Etnoflora de Morelos, which represents over 5-year effort of our research team in the Centro de Investigación en Biodiversidad y Conservación (CIByC – UAEM). The database summarizes and documents the use and management of plant resources in the state of Morelos.

Ethnobotany of the Sierra de Huautla Biosphere Reserve Ecological Context The REBIOSH is a natural protected area in southern Morelos that was decreed in 1999 because of its importance in the conservation of the tropical deciduous forest. It represents a reservoir of species endemic to Mexico due to its mountainous topography with elevations from 700 to 2400 m a.s.l., which favors the formation of microhabitats (Fig. 1). It has a surface of 59,000 ha and its climate is warm subhumid with annual mean temperature of 22.5  C and average yearly precipitation of 906 mm, mostly falling between June and September. The dry season occurs from October to May, with a driest and warmest period from March to May (Maldonado 1997; CONANP 2005; Bolongaro and Torres 2020). The water bodies in the region are mainly seasonal ephemeral streams. The most extended vegetation is tropical deciduous forest characterized by trees 5 to 10 m in height dominated by Amphipterygium adstringens (Schltdl.) Standl., Bursera aptera Ramírez, Bursera longipes (Rose) Standl., Bursera morelensis Ramírez, Conzattia multiflora (B.L. Rob.) Standl., Ceiba aesculifolia subsp. parvifolia (Rose) P.E. Gibbs & Semir, Ipomoea murucoides Roem. & Schult., Lysiloma acapulcense (Kunth) Benth., and Lysiloma divaricatum (Jacq.) J.F. Macbr. In the more humid zones like canyons and ravines, the most representative species are Astianthus viminalis (Kunth) Baill., Bursera grandifolia (Schltdl.) Engl., Enterolobium cyclocarpum (Jacq.) Griseb., Euphorbia tanquahuete Sessé & Moc., Ficus petiolaris Kunth, Guazuma ulmifolia Lam., Salix humboldtiana Willd., and Sapindus saponaria L. In zones disturbed by human activities, Acacia farnesiana (L.)

Ethnobotanical Knowledge and the Patterns of. . .

627

Fig. 1 Location of the Chichinautzin Biological Corridor (COBIO) and the Sierra de Huautla Biosphere Reserve (REBIOSH)

Willd., Acacia cochliacantha Humb. & Bonpl. ex Willd., Pithecellobium acatlense Benth., and Mimosa polyantha Benth. are dominant species (Maldonado 1997; Fig. 2).

Sociocultural Context of the Peoples in the Sierra de Huautla Biosphere Reserve The history of human settlements in the REBIOSH is comparatively more recent than in other regions of Morelos. Although human occupation in the zone dates back 3000 years, permanent settlements are 1000 years younger, as seen in the archaeological sites in the current localities of Chimalacatlán and Nexpa. At present, the southern region has lower population densities than the northern and central portions of the state (CONANP 2005). The permanence of the settlements in the REBIOSH that subsisted after the Spanish invasion, and those founded later, was linked to four main events: (i) the exploitation of gold and silver mines; (ii) the introduction of livestock; (iii) the cultivation of sugarcane; and (iv) the creation of Ejidos. Most of the current settlements in the

628

J. J. Blancas Vázquez et al.

Fig. 2 Main vegetation types in the Sierra de Huautla Biosphere Reserve. (a) and (b) Aspect of the Tropical Deciduous Forest in the rainy season; (c) and (d) contrast between the rainy and dry seasons in the REBIOSH; (e) Cerro Frío is the highest mountain in the Sierra de Huautla Biosphere Reserve with 2204 m.a-s.l.; and (f) Oak Forest and an ecotone between Oak Forest and Tropical Deciduous Forest are the predominant types of vegetation in Cerro Frío. José Blancas, with the exception of (b) Darely Acosta

REBIOSH were established in the 1930s, mainly because of the agrarian distribution. In many cases, the creation of Ejidos stimulated the immigration from other parts of the state or from neighboring states like Guerrero, Puebla, and the State of México. This fact caused that the demographic composition of these settlements is heterogeneous. Although most people consider themselves as mestizos, the Nahuatl cultural inheritance is prominent regarding the forms of vegetation management, gastronomy, and traditional botanical nomenclature. Within the REBIOSH, there are 31 rural communities with a population of 23,920 inhabitants organized into 27 Ejidos

Ethnobotanical Knowledge and the Patterns of. . .

629

(CONANP 2005; INEGI 2021). All these rural communities are classified as high marginalized because not all their inhabitants have access to public services like health care, education, sewage, drinking water, and roads. The main economic activities of people within the REBIOSH are rainfed agriculture of basic crops like maize (Zea mays L.), beans (Phaseolus vulgaris L.), chili pepper (Capsicum annuum L.), and squash (Cucurbita pepo L.), and cultivation of commercial crops like sorghum (Sorghum bicolor (L.) Moench), sugarcane (Saccharum officinarum L.), hibiscus tea (Hibiscus sabdariffa L.), sesame (Sesamum indicum L.), and pitaya (Stenocereus stellatus (Pfeiff.) Riccob.). People obtain supplementary income from gathering and selling firewood, medicinal plants, and other non-timber forest products like copal resin (Bursera spp.; Maldonado 1997; Trujillo and López-Medellín 2018; Fig. 3).

Patterns of Use and Management of the Flora of the REBIOSH Useful Flora The useful flora of the REBIOSH includes 1024 species in 153 families and 585 genera. The families with most ethnobotanical importance are Asteraceae, Fabaceae, Poaceae, Lamiaceae, Malvaceae, Solanaceae, Euphorbiaceae, Apocynaceae, Cactaceae, Boraginaceae, Rubiaceae, Asparagaceae, Crassulaceae, Moraceae, and Araceae. These 15 families concentrate 51.75% of the useful flora, the remaining 48.24% species are distributed among 138 families (Table 1). These numbers agree with Maldonado’s (1997, 2013) finding of the useful plants within the REBIOSH and the Upper Balsas Basin. The genera with more useful species are Bursera, Euphorbia, Ficus, Acacia, Ipomoea, Sedum, Senna, and Solanum (Table 2), which might be due to their ecological importance, ease for propagation, and high number of uses of their species. For example, 12 species of Bursera have natural distributions in the REBIOSH and all have at least one use, mostly as living fence, construction, medicinal, and ritual purposes. Also, 11 species of the genus Acacia are native to the REBIOSH and have medicinal, fodder, and fuel uses. In contrast, genera like Cochlospermum have only one useful species, perhaps because it is basically used as ornament. Provenance of Useful Species Most useful flora of the REBIOSH, 740 (72.27%) species are native and 281 (27.44%) are exotic, the remaining species having an uncertain origin. According to CONANP (2005), the total and useful flora in the region includes 936 and 602 species, respectively; however, this accounting only considered the native plant species, that is, those naturally distributed in Mexico. In the present research, we recorded 1024 useful species including the native and exotic floras. We estimated that 740 of these useful species are native to Mexico and 520 species are native to the REBIOSH. Our results show an increase (123%) in the useful flora of the REBIOSH, mainly due to the considerable increment in ethnobotanical studies conducted in the region during the last 17 years. The exotic useful flora plays an important role in meeting the needs of the local people. It may displace the native flora (Ramírez et al. 2020), but provides valuable

630

J. J. Blancas Vázquez et al.

Fig. 3 Main economic activities of the inhabitants of the Sierra de Huautla Biosphere Reserve. (a) Seasonal agriculture; (b) cattle raising; (c) propagation of fruit trees in homegardens; (d) firewood collection; (e) cultivation of commercial crops such as pitaya (Stenocereus stellatus); collection of non-timber forest products such as (f) cuachalalate bark (Amphipterygium adstringens); and (g) bonete fruit harvest (Jacaratia mexicana). José Blancas, with the exception of (a) Darely Acosta

resources, mainly for ornamental use. Our record of 281 exotic species in the REBIOSH reveals a potential high gene flow promoted by people for different purposes, which should call the attention of conservation regulation policies.

Growth Habit Regarding the growth habit, the useful plant species include 541 herbs (52.83%), 273 trees (26.66%), 161 shrubs (15.72%), 29 climbing plants (2.83%), 14 palms and rosettes (1.37%), and 6 epiphytes (0.59%). This pattern of growth habit proportions

Ethnobotanical Knowledge and the Patterns of. . .

631

Table 1 Most important plant families by number of useful species in the Sierra de Huautla Biosphere Reserve, Morelos, Mexico Family Asteraceae Fabaceae Poaceae Lamiaceae Malvaceae Solanaceae Euphorbiaceae Apocynaceae Cactaceae Boraginaceae Rubiaceae Asparagaceae Crassulaceae Moraceae Araceae Apiaceae Cucurbitaceae Verbenaceae Amaranthaceae Burseraceae Convolvulaceae Rosaceae Rutaceae Anacardiaceae Acanthaceae Arecaceae Bignoniaceae Piperaceae Amaryllidaceae Aristolochiaceae Euphorbiaceae Nyctaginaceae Urticaceae Bromeliaceae Caprifoliaceae Commelinaceae Lythraceae Onagraceae Passifloraceae Polygonaceae Vitaceae Brassicaceae

Number of useful plant species 122 105 40 34 31 28 27 23 21 18 18 16 16 16 15 14 14 13 12 12 12 12 12 11 9 9 9 9 8 8 8 8 8 7 7 7 7 7 7 7 7 6

Percentage 11.91 10.25 3.91 3.32 3.03 2.73 2.64 2.25 2.05 1.76 1.76 1.56 1.56 1.56 1.46 1.37 1.37 1.27 1.17 1.17 1.17 1.17 1.17 1.07 0.88 0.88 0.88 0.88 0.78 0.78 0.78 0.78 0.78 0.68 0.68 0.68 0.68 0.68 0.68 0.68 0.68 0.59 (continued)

632

J. J. Blancas Vázquez et al.

Table 1 (continued) Family Fagaceae Lauraceae Plantaginaceae Sapindaceae Scrophulariaceae Geraniaceae Malpighiaceae Oleaceae Orchidaceae Papaveraceae Begoniaceae Ericaceae Fabaceae Juglandaceae Musaceae Myrtaceae Plumbaginaceae Primulaceae Pteridaceae Ranunculaceae Rhamnaceae Sapotaceae Smilacaceae Annonaceae Equisetaceae Iridaceae Loranthaceae Oxalidaceae Pinaceae Polypodiaceae Salicaceae Araliaceae Asphodelaceae Balsaminaceae Boraginaceae Campanulaceae Cannabaceae Capparidaceae Caricaceae Caryophyllaceae Casuarinaceae Celastraceae Cleomaceae

Number of useful plant species 6 6 6 6 6 5 5 5 5 5 4 4 4 4 4 4 4 4 4 4 4 4 4 3 3 3 3 3 3 3 3 2 2 2 2 2 2 2 2 2 2 2 2

Percentage 0.59 0.59 0.59 0.59 0.59 0.49 0.49 0.49 0.49 0.49 0.39 0.39 0.39 0.39 0.39 0.39 0.39 0.39 0.39 0.39 0.39 0.39 0.39 0.29 0.29 0.29 0.29 0.29 0.29 0.29 0.29 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 (continued)

Ethnobotanical Knowledge and the Patterns of. . .

633

Table 1 (continued) Family Combretaceae Costaceae Dryopteridaceae Hydrangeaceae Hypericaceae Juncaceae Krameriaceae Lamiaceae Loasaceae Meliaceae Myrtaceae Orobanchaceae Petiveriaceae Polemoniaceae Portulacaceae Rutaceae Santalaceae Schisandraceae Selaginellaceae Theaceae Zygophyllaceae Acanthaceae Araucariaceae Betulaceae Bixaceae Burseraceae Cactaceae Calochortaceae Calophyllaceae Chrysobalanaceae Cistaceae Cupressaceae Cupressaceae Cycadaceae Cyperaceae Dennstaedtiaceae Dioscoreaceae Ebenaceae Ephedraceae Gelsemiaceae Gentianaceae Gesneriaceae Hypoxidaceae

Number of useful plant species 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

Percentage 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 (continued)

634

J. J. Blancas Vázquez et al.

Table 1 (continued) Family Liliaceae Malvaceae Martyniaceae Meliaceae Moringaceae Myristicaceae Nephrolepidaceae Opiliaceae Pedaliaceae Phrymaceae Phyllonomaceae Phytolaccaceae Piperaceae Polygalaceae Saxifragaceae Simaroubaceae Strelitziaceae Styracaceae Thymelaeaceae Tropaeolaceae Typhaceae Viburnaceae Violaceae Zamiaceae Zingiberaceae

Number of useful plant species 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

Percentage 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.10

of useful plants may be determined by several factors, one of them depending on the growth habits abounding in the vegetation types (Caballero et al. 1998). The high number of useful herbaceous species might be a consequence of the long history of human-plant interaction in the tropical deciduous forest the long history of management of which favors the development of herbaceous species (Maldonado 1997; Caballero et al. 1998; Maldonado et al. 2013).

Plant Use Categories We recorded 15 use categories in the REBIOSH, among which medicinal use is the most frequent (72.66%) with 744 plant species, followed by 128 ornamental species (12.5%), 120 species of edible plants (11.72%), and 78 species used as fuel (7.62%; Table 3). The use categories we recorded generally agree with previous reports of ethnobotanical studies made in Mexico and elsewhere in the world (Caballero et al. 1998). The predominance of medicinal use in the REBIOSH is related with the tropical deciduous forest being the vegetation type that provides the largest number of medicinal plants in Mexico (Beltrán-Rodríguez et al. 2017). We observed a growing number of plant species used as ornamental, a trend that has been occurring during the past two decades even above the increase of edible

Ethnobotanical Knowledge and the Patterns of. . . Table 2 Main genera by number of useful plant species in the Sierra de Huautla Biosphere Reserve, Morelos, Mexico. The 20 genera listed in the table grouped 236 useful plant species, the remaining 708 genera had between 1 and 7 useful species, and 452 genera were represented by only one useful plant species

Genera Bursera Euphorbia Ficus Acacia Ipomoea Sedum Senna Solanum Mimosa Aristolochia Citrus Salvia Opuntia Passiflora Agave Cordia Piper Quercus

635 Number of useful plant species 12 12 12 11 10 10 10 10 9 8 8 8 7 7 6 6 6 6

Table 3 Plant use categories reported for the flora of the Sierra de Huautla Biosphere Reserve, Morelos, Mexico. Percentages do not add to 100 because several species have more than one use Uses Medicinal Ornamental Edible Firewood Construction Fooder Tools and utensils Living fence Ritual Shade Wood Handicrafts Poison Dyer Aromatize

Number of species in category 744 128 120 78 77 39 25 20 19 15 8 6 3 2 1

Percentage 72.66 12.50 11.72 7.62 7.52 3.81 2.44 1.95 1.86 1.46 0.78 0.59 0.29 0.20 0.10

plant species, which might be due to the modification of plant-use patterns associated to processes of cultural change (Arjona-García et al. 2021). As stated by Villalpando-Toledo (2020), local culture changes have modified the composition and structure of agroecosystems, a transformation that is evident in homegardens. These were formerly conceived as production units aiming at supplementing the needs of households, but nowadays homegardens are mainly used

636

J. J. Blancas Vázquez et al.

for cultivating ornamental plants. Alternatively, the predominance of ornamental plants might be due to the present higher number of studies addressing this use category, which could influence a false perception of an increased frequency in their cultivation.

Ecological Status A total of 474 (46.29%) useful plant species are wild, 294 are cultivateddomesticated (28.71%), 284 are ruderal (27.73%), and 45 are weeds (4.39%). The cross-examination of ecological status and use category shows that 70% of the medicinal plants and 79.45% of the fuel plants are wild, 60.24% of the ornamental plants and 46.62% of the edible plants are domesticated, and 51.28% of the fodder plants are weeds or ruderal plants. Our results suggest that the suitability of using particular environments – both natural vegetation and anthropogenic areas – depends on the plant resources occurring in these environments (Toledo et al. 2008). Thus, it is common that people use forests for gathering firewood, construction materials, or medicinal plants, while anthropized spaces are preferred over forests to obtain edible plants that normally do not prosper in forests. There are evident exceptions to that, but, more frequently, edible plants from forests are seen as supplementary or emergency foodstuffs (Blancas et al. 2013). The latter considerations are strongly rooted in peoples with a strong agricultural tradition, as is the case of most Mesoamerican cultures. Management Categories We recorded four plant management categories in the REBIOSH that are not exclusive because a plant species can be managed in more than one way. Local people gather 718 (70.12%) plant species in the surrounding forest and anthropized areas, 293 (28.61%) plant species are either incipiently or intensely cultivated by several methods (through seed or vegetative propagation), 20 (1.4%) plant species are tolerated, and 3 (0.2%) plant species are protected. There is a noticeable lack of information about plant management in the REBIOSH and the ethnobotanical studies made in the region include few mentions about it. For many years ethnobotanical studies were limited to listing useful plants, it was until the late twentieth century when topics and methods applied by ethnobotanists became more diverse (Phillips and Gentry 1993). However, plant management in the REBIOSH is far from having been documented, and even plant gathering follows appropriation strategies beyond cropping from nature, involving complex decisions based on socioeconomic factors (González-Insuasti et al. 2008). The management of copal (Bursera spp.) provides an example of the lack of clarity in the documentation of plant management. The sources of information predating the works of Mena (2018) and Abad-Fitz et al. (2020) classified the extraction of copal products as simple gathering, but these studies documented that current interactions range from gathering from wild populations to harvesting resin from trees cultivated in agroforestry systems. Cultivation of these trees may be by means of cuttings, and more recently, from seeds. For many details, see the chapter Bursera spp. in this book.

Ethnobotanical Knowledge and the Patterns of. . .

637

Use of Medicinal Plants in the REBIOSH: Current Issues and Strategies for Their Sustainable Management As mentioned above, we have recorded 744 plant species with medicinal use in the REBIOSH, 581 of them are native to Mexico and 295 have natural distributions in the reserve. We also mentioned that medicinal use is predominant in the REBIOSH and argued that this preponderance is related to the richness of medicinal plants present in the tropical deciduous forests (Beltrán-Rodríguez et al. 2017). The abundance of medicinal plants in tropical deciduous forests may be due to the presence in them of structures that accumulate the secondary metabolites on which medicinal uses are based (Soto-Nuñez and Sousa 1995). In that regard, Albuquerque (2010) hypothesized that seasonal forests are preferred by people as sources of gathered medicinal plants. Another important factor is the medicinal plant gathering, storage, and commercialization tradition rooted in the Balsas River Basin, of which the REBIOSH is part. As noticed by Hersch-Martínez (2010), since pre-Hispanic times, this region has been a source of exchange and tribute of medicinal plant species, mainly in the form of barks and resins. Nowadays, important centers of accumulation and sale of medicinal plants exist near the REBIOSH, such as the market in Axochiapan and the fair in Tepalcingo, Morelos (Hersch-Martínez 2010; Argueta 2016; Blancas et al. 2020; Fig. 4). From these markets, wholesale traders send large quantities of plant materials to regional traditional markets (tianguis) and urban markets, such as the Sonora market in Mexico City. The latter market is ranked as the largest market of medicinal plants in Mexico because of the diversity and volume of medicinal plants traded in it (Linares and Bye 2010). The overharvesting of wild plants to meet urban market demands is one of the factors that endangers some species that are gathered from the REBIOSH. One of the most illustrative examples of overharvesting is the cuachalalate (Amphipterygium adstringens (Schltdl.) Standl.), a tree from which bark is gathered to treat gastric ulcers and to heal wounds (Beltrán-Rodríguez et al. 2021a). Cuachalalate bark has been used by rural communities since long time ago, but due to the surge of herbalism during the past 20 years, its use has expanded to the point of causing its local extinction due to inappropriate management of its wild populations (Solares et al. 2012). Solares et al. (1992) documented how the overdemand of cuachalalate bark prompts ways of harvesting that endanger the trees by extraction of large portions of bark from individuals, frequently cutting through the vascular cambium, and, in some populations, the trees are felled and debarked. Although the environmental regulation currently in force prohibits the latter destructive practice, it continues to be illegally practiced by inexperienced gatherers. Regional initiatives have been generated to address this issue. This is the case of the Cuachalalate Productive Chain, which groups 17 communities of the REBIOSH that has established links with social actors, including the scientific community, to create alliances to design sustainable management of this plant species (Solares et al. 2012). Much work needs to be made in that regard since the cuachalalate is a dioic

638

J. J. Blancas Vázquez et al.

Fig. 4 Aspect of the Tepalcingo Tianguis, the most important market in the region of the REBIOSH for the commercialization and exchange of plants and their derivatives. (a) Basketry products; (b) copal resin (Bursera spp.); (c) handicrafts made with bules (Lagenaria siceraria); (d) sale of a great diversity of medicinal plants; and (e) sale of edible plants such as Timbiriche (Bromelia spp.). (Photos: José Blancas)

species with a low proportion of female trees, low seed viability and germination rates, and a low seedling establishment rate. Overcoming these drawbacks obligates researchers to make ecological, genetic, and reproductive biology studies to provide tools for its sustainable use (Beltrán-Rodríguez 2018). The inadequate environmental regulations addressing the vulnerability of medicinal plants in Mexico is another issue that increases pressure over some species populations. The regulation of medicinal plant extraction has contradictions that make highly vulnerable species that are not priority for conservation programs. It assumes that these plants have an extensive distribution thus preventing species from being endangered. However, vulnerability of both wide or restricted distribution species is severely affected by overharvesting and high market demand (DelgadoLemus et al. 2014). For instance, the quina amarilla or copalchi (Hintonia latiflora Sessé & Moc. ex DC.) is widely distributed in Mexico, but has low population densities, lacks vegetative propagation structures and its sexual propagation is difficult. Bark of this species is overharvested for its hypoglycemic effects (Cristians et al. 2009); but

Ethnobotanical Knowledge and the Patterns of. . .

639

despite its vulnerability, it is not included in the list of endangered species neither actions are carried out for ensuring its reproduction in wild populations, therefore, the species is highly vulnerable (Beltrán-Rodríguez et al. 2021b). The opposite situation is illustrated by the palo dulce (Eysenhardtia polystachya (Ortega) Sarg.), whose bark is extracted to treat urinary conditions (BeltránRodríguez et al. 2017) and is included in the IUCN Red List (IUCN 2022); however, its populations seem to remain stable and its current appropriation rates do not represent a major risk. Hence, many plants in the REBIOSH require ecological, ethnobotanical, and general natural history studies to assess their vulnerability and degree of risk from local and regional harvesting. Another source of vulnerability of medicinal plants extraction is the loss of knowledge regarding its management (Blancas et al. 2020). The 32 communities within the REBIOSH experienced a true exodus of youngsters who emigrate to other areas in Mexico or to the USA (Arjona-García et al. 2021). The result of this emigration is that only elders remain in the communities who, due to the lack of laborers, frequently rent their forest land to regional medicinal plant hoarders, most of which look for immediate profit beyond the conservation of the resources (Blancas et al. 2020). Furthermore, many of these hoarders lack the knowledge and technical skills to ensure the reproduction of many plant species, therefore making the populations of medicinal plants to become vulnerable (Beltrán-Rodríguez et al. 2017). The loss of knowledge about medicinal plants is associated with changes in the cultural patterns of the local population. The media and some public policies have a high influence on the contempt and misvaluing of traditional medicine, which has contributed to the abandonment of management practices that allowed harvesting medicinal plants without causing harm to its local populations (Arjona-García et al. 2021). We consider that joint efforts involving different social sectors are required to address these issues and find solutions based on the complementarity of local and scientific knowledge. One of the efforts that is worth mentioning is the Program of Social Actors of the Medicinal Flora of Mexico, which has been working in the region for over 25 years. This program looks for dialoguing about medicinal plant use and conservation from a transdisciplinary perspective, and to foster and articulate researchers of medical anthropology and ethnobotany. In practice, this program has made outstanding efforts to harmonize domestic and commercial use of medicinal plants, consider the cultural context of traditional medicine, enhance training of health promoters and traditional therapists, value the role of gatherers in the system of wild medicinal flora supply, and promote the exchange of knowledge between traditional physicians, and researchers of the public health systems (Hersch-Martínez 2003).

The Traditional Management of Copal: A Case of Sustainable Appropriation of the Tropical Deciduous Forests in the REBIOSH The southeastern portion of Morelos is an extraction and trading zone of copal, which together with the neighboring zones in the states of Puebla and Guerrero, provide nearly one-third of the total copal resin commercialized in Mexico (Linares

640

J. J. Blancas Vázquez et al.

and Bye 2008). Copal resin is a non-timber forest product, which in the REBIOSH is extracted from two species of the family Burseraceae. The main uses of copal are as a ritual resin that is burned to produce fragrant smoke in magic-religious ceremonies and agricultural rituals, and for medicinal purposes to treat some muscular ailments. Perhaps its use began in prehistory, but became highly important during pre-Hispanic time, when several regions of the present Mexico paid it as tribute and traded it with the political power centers (Abad-Fitz et al. 2020). The Tepalcingo traditional market, located near the limits of the REBIOSH, is one of the most important trading centers of copal that opens just before the Day of the Dead, one of the most important traditional celebrations in Mexico (Cruz et al. 2006; Argueta 2016). A large percentage of the copal resin commercialized in the Tepalcingo market comes from communities within the REBIOSH that are specialized in the extraction of the resin of the copal chino (Bursera bipinnata (DC.) Engl.) and copal ancho (Bursera copallifera (DC.) Bullock). The former species is the most appreciated because of its white resin and intense fragrance (Mena 2018; Abad-Fitz et al. 2020). In the REBIOSH, the gathering season of copal resin begins in August and ends in late October, just before the Day of the Dead. The management of copal trees involves careful selection focused on the trees’ productivity of resin. The copal resin gatherers, called copaleros, choose the trees based on their vigor and height (AbadFitz et al. 2020). Afterwards, incisions are made on the trees from which the resin flows to be collected in leaves of maguey criollo (Agave angustifolia Haw.). Once the selected trees stop producing resin, they are not harvested any further, and may be eliminated (Mena 2018; Abad-Fitz et al. 2020). By that process, highly productive trees are favored in the gathering zones within wild vegetation; an example of in situ management of copal trees. Copal trees are also propagated in agroecosystems like milpas – a polyculture system with maize, bean, squash, and chili pepper – and grazelands, where the trees are tolerated, promoted, protected, and even vegetatively propagated. More recently copal are propagated through seed sowing (Mena 2018; Abad-Fitz et al. 2020). As mentioned above, the patterns of use and management of the flora in the REBIOSH have received little attention from researchers, who have considered the copal chino as a wild plant, but the copaleros are developing multiple management strategies. Abad-Fitz et al. (2020) observed copal tree selection patterns in the REBIOSH, finding significant differences in the production and chemical composition of resin between wild and managed trees. In agroecosystems, trees with high production of intensely fragrant resin are more abundant than in unmanaged vegetation, in which most trees produce less resin with less intense fragrance. These results confirm that the copal chino can be considered as a tree under incipient domestication process. Another consequence of the observations of Abad-Fitz et al. (2020) is that, despite the copal chino is a tree with low abundance in the tropical deciduous forests of the Balsas River Basin (Maldonado 2013), the density of the trees is five times higher in sites managed by the copaleros of the REBIOSH, compared with the wild populations of the area. Therefore, the copal chino can be a key species for the recovery of vegetation, at the same time it provides supplementary economic resources to the local households

Ethnobotanical Knowledge and the Patterns of. . .

641

in a high marginality region. The organization of the copaleros has allowed the copal-producing communities to increase the price per kilogram of the copal resin, improve the commercial networks, and have access to income for improving their productive land and establishing communitarian nurseries (Mena 2018). That also implies that the local communities are acquiring a high level of participation in the development of environmental policies and shows that the management of copal is a sustainable activity with social and cultural value that ensures both the conservation of the natural environment and the livelihoods of local people. Copal management is an example that can be replicated at REBIOSH with other non-timber forest products. However, its benefits must be recognized in public policies at the regional level. Hence, the copaleros should have an incentive, be it social or even monetary recognition for managing two species of copal that are key to the conservation of the Tropical Deciduous Forest.

Ethnobotany of the Chichinautzin Biological Corridor Ecological Context The COBIO is a 37,301 ha natural protected area located in the northern portion of the state of Morelos (Fig. 1) decreed in 1988 with the purpose of safeguarding the region’s wild flora and fauna (Diario Oficial de la Federación 1988). It forms part of the Trans-Mexican Volcanic Belt, a mountain area that crosses Mexico from east to west, which in the state of Morelos has elevations ranging from 1250 to 3450 m (Santillán-Alarcón et al. 2010). It encompasses three climates: (i) semi-cold (mean annual temperature from 8 to 12  C and average yearly precipitation between 1200 to 2000 mm), (ii) temperate (16–20  C and 1000–1200 mm), and (iii) semi-warm subhumid (18–22  C and 800–1000 mm; García 1988). The COBIO is an aquifer recharge zone receiving high annual precipitation, where bodies of water include a complex network of streams, rivers, and at least seven lakes (Santillán-Alarcón et al. 2010). The vegetation in the zone includes: (i) Pine Forest dominated by Pinus montezumae Lamb., Pinus hartwegii Lindl., Cupressus lindleyi Klotzsch ex Endl., Alnus jorullensis Kunth, and Abies religiosa (Kunth) Schltdl. & Cham.; (ii) Pine-oak Forest with Pinus leiophylla (Schiede ex Schltdl. & Cham.), P. montezumae, Pinus pringlei Shaw, Pinus pseudostrobus Lindl., Pinus teocote Schltdl. & Cham., Quercus castanea Née, Quercus crassifolia Bonpl., and Quercus laurina Bonpl., (iii) Oak Forests dominated by Quercus candicans Née, Q. castanea, Q. crassifolia, Q. laurina, Quercus obtusata Bonpl., and Quercus rugosa Née; and (iv) mountain cloud forest with Carpinus caroliniana Walter, Celastrus pringlei Rose, Clethra mexicana DC., Cornus disciflora DC., Meliosma dentata (Liebm.) Urb., Oreopanax peltatus Linden ex Regel, Styrax ramirezii Greenm., Symplocos citrea Lex. ex La Llave & Lex., and Ternstroemia pringlei (Rose) Standl. (Bonilla-Barbosa and Villaseñor 2003; Fig. 5).

642

J. J. Blancas Vázquez et al.

Fig. 5 Vegetation of the Chichinautzin Biological Corridor (COBIO). (a) and (b) Pine Forest; and (c) Cloud Forest. (Photos: Itzel Abad Fitz)

The known flora in the COBIO includes 1265 vascular plant species within 153 families and 516 genera. The families with most species are Orchidaceae, Poaceae, Asteraceae, Lamiaceae, and Fabaceae, while the more represented genera are Salvia, Malaxis, Cyperus, Cuphea, Cheilanthes, and Tillandsia. Nearly 20 species in the region are within an endangered species category and six are endemic to the COBIO (Flores-Castorena and Martínez-Alvarado 2010).

Rural Settlements The COBIO is distributed among 15 municipalities of the states of Morelos, México, and Mexico City, where an estimated number of 34 rural communities have a population of 31,653 inhabitants, 1354 (4.27%) of which consider themselves as indigenous, most of them speakers of Nahuatl (SIMEC 2022). The region has a poverty level from slight to moderate (CONEVAL 2020). Some of these settlements have existed since pre-Hispanic time, such as Huitzilac, Tepoztlán, Amatlán, Tlayacapan, and Tlalnepantla. The long history of population in the region, relative to human occupation in the southern portion of the state of Morelos, has had consequences in the infrastructure as in the availability of public services. For several centuries, the region has been an access zone for the neighboring Mexico City, and, through it, the commodities brought from Asia and

Ethnobotanical Knowledge and the Patterns of. . .

643

disembarked in the port of Acapulco made their way to the city. One of the main railroads in 1897 connected Mexico City with Cuernavaca, Yautepec, and Cuautla in the state of Morelos, connecting settlements now included in the COBIO (Semo 2020). The main economic activities in the region are trading, lumbering and extraction of forest topsoil, subsistence and commercial agriculture, and sheep husbandry. The crops grown in the region are those adapted to temperate zones and crops with tolerance to several climates like maize (Zea mays L.) and bean (Phaseolus vulgaris L.). Native and exotic crops produced in the region include oat (Avena sativa L.), peanut (Arachis hypogaea L.), cucumber (Cucumis sativus L.), green tomato (Physalis philadelphica Lam.), tomato (Solanum lycopersicum L.), squash (Cucurbita spp.), chilacayote (Cucurbita ficifolia Bouché), potato (Solanum tuberosum L.), and fruit trees like avocado (Persea americana Mill.), peach (Prunus persica (L.) Batsch), apple (Malus domestica (Suckow) Borkh.), prune (Prunus domestica L.), black cherry (Prunus serotina Ehrh.), fig (Ficus carica L.), pear (Pyrus communis L.), and Mexican hawthorn (Crataegus mexicana DC.) (INECC 2022).

Ethnobotany of the COBIO We recorded a total of 576 useful plant species grouped in 115 families and 378 genera. The families with most used plant species in the COBIO are Asteraceae (76 species), Fabaceae (42), Lamiaceae (29), Solanaceae (28), Rosaceae (18), Malvaceae (17), Amaranthaceae (15), Fagaceae (12), and Orchidaceae, Pinaceae, Rubiaceae, Rutaceae, and Verbenaceae with 11 species each (Table 4). The latter 13 families concentrated 50.7% of the useful species, the remaining 49.3% being distributed in 102 families. The more represented genera were Quercus (12 species), Pinus (10), Solanum (9), Citrus (8), Salvia (7), Tagetes (6), and Bursera, Chenopodium, Ipomoea, Oenothera, Passiflora, and Tournefortia with five species each (Table 5). The useful flora in the COBIO represented 30.7% of the total species we recorded in the region, which agrees with the pattern of the used flora in Mexico, where the useful flora represents between 30 and 45% of the total (Caballero et al. 1998). That the number of species in the useful flora we registered in the COBIO (576 species) was considerably lower than that we recorded in the REBIOSH might be the consequence of the lower diversity of temperate zones compared to dry and humid tropical zones. Of the total useful species, we recorded in the COBIO, 389 (67.63%) were native and 187 (32.47%) were exotic. Regarding growth habit, 301 (52.3%) species were herbs, 160 (27.8%) trees, 92 (16%) shrubs, and 12 (2.1%) species epiphytes, while the remaining species (1.9%) were climbers, rosettes, vines, and palms. We recorded in the COBIO, 399 (39.27%) plant species used for medicinal purposes, 79 (13.72%) as food, and 69 (11.98%) for ornamental purposes (Table 6). Of the total species of useful plants, we recorded in the COBIO, 258 (44.8%) wild, 241 (41.8%) domesticated, 83 (14.4%) ruderal, and 18 (3.1%) weedy plants.

644

J. J. Blancas Vázquez et al.

Table 4 Most important plant families by number of useful species in the Chichinautzin Biological Corridor Family Asteraceae Fabaceae Lamiaceae Solanaceae Rosaceae Malvaceae Amaranthaceae Fagaceae Orchidaceae Pinaceae Rubiaceae Rutaceae Verbenaceae Apocynaceae Boraginaceae Convolvulaceae Euphorbiaceae Onagraceae Apiaceae Bignoniaceae Cucurbitaceae Malpighiaceae Myrtaceae Poaceae Brassicaceae Cactaceae Anacardiaceae Asparagaceae Burseraceae Crassulaceae Lauraceae Moraceae Nyctaginaceae Passifloraceae Sapotaceae Scrophulariaceae Araceae Aristolochiaceae Ericaceae Plantaginaceae Acanthaceae Amaryllidaceae

Number of useful plant species 76 42 29 28 18 17 15 12 11 11 11 11 11 9 9 9 8 8 7 7 7 7 7 7 6 6 5 5 5 5 5 5 5 5 5 5 4 4 4 4 3 3

Percentage 13.2 7.3 5.0 4.9 3.1 3.0 2.6 2.1 1.9 1.9 1.9 1.9 1.9 1.6 1.6 1.6 1.4 1.4 1.2 1.2 1.2 1.2 1.2 1.2 1.0 1.0 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.9 0.7 0.7 0.7 0.7 0.5 0.5 (continued)

Ethnobotanical Knowledge and the Patterns of. . .

645

Table 4 (continued) Family Cupressaceae Geraniaceae Iridaceae Meliaceae Oleaceae Papaveraceae Piperaceae Polypodiaceae Santalaceae Sapindaceae Selaginellaceae Smilacaceae Urticaceae Annonaceae Arecaceae Betulaceae Bromeliaceae Caprifoliaceae Commelinaceae Dioscoreaceae Ebenaceae Juglandaceae Liliaceae Lythraceae Musaceae Polygalaceae Portulacaceae Ranunculaceae Zygophyllaceae Altingiaceae Asphodelaceae Balsaminaceae Campanulaceae Cannabaceae Cannaceae Caricaceae Celastraceae Cistaceae Equisetaceae Garryaceae Gentianaceae Guttiferae Hypericaceae

Number of useful plant species 3 3 3 3 3 3 3 3 3 3 3 3 3 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1 1 1 1 1 1 1 1 1 1 1 1 1 1

Percentage 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.3 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 (continued)

646

J. J. Blancas Vázquez et al.

Table 4 (continued) Family Linaceae Loranthaceae Marantaceae Mirsinaceae Monimiaceae Moringaceae Nephrolepidaceae Orobanchaceae Pentaphylacaceae Phytolaccaceae Plumbaginaceae Polemoniaceae Polygonaceae Polypodiaceae Polygonaceae Primulaceae Proteaceae Punicaceae Salicaceae Saxifragaceae Schisandraceae Simaroubaceae Sterculiaceae Styracaceae Tamaricaceae Theaceae Tropaeolaceae Viburnaceae Vitaceae Zingiberaceae

Number of useful plant species 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

Percentage 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2

In the REBIOSH, 317 (55%) useful plant species are gathered wild plants obtained from natural or anthropized vegetation, 243 (42.2%) are cultivated, both by seed or vegetatively, and 17 (3%) are tolerated.

Use of Medicinal Plants in the Chichinautzin Biological Corridor The more represented use category of plants in the COBIO was medicinal, as it was in the REBIOSH. However, in the former region that predominance appears to be due to the high percentage of indigenous people and the use of traditional medicine for first-level health care (Monroy-Ortiz and Monroy 2010), rather than

Ethnobotanical Knowledge and the Patterns of. . .

647

Table 5 Main genera by number of useful plant species in the Chichinautzin Biological Corridor Genera Quercus Pinus Solanum Citrus Salvia Tagetes Bursera Chenopodium Ipomoea Oenothera Passiflora Tournefortia Acacia Ageratina Aristolochia Artemisia Asclepias Cucurbita Mentha Montanoa Prunus Rosa Sedum Senna Baccharis Bougainvillea Brassica Brugmansia Buddleja Capsicum Cosmos Datura Euphorbia Opuntia Phaseolus Phoradendron Physalis Plectranthus Porophyllum Pouteria Pseudognaphalium Selaginella Smilax

Number of useful plant species 12 10 9 8 7 6 5 5 5 5 5 5 4 4 4 4 4 4 4 4 4 4 4 4 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 (continued)

648

J. J. Blancas Vázquez et al.

Table 5 (continued) Genera Verbena Acourtia Agastache Agave Allium Alnus Amaranthus Annona Argemone Cirsium Clinopodium Crataegus Dalea Dioscorea Diospyros Dyssodia Epidendrum Eucalyptus Ficus Galphimia Geranium Gladiolus Hibiscus Iresine Juglans Lantana Larrea Leucaena Lippia Mimosa Morus Musa Nicotiana Ocimum Origanum Peperomia Persea Phyla Plantago Randia Rhynchostele Rubus Ruta

Number of useful plant species 3 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 (continued)

Ethnobotanical Knowledge and the Patterns of. . .

649

Table 5 (continued) Genera Senecio Serjania Sida Sideroxylon Spermacoce Spondias Stevia Tecoma Verbesina Waltheria Abies Acalypha Achillea Agapanthus Alcea Alchemilla Aloe Alomia Aloysia Alternanthera Amphipterygium Anagallis Ananas Anoda Antigonon Apium Arachis Arbutus Arctostaphylos Ardisia Arnica Avena Bambusa Barkleyanthus Beaucarnea Beta Bidens Bocconia Boerhavia Borago Brahea Brickellia Brosimum

Number of useful plant species 2 2 2 2 2 2 2 2 2 2 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 (continued)

650

J. J. Blancas Vázquez et al.

Table 5 (continued) Genera Bunchosia Byrsonima Calandrinia Calathea Calea Calendula Calliandra Callistephus Camellia Canna Cannabis Carica Cascabela Casimiroa Cassia Castilleja Catharanthus Cecropia Cedrela Ceiba Celosia Cestrum Chamaecrista Chiococca Chiranthodendron Chlorophytum Chromolaena Cinnamomum Cissus Clematis Cnidoscolus Cocos Coffea Comarostaphylis Commelina Convolvulus Conyza Cordia Coriandrum Coryphantha Crescentia Crocosmia Crotalaria

Number of useful plant species 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 (continued)

Ethnobotanical Knowledge and the Patterns of. . .

651

Table 5 (continued) Genera Croton Cucumis Cuminum Cunila Cymbopogon Cynara Dahlia Daucus Delonix Desmodium Dichaea Didymaea Diphysa Discocnide Dodonaea Dysphania Enterolobium Equisetum Erigeron Eriobotrya Eriosema Eryngium Erythrina Eugenia Evolvulus Exogonium Eysenhardtia Fleischmannia Flourensia Foeniculum Fragaria Fraxinus Fuchsia Gardenia Garrya Gentiana Gnaphalium Gomphrena Guazuma Haematoxylum Hechtia Hedeoma Heimia

Number of useful plant species 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 (continued)

652

J. J. Blancas Vázquez et al.

Table 5 (continued) Genera Helianthemum Helianthus Heliocarpus Heliocereus Hesperalbizia Heteropterys Heterotheca Heuchera Hieracium Hintonia Hypericum Hypericum Illicium Impatiens Inga jinicuil Iostephane Jacaranda Jasminum Jatropha Juniperus Justicia Lactuca Laelia Lagerstroemia Lathyran Laurus Leonotis Lepechinia Lepidium Leucanthemum Lilium Linum Liquidambar Lithospermum Litsea Lobelia Loeselia Lopezia Ludwigia Lycopersicon Lysiloma Macadamia Malpighia

Number of useful plant species 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 (continued)

Ethnobotanical Knowledge and the Patterns of. . .

653

Table 5 (continued) Genera Malus Malva Malvaviscus Manfreda Mangifera Marrubium Matricaria Mecardonia Medicago Melia Milla Mirabilis Monstera Morinda Moringa Myrtus Nasturtium Nephrolepis Nerium Nicotiana Nopalea Olea Oncidium Oriza Pachira Pachyrhizus Packera Parmentiera Pelargonium Penstemon Petroselinum Peumus Phalaris Phlebodium Phytolacca Pimenta Pinaropappus Piper Piqueria Pisum Pithecellobium Pleopeltis Plumbago

Number of useful plant species 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 (continued)

654

J. J. Blancas Vázquez et al.

Table 5 (continued) Genera Plumeria Polygala Polygonum Polypodium Portulaca Potentilla Prosopis Prosthechea Pseudobombax Psidium Pteridium Punica Pyrostegia Pyrus Quamoclit Quassia Ranunculus Raphanus Rhododendron Ricinus Roldana Rosmarinus Ruellia Rumex Russelia Saccharum Salix Sambucus Sanvitalia Schinus Sechium Semialarium Sempervivum Sicyos Solandra Sonchus Spathiphyllum Spathodea Spinacia Stanhopea Stelis Stigmaphyllum Struthanthus

Number of useful plant species 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 (continued)

Ethnobotanical Knowledge and the Patterns of. . .

655

Table 5 (continued) Genera Styrax Swietenia Symphoricarpos Syzygium Tamarix Tanacetum Taraxacum Taxodium Ternstroemia Theobroma Thuja Thunbergia Thymus Tilia Tithonia Tradescantia Trichocentrum Trifolium Trixis Tropaeolum Uncaria Urtica Valeriana Verbascum Vicia Vigna Vinca Vitex Wigandia Xanthosoma Zantedeschia Zea Zingiber Zinnia

Number of useful plant species 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

to ecological predominance or intensive trade of medicinal plants as in the latter zone. Another probable reason for the high number of medicinal plant species in the COBIO could be the abundance of medicine men and women (curanderos), bonesetters (hueseros), healers (sanadores), herbalists, and other professional traditional medical practicioners. This fact has generated the proliferation of traditional medicine tourism in the region – particularly since the Mexican government’s tourism ministry assigned Tepoztlán the category of Pueblo Mágico, an assignment meant to favor

656 Table 6 Plant use categories reported for the flora of the Chichinautzin Biological Corridor. Percentages do not add to 100 because several species have more than one use

J. J. Blancas Vázquez et al.

Use Medicinal Edible Ornamental Firewood Ritual Wood Construction Handicrafts Living fence

Number of species in category 399 79 69 29 11 7 6 2 2

Percentage 69.27 13.72 11.98 5.03 1.91 1.22 1.04 0.35 0.35

regional tourism in indigenous and high biocultural richness. Consequently, temascals, massage spas, and many traditional medic offices have been established in the region, which offer services to national and foreign patients (Ruiz-López and Alvarado-Rosas, 2017).

Threats to the Traditional Knowledge About Plants in the COBIO Despite the current importance of plants in the life of the inhabitants of the COBIO, several threats exist to the traditional knowledge about plants. These threats might have negative impacts for conservation of natural vegetation and traditional agroecosystems like milpas and homegardens. Maybe, one of the strongest pressures on the traditional knowledge of plants in the COBIO could be land use change, which between 1973 and 2000 caused a 35% reduction of forest covers in the region, mostly due to the expansion of urban areas and agriculture (CONANP 2003). Because the northern portion of the state of Morelos is adjacent to Mexico City, land is highly attractive for the real estate markets. Consequently, it is frequent that illegal loggers and organized crime groups – in collusion with corrupt local authorities – promote forest fires with the purpose of favoring the urbanization of forest areas (López-Portillo-Vargas and Corral-Torres 2021). Deforestation modifies people’s habits by decreasing activities like plant gathering for several purposes, or by restricting these activities to the elder population (Ayala-Enríquez et al. 2020), while the loss of agricultural land changes household’s feeding habits by increasing the consumption of processed foodstuffs; a tendency occurring not only in the COBIO but throughout Mexico (Bertran-Vilá 2005). The study made by Tegoma-Coloreano (2019) in the community of Tres Marías – located within the COBIO – shows the deep consequences of the loss of natural vegetation and agroecosystems in the transmission of knowledge about medicinal plants to children. The author found that 11-year-old children in Tres Marías have less knowledge of the medicinal flora in their community than their parents and grandparents. Children do not know the plants occurring in their surroundings, and – maybe because of the influence of the media and school inputs little related to their

Ethnobotanical Knowledge and the Patterns of. . .

657

ecological and cultural contexts – have more knowledge about exotic than about native plants. She also documented that the transmission of knowledge to children in the community is mostly vertical, but also acknowledges a growing trend of the transversal and horizontal transmission of information. The study of Villalpando-Toledo (2020) in the home gardens of Tepoztlán, Morelos, is another example of how land use changes modify traditional knowledge and practices. The author documented how tourism in Tepoztlán has become a strong change agent in the modification in composition and structure of the local home gardens. Many homegardens, until 20 years ago, were sources of edible and medicinal plants and, nowadays, have been transformed to parking lots or restaurant gardens and terraces dominated by ornamental plants (Fig. 6).

Fig. 6 Changes in land use and cultural patterns modify the landscapes and homegardens in the Chichinautzin Biological Corridor (COBIO). (a) Progress of the urbanization in Tres Marías, Morelos, Mexico; (b) traditional milpa in Tres Marías, Morelos, Mexico; (c) and (d) homegardens in Tepoztlán, Morelos, Mexico. Some homegardens specilialized in aromatic and medicinal plants, others in edible plants; (e) and (f) the change in the homegardens is notorious, since ornamental plants are currently the dominant elements in these agroecosystems. (Photos: (a), (b), and (f) Araceli Tegoma Coloreano; (c), (d), and (e) María Idalia Villalpando Toledo)

658

J. J. Blancas Vázquez et al.

Land use changes might also accelerate other multifactorial processes like the abandonment of primary activities, migration, and loss of contact with the surrounding environment (Benz et al. 2000); these factors representing strong pressures that threaten the persistence of the traditional knowledge about plants in the COBIO pose a challenge for the compilation of a multidisciplinary agenda aiming at reversing this trend. Therefore, it is important to recognize that traditional knowledge is part of a biocultural heritage, which is strongly threatened, and measures should be taken for its protection. Some of them include: the incorporation in basic level schools of alternative forms of learning that encourage children’s contact with the surrounding environment. The ethnobotanical walks would not only allow the recognition of the plants, animals, and fungi that are distributed in the area, it is also a mode of coexistence with parents and grandparents, which strengthens social relationships and would help improve the deteriorated social environment. Another measure should aim at the conservation of homegardens, milpas, and other agroecosystems as providers of food in villages and cities. Efforts should be made to generate exchanges of successful experiences, where homegardens, in addition to being seen as productive units, can also serve as places of learning, experimentation, and communitarian coexistence.

Conclusions The information we systematized show that the REBIOSH and COBIO mountain zones of the state of Morelos, Mexico, harbor high biocultural richness that expresses in the local people’s ample knowledge about the use and management of native and exotic plants. In both regions, we documented the preponderance of the ethnobotanical knowledge about the medicinal use of plants, which in each region is due to different ecological, historical, and cultural factors. However, our results also show that we are yet far from completing the information about the useful flora in the REBIOSH. Therefore, more research needs to be made in the region to document processes beyond use, such as plant resource management and the factors affecting its temporal dynamics. Finally, in both regions, the persistence of the local ethnobotanical knowledge is threatened by strong pressures driven by local cultural change, which is frequently amplified by social phenomena like migration and land use changes. Our work in the main mountain zones of the state of Morelos should be seen as a first effort to compile information, and to find documentation voids, vulnerabilities, and pending research tasks. Our results might facilitate the building of an agenda to visualize the local traditional knowledge and practices at the regional level, aiming at establishing biocultural conservation priorities. Acknowledgments The authors thank financial support from the Consejo Nacional de Ciencia y Tecnología (CONACYT), Mexico (research projects 271837, 280901, 293914, 299274, 299287, 318799).

Ethnobotanical Knowledge and the Patterns of. . .

659

References Abad-Fitz I, Maldonado-Almanza B, Aguilar-Dorantes K, Sánchez-Méndez L, Gómez-Caudillo L, Casas A, Blancas J, García-Rodríguez Y, Beltrán-Rodríguez L, Sierra-Huelsz JA, Cristians S, Moreno-Calles AI, Torres-García I, Espinosa-García FJ. Consequences of traditional management in the production and quality of copal resin (Bursera bipinnata (Moc. & Sessé ex DC.) Engl.) in Mexico. Forests. 2020;11(9):991. https://doi.org/10.3390/f11090991. Albuquerque UP. Implications of ethnobotanical studies on bioprospecting strategies of new drugs in semi-arid regions. The Open Complementary Medicine Journal. 2010;2:21–3. https://doi.org/ 10.2174/1876391X01002010021. Argueta A. El estudio etnobiológico de los tianguis y mercados en México. Etnobiología. 2016;14(2):38–46. Arjona-García C, Blancas J, Beltrán-Rodríguez L, López Binnqüist C, Colín H, Moreno-Calles AI, Sierra-Huelsz JA, López-Medellín X. How does urbanization affect perceptions and traditional knowledge of medicinal plants? J Ethnobiol Ethnomed. 2021;17:48. https://doi.org/10.1186/ s13002-021-00473-w. Ayala-Enríquez MI, Maldonado-Almanza B, Blancas J, Román-Montes de Oca E, García-Lara F. Panorama general de la flora medicinal. In: CONABIO, editor. La biodiversidad en Morelos. Estudio de Estado 2. Vol. III. Ciudad de México: CONABIO; 2020. p. 69–76. Beltrán-Rodríguez L. Estructura, dinámica poblacional y regeneración del leño de Amphipterygium adstringens (Anacardiaceae) en el ejido El Limón, Morelos, México. PhD thesis. Colegio de Postgraduados; 2018. Beltrán-Rodríguez L, Manzo-Ramos F, Maldonado-Almanza B, Martínez-Ballesté A, Blancas J. Wild medicinal species traded in the Balsas basin, Mexico: risk analysis and recommendations for their conservation. J Ethnobiol. 2017;37(4):743–64. https://doi.org/10.2993/02780771-37.4.743. Beltrán-Rodríguez L, Valdez-Hernández JI, Saynes-Vásquez A, Blancas J, Sierra-Huelsz JA, Cristians S, Martínez-Ballesté A, Romero-Manzanares A, Luna-Cavazos M, de la Rosa MA B, Pineda-Herrera E, Maldonado-Almanza B, Ángeles-Pérez G, Ticktin T, Bye R. Sustaining medicinal barks: survival and bark regeneration of Amphipterygium adstringens (Anacardiaceae), a tropical tree under experimental debarking. Sustainability. 2021a;13:2860. https://doi.org/10.3390/su13052860. Beltrán-Rodríguez L, Cristians S, Bye R. Las cortezas como productos forestales no maderables en México: Análisis nacional y recomendaciones para su aprovechamiento sostenible. Ciudad de México: Instituto de Biología, Universidad Nacional Autónoma de México; 2021b. Benz B, Cevallos J, Santana F, Rosales J, Graf S. Losing knowledge about plant use in Sierra de Manantlan Biosphere Reserve, Mexico. Econ Bot. 2000;54(2):183–91. https://doi.org/10.1007/ BF02907821. Bertran-Vilá M. Cambio alimentario e identidad de los indígenas mexicanos. Ciudad de México: Universidad Nacional Autónoma de México; 2005. Blancas J, Casas A, Pérez-Salicrup D, Caballero J, Vega E. Ecological and socio-cultural factors influencing plant management in Náhuatl communities of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013;9(39) https://doi.org/10.1186/1746-4269-9-39. Blancas J, Beltrán-Rodríguez L, Maldonado-Almanza B, Sierra-Huelsz JA, Sánchez L, MenaJiménez F, García-Lara F, Abad-Fitz I, Valdez-Hernández JI. Comercialización de especies arbóreas utilizadas en medicina tradicional y su impacto en poblaciones silvestres. In: CONABIO, editor. La biodiversidad en Morelos. Estudio de Estado 2. Vol. III. Ciudad de México: CONABIO; 2020. p. 215–23. Bolongaro A, Torres V. El clima como integrador de ecosistemas. In: CONABIO. La biodiversidad en Morelos. Estudio de Estado 2. Volumen I. Ciudad de México: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2020. Bonilla-Barbosa JR, Villaseñor JL. Catálogo de la flora del estado de Morelos. Cuernavaca: Centro de Investigaciones Biológicas, Universidad Autónoma del Estado de Morelos; 2003.

660

J. J. Blancas Vázquez et al.

Caballero J, Casas A, Cortés L, Mapes C. Patrones en el conocimiento, uso y manejo de plantas en pueblos indígenas de México. Estudios Atacameños. 1998;16:181–95. CONANP. Estimación de la tasa de transformación del hábitat en el “Corredor Biológico Chichinautzin”. Periodo 1973–200. Informe final. Ciudad de México: Comisión Nacional de Áreas Naturales Protegidas – Fondo Mexicano para la Conservación de la Naturaleza A.C; 2003. https://simec.conanp.gob.mx/TTH/Chichinautzin_Zempoala_Tepozteco/Chichinautzin_ Zempoala_Tepozteco_TTH_1973_2000.pdf. Accessed 23 Apr 2022. CONANP. Programa de Conservación y Manejo Reserva de la Biosfera Sierra de Huautla. Ciudad de México: Comisión Nacional de Áreas Naturales Protegidas; 2005. CONEVAL (Consejo Nacional de Evaluación de la Política de Desarrollo Social). Informe de pobreza y evaluación. Morelos. Ciudad de México: CONEVAL; 2020. Corona E. Escenarios paleobiológicos para las interacciones entre las sociedades y el medio ambiente en la región de Morelos. In: López-Varela SL, editor. La arqueología en Morelos: dinámicas sociales sobre las construcciones de la cultura material. Tomo II. Cuernavaca: Universidad Autónoma del Estado de Morelos; 2018. p. 31–42. Crespo H. Los inicios de la agroindustria azucarera en la región de Cuernavaca y Cuautla. In: Crespo H, editor. Historia de Morelos. Tierra, gente, tiempos del Sur. Tomo III. Cuernavaca: Universidad Autónoma del Estado de Morelos; 2018. p. 739–66. Cristians S, Guerrero-Analco JA, Pérez-Vásquez A, Palacios-Espinosa F, Ciangherotti C, Bye R, Mata R. Hypoglycemic activity of extracts and compounds from the leaves of Hintonia standleyana and H. latiflora: potential alternatives to the use of the stem bark of these species. J Nat Prod. 2009;72(3):408–13. https://doi.org/10.1021/np800642d. Cruz A, Salazar ML, Campos OM. Antecedentes y actualidad del aprovechamiento de copal en la Sierra de Huautla, Morelos. Revista de Geografía Agrícola. 2006;37:97–116. Delgado-Lemus A, Torres I, Blancas J, Casas A. Vulnerability and risk management of Agave species in the Tehuacán Valley, México. J Ethnobiol Ethnomed. 2014;10(1):53. https://doi.org/ 10.1186/1746-4269-10-53. Diario Oficial de la Federación. Decreto por el que se declara el área de protección de flora y fauna silvestre, ubicada en los municipios de Huitzilac, Cuernavaca, Tepoztlán, Jiutepec, Tlalnepantla, Yautepec, Tlayacapan y Totolapan, Morelos. Órgano del Gobierno Constitucional de los Estados Unidos Mexicanos. 1988;CDXXII(22):27–48. Escandón-Calderón J, Ordoñez-Díaz JAB, Nieto MC, Ordóñez-Díaz MJ. Cambio en la cobertura vegetal y uso del suelo del 2000 al 2009 en Morelos, México. Revista Mexicana de Ciencias Forestales. 2018;9(46):27–53. https://doi.org/10.29298/rmcf.v9i46.135. Flores-Castorena A, Martínez-Alvarado D. Sinopsis florística. In: Bonilla-Barbosa JR, Mora VM, Luna-Figueroa J, Colín H, Santillán-Alarcón S, editors. Biodiversidad, conservación y manejo en el Corredor Biológico Chichinautzin. Condiciones actuales y perspectivas. Cuernavaca: Centro de Investigaciones Biológicas, Universidad Autónoma del Estado de Morelos; 2010. p. 69–97. García E. Modificaciones al sistema de clasificación climática de Köppen (para adaptarlo a las condiciones de la República Mexicana). Ciudad de México: Instituto de Geografía, Universidad Nacional Autónoma de México; 1988. García J. Hernán Cortés empresario: el papel económico Cuauhnáhuac en las empresas cortesianas. In: Crespo H, editor. Historia de Morelos. Tierra, gente, tiempos del Sur. Tomo III. Cuernavaca: Universidad Autónoma del Estado de Morelos; 2018. p. 648–713. Garza S, Palavicini B. Xochicalco. La serpiente emplumada y Quetzalcóatl. Arqueología Mexicana. 2002;53:42–5. https://arqueologiamexicana.mx/mexico-antiguo/xochicalco-la-serpienteemplumada-y-quetzalcoatl. Accessed 23 Apr 2022. González-Insuasti MS, Martorell C, Caballero J. Factors that influence the intensity of non-agricultural management of plant resources. Agrofor Syst. 2008;74(1):1–15. https://doi. org/10.1007/s10457-008-9148-z. Grove DC. Chalcatzingo. Excavations on the Olmec Frontier. Hungary: Thames and Hudson; 1984.

Ethnobotanical Knowledge and the Patterns of. . .

661

Hersch-Martínez P. Actores sociales de la flora medicinal en México. Revista de la Universidad de México. 2003;629:30–6. Hersch-Martínez P. Plantas medicinales silvestres del suroccidente poblano y su colindancia en Guerrero, México: Rutas de comercialización, antecedentes y dinámica actual. In: Long J, Attolini A, editors. Caminos y mercados de México: por los caminos del sur. Ciudad de México: Instituto de Investigaciones Históricas-UNAM; 2010. p. 665–86. INAH. Zona Arqueológica Teopanzolco. 2022. https://www.inah.gob.mx/zonas/127-zonaarqueologica-tepozteco. Accessed 25 Jun 2022. INAH. Zona Arqueológica Tepozteco. 2022. https://www.inah.gob.mx/zonas/127-zonaarqueologica-tepozteco. Accessed 16 Jun 2022. INECC. Instituto Nacional de Ecología y Cambio Climático. Área de Protección de Flora y Fauna Corredor Biológico Chichinautzin. http://www2.inecc.gob.mx/publicaciones2/libros/2/ chichinau.html. Accessed 25 Jun 2022. INEGI. Panorama sociodemográfico de Morelos. In: Censo de Población y Vivienda 2020. 2021. https://www.inegi.org.mx/contenidos/productos/prod_serv/contenidos/espanol/bvinegi/pro ductos/nueva_estruc/702825197896.pdf. Accessed 30 october 2022. IUCN (International Union for Conservation of Nature and Natural Resources). The IUCN Red List of Threatened Species. 2022. https://www.iucnredlist.org. Accessed on 21 Mar 2022. Juárez-Mondragón A. Prácticas de aprovechamiento en unidades de manejo para la conservación de la vida silvestre (UMA) del sur de Morelos. Cuernavaca: Universidad Nacional Autónoma de México, Centro Regional de Investigaciones Multidisciplinarias; 2019. Linares E, Bye R. El copal en México. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad. Biodiversitas. 2008;78:8–11. Linares E, Bye R. La dinámica de un mercado periférico de plantas medicinales de México: el tianguis de Ozumba, Estado de México, como centro acopiador para el mercado de Sonora (mercado central). In: Long J, Attolini A, editors. Caminos y mercados de México: por los caminos del sur. Ciudad de México: Instituto de Investigaciones Históricas-UNAM; 2010. p. 631–64. López-Portillo-Vargas A, Corral-Torres Y. 2021. Atención a un incendio forestal en el Parque Nacional Lagunas de Zempoala, Morelos. Comisión Nacional de Áreas Naturales Protegidas. Parque Nacional Lagunas de Zempoala Informe final SNIB-CONABIO, proyecto No. TR002. Ciudad de México: CONANP; 2021. http://www.conabio.gob.mx/institucion/proyectos/ resultados/InfTR002.pdf. Accessed 29 Aug 2022. Maderey LE. Evapotranspiración real en Hidrogeografía IV.6.6. Atlas Nacional de México. Vol. II Escala 1 4000000. Instituto de Geografía UNAM. México. 2002. http://www.conabio.gob.mx/ informacion/metadata/gis/evapr4mgw.xml?_httpcache¼yes&_xsl¼/db/metadata/xsl/fgdc_ html.xsl&_indent¼no. Accessed 27 Jul 2022. Maldonado B. Aprovechamiento de los recursos florísticos de la Sierra de Huautla, Morelos, México. MSc thesis. Universidad Nacional Autónoma de México; 1997. Maldonado B. Patrones de uso y manejo de los recursos florísticos del Bosque Tropical Caducifolio en la Cuenca del Balsas, México. PhD thesis. Universidad Nacional Autónoma de México; 2013. Maldonado D. La producción agrícola en el Morelos prehispánico. In: Crespo H, editor. Historia de Morelos. Tierra, gente, tiempos del Sur. Tomo III. Cuernavaca: Universidad Autónoma del Estado de Morelos; 2018. p. 115–37. Maldonado B, Caballero J, Delgado-Salinas A, Lira R. Relationship between use value and ecological importance of floristic resources of seasonally dry tropical forest in the Balsas River Basin, Mexico. Econ Bot. 2013;67(1):17–29. Mena F. Estrategias ecológicas y culturales para garantizar la disponibilidad de Productos Forestales No Maderables en la Selva Baja del sur de Morelos. MSc Thesis. Cuernavaca: Universidad Autónoma del Estado de Morelos; 2018. Monroy-Ortiz C, Castillo-España P. Plantas medicinales utilizadas en el estado de Morelos. Cuernavaca: Universidad Autónoma del Estado de Morelos; 2007.

662

J. J. Blancas Vázquez et al.

Monroy-Ortiz C, Monroy R. Plantas medicinales utilizadas en el COBIO: listado preliminar. In: Bonilla-Barbosa JR, Mora VM, Luna-Figueroa J, Colín H, Santillán-Alarcón S, editors. Biodiversidad, conservación y manejo en el Corredor Biológico Chichinautzin. Condiciones actuales y perspectivas. Cuernavaca: Centro de Investigaciones Biológicas, Universidad Autónoma del Estado de Morelos; 2010. p. 211–28. Parodi BG. La exploración botánica del doctor Francisco Hernández en el Marquesado del Valle (1573). In: Goldsmit S, Lozano R, editors. España y Nueva España: sus acciones transmarítimas. Memorias del I Simposio Internacional. Ciudad de México: Universidad Iberoamericana; 1991. p. 39–70. Phillips O, Gentry AH. The useful plants of Tambopata, Peru: II. Additional hypothesis testing in quantitative ethnobotany. Econ Bot. 1993;47:33–43. Ramírez R, Ocampo F, Rojas B, Flores G, Tovar E, Sánchez A. Flora arbórea no nativa, un potencial riesgo para la biodiversidad. In: CONABIO, editor. La biodiversidad en Morelos. Estudio de Estado 2. Volumen III. Ciudad de México: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2020. Ruiz-López CF, Alvarado-Rosas C. Los falsos escenarios turísticos y la reconfiguración del territorio en Tepoztlán, Morelos. El Periplo Sustentable. 2017;33:291–329. SADER. Secretaría de Agricultura y Desarrollo Rural. Arroz a la mexicana, sólo el de Morelos ¡sí señor! 2021. https://www.gob.mx/agricultura/articulos/arroz-a-la-mexicana-solo-el-de-morelossi-senor. Accessed 27 Jul 2022. Santillán-Alarcón S, Sorani-Dalbón V, Bonilla-Barbosa JR, Luna-Figueroa J, Colín H. Escenario Geográfico. In: Bonilla-Barbosa JR, Mora VM, Luna-Figueroa J, Colín H, Santillán-Alarcón S, editors. Biodiversidad, conservación y manejo en el Corredor Biológico Chichinautzin. Condiciones actuales y perspectivas. Cuernavaca: Centro de Investigaciones Biológicas, Universidad Autónoma del Estado de Morelos; 2010. p. 3–20. Semo A. El ferrocarril en México (1880–1900). Tiempo, espacio y percepción. Ciudad de México: Centro Nacional para la Preservación del Patrimonio Cultural Ferrocarrilero; 2020. SIMEC (Sistema de Información, Monitoreo y Evaluación para la Conservación). Corredor Biológico Chichinautzin. Comisión Nacional de Áreas Naturales Protegidas. https://simec. conanp.gob.mx/ficha.php?anp¼39®¼7. Accessed 27 Jul 2022. Solares F, Díaz V, Boyás JC. Avances del estudio sobre efecto del descortezamiento en la capacidad de regeneración de corteza de cuachalalate (Amphipterygium adstringens Schiede ex Schlecht.) en el estado de Morelos. Memoria INIFAP-SARH. Campo Experimental “Zacatepec” Publicación especial. 1992;7:91–8. Solares F, Vázquez JMP, Gálvez MC. Canales de comercialización de la corteza de cuachalalate (Amphipterigium adstringens Schiede ex Schlecht.) en México. Revista Mexicana de Ciencias Forestales. 2012;3(12):29–42. Soto-Nuñez J, Sousa M. Plantas Medicinales de la Cuenca del Río Balsas. Ciudad de México: Instituto de Biología-Universidad Nacional Autónoma de México; 1995. Tegoma-Coloreano A. Factores que afectan la transmisión del conocimiento de la flora medicinal entre generaciones: estudio de caso en Tres Marías, Morelos, México. Bachelor thesis. Cuernavaca: Universidad Autónoma del Estado de Morelos; 2019. Toledo V, Barrera-Bassols N, García-Frapolli E, Alarcón-Chaires P. Uso múltiple y biodiversidad entre los Mayas yucatecos (México). Interciencia. 2008;33(5):345–52. Trujillo ML, López-Medellín X. ¿Qué es la conservación desde el punto de vista de los campesinos? Condiciones productivas en un área natural protegida, Morelos, México. Etnobiología. 2018;16(1):58–72. Villalpando-Toledo MI. Cambio cultural y su impacto en la estructura y composición de los huertos familiares de Tepoztlán, Morelos. MSc thesis. Cuernavaca: Universidad Autónoma del Estado de Morelos; 2020. Zolla C. Epidemias, epidemiologías y herbolaria medicinal en el Morelos del siglo XVI. In: Crespo H, editor. Historia de Morelos. Tierra, gente, tiempos del Sur. Tomo III. Cuernavaca: Universidad Autónoma del Estado de Morelos; 2018;309–354.

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management, and Local People Subsistence Berenice Farfa´n-Heredia and Alejandro Casas

Abstract

This study documents the Mazahua ecological knowledge of plants, their nomenclature and classification, and their role in subsistence of people in a village of the Monarch Butterfly Biosphere Reserve, Mexico. We registered 213 useful plant species within the territory studied. Prunus serotina, Rubus liebmannii, and Crataegus mexicana were the main species providing wild fruits gathered by the Mazahua people, whereas Brassica campestris, Rorippa nasturtiumaquaticum, Chenopodium berlandieri, and Amaranthus hybridus were the principal non-crop greens locally consumed. Extraction of medicinal plants is low but gathering of flowers of Ternstroemia spp. for commercialization involves practices that endanger local populations of these plants. All households of the village make use of fuelwood, mainly of pine and oak species; in addition, they practice livestock, mostly extensive free raising of cows and sheep, but commonly people gather some wild and weedy plants for feeding their animals. Spatial and temporal availability of useful plants were investigated to determine their abundance and relation to their role in people subsistence throughout the year. The information was compared with data on extraction rates of the main useful plants to analyze conditions for sustainable use of plant resources. Nearly half of the territory of the village was covered by forest areas, including different types of pine-oak-fir forests and riparian vegetation. The other half of the territory has been transformed, including agricultural areas and secondary scrub B. Farfán-Heredia (*) Universidad Intercultural Indígena de Michoacán (UIIM), Pichátaro, Michoacán, Mexico A. Casas Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_8

663

664

B. Farfa´n-Heredia and A. Casas

grasslands. Although the village studied (Francisco Serrato, Michoacan) is part of the core zone and the buffer zone of the Monarch Butterfly Biosphere Reserve, during the period of 2001–2006 its territory suffered a drastic deforestation caused by the influence of outsider organized crime. Such action decreased the forest area of the community nearly 350 ha of the Abies forest. The highest diversity of useful trees and shrubs was found in the riparian vegetation, where impact of extraction is relatively low. Agricultural areas lack arboreal vegetation but have the highest diversity of herbaceous useful plants mainly including weedy plants gathered for human food as greens and fodder. Pine-oak associations have intermediate diversity of useful trees and shrubs and herbs but are the main reservoirs of biomass of useful plants and are also the most used areas by people. Non-timber plant resources are relatively abundant and extraction rates did not appear to endanger their populations. However, the extraction of Ternstroemia lineata flowers for commercialization as medicine, and fuelwood of Pinus pseudostrobus, Quercus spp., Abies religiosa, Alnus spp., and Comarostaphylis spp. used and commercialized, may represent serious risks to sustainable maintenance of their populations. In addition, the timber extraction to which the forest region has been subject for decades severely threatens the integrity of the forest ecosystems of the Monarch Butterfly Biosphere Reserve. The bases for sustainable use of forests are the traditional forest management practices and some aspects about these practices are discussed.

Introduction This research was conducted to document the botanical knowledge of the Mazahua, an indigenous group belonging to the Otomanguean language family with two linguistic variants, the eastern or jnatrjo and the western one or jnatjo. In the Mazahua communities, the language, the traditional dress for women, the customs, and traditions are preserved (Forero 1997; Hopkins and Josserand 1979). In 2015, a total of 147,088 speakers of the Mazahua language were registered by the governmental census (INEGI 2015). The Mazahua communities live mostly in Central Mexico, in ten municipalities of the state of Mexico and seven of the state of Michoacán, and this area includes the Monarch Butterfly Biosphere Reserve (CONANP 2001). People in some Mazahua villages mainly practice productive activities such as agriculture, cattle and sheep raising, extraction of forest products, and elaboration of diverse handicrafts made of wool, mud ceramics, and wood. The region covers about 56,260 ha (CONANP 2001) and has a high biological diversity that includes more than 400 plant species within the core zones of the Biosphere Reserve (CornejoTenorio et al. 2003). The Mazahua, Matlatzinca, Otomí, Nahua, and Purépecha groups have occupied and managed the Michoacán territory of the Biosphere Reserve for centuries. Forests have been cleared up progressively to support agriculture and domestic animals to sustain the settlements as well as to provide wood for the establishing of villages and

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . .

665

people subsistence. Clearing of lands for agriculture has determined a continuous pressure on local ecosystems since pre-Hispanic times. However, pressure has been particularly strong during the last decades, mainly because of illegal logging carried out by organized crime, and partly also because the communitarian exploitation of natural resources increased with the population growth. A dynamic of forest loss from Abies and Pinus 1,200–2,500 ha was recorded for the Monarch Butterfly Biosphere Reserve during the period 2000–2006, and an annual rate of 329–500 ha per year during the period from 2006–2010 (López-García 2007; Champo-Jiménez et al. 2012), and the process continues until the present. Deforestation has been caused by agricultural clearing and domestic legal timber extraction and, therefore, alternative land use is necessary to conserve the regional forests and recognize their contribution to local subsistence of non-timber forest resources occurring in forest areas. However, the main destructive activity is the illegal extraction of wood by the organized crime, a great and complex problem that requires special attention by the Mexican authorities. Here we will analyze the possible plans and activities that can be promoted in association with local people, but the control of the timber extracting mafias by the government is the most imperative action. Worldwide, there is an increasing interest in using and commercializing non-timber forest resources. Some of them have high cultural and economic value and are considered as actual or potential bases for practices of low impact on ecosystems that could substitute timber extraction and would contribute to conservation of forests (Balick and Mendelsohn 1992; Salick et al. 1995). Such resources offer the possibility of finding valuable options of non-timber forest products for both conservation and wealth of local peoples. Searching for such options requires documenting cultural and economic aspects in relation to use of plant resources and traditional management techniques practiced on them, as well as ecological information of plant populations and communities, which may make valuable contributions for defining criteria to sustainable management of forests. These are currently among the main challenges of ethnobotanical and ecological research that can contribute with proposals for sustainable use and conservation of plant resources and the ecosystems these belong to (Salick et al. 1999; Shanley et al. 2002). Use and commercialization of non-timber plant resources are potential alternative practices with low impact on ecosystems in some areas (Shanley et al. 2002). Non-timber forest resources offer products such as food, fodder, medicine, materials for construction, handicrafts, and fuel, and the indigenous peoples of the region we studied have used these resources for centuries. Therefore, identifying useful resources, evaluating their economic potential, and designing forms of utilization to ensure their future maintenance should be the first aim of ethnobotanical research to contribute to their conservation and sustainable management. This information may be a valuable support, based on the local knowledge, forms of use and management, and helping to estimate the abundance and the rates of extraction of wild and weedy plant resources, especially those that are more important for people’s subsistence. This information shared with local authorities of the communities, peasant households, and schools may be useful to contribute to design programs for preservation of resources and maintaining the biocultural memory.

666

B. Farfa´n-Heredia and A. Casas

For designing strategies of resource management congruent with the principles of sustainability, it is firstly necessary to document what biotic resources are available in an area, those that are wild, weedy, and domesticated, and identifying which are considered more important resources for subsistence by local people. Then, it is helpful to investigate how is the distribution and abundance of the principal resources, and what is the biodiversity they are associated to. In addition, what are the current rhythms of extraction of useful products and those that would be the appropriate to ensure the long-term maintenance of plant populations and biotic communities they belong to. Ecological studies of populations and communities of plant resources, especially of those species with commercial value and more intensively extracted, could establish the bases to design strategies of sustainable management and techniques to increase their availability (Flachsenberg and Galletti 1998; Von Gadow et al. 2012). For such purpose, it is particularly helpful to characterize the diversity of useful plants existing in the different local ecosystems, as well as their abundance and temporal availability. This information may help identifying those species endangered under the present patterns of extraction, and may be the basis for planning extraction of products under sustainable criteria. We conducted our research having these premises in mind. In this chapter, we summarize the information we documented on the Mazahua traditional ecological knowledge of plants, including nomenclature, classification, forms of use and management, and their role in local people subsistence. We aimed, in principle, to generate an inventory of the useful plants locally available but thinking in the socioecological information needed to support sustainable management plans. We emphasized our efforts in identifying those species with higher risk to disappear under the current extraction rates, and those species offering potential opportunities for production. We finally discuss the management requirements for local sustainable use of plant resources.

Materials and Methods Study Site The research was conducted in the village of Francisco Serrato, municipality of Zitácuaro, Michoacán (Fig. 1). The region has an elevations range between 2,400 and 3,000 m with temperate climate. Natural vegetation with areas of pine forests (mainly Pinus pseudostrobus) and pine in association with oaks (Quercus laurina, Q. rugosa, and Q. crassifolia), with oaks and firs (Abies religiosa), and with the Mexican alder Alnus jorullensis and A. acuminata. People maintain the Mazahua language and cultural customs, including the use of traditional clothes among women (Fig. 2). Agriculture is the main economic activity, involving the seasonal cultivation of maize, beans, wheat, barley, and potatoes. Moreover, raising of chickens, turkeys, pigs, and sheep, as well as the extraction of wood for the construction of houses, fences, tools, and for commercialization are also relevant activities. Gathering of plants and other resources complements the

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . .

667

Fig. 1 Geographical location of the village of Francisco Serrato, in the municipality of Zitácuaro, Michoacán, Central Mexico. (Geographic information source Esri Satellite imagery)

peasant economy, contributes to the fulfillment of the needs for food, medicine, and fodder, and is a source of monetary income from the commercialization of useful products (Farfán et al. 2007). The land tenure is collective, mostly communal (89.8% of the total area) and ejidal (10.2% of the total area).

Data Collection Ethnobotanical research included the documentation of plant names, uses, and forms of management, with the collaboration of 30 Mazahua people. For documenting common names of plants, lists of names were audio-recorded in the Mazahua language, then transcribed with the help of a person speaking the Mazahua language with knowledge in Mazahua linguistics and expertise in editing educational books in this language. Mazahua classification of plants was analyzed based on nomenclature and through close-ended questions showing voucher specimens to five people. Botanical specimens were collected in different types of forest vegetation as well as in agricultural systems, disturbed areas, and roadsides. Voucher specimens of plants were deposited into the National Herbarium of Mexico (MEXU), as well as in the herbaria of the Universidad Michoacana de San Nicolás de Hidalgo (EBUM), and the Instituto de Ecología, A. C. at Pátzcuaro, Michoacán, Mexico (IEB).

668

B. Farfa´n-Heredia and A. Casas

Fig. 2 Mazahua people: (a) Mazahua women in traditional dress, (b) traditional food. (Photo Aviña, 2014 http://jeseltorres.blogspot.com/2014/11/mazahuas.html)

The interviews were used to record information on the production and consumption of forest products to establish the basic subsistence pattern of local people and to analyze the role of forest products in subsistence. This information was compared with ecological information on distribution and abundance of plant resources.

Mapping of Environmental Units General environmental units analyzed included the natural vegetation types, and the agricultural areas. These environmental units were mapped by using geographical information system, and field observations georeferencing points through a GPS, which allowed the estimation of the extent of areas of the environmental units.

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . .

669

Distribution and Abundance of Useful Plants Vegetation was sampled in each environmental unit through 500 m2 squares, subdivided into five 100 m2 squares. A total of 1,500 m2 were sampled in each of the different vegetation types. All individuals of shrubs and trees species included in the squares sampled were counted to estimate their density (total number of individuals of each species per sampled unit). Herbaceous plants were sampled through 1 m2 square, which were placed once, at random, within each of the 100 m2 squares. Agricultural fields comprised only herbaceous plants and these plants were sampled similarly as in forest units. Plant diversity of the different vegetation types was calculated based on data of density of tree and shrub species in the sites sampled, pooled per vegetation type. Diversity of herbaceous plants in the different environmental units studied was calculated based on data of presence/absence of plant species.

Temporal Availability of Useful Plants Information on temporal availability of useful parts of all useful plant species was obtained through structured interviews with people of 30 Mazahua households.

Results Traditional ecological knowledge. Mazahua nomenclature and classification of plants. The Mazahua system of classification of the natural world considers as first level of inclusion the moving elements of nature (tenxe yo nyomb’u), a group that comprises all biotic and abiotic elements that have movement such as animals, water, and wind. The other group at this first level of inclusion clusters unmoving elements (tenxe yo dya nyomb’u), which include plants, rocks, and mushrooms. The Mazahua system of classification of plants uses five life-form categories: xizaa meaning tree and stick, which includes trees and woody tall shrubs, pzijño that includes small shrubs and tall herbaceous woody plants, ts’i pzijño grouping small herbs, and ngüeí that comprises some ferns, since some of them are grouped with the small herbs (i.e., Adiantum andicola is called ts’i banjua and Cheilanthes chaerophylla is called ts’ngüeí; the prefix ts’ is used to name herbs). The prefix nrr includes plants producing large, beautifully colored flowers, no matters if they are herbaceous or woody plants. Mazahua nomenclature uses generic names and commonly includes a prefix or suffix indicating the “life-form” category of a plant. For instance, the prefix ts’ indicates that the plant is a small herb (Table 1), while the suffix pzijño includes shrubs and tall herbaceous plants. Thus, in the Mazahua nomenclature the plants of the Poaceae and Cyperaceae families are considered small herbaceous plants and are grouped under the word ts’nrreb’í. Trees are named by using the prefix xi (apparently a contraction of the term xizaa); for instance, Pinus pseudostrobus is named xivatí, Quercus laurina is xizhaa, and Q. rugosa and Q. crassifolia are named using the term xibatr’i.

670

B. Farfa´n-Heredia and A. Casas

Table 1 Mazahua nomenclature at the form level (In the Mazahua names, letters in bold indicate the classifying terms) Life form Trees

Small herbs

Shrubs and large herbs

Ferns

Scientific name Pinus spp. Buddleia cordata Quercus spp. Iresine celosia Euphorbia dentata Salvia fulgens Eupatorium spp. and Stevia spp. Salvia mexicana Polygonum aviculare Chelianthes chaerophylla Adiantum andicola

Mazahua name xivatí xij’ta xilojo Ts’inguitz’ajna ts’inrroí ts’inbarenzé Pepziño mb’opziñotr’eje pziñodyotr’eje ngueí ts’ingüeítr’eje

The Mazahua nomenclature is commonly binomial, with generic categories of plants, named considering the morphological similarities among them. For instance, most plants of the genus Salvia are classified under the term k’anrrejna, whereas plants of the genera Tagetes and Senecio are grouped together under the term k’axtr’u nrran-j’a. Additionally, names include specific epithets indicating the place where the plants are found or distinctive morphological characteristics. For instance, people use the word tr’eje to indicate that plants are from the forest, as in the case of Cheilanthes chaerophylla which is called ts’ingüeí tr’eje, a term indicating that these plants are herbaceous (ts’) ferns (ngüeí) and occurring in the forest (tr’eje). Use forms of plants. A total of 213 useful plant species were identified in the community studied (Table 2). Most useful plant species are wild (164 species), mainly occurring in pine-oak forests (94 species) and pine-oak-fir forests (81 species), with species overlapping their distribution in both types of forests. Disturbed habitats support 52 useful species of weedy plants and 29 species of useful ruderal plants. A total of 30 plant species used in the village were cultivated and six of them (Crataegus mexicana, Cupressus lusitanica, Phaseolus coccineus, Prunus serotina subsp. Capuli, Rubus liebmannii, and Tagetes erecta) occur also in wild populations in the forests surrounding the village (Fig. 3). Most plant species are used as fodder, human food, medicine, as well as commercial and fuel wood (Table 2). A total of 34 plant species (20 edible species, 7 species used as wood or fuel wood, 5 medicinal plant species, and 2 species of Agave used for the production of fermented beverage called “pulque,” which results from the fermentation of the sap extracted from mature agaves) are sold in local and regional markets (Appendix 1). Role of extraction and consumption of wild fruit in peasant subsistence. To complement their subsistence, the Mazahua families collect fruits of 17 species that are consumed throughout the year. Some of them are in addition sold in local and regional markets, allowing income; these are for instance the cases of blackberries (R. liebmannii), “capulín” (P. serotina subsp. capuli), and “tejocote” (C. mexicana), which are the species with the highest volume of fruits extracted, consumed, and

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . .

671

Table 2 Use form of plant species from Mazahua community Francisco Serrato Use form Fodder Medicine Food Ornamental Fuelwood Handcrafts Construction Soap

Number of species (n ¼ 213 spp.) 141 59 56 18 16 9 8 6

Percentage (%) 66.20 27.70 26.29 8.45 7.51 4.23 3.80 2.82

marketed in the closest city of Zitácuaro, Michoacán. Fruits of R. liebmannii were used by 90.5% of households, mainly fresh or prepared in flavored water, and in “atole,” that is a traditional beverage prepared with maize dough and water, consumed for breakfast and dinner. It was estimated that on average 6 kg of blackberries are consumed per family per week during 12 weeks of the production season, between March and August every year. And nearly 19% of households of the village sell between 2.3 and 8.0 kg of blackberries per week in the market of the city of Zitácuaro. We estimate that 1.04 t of blackberries were consumed in the village per season and that 3.36 ton marketed per year (in total 4.40 ton harvested per year). Similarly, we calculate that the fruits of C. mexicana were consumed by 81% of the households interviewed. The average consumption was 1.0–0.20 kg per household per week during 12 weeks of the production season, estimating 1.82 ton consumed per year in the village. All people of the village consume fruits of P. serotina fresh because they are highly appreciated for their flavor. These trees are commonly distributed in the borders of roads, gullies, small rivers, and homegardens. Fruits of P. serotina are available for 12 weeks from April to June. It is difficult to estimate the amount of capulín fruits used by the local population, since people not only collect them to consume at home, but also consume the fruits in the field, when carrying out agricultural or forestry activities. According to the interviews we carried out, the average consumption of fruit of this species was 1.8–0.4 kg per household per week, which allowed estimating a total of 4.04 ton per season by the whole community. Nearly 14% of the households sell fruits of P. serotina, on average of 10–12 kg of fruit per household per week, which allows estimating 3.43 ton of fruit harvested per season. Finally, the total amount of the capulín fruit harvested in the village is 7.47 ton. Other fruits are commonly used in smaller amounts than those mentioned above, generally in the sites where they are collected. Solanum appendiculatum, Cestrum tyrsoideum, Prunus brachybotrya, Comarosthaphylis longifolia, the achenes of Cirsium anartiolepis, and acorns of Q. crassifolia are among the most relevant. The Mazahua people use plants whose parts are consumed as greens, which are traditionally called “quelites.” The quelites mainly used by the Mazahua are Amaranthus hybridus, Brassica campestris, Chenopodium berlandieri, Dysphania ambrosioides, Cucurbita spp., Drymaria cordata, Galinsoga parviflora, Malva parviflora, Oxalis alpina, Phytolacca icosandra, Portulaca oleracea, Rorippa

672

B. Farfa´n-Heredia and A. Casas

Fig. 3 Diversity of plant species used in the Mazahua Francisco Serrato Michoacán community: edible fruits of (a) Prunus serotina, (b) Rubus liebmannii, (c) Crataegus mexicana, (d) medicinal plant Clinopodium macrostemum, (e) plant use as a condiment Dysphania ambrosioides, (f) Agave sp., “quelites” edible plants, (g) Amaranthus hybridus, (h) Chenopodium berlandieri, (i) dry flowers and fruits of Ternstroemia lineata, and (j) tree trunks for firewood. (Photos: Berenice Farfán-Heredia)

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . .

673

nasturtium-aquaticum, Solanum nigrescens, and Stellaria cuspidata (Appendix 1). All of them are collected during the rainy season, except R. nasturtium-aquaticum that is available the year round. Some collected plants are used to prepare beverages, one of the most important is Clinopodium macrostemum, whose leaves are boiled to prepare a stimulating infusion that is drunk in breakfast and dinner. Nearly 65.4% of the households interviewed consumed leaves of this species, with a frequency of one to 3 days per week, which is equivalent to 150–50 g per household per week, that is, nearly 1,078 ton in the whole town per year. The rest of the households used the infusion occasionally. The alcoholic beverage “pulque” is occasionally prepared and consumed with sap extracted from Agave atrovirens and A. salmiana. Approximately 28.6% of the families interviewed used medicinal plants, mostly cultivated specimens. The volume of extraction of plants or plant parts for this purpose is low, therefore the impact seems to be insignificant. However, T. lineata, Cuphea aequipetala, Equisetum sp., and Salvia lavanduloides, which are collected from the wild, are commercialized and their extraction may be significant. For instance, the extraction of flowers of Ternstroemia spp. was motivated by their market value. A medium-sized tree of Ternstroemia spp. may produce 20–35 bags of dry flowers, weighing 150 g per bag, which is approximately $1.30 US dollars per bag and US $45.50 per tree. Unfortunately, the effect of this trade represents a risk since it motivates people to commonly cut the main stems of trees to collect flowers more easily. Although some trees survive and sprout, more than 20% die. All households of the village cook with firewood. Some families have gas stoves but these are used only occasionally. The plants most commonly used as fuelwood are Alnus acuminata subsp. arguta, Q. crassifolia, Q. laurina, Buddleia cordata, P. pseudostrobus, T. lineata, C. longifolia, and Baccharis conferta. The average amount of fuelwood used per family per week is 196.9–45.3 kg, which means approximately 9.4 ton per year per household. The entire village consumes nearly 1,767 ton of firewood per year. Use of fuelwood is perhaps one of the major impacts on forest ecosystems, since unlike the timber extraction, obtaining firewood is a daily activity. The raising of domestic animals is usually through free grazing in fields, but 9.5% of households gather wild plants to feed domestic animals managed in pens. The plant species used as fodder are particularly important in peasant economy as long as raising of domestic animals is a source of monetary income, labor animals, meat for direct consumption, and wool for textile manufacturing. At least 68 plant species are used to raise animals in the wild, but also some of them are collected to directly feed the animals in small stables. On average, they collect 5.5–1.25 kg of plants to feed three lambs per day, and 5.0–1.5 kg to feed one horse or donkey per day.

Environmental Units The territory of the Mazahua community of Francisco Serrato is constituted by an area of 2,039.83 ha, of which 1,831.95 ha are under communal property regime, which represents 89.8% of the total area, and 207.88 ha under ejido property regime, which represents 10.2% of the total area.

674

B. Farfa´n-Heredia and A. Casas

Nearly one half of the territory of this community was covered with forests in the year 2000, whereas the other half of the territory was covered by transformed landscapes, including agricultural areas and shrub grasslands. However, during the period 2001–2003, a forest area greater than 350 ha was lost (more than 17% of the total area owned by the communal and ejidal properties), mainly on ejidal property, due to the intensive extraction of wood illegally in Abies forest (Fig. 4) (Ramírez et al. 2008; W.W.F. 2004). By the year 2020 nearly half of the forest area became transformed. By 2013 the transformed areas had secondary shrub vegetation, without clear ecological recovery processes (Salas 2013), and such condition is similar at present. In the year 2000 the fir (pine-oak) forest covered 23.43% of the territory of the village and it was characterized by the dominance of Q. laurina, P. pseudostrobus, and the fir A. religiosa. Other arboreal species that occurred in this vegetation type were P. brachybotrya, Alnus spp., Cornus disciflora, Oreopanax xalapensis, and Clethra mexicana. Among the shrubs, Cestrum nitidum, Solanum argentinum, and Arracacia rigida were the most common, whereas in the herbaceous stratus Solanum appendiculatum, Adiantum andicola, Tradescantia commelinoides, Salvia mexicana and Stachys sp. were the most abundant. The pine-oak forest occurs in elevations of 2,600–2,900 m. It covered 15.64% of the territory of the village and was dominated by P. pseudostrobus in association with Q. laurina, A. acuminata, P. brachybotrya, B. cordata, C. disciflora, and

Fig. 4 Common use areas of the Mazahua Francisco Serrato community, Michoacán, México. Visualization of deforested areas of the Francisco Serrato territory

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . .

675

A. religiosa. Shrubs were represented by C. thyrsoideum, A. rigida, S. argentinum, Fuchsia thymifolia, and Cirsium anartiolepis, whereas the representative herbaceous plants were Salvia fulgens, Salvia mexicana, Geranium seemannii, Lopezia racemosa, and Acaena elonganta. The pine-oak-alder forest covered approximately 5.24% of the territory in elevations between 2,500 and 2,720 m. It was dominated by Q. laurina, Q. rugosa, Q. crassifolia, P. pseudostrobus, Populus sp., Alnus acuminata, T. lineata, Comarostaphylis discolor, C. longifolia, and Oreopanax xalapensis. Among the shrubs, Baccharis conferta, Euphorbia dentata, Fuchsia thymifolia, Cestrum nitidum, and Solanum argentinum were the most abundant, whereas the main herbaceous plants were Lopezia racemosa, Salvia mexicana, and Solanum appendiculatum. The riparian vegetation covered nearly 10.30% of the territory of the village, including gullies with pronounced inclination and small rivers. The arboreal stratus included A. acuminata, B. cordata, Prunus sp., P. brachybotrya, T. lineata, Fraxinus uhdei, Q. laurina, C. mexicana, O. xalapensis, and P. pseudostrobus. The shrubby stratus was dominated by F. thymifolia, A. rigida, R. liebmannii, S. appendiculatun, Solanum nudum, Salvia elegans, Iresine diffusa, Salvia mexicana, Salvia fulgens, S. argentinum, and B. conferta. The agricultural areas comprise nearly 43.5% of the territory, mainly included fields of seasonal agriculture of maize and wheat, and fallow agricultural fields. The most common herbaceous plants occurring in agricultural areas were Tradescantia commelinoides, Prunella vulgaris, Reseda luteola, Cyclanthera ribiflora, Acaena elongata, B. campestris, Phytolacca icosandra, Oxalis alpina, Piqueria pilosa, L. racemosa, Lepidium virginicum, Melampodium divaricatum, A. hybridus, Galinsoga parviflora, Simsia amplexicaulis, Lupinus elegans, Jaltomata procumbens, Gnaphalium attenuatum, and Bidens odorata. The scrub grasslands were in elevations between 2,500 and 2,900 m, covered nearly 6.66% of the territory of the village. Trees were absent and the dominant shrub was B. conferta. A high diversity of herbaceous plants were found, among them Lepechinia caulescens, A. elongata, L. racemosa, R. nasturtium-aquaticum, Tradescantia commelinoides, Trifolium amabile, Tinantia erecta, C. aequipetala, Fragaria mexicana, and Mimulus glabratus. During the period from 2001 to 2003, the illegal logging decreased the forest area. The entire area of ejidal-owned Abies forest (207.88 ha) and 146.48 ha of communal owned Abies forest were lost, which represented a loss of 17.4% of the total area of the territory of the Mazahua community of Francisco Serrato (Table 3; Fig. 5).

Distribution of Useful Plants in the Environmental Units A total of 72 useful plant species were recorded in the sampled sites, 52 of them (16 trees, 14 shrubs, and 22 herbs species) in forest areas and 20 in agricultural field. Plant species recorded included plants used as fodder, food, fuelwood, medicine, manufacture of tools, ornamental, and soap.

B. Farfa´n-Heredia and A. Casas

676

Table 3 Area by type of vegetation and deforested area during the period from 2001–2003 Vegetation type Agricultural areas Fir forest

Total área (Hectares) 897

Total área (Percentage) 43.97

Deforested area in 2001–2003 (Hectares)

Total área (Percentage)

477.9

23.43

207.88 Ejidal property 146.48 Comunal property

10.1 7.1

Pine-oak (Alder) forest Riparian vegetation Scrub grasslands

319

15.64

210.2

10.30

135.8

6.66

The riparian vegetation had the highest richness and diversity of useful plant species, including the 53.8% of the useful plant species that recorded 62.5% of the useful trees, 78.6% of the useful shrubs, and 31.8% of the useful herbs in all units sampled. However, access to plant resources of this vegetation types is relatively difficult since it mostly occurs in areas of abrupt terrain. The different types of pine-oak forests have been the most important areas supplying plant resources to the Mazahua of Francisco Serrato since they covered nearly one-third of the territory of the village, and density and biomass of useful species in this vegetation type is higher than in others. The association of pine-oak-fir, for instance, included 46.2% of the useful plant species recorded (50% of the useful trees, 57.1% of the useful shrubs, and 36.3% of the useful herbs recorded in all units sampled), covering 17.8% of the surface of the territory of the village. However, because of the abundance of important timber producing species P. pseudostrobus and A. religiosa, this vegetation type has been the area with the highest intensity of timber extraction and, for this reason, the abundance of non-timber products in this vegetation type offered in the year 2000 possible options for sustainable use, but after their severe transformation the recovery of the forest is a priority. The different types of pine-oak forest included also a broad spectrum of useful plants in areas close to the settlements of the village and, therefore, it was an area of easy and continuous access. Unfortunately, this area has been under continuous extraction of wood and fuelwood and, therefore, densities of timber species such as P. pseudostrobus, Q. laurina, Q. crassifolia, Q. rugosa, and A. acuminata are relatively low (Table 4). Useful trees and shrubs were relatively scarce in the pine-oak-alder forest, but in this vegetation type species of useful herbs were rich, including 50% of the useful herbs recorded (Table 4). The riparian vegetation supplies more than 60% of useful plants for different purposes, including all edible plants recorded, whereas the pine-oak forest supplied more than 60% of edible plants. The agricultural areas were also important reservoirs of useful noncultivated plants. Corn fields contained nearly 30% of useful plants, especially weedy plants. Among the weedy plants, the “quelites” were consumed from June to August. In addition, 13 species used as fodder gathered to feed chickens were available in maize

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . .

677

Fig. 5 Deforested area in the ejidal and communal property of the Francisco Serrato community: (a) forest area in 1994, (b) deforested area in the period 2001–2003

fields. The wheat fields contained nearly 15% of the useful species. In the fallow agricultural fields plant resources were mostly fodder and these were commonly used for free raising cattle, sheep, and donkeys.

B. Farfa´n-Heredia and A. Casas

678

Table 4 Vegetation types their extent within the village of Francisco Serrato, and the pooled diversity of trees and shrubs and herbs

Vegetation type Fir (Pine-oak) forest Pine-oak (alder) forest Riparian vegetation Scrub grass area Agricultural fields Total area

Area (ha) 477.9

Percentage of the total 23.43

Diversity of trees and shrubs Shannon Simpson H 1/D 0.950 6.794

319

15.64

1.004

6.307

0.778

210.2

10.30

1.114

10.474

0.903

135.8 897 2,039.9

6.66 43.97 100

0.061 0.000 –

1.055 0.000

1.000 1.342 –

Diversity of herbs Shannon H 1.041

Temporal Availability of the Main Useful Plants Perennial plants used as fodder, mainly shrubs, are available throughout the whole year, but annual plants are available mainly during the rainy season from June to September, and some of them until October. Availability of medicinal plants is variable, depending on the plant part used. For instance, perennial plants whose cortex is the useful part, as it is the case of B. cordata, are available the year round; it is also the case of perennial plants whose leaves continually produced are the useful part, as in S. argentinum, Salvia lavanduloides, or Lippia umbellata. However, some species produce seasonally useful products, as in the case of the flowers of T. lineata, available from November to January or, in the case of annual plants, the season they are available, as in the case of G. attenuatum, Tagetes lucida, and Verbena litoralis, among others. Availability of edible plants depends on the production season of the edible parts. Some of them are available throughout the year as are the cases of Stellaria cuspidata and Salvia fulgens. In these cases, the whole plant and the stems, respectively, are consumed. Availability of edible fruits generally occurs during relatively short times, but there is availability of products of at least one species the year round. Greens or quelites are seasonally available, mainly during the rainy season, but others such as R. nasturtium-aquaticum, which are aquatic plants, are available in January and February. In the agricultural areas, availability of herbaceous plants is restricted to the growth periods of crop plants in the case of corn fields from April to September whereas in the case of wheat fields from August to January. B. campestris, A. hybridus, O. alpina, and J. procumbens are available from June to August, when staple crops have not been still harvested, and for that reason these plants complement importantly the diet of local people. Plants used as fuelwood are available the whole year.

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . .

679

Conclusions and Perspectives The Mazahua system of classification and nomenclature of plants follows the general principles of classification presented by Berlin (1992), which consists of hierarchical systems with inclusion levels, which include not only plants with some form of use, but also unused plants (Berlin 1992). In general, the Mazahua classification system is based on the affinities of morphological characteristics, habits and habitats, independently of their cultural importance. The Mazahua use a wide spectrum of resources, we recorded 213 species of plants that use them in their daily life. This pattern of diversified use of plant resources is consistent with the pattern defined as “multiple utilization of natural resources and ecosystems” that generally characterizes the form of subsistence of indigenous communities in Mesoamerica (Farfán-Heredia et al. 2018a, b; RangelLanda et al. 2016; Blancas et al. 2010; Casas et al. 1994, 2014; Caballero et al. 1998; Caballero and Mapes 1985; Toledo et al. 1978). The Mazahua include wild plants and fruits in their diet. For instance, the “quelites” are the non-crop plants mostly used as food. These plants are available during the period from June to August, beginning the rainy season, at the time when the basic grains have just been sown and the harvest of the previous year has decreased. During this period, wild and weedy edible plants play an important role in the diet of Mazahua households. Agriculture is not sufficient to satisfy the needs of Mazahua households studied. To compensate, they obtain a significant number of species and biomass of products for both direct consumption and commercialization. They regularly sell wild fruits such as P. serotina and R. liebmannii, to buy basic grains. They also collect and consume wild edible plants to complement their diet. The volume of extraction of medicinal plants and the impacts due to this activity are mostly insignificant. However, the destructive form of extraction of Ternstroemia spp. is relevant. The occurrence of 150 individuals of Ternstroemia spp. per hectare in riparian vegetation was recorded, covering an area of 210 ha of the territory of the community. Almost 30% of these trees have signs of having been cut for harvest, with evidence of buds that arise from the main trunk. This practice endangers the populations of Ternstroemia since nearly 20% of the trunks tamed do not regenerate, which indicates the need to stop this form of extraction by constructing local regulations, designing appropriate gathering tools, and conducting environmental education programs. The species of plants used as firewood are indispensable in the local economy, since all people use firewood for cooking, as well as a source of monetary income. According to the people interviewed, the extraction of firewood does not involve the cutting down of trees, but only the harvesting of dead branches and trees. Nevertheless, the total annual consumption of fuelwood is high and the promotion of strategies for more efficient consumption of fuelwood could contribute to the reduction of wild fuel consumption. Programs of efficient stoves may be highly helpful for reaching successfully such purpose.

680

B. Farfa´n-Heredia and A. Casas

Forage plants are important in the peasant economy, as raising domestic animals is a source of meat, work animals, and monetary income. Therefore, designing sustainable patterns of livestock raising, including silvo-pastoralism and agrosilvo-pastoralist systems are a priority. “Quelites” and wild fruits are a direct source of food and monetary income throughout the year. For maintaining or increasing their availability, it is essential to design sustainable management strategies of use, production, and commercialization of non-timber forest. For maintaining sustainable resource management strategies, it is necessary to analyze the spatial and temporal availability of plants and extraction rates to ensure their long-term maintenance. For instance, we documented that the spatial availability of P. serotina is 126 individuals per hectare in the riparian vegetation, covering 210 ha of the territory of the village. This species produces on average 6,200–447 fruits per adult tree, equivalent to 11.4–0.8 kg of fruits per tree, and approximately 302.6 ton available in the village. Considering that people consume nearly 7.47 ton per year, the ecological information suggests that the impact of harvesting these fruits is relatively low and that it would be possible to increase this practice to develop alternative commercial products. In a similar way, we estimated that approximately 34.18 ton of fruits of C. mexicana were produced annually in the village, but people consume 1.82 ton per year. However, R. liebmannii produces nearly 6 ton of fruit per year and people ingest 4.4 ton yearly, nearly 70% of the total production. This information suggests that the extraction rate could affect the maintenance of the populations of this species in the long term. We estimate that the agricultural area of the village, which covers an area of 902 ha, produces 23.6–46 ton of the “quelites” B. campestris and A. hybridus, respectively; and that the annual consumption of these plants was 4.3 ton and 0.9 ton, respectively, which indicates that the harvest rates of these resources are relatively low. In this study, we inventoried information on use forms, distribution, abundance, and availability of non-timber plant resources with a potential for exploitation that allows the social and economic development of the community. This information may be a basis for establishing strategies for extraction, processing, and marketing of these products in a sustainable way, for the diversification of the use of wild resources in the community. Although in the current scenario of dynamics of forest cover loss, the wide range of non-timber forest resources is not considered (Fig. 4). Those that provide local means of subsistence to peasant communities, for food, for health, as biofuels, as ritual and ceremonial elements, for the elaboration of utensils, crafts and for smallscale livestock. In addition, forests are the setting to generate and transmit traditional knowledge, worldviews, ecological ethics, and traditional management practices from the worldview of indigenous and peasant communities. The information reported has been the product of the articulation of different methods and techniques of research, description, systematization, and data analysis, under the focus of several disciplines such as botany, floristic, ecology, cartography,

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . .

681

sociology, and linguistics, all of them around the basics of ethnobotany. Such a multidisciplinary methodological framework allowed having a set of elements and aspects around the plant resources that have allowed, in the first instance, to describe the knowledge, use, and management of the plant resources of the Mazahua. In addition, to note the importance of the units of vegetation in terms of diversity, density, and biomass of non-timber plant resources used by the community and susceptible to use in the short, medium, and long term, all these information are in relation to the dynamics of consumption and temporary availability of plant resources. These elements of analysis appear to provide criteria for designing sustainable forms of using both resources and ecosystems (Bray et al. 2003). Given the processes of change in current ecosystems, it is of great value to have information on ecological, cultural, and economic aspects of ecosystems, which contribute to the subsistence of people from rural communities, recognizing the role of timber, non-timber forest resources and ecosystem services as fundamental to human well-being and the maintenance of diversity and biocultural heritage. Acknowledgments The authors thank financial support from the Consejo Nacional de Ciencia y Teconología (CONACYT, research project AS-1-14306), Mexico, the Dirección General de Asuntos del Personal Académico (DGAPA-PAPIIT, research project IN206520 and IN224023)), UNAM, and the Comisión para el Conocimiento y Uso de la Biodiversidad (CONABIO/GEF/FAO, GEF proyect ID9380 770j, research project RG-023), Mexico.

Appendix 1 Useful plants of the Mazahua village of Francisco Serrato. Habitats: 1 5 ruderal; 2 5 agricultural areas as weeds; 3 5 agricultural area as crops; 4 5 pine-oak forest; 5 5 pine-oakfir forest; 6 5 riparian vegetation; 7 5 shrub and grassland. Voucher specimens correspond to Farfa´n-Heredia collection numbers Plant family Pteridophyta Aspleniaceae

Species

Asplenium monanthes L. 4, 5, 6

ngüeí

Dryopteridaceae

Polystichum distans Fourn. Pleopeltis mexicana (Fée) Mickel & Beitel Polypodium madrense J. Sm Adiantum andicola Liebm. Cheilanthes chaerophylla (M. Martens & Galeotti) Kunze

2, 6

Polypodiaceae

Pteridaceae

Habitat Mazahua name Use

Voucher

ngüeí

Ornamental, medicinal Ornamental

9 342

4, 5

ngüeí

Ornamental

169

4

ngüeí

Medicinal

20,173

4, 5, 6

ts’ibankjua

Medicinal

8

4, 5

ts’ingüeí tr’eje Ornamental

200

(continued)

B. Farfa´n-Heredia and A. Casas

682

Plant family Monilophyta Equisetaceae Coniferophyta Cupressaceae

Pinaceae

Magnoliophyta Magnoliopsida Amaranthaceae

Apiaceae

Apocynaceae Araliaceae

Asteraceae

Species

Habitat Mazahua name Use

Voucher

Equisetum sp. Cupressus lusitanica Mill.

2, 6 3, 5

122 108

Abies religiosa (Kunth) Schltdl. & Cham. Pinus hartwegii Lindl.

5

Medicinal Handcrafts, construction, fuelwood, ornamental Construction, firewood Construction, firewood Construction, firewood Construction, firewood, medicinal Fodder

71

Food, fodder, medicinal Medicinal, fodder Fodder

Photo record 42

2 5

Ornamental Fodder

121 357

2, 4

Fodder

311

2, 4

Fodder

279

2, 5

Fodder

316

kox’y zez’ojnu

4, 5

xivatí

Pinus leiophylla Schiede 4 ex Schltdl. & Cham. Pinus pseudostrobus 4, 5 Lindl.

xivatí

Iresine diffusa Humb. & 6 Bonpl. ex Willd. Amaranthus hybridus L. 2

ts’inguitz’ ajna

Arracacia rigida J.M. Coult. & Rose Eryngium columnare Hemsl Vinca major L. Oreopanax xalapensis (Kunth) Decne. & Planch. Ageratum corymbosum Zuccagni ex Pers. Archibaccharis schiedeana (Benth.) J.D. Jacks. Archibaccharis hirtella (DC.) Heering Artemisia ludoviciana Nutt. Baccharis conferta Kunth

4, 5

z’apzantreje

5

nrab’idyi

Baccharis pteronioides DC. Barkleyanthus salicifolius (Kunth) H. Rob. & Brettell Bidens odorata Cav.

xivatí

349 76 198 93

24

2, 4

j’mipzi

Medicinal

127

1, 5

mb’axu

14

1, 2, 4

mb’axu

5

dyecha

Fodder, handcrafts, fuelwood Fuelwood, handcrafts Fodder

1,2, 4

ñinch’ujnu

Fodder

283

111 45

(continued)

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . . Plant family

Species Brickellia squarrosa B.L. Rob. & Seaton Cirsium anartiolepis Petr. Conyza coronopifolia Kunth Erigeron galeottii (A. Gray) Greene Erigeron karvinskianus DC.

683

Habitat Mazahua name Use 4 mb’opziño Fodder

Voucher 56

5

Food, fodder

91

Fodder

163

nrrab’idyi

2, 4 2, 4

nrrab’a

Fodder

144

4

nrrab’a

180

Erigeron longipes DC. Ageratina areolaris (DC.) Gage ex B.L. Turner Eupatorium sp. Eupatorium sp. 1 Galinsoga parviflora Cav. Gnaphalium attenuatum DC. Gnapahlium sp.

4 5

nrrab’a nupziño

Fodder, medicinal, ornamental Fodder Fodder

5 5 2, 4

kjoñizha pepziño

Fodder Fodder Food, fodder

22 4 323

5

gordolobo

Medicinal

5

gordolobo

Medicinal

Jaegeria hirta (Lag.) Less. Melampodium divaricatum (Rich. ex Rich.) DC. Melampodium perfoliatum (Cav.) Kunth Pinaropappus roseus (Less.) Less. Pinaropappus sp. Piqueria pilosa Kunth Roldana hederifolia (Hemsl.) H. Rob. & Brettell Roldana albonervia (Greenm.) H. Rob. & Brettell Roldana angulifolia (DC.) H. Rob. & Brettell Roldana barba-johannis (DC.) H. Rob. & Brettell Roldana suffulta (Greenm.) H. Rob. & Brettell

2, 4

Photo record Photo record 322

Fodder

179 299

4, 6

nrrab’a

Fodder, medicinal

214

4

nrrab’a

Fodder

293

4

pob’a

Fodder

113,162

5 4 5

pob’a mb’opziño

Medicinal Fodder Fodder

43 195 363

4, 5

mb’opziño

Fodder

21

2, 4

mb’opziño

Fodder

334

1. 2, 4, mb’opziño 5, 6 4 mb’opziño

Fodder

344

Fodder

361

(continued)

B. Farfa´n-Heredia and A. Casas

684

Plant family

Betulaceae

Brassicaceae

Species Senecio callosus Sch. Bip. Senecio stoechadiformis DC. Senecio toluccanus DC.

Habitat Mazahua name 2, 4, 6 k’axtr’u nrranj’a 1, 4 k’axtr’u nrranj’a 4, 5 k’axtr’u nrranj’a 4

Use Fodder, ornamental Fodder

Voucher 343

Fodder, ornamental Fodder

27

2, 4

Fodder

326

295

Sigesbeckia jorullensis Kunth Simsia amplexicaulis (Cav.) Pers. Stevia lucida Lag. Stevia serrata Cav. Stevia origanoides Kunth Stevia subpubescens Lag. Tagetes erecta L.

5 1, 4 4

pe’pziño nrr’axikjua pe’ pziño

Fodder Fodder Fodder

25,49 296 297

4, 6

pe’pziño

Fodder

13,288 192

Tagetes lucida Cav.

1, 2, 4

Tagetes micrantha Cav.

4, 7

Ornamental, medicinal, fodder Medicinal, fodder Medicinal, food Fodder

pe’pziño

Fodder

359

mb’e’e

Handcrafts, fuelwood, construction Handcrafts, fuelwood, construction Fodder

54

Fodder Food, commerce

307 33

Food, fodder, medicinal, commerce Food, fodder

Photo record

2, 3, 4, k’axtr’u 7 nrranj’a

Taraxacum officinale 1, 2 F.H. Wigg. Taraxacum officinale (L.) Weber ex F.H.Wigg. Verbesina klattii 5 B.L. Rob. & Greenm. Alnus acuminata subsp. 5, 6 arguta (Schltdl.) Furlow

k’axtr’u nrranj’a k’axtr’u nrranj’a ts’ik’axtr’un ajnua

Alnus jorullensis Kunth

5, 6

mb’e’e

Lepidium sordidum A. Gray Lepidium virginicum L. Rorippa nasturtiumaquaticum (L.) Schinz & Thell. Rorippa nasturtium-aquaticum (L.) Hayek Brassica rapa L.

2

ts’inrroí

2 6

ts’inrroí

Brassica campestris L.

2

2

346

385 337 149

159

106

480 (continued)

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . . Plant family Buddleiaceae Cactaceae

Species Habitat Mazahua name Buddleia cordata Kunth 5 xij’ta Opuntia ficus-indica 3 kij’i’i (L.) Mill.

Use Fodder Food, medicinal, fodder Fodder Fodder Fodder

44 252 148

5

Food, fodder

32

4 5

Fodder Food, fodder

88 92

Medicinal, fodder

105

333

Campanulaceae Lobelia laxiflora Kunth Caryophyllaceae Cerastium nutans Raf. Cerdia virescens Moc. & Sessé ex DC. Drymaria cordata (L.) Willd. ex Schult. Drymaria excisa Standl. Stellaria cuspidata Willd. ex Schltdl. Cistaceae Helianthemum glomeratum (Lag.) Lag. ex Dunal Clethraceae Clethra mexicana DC.

1, 5 4, 5 2, 4

Convolvulaceae Cornaceae

2 5

nrrempe

Handcrafts, soap Ornamental Fodder

4

zatrjo

Soap, fodder

Cucurbitaceae

Ipomoea sp. Cornus disciflora Moc. & Sessé ex DC. Cyclanthera ribiflora (Schltdl.) Cogn. Cucurbita argyrosperma C.Huber

2, 4

ts’irr’oztejna

xits’ajna

2, 4, 5

3

Cucurbita ficifolia Bouché Cucurbita moschata Duchesne

3

Cucurbita pepo L.

3

Chenopodiaceae Dysphania ambrosioides (L.) Mosyakin & Clemants Chenopodium berlandieri Moq. Ericaceae Comarostaphylis discolor (Hook.) Diggs

685

3

2

Xujmu

2 4

trjoxu

4

penchoxu

Voucher 48 131

269 247, 340 84

Food, fodder, soap, medicinal, commerce Food, fodder, commerce Food, fodder, soap, medicinal, commerce Food, fodder, soap, medicinal, commerce Food, fodder, medicinal, commerce Food, fodder, commerce Fuelwood

Photo record

Food, fuelwood

59

Photo record Photo record

Photo record

135

Photo record 478

(continued)

B. Farfa´n-Heredia and A. Casas

686

Plant family

Euphorbiaceae

Fagaceae

Species Comarostaphylis longifolia (Benth.) Klotzsch Comarostaphylis discolor subsp. rupestris (B.L.Rob. & Seaton) Diggs Euphorbia dentata Michx. Euphorbia pulcherrima Willd. ex Klotzsch Quercus crassifolia Bonpl.

Habitat Mazahua name Use

Voucher

5

trjoxu

Fuelwood

70

1

ts’inrroí

Ornamental

287

3

pascua

Ornamental

4, 6

Xilojo

Quercus laurina Bonpl.

4, 5

xilojo

Quercus rugosa Née

4, 5

xilojo

Handcrafts, food, fuelwood, medicinal Food, handcrafts, fuelwood Food, handcrafts, fuelwood, fodder, medicinal, construction Fodder

Photo record 203

Gentianaceae

Halenia brevicornis (Kunth) G. Don Geraniaceae Geranium seemannii Peyr. Hydrophyllaceae Nama prostrata Brand Phacelia platycarpa (Cav.) Spreng. Lamiaceae Lepechinia caulescens (Ortega) Epling Prunella vulgaris L. Salvia concolor Lamb. ex Benth. Salvia elegans Vahl Salvia fulgens Cav. Salvia carnea Kunth. Salvia helianthemifolia Benth. Salvia lavanduloides Kunth Salvia mexicana L. Salvia sp. Salvia sp.

7

53

51

351

5

xu’u

Medicinal

39

4, 5 4, 5

pzochavo

Fodder Fodder

80 183

2, 5

k’anrrejna

Fodder

318

4, 5 5

k’anrrejna k’anrrejna

Fodder Fodder

237 244

1, 4 5 5 5

k’anrrejna ts’imbarenzé ts’ixitz’ajna k’anrrejna

282 40 245 3

4, 6

k’anrrejna

2, 4, 5

mb’opziño tr’eje k’agrejne ts’ik’agrejne

Fodder Fodder Fodder Medicinal, commerce Food, medicinal Fodder Fodder Fodder Medicinal

208 60

6 6 4, 5

73 117

(continued)

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . . Plant family

Lauraceae

Leguminosae

Species Clinopodium mexicanum (Benth.) Govaerts Clinopodium macrostemum (Moc. & Sessé ex Benth.) Kuntze Stachys coccinea Ortega Persea americana Mill.

Habitat Mazahua name Use

Voucher Photo record

4, 5

485

Astragalus lyonnettii Barneby Dalea thouinii Schrank

4

Desmodium aparines (Link) DC. Desmodium densiflorum Hemsl. Lupinus elegans Kunth Phaseolus coccineus L. subsp. coccineus Phaseolus coccineus L. subsp. formosus (Kunth) Maréchal, Mascherpa & Stainier Phaseolus coccineus subsp. formosus (Kunth) Marechal & al. Phaseolus vulgaris L.

1

4 3

Medicinal

mbarejnatr’eje Fodder Food, medicinal, commerce kjungo’o Fodder

1 pe’pziño

2

Malvaceae

Onagraceae

146 97 Photo record 261

6

patol

Food, fodder

5

Cuphea jorullensis Kunth Kearnemalvastrum subtriflorum (Lag.) D.M. Bates Malva parviflora L.

4, 6

Fuchsia thymifolia Kunth

Fodder Fodder Food

Cuphea aequipetala Cav.

Food, commerce Fodder Food, fodder ngüechuxu

1, 5

nrr’akojnu

1

ts’inrr’ axkojnu

5

186 290

nrramona

3

85 Photo record

Fodder, medicinal Fodder

5 3

Trifolium amabile Kunth 2 Vicia faba L. 3 Lythraceae

687

Medicinal, food, commerce Medicinal, commerce Fodder

Food, ornamental, fodder, medicinal Fodder

289

Photo record 321 Photo record 36

222 285

305,306

13 (continued)

B. Farfa´n-Heredia and A. Casas

688

Plant family

3, 5 4, 5 2, 4 1, 6

Portulacaceae

Species Lopezia racemosa Cav. Oenothera deserticola (Loes.) Munz Oxalis alpina (Rose) Rose ex R. Knuth Oxalis sp. Phytolacca icosandra L. Peperomia galioides Kunth Peperomia quadrifolia (L.) Kunth Plantago australis Lam. Monnina ciliolata DC. Polygonum aviculare L. Persicaria lapathifolia (L.) Delarbre Polygonum sp. Portulaca oleracea L.

Primulaceae Punicaceae

Ranunculaceae Resedaceae Rhamnaceae

Oxalidaceae

Phytolaccaceae Piperaceae

Plantaginaceae Polygalaceae Polygonaceae

Rosaceae

Habitat Mazahua name Use 2, 4, 5 pe’jnche Fodder 4, 5 Fodder

Voucher 145 29,157

4, 6

k’opi

Food, fodder

213

4, 6 1, 2 1, 6

ixí cangara

Fodder Food, fodder, Soap, medicinal Medicinal

68 124 331

lant’a ts’ib’orencé pziñotr’eje

Medicinal Food, fodder Fodder Fodder

151 6 139 308

2 2

xu’u

Anagallis arvensis L. Punica granatum L.

4 3

pzochavo

133 Photo record 95 Photo record

Clematis dioica L. Reseda luteola L. Ceanothus caeruleus Lag. Frangula mucronata (Schltdl.) Grubov. Acaena elongata L. Alchemilla pringlei Fedde Alchemilla procumbens Rose Crataegus mexicana DC.

4 2 4, 5

kjopziodyo ts’ik’agrejne

Medicinal Food, fodder, medicinal Fodder Food, medicinal, commerce Fodder Fodder Medicinal

200 138 174

4, 6

kjonz’a

Fuelwood

75,138

4, 5 2

ñib’ota ñib’ota

Fodder Fodder

18 147

4, 6

ñib’ota

Fodder

230

3, 5

xinpedyi

99

Eriobotrya japonica (Thunb.) Lindl. Fragaria mexicana Schltdl. Prunus brachybotrya Zucc. Prunus domestica L.

2, 3

níspero

Food, medicinal, commerce Food

5, 7

muxatr’u

Food, fodder

Photo record 159

5

xínchoparu

Medicinal

90

3

endrina

132

1, 3, 4

xinrrenz’e

Food, commerce Food, commerce

4, 6

348

86 (continued)

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . . Plant family

Rubiaceae

Rutaceae Salicaceae

Species Habitat Mazahua name Prunus serotina subsp. capuli (Cav. ex Spreng.) McVaugh. Prunus persica (L.) 3 nrora Batsch Rubus liebmannii Focke 1, 3, 5 xarrtr’u Bouvardia ternifolia (Cav.) Schltdl. Didymaea floribunda Rzed. Ruta chalepensis L. Salix paradoxa Kunth

Scrophulariaceae Calceolaria mexicana Benth. Mimulus glabratus Kunth Penstemon campanulatus (Cav.) Willd. Veronica americana Schwein. ex Benth. Solanaceae Cestrum nitidum M. Martens & Galeotti Cestrum thyrsoideum Kunth Jaltomata procumbens (Cav.) J.L. Gentry Lycianthes rzedowski E. Dean Lycopersicon esculentum Mill.

Use

689

Voucher

142

4, 6

Food, commerce Food, commerce rnanta Fodder, medicinal xitz’ajna tr’eje Fodder

3 4, 5, 6

lota xijño

134 250

5

Medicinal Fodder, fuelwood Food

239,345

5, 7

Food

240

5

Fodder

337

5

Fodder

235

1

Ornamental

312

4

23 166 61

4, 5

rr’oxululu

Food

15

2, 4

nrrempe

Food, fodder

263

4

onglaba

Food, fodder

165,191 Photo record

3

Physalis volubilis Waterf. Solanum appendiculatum Dunal Solanum cardiophyllum Lindl. Solanum argentinum Bitter & Lillo Solanum nigrescens M. Martens & Galeotti

4, 6

ongo’o

Food, fodder, medicinal, commerce Food, fodder

4, 5

nts’imbalixi

Food, fodder

5

Fodder

336

Solanum nudum Dunal

1, 6

210

5

xíncho paru

Medicinal

41

2, 5

ongo’o

123

4

xínchoparu

Food, medicinal, fodder Ornamental

211 (continued)

B. Farfa´n-Heredia and A. Casas

690

Plant family

Symplocaceae Theaceae

Valerianaceae

Verbenaceae

Violaceae

Viscaceae Magnoliophyta Liliopsida Agavaceae

Commelinaceae

Cyperaceae

Species Solanum tuberosum L.

Habitat Mazahua name Use 3 Food, commerce Symplocos prionophylla 6 Fodder Hemsl. Cleyera integrifolia 5 nrrensétr’eje Medicinal, (Benth.) Choisy food Ternstroemia lineata 4, 6 mb’áza Medicinal, DC. subsp. lineata commerce Ternstroemia lineata 4, 5 mb’a’za Medicinal, DC. commerce Valeriana barbareifolia 1 ts’ir’ostejna Medicinal M. Martens & Galeotti Valeriana clematitis 5, 6 ts’ir’ostejna Medicinal Kunth Lippia umbellata Cav. 4 zalberia Medicinal Verbena gracilis Desf. 2, 4, 7 kjestexu Fodder Verbena litoralis Kunth 2 kjestexu Fodder Verbena recta Kunth 2, 4 kjestexu Fodder Hybanthus attenuatus 6 Fodder (Humb. & Bonpl. ex Schult.) Schulze-Menz Viola humilis Kunth 6 Fodder Phoradendron 4, 5 búngu Fodder velutinum (DC.) Oliv. Agave atrovirens Karw. 3 Food, ex Salm-Dyck alcoholic beverage, commerce Agave salmiana Otto ex 3 Food, Salm-Dyck alcoholic beverage, commerce Furcraea parmentieri 3 Ornamental (Roezl) García-Mend. Commelina elliptica 4, 6 Fodder Kunth. Tinantia erecta (Jacq.) 1 Fodder Fenzl. Tradescantia 4 kjua’a Fodder commelinoides Schult. & Schult.f. Bulbostylis sp. 2, 7 nrreb’i Fodder Cyperus 7 nrreb’i Fodder, hermaphroditus (Jacq.) medicinal Standl.

Voucher Photo record 205 1,332 76 47 292 218 58 164 129 265 286,341

150,193 2 PhR

PhR

PhR 320 294,325 182,238

153 267

(continued)

Mazahua Ethnobotany: Traditional Ecological Knowledge, Management,. . . Plant family Juncaceae Orchidaceae Poaceae

691

Species Juncus effusus L. Malaxis macrostachya (Lex.) Kuntze Avena sativa L.

Habitat Mazahua name Use 5, 6 nrreb’i Fodder 4, 5 Ornamental

Voucher 31 188

3

avena

Briza minor L. Cynodon dactylon (L.) Pers. Festuca breviglumis Swallen Hordeum vulgare L.

1 2, 7

nrreb’i nrreb’i

Photo record 313 140

2, 7

nrreb’i

Food, fodder, commerce Fodder Fodder, medicinal Fodder

3

zebara

110

Muhlenbergia macroura (Humb., Bonpl. & Kunth) Hitchc. Sporobolus indicus (L.) R. Br. Trisetum virletii E. Fourn. Triticum aestivum L.

5, 7

nuxnrreb’i

Food, fodder, medicinal, commerce Fodder, medicinal

2, 7

nuxnrreb’i

Fodder

141

5, 7

nrreb’í

Fodder

16

3

nrreju

109

Zea mays L.

3

Food, fodder, commerce Food, fodder, commerce

350

55

Photo record

References Balick MJ, Mendelsohn R. Assessing the economic value of traditional medicines from tropical rain forests. Conserv Biol. 1992;6:128–30. Berlin B. Ethnobiological classification. Principles of categorization of plants and animals in traditional societies. Princeton Legacy Library: Princeton University Press; 1992. p. 354. Blancas J, Casas A, Rangel-Landa S, Moreno-Calles A, Torres I, Pérez-Negrón E, et al. Plant management in the Tehuacán-Cuicatlán Valley, México. J Econ Bot. 2010;64(4):287–302. Bray D, Merino-Pérez L, Negreros-Castillo P, Segura-Warnholtz G, Torres-Rojo J, Vester H. Mexico’s community-managed forests as a global model for sustainable landscapes. J Conserv Boil. 2003;17(3):672–7. Caballero J, Mapes C. Gathering and subsistence patterns among the P’urhepecha Indians of Mexico. J Ethnobiol. 1985;5:31–47. Caballero J, Casas A, Cortes L, Mapes C. Patrones en el conocimiento, uso y manejo de plantas en pueblos indígenas de México. Revista de Estudios Atacameños. 1998;16:181–95. Casas A, Viveros JL, Caballero J. Etnobotánica mixteca: Sociedad, cultura y recursos naturales en la Montaña de Guerrero. México: Consejo Nacional Para la Cultura y las Artes: Instituto Nacional Indigenista; 1994. Casas A, Camou A, Rangel-Landa S, Solís L, Torres I, Delgado-Lemus A, Moreno-Calles AI, Vallejo M, Guillén S, Blancas J, Parra F, Aguirre X, Farfán-Heredia B, Arellanes Y, PérezNegrón E. Manejo tradicional de biodiversidad y ecosistemas en Mesoamérica: El Valle de Tehuacán. J Inv Amb Cien Pol Públ. 2014;6(2):23–44.

692

B. Farfa´n-Heredia and A. Casas

Champo-Jiménez O, Valderrama-Landeros L, España-Boquera ML. Pérdida de cobertura forestal en la Reserva de la Biósfera Mariposa Monarca, Michoacán, México (2006–2010). Revista Chapingo Serie Ciencias Forestales y del Ambiente. 2012;18(2):143–57. Comisión Nacional de Áreas Naturales Protegidas (CONANP). Programa de manejo de la Reserva de la Biosfera Mariposa Monarca, México. Secretaría del Medio Ambiente y Recursos Naturales. 2001. Cornejo-Tenorio G, Casas A, Farfán-Heredia B, Villaseñor JL, Ibarra-Manríquez G. Flora y vegetación de las zonas núcleo de la Reserva de la Biosfera Mariposa Monarca, México. J Bol Soc Bot México. 2003;73:43–62. Farfán B, Casas A, Ibarra G. Plant resources in the Monarch Butterfly Biosphere Reserve, Mexico: Mazahua ethnobotany and Peasant Subsistence. J Econ Bot. 2007;61(2):173–91. Farfán-Heredia B, Casas A, Rangel-Landa S. Cultural, economic, and ecological factors influencing management of wild plants and mushrooms interchanged in Purépecha markets of Mexico. J Ethnobiol Ethnomed. 2018a;14(1):68. Farfán-Heredia B, Casas A, Moreno-Calles AI, García-Frapolli E, Castilleja A. Ethnoecology of the interchange of wild and weedy plants and mushrooms in Phurépecha markets of Mexico: economic motives of biotic resources management. J Ethnobiol Ethnomed. 2018b;14(1):5. Flachsenberg H, Galletti H. Forest management in Quintana Roo, Mexico. In: Timber tourists and temples: conservation and development in the Maya Forest of Belize Guatemala and Mexico. Washington, DC: Island Press; 1998. p. 47–60. Forero E. Población y cultura en la etnorregión mazahua ( jañtjo). México: Universidad Autónoma del Estado de México; 1997. Hopkins NA, Josserand JK. Estudios lingüísticos en lenguas otomangues. México: Instituto Nacional de Antropología e Historia; 1979. INEGI. Principales resultados de la Encuesta Intercensal. 2015. López-García J. Análisis de cambio de la cobertura forestal en la Reserva de la Biosfera Mariposa Monarca (2006–2007). México: Fondo para la Conservación de la Mariposa Monarca. WWF-FMCN; 2007. Ramírez M, Miranda R, Zubieta R. Vegetación y cubiertas del suelo 2006 (1: 75000). Serie cartográfica Monarca. 2008. Rangel-Landa S, Casas A, Rivera-Lozoya E, Torres-García I, Vallejo-Ramos M. Ixcatec ethnoecology: plant management and biocultural heritage in Oaxaca, Mexico. J Ethnobiol Ethnomed. 2016;12:30. Salas L. Causas directas del cambio en las cubiertas del suelo en la región Mariposa Monarca: un análisis multiescalar. México: Universidad Nacional Autónoma de México; 2013. Salick J, Mejía A, Anderson T. Non-timber forest products integrated with natural management, Río San Juan, Nicaragua. J Ecol Appl. 1995;5:878–95. Salick J, Bien A, Martin G, Apin L, Beaman R. Whence useful plants? A direct relationship between biodiversity and useful plants among the Dusun of Mt. Kinabalu. J Biodivers Conserv. 1999;8:797–818. Shanley P, Pierce AR, Laird SA, Guillén A. Explorando el mercado verde. Certificación y manejo de productos forestales no maderables. Uruguay: World Wildlife Fund/Nordan-Comunidad; 2002. Toledo VM, Caballero J, Argueta A, Rojas P, Aguirre E, Viccon J. El uso múltiple de la selva basado en el conocimiento tradicional. Biótica. 1978;3(2):85–101. Von Gadow K, Pukkala T, Tomé M, editors. Sustainable forest management. Springer Science & Business Media; 2012. p. 368. W. W. F. La tala ilegal y su impacto en la Reserva de la Biosfera Mariposa Monarca. México: WWF; 2004.

Risk Management of Availability of Plant and Fungi Resources Among the Pure´pecha in Michoaca´n, Central-Western Mexico Berenice Farfa´n-Heredia and Alejandro Casas

Abstract

Management is commonly directed to decrease risk and increase certainty in resources availability. Risk in access to resources may be influenced not only by ecological factors, but also by human pressures which may be higher with the cultural and economic values of resources. People manage plants and fungi according to their role in subsistence, amounts available, and their quality and demand for their direct consumption and commercialization. This study aimed to analyze the management strategies and intensity, in relation to their distribution, abundance, and demand in the Purépecha community of Cuanajo, Michoacán. We expected more complex and intense management practices in more valued but scarce species. Through 25 in-depth interviews to agriculturalists and gatherers of wild plants and fungi from Cuanajo, Michoacán, we documented management intensity and risk of availability. Then, through categorizing all variables documented and using a Principal Component Analysis (PCA) we estimated (using values of the first principal component) an index of management intensity (IMI) and an index of ecological risk (IER) for different use forms of the species of plants and mushrooms. We evaluated the relation between IMI and IER through regression analysis. We documented 50 species of plants and 23 species of mushrooms used, managed, and commercialized; nearly 71.2% of these species are obtained through simple gathering, and the rest involve silvicultural management in agroforestry systems and homegardens. A total of 42 species are B. Farfán-Heredia (*) Universidad Intercultural Indígena de Michoacán (UIIM), Pichátaro, Michoacán, Mexico A. Casas Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_9

693

694

B. Farfa´n-Heredia and A. Casas

commercialized in regional markets, and 31 are consumed in the community. Plants with higher management intensity are those receiving higher number of practices with higher complexity, managed in several types of forest and agroforestry systems. More complex practices are carried out in plants with higher risk, restricted spatial and temporal availability; these species are in homegardens. Regression analyses were significant among higher risk and management intensity in edible, medicinal, ceremonial, and ornamental plants. Our chapter shows that the main motives for managing wild plants and mushrooms are related to pressures associated to demand in markets and for direct consumption by households, a pressure that is more pronounced in resources that are scarce. Using scarce resources causes risk, and management is a response to such risk.

Introduction Throughout history, humans have continually faced uncertainties in the availability of resources through practicing diverse strategies (Casas et al. 1997, 2007, 2017; Blancas et al. 2010, 2013). For hunter-gatherers, knowing habitats, behavior, life cycle and annual seasons of events like migration, reproduction and phenological stages, and interactions with other organisms, among other facts of the resources used, was and still is crucial for designing such strategies. For agriculturalists and pastoralists, similarly, decision-making on dates for land tilling, seed sowing, harvesting and other agricultural labors, the routes, and seasons for optimizing grazing, base on knowing the seanality of phenomena in nature. The reproduction and birth seasons, the management of cycles of use, fallow and rotation of land and grasslands, and the multiple use of areas within territories are all aspects related to preventing and mitigating uncertainty or for ensuring availability of resources (Casas et al. 2016, 2017; Blancas et al. 2016). Uncertainty in resources availability increases with risks on individuals used, or the populations, communities, and ecosystems they belong to, which may be put in critical situations by natural factors and human actions. Frost, fires, extreme rain, drought, and temperatures, and unusual pests are all examples of natural phenomena that may affect the availability of resources and that people perceive in their interannual experiences. Similarly, overextraction of plants, mushrooms, or animals in forests and overexploitation and contamination of soils, water, and grasslands are all human-determined factors guided by social and economic processes that compromise the current and future availability of resources. Therefore, management of uncertainty is risk management. Agriculture, pastoralism, aquaculture and fisheries, silviculture, and domestication are all expressions of risk management which derived in continually new aspects of innovation and inventive of humans and their dynamic culture. The level of risk associated to human pressure on resources and ecosystems, in rural contexts, is strongly influenced by the cultural and economic values of the useful products, as well as by their spatial and temporal availability. An abundant resource with low cultural or economic value will have a low risk; for the contrary, a scarce resource with high cultural or economic value will be under high pressure and risk. And the spectrum of risk may be continuum as continuum is the spectrum of

Risk Management of Availability of Plant and Fungi Resources Among. . .

695

conditions of availability and the states of cultural and economic values. Management practices may decrease risk and increase certainty in resources availability, and these practices may be more effective as they are more careful and involve several techniques and strategies (higher complexity). Several authors have referred to management intensity as a function of the amount of energy invested in a practice, the degree of complexity of both strategies and tools used, and the amount of products obtained per area unit, among other indicators (Blancas et al. 2010, 2013; Casas et al. 2017). The more intense or complex the management is, this would be more efficient to achieve certainty in the availability of resources. Contrarily, the absence or insufficient management on plants or other biotic resources endangered may favor their loss (Rangel-Landa et al. 2017; Farfán-Heredia et al. 2018). People manage plants and mushrooms according to their role in household subsistence, their amounts available, their quality, the balance between their availability and demand for direct use, or their commercialization (González and Caballero 2007; González et al. 2008; Blancas et al. 2013; Rangel-Landa et al. 2016, 2017; Farfán-Heredia et al. 2018). Risk means threat or probability that a resource becomes lost or endangered to be lost. Such risk depends on the following diverse social and ecological factors, including the vulnerability that may vary spatially and temporarily depending on the nature of each species: biological aspects like length of life cycle, life form, type of reproduction and breeding system, vulnerability to disturbance, frost, or fire; ecological aspects like distribution, abundance, generalist or specialized interactions with other organisms for pollination, seed dispersal, tolerance or protection against herbivores, and resilience capacity are all aspects relevant to analyze intrinsic features associated to risk. But in addition, human, social, cultural, and economic factors are relevant. Among them, practices associated to intensity of extraction of resources, the scarce or null management practices, and social regulations for organizing use of resources are all determinants for driving the collapse or sustainability in using biotic resources and ecosystems. Risk may vary in magnitude depending on natural and cultural factors, and some risk indices that have been constructed to evaluate such variation of conditions, generally, include biological and ecological information of a species, its use and interchange values, aesthetic appreciation, management type, and information on the capacity of resources and/or ecosystems they form part of to recover after disturbance caused by humans (Blancas et al. 2013; Delgado et al. 2014; Torres et al. 2015; Farfán-Heredia et al. 2018a). Our previous studies in the regional traditional markets of the Purépecha region (Farfán-Heredia et al. 2018a, b) analyzed the relation between supply and demand of forest products in markets as factors motivating management. In those studies, we hypothesized that markets generate current processes of pressure on resources and their increasing demand push people to increase their extraction, and if the resources are scarce, they may become collapsed or maintained if people develop technologies to prevent it and even increase their availability. In other words, we propose that to ensure the availability of the most valuable resources in markets, people develop management techniques on them (Arellanes et al. 2013; Farfán-Heredia et al. 2018a, b). If the rhythm of technological innovation is according with that of the demand, the innovation would be successful, otherwise the resource would become locally

696

B. Farfa´n-Heredia and A. Casas

extinct. Such pressure could be buffered through interchange since those communities where a resource is absent or scarce may be obtained from other territories where it is abundant. Our previous studies (Farfán-Heredia et al. 2018a, b) identified that analyzing regional markets allows a better view about the pressure on resources than the local studies, which in turn makes possible identifying those resources in high risk of disappearing and those on which people have developed successful management techniques. However, in those studies we also recognized that some aspects cannot be visualized at regional level and should be analyzed at local level. For instance, the detailed information about scarcity or abundance of resources difficultly can be documented in markets since the resource managers are not always present in the markets and the intermediaries of products have a higher presence. Consequently, this study was conducted in the Purépecha community of Cuanajo, Michoacán, which is the main provider of wild plants and mushrooms in the markets of the Pátzcuaro Lake region in Michoacán. And therefore, the study of factors influencing the extraction of resources and their availability and views of the resource managers on the motivation to manage them are valuable sources of information complementing that obtained in markets for constructing a view on factors influencing resource management, particularly how markets may decrease pressure on resources but also increase it. This study aimed to analyze the management strategies and their intensity in relation to pressures associated to their interchange at local and regional levels. We hypothesized that people invest higher management effort on those species more valued and demanded in markets and that are under risk to remain available because of their scarcity or uncertainty in their availability.

Methods Study Area Cuanajo is a Purépecha community of the municipality of Pátzcuaro, in the state of Michoacán, central western Mexico. The last National census of population recorded 4850 people living in the community, adults and young people speakers of the Purépecha language (INEGI 2020). The communal property territory is approximately 9434 ha of common use. The village is settled at 2100 m, but the territory is surrounded by several volcanic cones from La Cantera (2860 m) to Los Puercos (3300 m). Climate is temperate subhumid (Fig. 1) (SPP 1979a); the vegetation includes pine, pine-oak, oak, and fir forests, combined with areas cleared for agricultural and livestock practices (Fig. 2). Soils are mainly humic andosol in the higher elevation managed with silvicultural practices and haplic feozem in lower elevations used for agriculture. Since pre-Columbian times, Cuanajo and other Purépecha communities based their subsistence on a diversified economy with cultivation of maize, beans, squashes, and peppers, together with gathering of multiple plant and animal resources and the multiple use of ecosystems, all of which allowed self-sufficiency (Caballero 1982; Nuño 1996). Historical documents show that people from Cuanajo worked carving wood and fabricating furniture, boxes, boats, and trunks since the

Risk Management of Availability of Plant and Fungi Resources Among. . .

697

Fig. 1 Location of the Purépecha community of Cuanajo, Michoacán (the red line indicates the polygon of the communal territory according to INEGI (2005) and RAN (2018))

early seventeenth century (Téllez 2017). West and Serra (2013) found records corroborating these activities by the end of the eighteenth century, and this activity continues until present.

698

B. Farfa´n-Heredia and A. Casas

Fig. 2 Map of vegetation and land use in Cuanajo, Michoacán (in red the polygon of the communal territory; source of information, INEGI 2005, 2010; RAN 2018).

In the nineteenth century, a railroad was constructed and tranis started to pass through Cuanajo, enhancing the fabrication and commercialization of wooden handicrafts, thus increasing an overuse of forests (Garibay and Álvarez-Icaza 1992), which continues until the present. Until 1940, the main activities were agriculture, trade of wood, fuelwood, carbon, wooden furniture and handcrafts, and gathering of plants and other nontimber forest products. However, agricultural practices have decreased progressively in the last decades mainly because they require inputs that lower the production rentability (Nuño 1996; Téllez 2017). Until 1960, most men were dedicated to agriculture in summer and fabricating furniture during winter, but from 1960 to 1972 fabrication of furniture became a mechanized industry offering employments the whole year, which enhanced agriculturalists to sell their land (Nuño 1996). Currently, 90% of people have as main economic activity fabricating furniture. Women work embroidering blouses and other textile handicrafts. Agriculture of maize, wheat, oat, fava and common beans, red potato, and amaranth is now less important than in the past, but it is still alive. In addition, the majority of households are dedicated to the interchange of edible, medicinal, and ornamental plants and edible mushrooms, which are propagated in their homegardens. These products are exhibited in stores together with furniture and textiles and are in addition traded in the regional markets (Farfán-Heredia et al. 2018). The economy of households is significantly complemented by monetary remittances from relatives working in the USA (Téllez 2017).

Risk Management of Availability of Plant and Fungi Resources Among. . .

699

Due to the Agrarian Reform, since the 1940s Cuanajo is surrounded by new villages that have increased pressure on forests though clandestine illegal wood extraction and forest degradation (Nuño 1996).

Ethnobotanical Studies We conducted 25 in-depth interviews to producers and gatherers of wild and semicultivated plants and mushrooms of Cuanajo (23 women and 2 men). Interviews were directed to document local ecological knowledge and management practices of plants and mushrooms, recording use forms, parts used, gathering effort, temporal and spatial availability, perception of abundance, management techniques and systems, criteria of selection, and methods of propagation and maintenance, among other issues. We systematized information on management following the typologies developed by Casas et al. (1999, 2017), Blancas et al. (2010, 2013), and RangelLanda et al. (2016, 2017), ordering practices from lower to higher complexity and intensity as follows: (1) simple or planned gathering, (2) let standing or tolerance, (3) enhancement or promotion of their abundance, (4) special practices of protection and care, (5) transplanting of whole individuals, and (6) propagation through vegetative and sexual structures. In each practice mentioned, we recorded the occurrence or not of selection and the features preferred and favored. We collected and herborized botanical and fungal material, and the specimens were deposited in the herbaria EBUM and IEB; in addition, we made a collection of photographs of fresh fruits, mushrooms, Opuntia cladodes, and orchids. Species reported are named according to APG III (el www.theplantlist.org), while those of mushrooms according to Index Fungorum (www.indexfungorum.org).

Data Analyses Based on the responses of the interviews, we constructed two matrices, one with variables related to management practices, systems and purposes, and mechanisms of selection by people, which was used to determine an index of management intensity (IMI); the other matrix was constructed based on information on used part, life form, distribution, abundance, temporal availability, management practices, and propagation techniques, to calculate an index of ecological risk (IER). The variables were categorized assigning values following a gradient from lower to higher management intensity and risk based on ecological indicators (Tables 1 and 2). Values recorded for each species were averaged (Table 3). These indices were calculated based on principal component analyses (PCA) per use form for plant species that are commercialized in the markets. Mushrooms were not analyzed in this way since no differences were documented in management forms among species. Based on PCA, we classified the species with higher to lower management intensity and risk, we considered as IMI and IER the values of the first principal component, which summarizes the information on the highest variation of all factors analyzed, and we

B. Farfa´n-Heredia and A. Casas

700

Table 1 Matrix of variables considered to estimate the index of management intensity Variables Management practices

Description Type of management practices used to obtain resources and/or increase their availability

Management system

Types of systems where species are managed, in a gradient form lower to higher management intensity

Human selection

Practice or not of human selection

Categories Simple or planned gathering Tolerance or let standing Enhancement of their abundance Protection Transplanting of complete individuals Vegetative and/or sexual propagation Silvicultural Agricultural Homegardens No selection Selective gathering Selective tolerance Selective propagation

Value 1 2 3 4 5

6 1 2 3 0 1 2 3

identified the variables with higher weight on them. Species with the highest values are under higher management intensity and risk (Blancas et al. 2013; Delgado et al. 2014; Rangel-Landa et al. 2016; Farfán-Heredia et al. 2018b). We finally conducted regression analyses between IMI and IER through JMP 11 (SAS 2013).

Results Local Ecological Knowledge of Wild Plants and Mushrooms Based on interviews, we identified 50 species of plants and 23 of mushrooms that people said to know and use or remember that were used. We documented cultural, biological, ecological, and management aspects and factors influencing their commercialization (Appendix 1). Table 4 summarizes information on the use types recorded. For 60% of plant species recorded, people use the entire plant, for 24% they use fruits, for 8% flowers, for 6% leaves, and for 2% stems; the collected part from mushrooms is fructiferous body. Most wild plants are available from June to August, the rainy season, when also all mushroom species are available, but some resources are available the year round (Fig. 3).

Risk Management of Availability of Plant and Fungi Resources Among. . .

701

Table 2 Matrix of variables of ecological risk (from lower to higher impact) Variable Used part

Life form

Distribution

Description Risk is proportional to the impact of harvesting parts affecting survival, sprouting, and reproduction of plants managed. We consider that, in general, a higher impact is involved when collecting complete individuals of long-lived perennials than those of annual plants. The extraction of vegetative parts causing death of individuals involves higher risk than collecting vegetative parts decreasing fitness of herbs or shrubs; in addition, higher risk than collective sexual reproductive parts Impact is generally higher collecting long-lived perennial individuals than annual individuals

A species with wide distribution has less risk than a species with restricted distribution. Species occurring in human managed areas have less risk than others absent. Homegardens are areas with more intensive care than other agricultural areas. Ruderal areas are common but without care

Abundance

The abundance of plants perceived by people interviewed was recorded

Temporal availability

Period of availability of the used part. In general, collecting of reproductive parts during restricted seasons involves higher risk than those available longer time

Management practices

Type of practices used to obtain a resource and/or increase its availability. Intensive extraction without management or planned strategies are more risky that others favoring recovering of parts or populations affected by extraction

Category Sexual reproductive parts Vegetative parts Whole individuals

Herbs Shrubs Epiphytes Trees Homegarden Agricultural areas Ruderal areas Wide distribution in forests Restricted distribution in forests Abundant Regular abundant Scarce The whole year Six months Three months or less Simple or planned gathering Tolerance or let standing Enhancing abundance of a species Protection Transplanting of entire individuals Propagation

Value 1

2 3

1 2 3 4 1 2 3 4

5

1 2 3 1 2 3 1

2 3

4 5

6

(continued)

B. Farfa´n-Heredia and A. Casas

702 Table 2 (continued) Variable Propagation

Description Propagation of sexual propagules involving selection involves the most complex practices

Category Vegetative propagation Seeds propagation

Value 1 2

Most wild plant and mushroom species occur in forest ecosystems, some of them reaching ruderal and weedy environments. Some species are exclusive of forests, agricultural field, or homegardens (Appendix 1, Table 5). Based on the in-depth interviews, 27.4% of the species of plants and mushrooms recorded are collected through field journeys exclusive for this purpose, while 67.1% are collected “opportunistically,” when people carry out other activities. Species that are markedly scarce are only occasionally collected.

Perception on Distribution, Abundance, Extraction, and Motivations for Management Distribution. According to people interviewed, plants are distributed only where their seeds are, while mushrooms occur only in specific sites; they say that “. . .plants and mushrooms in Cuanajo are here because there are seed in the field, seeds that were in the basket of the Virgin of the Nativity.” Based on this view, this Virgin is celebrated on the 2nd of February, when people carry a big offer of different types of seeds for agriculture, including maize, squash, beans, amaranth, and other domesticates, in addition, seeds of pines and other plants from forests. Local people recognize that some plants distributed in specific sites, for instance, the tarepenï (Clinopodium macrostemum) and the toronjil (Agastache mexicana), which are plants of restricted distribution. For the toronjil, local people recognize two types, one with purple flower and the other with white flower, which do not occur together in the field; the purple-flowered is in the highlands of the hills, while the white-flowered occurs in the lowlands. People say that vegetation, soil, and type of rocks of the hill Cerro de la Cantera are different to other hills of the area. They believe, as their ancestors told them, that the hill “came from warm land,” which is related to the fact that this is the exclusive area of distribution of some species like Quercus deserticola. Abundance. People said that the preferred mushrooms are not abundant; they affirm that “. . .a thing that is delicious is scarce while what is not very tasty is abundant and abundant for longer time.” They consider that some years ago several species of plants and mushrooms that now are scarce were more abundant and have decreased their numbers in populations because of the removal of forest for planting avocado trees. They say that abundance of plants and mushrooms has annual cycles; some years, some species may be abundant while in other years others are abundant, alternating their availability.

Clinopodium macrostemum (Moc. and Sessé ex Benth.) Kuntze

Plants Crataegus mexicana Moc. and Sessé ex DC Dysphania ambrosioides (L.) Mosyakin and Clemants Opuntia atropes Rose Prunus serotina subsp. capuli (Cav. ex Spreng.) McVaugh Rubus liebmannii Focke Tagetes micrantha Cav. Acalypha phleoides Cav. Agastache mexicana (Kunth) Lint and Epling Artemisia ludoviciana Nutt. Chenopodium graveolens Lag and Rodr. Chenopodium sp.

Species

Istafiate

Artlu

Edible Edible Medicinal Medicinal/ edible Medicinal

Clima

Epazote de perro Tarepenï

Medicinal/ edible

Medicinal

Zarza Anís Kanserini Toronjil

Ruli Tami Acaph Agame

Edible Edible

Chen

Paré Xenhua

Opat Pruse

Edible

Medicinal

Epazote

Dyam

Edible

Use form

Chenpo Epazote de zorrillo

Karasï

Common name

Crame

ID

6

7

7

1

3 1 6 12

9 21

12

3

Management practices

4

3

3

1

3 3 4 4

6 6

3

3

Management system

0

0

0

0

0 1 0 0

0 1

0

0

2

2

2

2

3 2 2 2

1 3

2

3

Human Used selection part

Table 3 Parameters used to estimate the indices of management intensity and ecological risk

2

1

1

1

2 1 1 2

6 4

1

4

6

4

4

4

9 9 6 6

10 10

1

9

3

3

3

2

2 1 3 3

2 1

3

1

3

1

1

3

3 3 3 2

2 3

2

2

Life Temporal form Distribution Abundance availability

(continued)

1

2

2

0

0 0 0 3

1 3

3

0

Propagation

Risk Management of Availability of Plant and Fungi Resources Among. . . 703

Espinosilla

Manrubio Mirto

Tila

Aparhecua, ortiga Flor de terciopelo Dalia

Orquídea

Gnap Hetin

Loeme

Marvu Salmi

Terli

Urch

Daco

Goca

Castilleja scorzonerifolia Kunth Dahlia coccinea Cav.

Govenia capitata Lindl.

Laelia speciosa (Kunth) Laesp Schltr. Tigridia pavonia (L.f.) Tipa DC.

Casc

Flor de Corpus Pañuelo

Gordolobo Árnica

Erca

Eryngium carlinae F. Delaroche Gnaphalium sp. Heterotheca inuloides Cass. Loeselia mexicana (Lam.) Brand Marrubium vulgare L. Salvia microphylla Kunth Ternstroemia lineata DC Urtica chamaedryoides Pursh Begonia gracilis Kunth

Begr

Cola de caballo Cuanasï

Equi

Equisetum sp.

Common name

ID

Species

Table 3 (continued)

Ceremonial/ ornamental Ceremonial/ ornamental Ceremonial/ ornamental Ceremonial/ ornamental Ceremonial/ ornamental Ceremonial/ ornamental

Medicinal

Medicinal

Medicinal Medicinal

Medicinal

Medicinal Medicinal

Medicinal

Medicinal

Use form

4 6

12

1

3

3

3

5

1

3 3

1

3 6

3

1

Management system

6

1

2

1

2

17

1

12 12

1

1 7

1

1

Management practices

0

0

0

0

0

0

0

0

0 0

0

0 0

0

0

2

5

2

2

3

2

2

3

2 2

2

2 2

2

2

Human Used selection part

1

3

1

1

1

1

1

4

1 2

1

1 1

1

1

10

10

4

5

9

5

6

9

1 4

7

9 10

9

4

2

3

1

2

2

2

3

3

3 2

1

1 1

2

3

3

3

3

3

3

3

3

3

3 1

3

2 3

3

1

Life Temporal form Distribution Abundance availability

Propagation

3

1

0

0

0

0

3

0

3 1

0

0 2

0

0

704 B. Farfa´n-Heredia and A. Casas

Patita de pájaro Patita de pájaro Patita de pájaro Patita de pájaro

Rabo

Rafli

Rafla

Rafe

Guachitas

Lyde

Lyco

Lasq

Laccaria squarrosa Bandala, Montoya and A. Ramos Lyophyllum connatum (Schumach.) Singer Lyophyllum decastes (Fr.) Singer Ramaria botrytis (Pers.) Ricken Ramaria fenica (P. Karst.) Ricken Ramaria flava (Schaeff.) Quél. Ramaria flavigelatinosa Marr and D.E. Stuntz

Trompa café Kiripe (hongo de jícara) Kiripe (hongo de jícara) Kiripe (hongo de jícara) Guachitas

Hysp

Lacla

Trompa

Hylac

Laccaria laccata (Scop.) Cooke

Pansa

Boae

Lacam

Tecomate

Amca

Laccaria amethystina Cooke

Mushrooms Amanita caesarea (Scop.) Pers. Boletus aestivalis (Paulet) Fr. Hypomyces lactifluorum (Schwein.) Tul. and C. Tul. Hypomyces sp.

Edible

Edible

Edible

Edible

Edible

Edible

Edible

1

1

1

1

1

1

1

1

1

Edible

Edible

1

1

1

1

Edible

Edible

Edible

Edible

1

1

1

1

1

1

1

1

1

1

1

1

1

0

0

0

0

0

0

0

0

0

0

0

0

0

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

5

5

5

5

5

5

5

5

5

5

5

5

5

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

3

0

0

0

0

0

0

0

0

0

0

0

0

0

Risk Management of Availability of Plant and Fungi Resources Among. . . 705

B. Farfa´n-Heredia and A. Casas

706 Table 4 Use forms of wild plants and mushrooms in Cuanajo

Use form Edible Medicinal Ceremonial/ornamental Edible/medicinal Edible live fences Food for birds Utilitarian Utilitarian/medicinal

Number of species 35 20 10 4 1 1 1 1

Scarcity. Local people sadly express their perception that many plant and mushroom species with edible, medicinal, and other uses are not available or are scarce. According to them, the current scarcity of plants and mushrooms is caused by their excessive gathering, as well as the extraction of forest ground, which is commercialized into cities, and in which are seeds of plants and mushrooms, also the removal of forest and its substitution for avocado plantations. In addition, they mention another important cause the extraction of water through well drilling, which has determined the scarcity of water stored in hills, particularly important for them the hill called Cantera which is the main source of water for the village. They consider that these activities cause dryness of forest and plants wither. Some plants that in the past were available the year round, like toronjil and tarepenï, are now available only during the rainy season. According to people, if they recovered the ancient Virgin of the Natividad, plenty of seeds of all plants and mushrooms in her basket, the species lost and scarce would come back. Management. According to the local customs, in Cuanajo all forest resources are for common use; all people have the right to take them. People interviewed said that during times of needs plants and mushrooms are there in forests to provide goods to people, preventing worries since they may collect and sell them in the market, thus satisfying their needs, at least for some months of the year. From plants like tarepenï (Clinopodium macrostemum) and toronjil (Agastache mexicana), people use extracting young plants and transplanting them to their homegardens to have them available at home and to prevent their scarcity or unavailability. They consider that these plants are progressively scarcer because of their intense collection for commercialization and direct use by households. However, people say that transplanted plants lose their medicinal properties and should be collected in their natural sites; therefore, they prefer using wild plants and take care of them in the wild. They think that plants acquire properties to heat or cold the body according to their distribution sites. For instance, they consider that tarepenï may be “cold” or “warm” depending on where it is collected. People interviewed said that in some areas plants cannot be collected where rattle snakes are abundant; they say “. . .¡no, they do not allow us collecting!, you should be brave and have a great heart and a machete to defend you against the snakes,” which in some way contribute to regulate and conserve some plant populations.

Risk Management of Availability of Plant and Fungi Resources Among. . .

Fig. 3 Temporal availability of wild plant and mushroom species in Cuanajo, Michoacán

707

708

B. Farfa´n-Heredia and A. Casas

Table 5 Distribution of wild plants and mushrooms in Cuanajo Distribution area Forest Forest, agricultural, ruderal, and homegardens Forest, agricultural, and ruderal Agricultural areas Homegardens

Percentage of species (%) 46.6 19.2 17.8 12.3 4.1

People said that many plants are not propagated since they can be found throughout the year in the forest; they know where these plants occur and the season when they or their useful parts are available. Some others, whose availability is restricted, can be dehydrated and stored and do not need to be planted. In addition, they say that mushrooms cannot be planted since they do not know their seeds and they therefore do not know the way to propagate them. They have the perception that tejocotes (hawthorn) do not produce good fruit since people have disregarded these trees; in the past, they pruned their branches, removed some competing plants, and collected good fruits to consume and sell them, and trees were more productive. Now, because these fruits are not commercialized as in the past, and because people dedicated to agricultures are fewer, trees are not managed, they got pests, and it has not been possible to use them in the last years. Consumption. Knowledge on edible, medicinal, ceremonial, and ritual plants is acquired from parents and great parents; people say they use and consume plants and mushrooms their parents and grandparents teach them that “serve,” as well as those that “do not serve,” to be used. People recommend to be aware of the availability season of each plant and mushroom species, because for many species the season is short. For mushrooms, the rainy season is the time for consuming them. They like mushrooms and consider them substitute of meat; they said that in the past chicken, cow, or pig meat was consumed occasionally; rather, they waited for the mushroom season. Perceptions on mushrooms and plants. Local people consider mushrooms as delicious food, not available all the time, attractive for their forms and colors, and suddenly emerging from the ground. They said it is exciting to find and collect them. They think that when the tejocote trees are sad they do not produce neither flowers nor fruit. They consider that the availability of plants and mushrooms when there are economic necessities gives peace of mind, and they may be happy because they directly consume obtained products or interchange them for other products. They also consider that flowers not only embellish the field, their homes, and homegardens, but also are offerings. Flowers are considered the way to communicate with saints and God.

Management of Wild Plants and Mushrooms Nearly 71.2% of mushroom and plant species are obtained through simple gathering, but on the rest people carry out practices of selective and nonselective silvicultural

Risk Management of Availability of Plant and Fungi Resources Among. . .

709

Fig. 4 Management practices of wild plants and mushrooms

management (Appendix 1, Fig. 4); 17 plant species that are obtained through simple gathering are also managed through silvicultural practices like tolerance, protection, promotion, transplanting, and propagation (through sexual or asexual propagules). All these plant species occur in areas of forest management. On capulín trees (P. serotina) and herbs like anise (Tagetes micrantha), local people practice selective management, since they recognize varieties differently appreciated. For instance, in capulín they recognize the purple large, medium, and small fruit varieties; in addition, they recognize the red, large, and medium and the white, large, and medium fruits. In the case of anise, they recognize large and small varieties; they prefer the large anise for direct consumption, but they carry the small variety to the regional markets since people prefer it. People propagate in homegardens a great variety of ornamental plants, which are commercialized in the shops of furniture and textiles in Cuanajo, or in the markets of Pátzcuaro, Morelia, Tzintzuntzan, Quiroga, or Tacámbaro. The commercialization of edible, ornamental, and medicinal plants is an activity carried out by most women of Cuanajo, within the town or in the regional markets. Young plants of tarepenï and toronjil are extracted from forests and transplanted into pots or in the ground in homegardens or patios. Toronjil is also propagated through seeds; both species are commercialized in pots as live plants or in bunches for their use. Begonia gracilis and Dahlia coccinea are wild ornamental plants commercialized as live plants, but according to people, these plants are not planted by people but emerge spontaneously from the ground collected in the forest; for this reason, these plants are considered tolerated. Other plants are propagated through their seeds into homegardens. These are the cases of Chenopodium graveolens, Dysphania ambrosioides, Heterotheca inuloides, Marrubium vulgare, Opuntia

710

B. Farfa´n-Heredia and A. Casas

atropes, Salvia microphylla, Urtica chamaedryoides, and Tigridia pavonia and are commercialized as live plants in pots or in bunches for their use. In agricultural systems, people manage 24% of the recorded plants, while all mushrooms are managed in forest systems, and 22% of plants in both agricultural and forest systems and homegardens; nearly 6% are exclusive of homegardens (Figs. 5, 6, and 7). From the plant and mushroom species recorded, 42 are commercialized in the regional markets, while 31 species are only directly consumed by households in the community. Plants under higher management intensity are those receiving the more and more complex management practices and are managed in several management systems, mainly homegardens. Such management intensity is influenced by its recognition as resource, its preference, interchange value, demand in markets, and spaciotemporal availability. Management intensity is explained by the different types of practices and management systems (Table 6). According to the index of management intensity, the capulín trees (P. serotina) are the plants more intensely managed, followed by the ortiga (Urtica chamaedryoides), pañuelo (Tigridia pavonia), and toronjil (Agastache mexicana) (Table 7).

Fig. 5 Forest management systems (a) pine forest, (b) Boletus aestivalis, (c) Erigeron galleoti, (d) Clinnopodium macrostemun, (e) oak-pine forest, (f) Laccaria squarrosa, (g) Ramaria botritys, and (h) Agastache mexicana. (Photos: Berenice Farfán-Heredia)

Risk Management of Availability of Plant and Fungi Resources Among. . .

711

Fig. 6 Milpa-fruit trees management system: (a) milpa with capulín, (b) Agave inaequidens and Crataegus mexicana, (c) Prunus serotina subsp. Capuli, (d) Jaltomata procumbens, (e) Brassica rapa, (f) milpa-frutal, (g) Physalis chenopodifolia, (h) Pyrus communis, and (i) pig raising system. (Photos: Berenice Farfán-Heredia)

Capulín trees (Prunus serotina), prickly pears (Opuntia atropes), and epazote (Dysphania ambrosoides) are the edible plants more intensely managed, receiving the more complex practices and found in several management systems (green circles in Fig. 5a). Tejocote trees (Crataegus mexicana), zarzamoras (Rubus liebmannii), and anise (Tagetes micrantha) are under few management practices and in the least intensely systems of management (yellow circles in Fig. 5a). Medicinal plants like ortiga (Urtica chamaedryoides), toronjil (Agastache mexicana), manrubio (Marrubium vulgaris), mirto (Salvia microphylla), árnica (Heterotheca inuloides), epazote de perro (Chenopodium graveolens), epazote de zorrillo (Chenopodium sp.), hierba del cáncer (Acalypha phleoides), and tarepenï (Clinopodium macrostemun) are considered under high management intensity since they receive several practices and are found in the more intensive management systems (green circle in Fig. 5b). The hierba del sapo (Eryngium carlinae) and gordolobo (Gnaphalium spp.) are considered in the category of low-management intensity since they are obtained through simple gathering and are found in two managed systems (yellow circles in Fig. 5b). The istafiate (Artemisia ludoviciana), tila (Ternstroemia lineata), espinosilla (Loezelia mexicana), and cola de caballo (Equisetum sp.) are considered in the category of low-intensity management since they are collected through simple gathering and are found only in forest systems (blue circles in Fig. 5b).

712

B. Farfa´n-Heredia and A. Casas

Fig. 7 Homegarden or patio: (a) ornamental flowers, (b) Laelia speciosa, (c) Acalypha phleoides, (d) wasp diaper, (e) Yucca sp., (f) chile, (g) A. mexicana, (h) hortaliza, (i) Dysphania sp., (j) Urtica chamaedryoides, (k) medicinal plants, (l) D. ambrosioides, and (m) ornamental flowers. (Photos: Berenice Farfán-Heredia)

Risk Management of Availability of Plant and Fungi Resources Among. . .

713

Table 6 Contribution of management variables explaining variation of wild plants according to their use form (PC components 1 and 2) Variable Management practice Management system Human selection Percentage of variation

Edible PC1 0.989 0.145 0.019 97.6

PC2 0.146 0.988 0.034 2

Medicinal PC1 PC2 0.981 0.191 0.191 0.981 0.000 0.000 95.6 4.4

Ceremonial-ornamental PC1 PC2 0.944 0.329 0.329 0.944 0.000 0.000 98 2

Table 7 Index of Management Intensity (IMI) and Index of Ecological Risk (IER) of the most economically important wild plants in Cuanajo, Michoacán Species Edible Prunus serotina subsp. capuli (Cav. ex Spreng.) McVaugh Dysphania ambrosioides (L.) Mosyakin and Clemants Opuntia atropes Rose Crataegus mexicana Moc. and Sessé ex DC Rubus liebmannii Focke Tagetes micrantha Cav. Medicinal Urtica chamaedryoides Pursh Agastache mexicana (Kunth) Lint and Epling Marrubium vulgare L. Salvia microphylla Kunth Heterotheca inuloides Cass. Chenopodium graveolens Lag and Rodr. Chenopodium sp. Clinopodium macrostemum (Moc. and Sessé ex Benth.) Kuntze Acalypha phleoides Cav. Eryngium carlinae F. Delaroche Gnaphalium sp. Equisetum sp. Artemisia ludoviciana Nutt. Loeselia mexicana (Lam.) Brand Ternstroemia lineata DC Ceremonial-ornamental Tigridia pavonia (L.f.) DC. Laelia speciosa (Kunth) Schltr. Begonia gracilis Kunth Dahlia coccinea Cav. Castilleja scorzonerifolia Kunth Govenia capitata Lindl.

IMI

IER

12.998 3.640 1.109 5.263 5.263 7.221

12.799 4.471 0.853 5.272 5.402 7.450

11.048 5.949 5.757 5.757 1.425 0.850 0.850 0.060 0.060 5.038 5.038 5.421 5.421 5.421 5.421

10.729 5.979 7.084 5.981 0.006 1.501 1.501 0.150 0.336 5.827 5.871 4.607 4.714 5.447 5.816

8.432 2.108 1.998 1.998 2.943 3.601

8.591 3.384 2.912 2.912 1.869 4.282

B. Farfa´n-Heredia and A. Casas

714

The ceremonial and ornamental plants Laelia speciosa and Tigridia pavonia are obtained through simple gathering but in addition receive other intensive management practices in homegardens; therefore, they have intermediate values of management intensity (green circles in Fig. 5c). Oher species have low-management intensity, receiving one single practice and one or two management systems, among them homegardens (yellow circle in Fig. 5c). Finally, one species with very low-management intensity obtained through simple gathering from forests was identified (blue circle in Fig. 5c).

Ecological Risk of Wild Plants and Mushrooms Plants with higher ecological risk are those with restricted spatial and temporal availability and receive more complex management practices, mainly in homegardens. The ecological risk of resources is also influenced by their preference, interchange value, and demand in markets. The species with higher ecological risk are P. serotina, U. chamaedryoides, T. pavonia, A. mexicana, M. vulgare, S. microphylla, D. ambrosioides, L. speciosa, O. atropes, and H. inuloides. The variables with higher influence on ecological risk are the inappropriate management practices and narrow distribution (Tables 7 and 8). The regression analysis between ecological risk and management intensity of edible, medicinal, ceremonial, and ornamental plants indicates significant values (Fig. 6), with higher ecological risk influencing higher management intensity, involving several practices in different management systems (Fig. 6).

Distribution, Regulations, and Access to Wild Plants and Mushrooms Nearly 70% of the wild plant species and all mushroom species recorded occur in forest areas; about 32.9% of the wild plants are in ruderal areas, and 24.7% are in homegardens (Appendix 1). Table 8 Contribution of variables of ecological risk explaining variation of management of wild plants per use form (PC, principal components 1 and 2) Variable Use part Life form Distribution Perceived abundance Temporal availability Propagation Management practices Percentage of variation

Edible PC1 0.004 0.072 0.074 0.0147 0.0001 0.177 0.978 77.4

PC2 0.021 0.380 0.895 0.158 0.044 0.141 0.068 18.6

Medicinal PC1 PC2 0.013 0.025 0.013 0.088 0.222 0.944 0.065 0.140 0.021 0.164 0.199 0.033 0.951 0.228 79.6 14.5

Ceremonial-ornamental PC1 PC2 0.066 0.476 0.054 0.241 0.456 0.704 0.061 0.184 0.000 0.000 0.235 0.131 0.851 0.408 81.2 15.4

Risk Management of Availability of Plant and Fungi Resources Among. . .

715

According to the customs of the community of Cuanajo, wild plants and mushrooms occurring in forest areas are of open access since they are in lands of common tenure regime. There are no special rules or agreements to gather or conserve these resources, which in the opinion of people interviewed has favored the decreasing availability of the species more intensely extracted or those for which the form of extraction is destructive. These are, for instance, the cases of tarepenï (C. macrostemun), toronjil (A. mexicana), epazote (D. ambrosoides), flor de corpus (Laelia speciosa), manrubio (M. vulgare), tila (T. lineata), and pashacuas (Lyophyllum spp.) and the mushrooms trompas de puerco (H. lactifluorum), tiripiti (A. caesarea sensu lato), and patitias de pájaro (Ramaria spp.). Some owners of communal and private land restrict the extraction of ground from forests; they confiscate the oak land that people extract to return it to the places where it was extracted. Other owners condition people who collect mushrooms to maintain the ground removed in its place since they consider the mushrooms will be maintained proceeding in this way, while others restrict or forbid the destructive harvest of tila flowers and fruits (T. lineata), since some people cut the entire trees to make the collection of these parts easier. Another important problem is the selling of communal land, which is legal since the early 1990s. This fact has increased, and the previous communal land is now private and the access to forest resources is closed; some ancient paths have been even canceled because of the fences established around avocado plantations. Some avocado plantations are in addition guarded by armed people. Plants of cultural and economic importance, especially tarepenï (C. macrostemum), toronjil (A. mexicana), flor de corpus (L. speciosa), zarzamora (R. liebmannii), tila (T. lineata), and all species of mushrooms, are now of difficult access, and people have to invest long journeys and time to find them. Local people interviewed said that now they have to get mushrooms in the market of Pátzcuaro, since local collectors prefer to bring the progressively scarcer mushrooms to the markets of Pátzcuaro, Quiroga, and Tacámbaro, where they get higher incomes.

Discussion Several authors have mentioned that management includes diverse practices or forms of interrelationships between humans and components of biodiversity, directed to transform, conserve, recover, or adapt ecosystems, their elements, and processes (Casas et al. 1999, 2001, 2007, 2008, 2014, 2016; González and Caballero 2007, González et al. 2008; Blancas et al. 2010, 2013; Torres et al. 2015; RangelLanda et al. 2017). All these practices are modulated by human needs and purposes as function of economic, cultural, and ecological aspects. Management practices are carried out with different intensity and complexity depending on the role ecosystems and resources play in people life, and these authors have proposed tipologies of forms and degrees of intensity of management determined by the interaction of

716

B. Farfa´n-Heredia and A. Casas

ecological, social, cultural, economic, and technological aspects (Casas et al. 1999, 2001, 2007, 2008, 2014, 2016; González and Caballero 2007, González et al. 2008; Blancas et al. 2010, 2013; Torres et al. 2015; Rangel-Landa et al. 2017). The wide variety of management strategies is an expression of the vast and deep traditional ecological knowledge developed by human groups in interaction with ecosystems, their resources, and processes (Toledo 1994, 2002; Berkes 1993; Berkes and Folke 1994; Berkes et al. 1995; Caballero et al. 1998). Our study describes the diversity of management practices of plants and mushrooms in Cuanajo and contributes to document patterns and to the construction of theoretical frameworks initiated by the authors referred to above. We hypothesized that intensity and complexity of management practices of wild plants and mushrooms are directly related to their spaciotemporal availability, cultural value, and role in people’s life and their demand in markets. In turn, in absence of management, cultural, ecological, and economic aspects are factors determining risks in the availability of resources. The results of our study show that the main motivation to manage wild plants and mushrooms is to ensure their availability, particularly those with higher value and demand in markets for which people practice different strategies directed to diminish the risk to lose them.

Management and Management Intensity The variety of management strategies used with wild plants and mushrooms directed to ensure their availability, is part of a general strategy of multiple use of ecosystems and resources documented previously in the P’uhrépecha region (Rendón 1947; Caballero 1982; Caballero and Mapes 1985; West and Serra 2013; Rodríguez 2012, 2016; Santos 2013, 2014, 2017). These are practices coexisting with agricultural management of crops, and forest and livestock practices, involving multiple and diversified strategies (Rendón 1947; Caballero 1982; Caballero and Mapes 1985; Blancas et al. 2010; West and Serra 2013; Rodríguez 2012, 2016; Santos 2013, 2014, 2017). The results of this study coincide with those of other researches that documented that simple gathering is the most common practice to obtain forest products; in the case studied, nearly 71.2% of the species was recorded. But for the 28.8% of the species studied, people carry out management practices like tolerance, promotion, protection, transplanting, and propagation mainly directed to ensure or increase their availability (Caballero 1982; Caballero and Mapes 1985; Casas et al. 1999, 2016; González and Caballero 2007; González et al. 2008; Blancas et al. 2010, 2013; Arellanes et al. 2013; Santos 2014; Rangel-Landa et al. 2016, 2017; Rodríguez 2016). The species with high cultural and economic value are under higher pressure due to their extraction, and we expected that those scarcer would be special targets of management. Several authors have documented similar patterns in other regions of Mexico (Blancas et al. 2010, 2013; Delgado et al. 2014; Blancas et al. 2010, 2013; Arellanes et al. 2013; Casas et al. 2017; Farfán-Heredia et al. 2018b).

Risk Management of Availability of Plant and Fungi Resources Among. . .

717

Rules of Access to Resources as a Factor Motivating Management The common use of ecosystems and resources prevailing in the past in the Púhrépecha region guaranteed open access to them by households and communities (Nuño 1996). At regional level, this land tenure regime favors the acquisition and complementarity of resources to satisfy needs through interchange (either by commercialization or barter) in the traditional markets of the region (Vera 2013; Arellanes and Ayala 2016; Farfán-Heredia et al. 2018a, b). In Cuanajo, however, this study documented how the privatization of land has had consequences in the rights of common use of resources and the complementarity achieved in the past from the different ecosystems conforming their territory. At local level, this change affects the complementarity among different ecosystems and diminishes the capacity of the community to interchange products with other communities of the region. This fact reduces the availability of resources and motivates management practices on some resources, or their abandonment as part of their life forms (Farfán-Heredia et al. 2018a).

Conclusion The main motivations to manage wild resources with higher cultural and economic importance ensure or increase their availability through strategies directed to decrease their inaccessibility and risk of local disappearance or extirpation. Selection directed to increase availability of favorable phenotypes has a similar principle but results in changes of variation in populations and incipient domestication. Practices like tolerance, promotion, protection, transplanting, and propagation are directed to ensure or increase the resources availability and are specially operating on the most important but scarce species and phenotypes. Acknowledgments The authors thank the financial support from the Consejo Nacional de Ciencia y Tecnología (CONACYT, research project A1-S-14306), Mexico, the Dirección General de Asuntos del Personal Académico (DGAPA-PAPIIT, research project IN206520 and IN224023), UNAM, and the Comisión Nacional para el Conocimiento y Uso de la Biodiversiodad (CONABIO/ GEF/FAO, GEF project ID 9380 770, research project RG023), Mexico.

References Arellanes Y, Ayala D. Tradición y sobrevivencia del trueque como alternativa de abasto y subsistencia: una mirada al tianguis de cambio de Pátzcuaro, Michoacán. Rev Etnobiol. 2016;14(2): 56–65. Arellanes Y, Casas A, Arellanes A, Vega E, Blancas J, Vallejo M, et al. Influence of traditional markets on plant management in the Tehuacán Valley. J Ethnobiol Ethnomed. 2013;9:38. Berkes F. In: Inglis JT, editor. Traditional ecological knowledge in perspective. Ottawa: International Development Research Centre; 1993. Berkes F, Folke C. Linking social and ecological systems for resilience and sustainability. Stockholm: Beijer International Institute of Ecological Economics, The Royal Swedish Academy of Sciences; 1994.

718

B. Farfa´n-Heredia and A. Casas

Berkes F, Folke C, Gadgil M. Traditional ecologycal knowledge, biodiversity, resilience and sustainability. Biodivers Conserv. 1995;63:281–99. Blancas J, Casas A, Rangel-Landa S, Moreno-Calles A, Torres I, Pérez-Negrón E, et al. Plant management in the Tehuacán-Cuicatlán Valley, Mexico. Econ Bot. 2010;64:287–302. Blancas J, Casas A, Pérez-Salicrup D, Caballero J, Vega E. Ecological and socio-cultural factors influencing plant management in Náhuatl communities of the Tehuacán Valley, Mexico. J Ethnobiol Ethnomed. 2013;9:39. Blancas J, Casas A, Moreno CA, Caballero J. Cultural motives of plant management and domestication. In: Ethnobotany of Mexico interactions of people and plants in Mesoamerica. New York: Springer; 2016. p. 233–55. Caballero J. Notas sobre el uso de los recursos naturales entre los antiguos purépecha. Biótic. 1982;7:31–42. Caballero J, Mapes C. Gathering and subsistence patterns among the P’urhepecha Indians of Mexico. J Ethnobiol. 1985;5:31–47. Caballero J, Casas A, Cortés L, Mapes C. Patrones en el conocimiento, uso y manejo de plantas en pueblos indígenas de México. Estud Atacameños. 1998;16:181–96. Casas A, Caballero J, Mapes J, Zárate S. Manejo de la vegetación, domesticación de plantas y origen de la agricultura en Mesoamérica. Bol Soc Bot Méx. 1997;61:31–47. Casas A, Caballero J, Valiente-Banuet A. Use, management and domestication of columnar cacti in south-central Mexico: a historical perspective. J Ethnobiol. 1999;19:71–95. Casas A, Valiente-Banuet A, Viveros JL, Caballero J, Cortés L, Dávila P, et al. Plant resources of the Tehuacán-Cuicatlán Valley, Mexico. Econ Bot. 2001;55:129–66. Casas A, Otero-Arnaiz A, Pérez-Negrón E, Valiente-Banuet A. In situ management and domestication of plants in Mesoamerica. Ann Bot. 2007;100:1101–15. Casas A, Rangel LS, Torres I, Pérez-Negrón E, Solís L, Parra F, et al. In situ management and conservation of plant resources in the Tehuacan-Cuicatlan Valley, Mexico: an ethnobotanical and ecological perspective. In: de Albuquerque UP, Alves M, editors. Current topics in ethnobotany. Thiruvananthapuram: Research Signpost; 2008. p. 1–23. Casas A, Camou-Guerrero A, Otero-Arnaiz A, Rangel-Landa S, Cruse-Sanders J, Solís L, et al. Manejo tradicional de biodiversidad y ecosistemas en Mesoamérica: el Valle de Tehuacán. Investig Ambient Cienc Polít Pública. 2014;6:23–44. Casas A, Lira R, Torres I, Delgado A, Moreno-Calles A, Rangel-Landa S, et al. Ethnobotany for sustainable ecosystem management: a regional perspective in the Tehuacán Valley. In: Ethnobotany of Mexico. New York: Springer; 2016. p. 179–206. Casas A, Parra RF, Rangel LS, Blancas J, Vallejo M, Moreno CA, Guillén S, et al. Manejo y domesticación de plantas en Mesoamérica. Una estrategia de investigación y estado del conocimiento sobre los recursos genéticos en México. In: Casas A, Torres GJ, Parra F, editors. Domesticación en el Continente Americano. Morelia: UNAM–UNALM; 2017. p. 537–69. Delgado LA, Torres I, Blancas J, Casas A. Vulnerability and risk management of Agave species in the Tehuacán Valley, México. J Ethnobiol Ethnomed. 2014;10(1):53. Farfán HB, Casas A, Moreno CA, García-Frapolli E, Castilleja A. Ethnoecology of the interchange of wild and weedy plants and mushrooms in Phurépecha markets of Mexico: economic motives of biotic resources management. J Ethnobiol Ethnomed. 2018;14(1):5. Farfán-Heredia B, Casas A, Rangel-Landa S. Cultural, economic, and ecological factors influencing management of wild plants and mushrooms interchanged in Purépecha markets of Mexico. J Ethnobiol Ethnomed. 2018a;14(1):68. Farfán-Heredia B, Casas A, Moreno-Calles AI, García-Frapolli E, Castilleja A. Ethnoecology of the interchange of wild and weedy plants and mushrooms in Phurépecha markets of Mexico: economic motives of biotic resources management. J Ethnobiol Ethnomed. 2018b;14(1):5. Garibay C, Álvarez Icaza P. Producción agraria y forestal. In: Plan Pátzcuaro 2000. Ciudad de México: Friedrich Ebert Stistung; 1992. González IM, Caballero J. Managing plant resources: how intensive can it be? Hum Ecol. 2007;35(3):303–14.

Risk Management of Availability of Plant and Fungi Resources Among. . .

719

González IM, Martorell C, Caballero J. Factors that influence the intensity of non-agricultural management of plant resources. Agrofor Syst. 2008;74(1):1–15. INEGI. Conjunto de datos vectoriales de uso del suelo y vegetación, escala 1:250,000, Serie 3 (conjunto nacional); 2005. INEGI. México en cifras: Cuanajo Michoacán. 2010. www.inegi.org.mx. Acceso agosto 2018. INEGI. Panorama sociodemográfico de Michoacán de Ocampo. Censo de población y vivienda 2020. Instituto Nacional de Estadística y Geografía, México; 2021 Nuño GM. La relación naturaleza cultura en una comunidad purépecha a través de sus expresiones orales. In: El ropaje de la tierra. Naturaleza y cultura en cinco zonas rurales. Mexico City: UNAM-PyV; 1996. p. 29–82. Rangel LS, Casas A, Rivera LE, Torres GI, Vallejo RM. Ixcatec ethnoecology: plant management and biocultural heritage in Oaxaca, Mexico. J Ethnobiol Ethnomed. 2016;12(1):30. Rangel LS, Casas A, García FE, Lira R. Sociocultural and ecological factors influencing management of edible and non-edible plants: the case of Ixcatlán, Mexico. J Ethnobiol Ethnomed. 2017;13(1):59. Registro Agrario Nacional. 2018. https://www.gob.mx/ran#709 Rendón S. La alimentación tarasca. An Inst Nacl Antropol Hist. 1947;30:207–28. Rodríguez ML. Manejo traidiconal y potencialidades de aprovechamiento de Prunus serotina spp. Capuli en la comunidad de Pichátaro, Michoacán (Tesis de licenciatura). Pátzcuaro: Universidad Intercultural Indígena de Michoacán; 2012. Rodríguez ML. Conocimiento, disponibilidad y manejo de las plantas comestibles de recolección de San Francisco Pichátaro, Michoacán (Tesis Maestría). Texcoco: Universidad Autónoma de Chapingo; 2016. Santos RM. Etnobotánica, aspectos ecológicos y difusión de los quelites de la comunidad de San Juan Carapan: bases para su aprovechamiento sustentable (Tesis de Licenciatura). Pátzcuaro: Universidad Intercultural Indígena de Michoacán; 2013. Santos EM. Etnoecología, etnobotánica y aspectos ecológicos de plantas útiles de la comunidad p’urhepecha San Juan Carapan, Michoacán (Tesis de Licenciatura). Pátzcuaro: Universidad Intercultural Indígena de Michoacán; 2014. Santos RM. Descripción, manejo e importancia de los sistemas agroforestales de la comunidad p’urhepecha de San Juan Carapan, Michoacán, México (Tesis Maestría). Texcoco: Universidad Autónoma de Chapingo; 2017. SAS Institute. JMP 11 scripting guide. Cary: SAS Institute; 2013. Secretaría de Programación y Presupuesto. Carta edafológica. Pátzcuaro: Secretaría de Programación y Presupuesto; 1979a. Secretaría de Programación y Presupuesto. Carta edafológica. Villa Escalante: Secretaría de Programación y Presupuesto; 1979b. Téllez JE. Diálogo de saberes en torno a la kosmopercepción de los cuerpos de agua y la educación ambiental; wantanskuecha cuanaxo anhatapu (relatos orales de cuanajo) (Tesis de Maestría en Educación Ambiental). Pátzcuaro: Universidad Intercultural Indígena de Michoacán; 2017. Toledo V. Apropiación campesina de la naturaleza: un análisis etnoecológico (Tesis de doctorado). Mexico City: Universidad Nacioonal Autónoma de México; 1994. Toledo V. Etnoecology: a conceptual framework for the study of indigenous knowledge of nature. In: Steep JR, editor. Ethnobiology and cultural diversity. Bristol: International Society of Ethnobiology; 2002. p. 511–22. Torres I, Blancas J, León A, Casas A. TEK, local perceptions of risk, and diversity of management practices of Agave inaequidens in Michoacán, Mexico. J Ethnobiol Ethnomed. 2015;11(1):61. Vera GR. Trueque en la cuenca de Pátzcuaro: significaciones sociales de una práctica económica ambigua (Tesis de maestría). Zamora: El Colegio de Michoacán; 2013. West R, Serra L. Geografía cultural de la moderna área tarasca. Zamora: El Colegio de Michoacán; 2013.

Traditional Ecological Knowledge and Biodiversity Conservation in the Tierra Caliente Region of Michoaca´n Selene Rangel-Landa, María Elizabeth Saucedo-Gudin˜o, Erandi Lizbeth Guzma´n-Go´mez, Maria Fernanda Salazar-Ramirez, Arnulfo Blanco-García, Delia Caldera-Cano, Aglaen Lucero Carbajal-Navarro, Rosendo Caro-Go´mez, Andrea Ponce-Rangel, Jose´ Isabel Texta-Herna´ndez, and Xavier Madrigal-Sa´nchez

S. Rangel-Landa (*) Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico Consejo Nacional de Ciencia y Tecnología (CONACYT) – Escuela Nacional de Antropología e Historia (ENAH), Mexico City, Mexico MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico e-mail: [email protected] M. E. Saucedo-Gudiño · M. F. Salazar-Ramirez Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Campus Morelia, Morelia, Michoacán, Mexico Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Michoacán, Mexico e-mail: [email protected] E. L. Guzmán-Gómez Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Campus Morelia, Morelia, Mexico Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Mexico A. Blanco-García · D. Caldera-Cano · A. L. Carbajal-Navarro Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Mexico e-mail: [email protected] R. Caro-Gómez Independent Services Provider, Morelia, Mexico A. Ponce-Rangel · J. I. Texta-Hernández Guacamayas Calentanas A.C., Churumuco, Michoacán, Mexico © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_11

721

722

S. Rangel-Landa et al.

Abstract

Traditional ecological knowledge has been the product of human-nature interactions. Its potential for developing technologically viable alternatives toward sustainable management is at present recognized by different sectors involved in the study of biodiversity and the implementation of actions for its conservation. In such context, ethnobotany faces the challenge of documenting and understanding how and why people-plant interactions are, and how these experiences may be inserted in conservation strategies. The specialized literature provides insufficient knowledge about the state and risks affecting some regions like the Tierra Caliente of the state of Michoacán, where several priority conservation areas have been identified. One of these areas was decreed in 2007 as the Zicuirán-Infiernillo Biosphere Reserve (ZIBR). Our chapter reviews the state of ethnobotanical information documented for the region. We recovered 4766 ethnobotanical records for 910 plant species of the total 2634 species identified in the region. Most (83%) ethnobotanical records were found in herbarium specimens and a technical report, and nearly 27% from 29 dissertations, scientific papers, and books. Of these works, 22% were specialized ethnobotanical studies, and the rest were the product of research in different disciplines, including zootechnical studies (28%). Only 23% of the species with ethnobotanical records contained information additional to common names and use type. Despite the early stage of ethnobotanical research in the region, we identified that the botanical knowledge by the communities in the Zicuirán-Infiernillo area has been essential for the implementation of conservation and sustainable management projects. One of them was implemented in 2002 and 2013 by the project Conservation of biodiversity in Indigenous Communities of Oaxaca, Michoacán, and Guerrero (COINBIO for its acronym in Spanish), as well as the Zicuirán-Infiernillo Biosphere Reserve (ZIBR) since its creation in 2007. The communities have assumed the task of conserving biodiversity and valuing and maintaining traditional ecological knowledge. We highlight the potential role played by ethnobotanical research for contributing to conservation of biodiversity and traditional knowledge, and to local development actions.

Introduction The traditional ecological knowledge (TEK) about living beings and their environment have guided the way traditional societies interact with them to satisfy their material and spiritual. The interaction of people with nature has generated intricate links, which in many cases allowed the conservation of species and ecosystems, and the creation of agrobiodiversity and landscapes (Boege 2008; Toledo and BarreraBassols 2008). The TEK systems are made up of the accumulated knowledge, practices, and worldviews, which have been tested, adapted, and orally transmitted X. Madrigal-Sánchez Facultad de Biología, Universidad Michoacana de San Nicolás de Hidalgo, Morelia, Michoacán, Mexico

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

723

through generations (Berkes et al. 2000). These systems are based on individual and collective observations and practical experiences and operate modulated by beliefs and values that determine how the environment is perceived and how to interact with it (Berkes 1999). These characteristics of TEK systems confer them an adaptive character, which adds to being environmentally friendlier than the intensive management ways. Nowadays, their capability for incorporating information or coupling to other types of knowledge has led to the recognition that TEK is an essential foundation for designing viable sustainable management alternatives (Altieri and Toledo 2011; Berkes et al. 1995, 2000; Casas et al. 2016a; Gómez-Baggethun et al. 2013). A proof of these deeply rooted interactions and the benefits they have generated is provided by the fact that, in Mexico, over 60% of the forests are located within land owned by Indigenous and mestizo communities and ejidos, where people make decisions about their territories based on TEK (Boege 2008). TEK about plants in Mexico has enabled the use of more than 7000 species, of which between 800 and 1200 are managed through practices addressed to assure their continued availability. In addition, peoples from Mexico have domesticated more than 200 plant species that are important at the regional or national scale, among them 20 plant species of global agri-food importance (Caballero and Cortés 2001; Casas et al. 2016b). The interaction of people with plants in the region during the past 10,000 years led to the development of a center of origin of agriculture, which was practiced in a high diversity of traditional agroforestry systems (MacNeish 1992; Casas et al. 1997; Moreno-Calles et al. 2013). Most of the agricultural production that satisfies the local and regional food needs to be carried out in these agroforestry systems, which, at the same time, preserve a large fraction of the country’s biodiversity and agrobiodiversity. All these facts make agroforestry systems highly important reservoirs of genetic diversity (Casas et al. 2007; Parra et al. 2010; Cruse-Sanders et al. 2013; Moreno-Calles et al. 2016). Despite TEK and the management systems derived from it are essential and highly beneficial, their maintenance is threatened by the numerous socio-ecological problems that the communities have to face, and the rate at which these problems occur make the context very challenging. Among the main threats are (1) the abandonment of traditional practices, which are being replaced by intensive production methods, (2) the overexploitation of some forest resources in response to market demand, (3) changes in local lifestyles, influenced by massive media, migration, among other factors, (4) high rates of migration to cities of Mexico and the USA, (5) loss of local languages and knowledge about the natural environment, and (6) the negative effects of global change on management and production systems (Aswani et al. 2018; Gómez-Baggethun et al. 2013; Saynes-Vásquez et al. 2013), among others. In these scenarios, identifying and understanding the complexity of the human-plant interactions might provide a solid foundation for the search of alternatives allowing to overcome the above-mentioned problems (Camou-Guerrero et al. 2016). In this context, despite the efforts made to document the country’s biocultural diversity, ethnobotanical research in Mexico faces a difficult challenge due to the large information gaps existing both in unknown regions and in the limited covered subjects.

724

S. Rangel-Landa et al.

One of the regions harboring vast biological and cultural richness, but little known, is the Tierra Caliente region in the state of Michoacán. The limits of the region may vary according to the criteria used by different studies; but researchers agree that it is formed by the settlements and municipalities located within the Balsas River basin (Villegas-Santibañez 2007). The Balsas River basin is a strongly-accidented depression located between the Mexican Volcanic Belt and the Sierra Madre del Sur, having a 3800 m elevation range with an average of 1000 masl (Martínez et al. 2016). The region is characterized by warm-weather valleys in which the main localities are settled surrounded by mountainous zones – like Tepalcatepec, Apatzingán, Nueva Italia, La Huacana, Churumuco, Nocupétaro, and Huetamo (González 2001). It is common that the topographical and ecological diversity that characterizes the Tierra Caliente region is represented in many of the settlements, with tropical dry forests in the lower elevations and oak and pine forests in the higher zones (Fernández et al. 1998). The presence of numerous, but little studied archaeological remains accounts for the antiquity of the interactions of people with the region’s natural environments (Oliveros 2011). In different historical periods, the Purépecha, Náhuatl, Otomí, Matlatzinca, Afro-Mexican, and European peoples have inhabited in the region, together with migrants from other regions of Mexico, which has generated numerous cultural expressions and lifestyles (Delfín 2011; González 2001; Orozco and Odenthal 2016). The main productive activities in the region include irrigated agriculture in the lowlands – having economic importance since pre-Hispanic time, self-subsistence minimumtillage agriculture in the mountain slopes, and livestock grazing throughout all the altitudinal steps, which, since its introduction in the late sixteenth century, uses the natural vegetation as a source of fodder (Léonard 1995; Martínez et al. 2016; Thiebaut 2011). The intensification of these productive activities, and social issues – like high socioeconomic marginalization, migration, and insecurity – pose a challenge to biodiversity conservation (Ihl et al. 2017; Burgos et al. 2010; Villegas-Santibañez 2007). The region’s environmental complexity and rich cultural history lead to assume the presence in it of both a high biodiversity and an ample knowledge about the environment in its communities; however, both matters have been little studied. The zone attracted the interest of early explorers and naturalists like Francisco Hernández, who in the sixteenth century studied specimens from La Huacana and Apatzingán; Martín Sessé and José Mariano Mociño, who in the eighteenth century recorded the noticeable biodiversity present in the latter locality; Aimé Bonpland and Alexander von Humboldt, who during their emblematic expedition to the Jorullo volcano in the early nineteenth century reported the biological, cultural, and geographic diversity in the region; George Boole Hinton and James Hinton, who in the 1930s and early 1940s settled in Aguililla, from where they explored the region’s remote mountains; and William C. Leavenworth, who in 1941 travelled through Apatzingán and its surroundings as part of his study of vegetation in Cerro Tancítaro and the Tepalcatepec river (Álvarez 1965; Hinton and Rzedowski 1972; Leavenworth 1946; Sánchez and Chávez 2012; Urquijo 2010). Since the mid-twentieth century, the efforts to document the regional flora were scattered among some municipalities, as can be seen in the number of voucher specimens deposited in the EBUM herbarium of the Universidad Michoacana de San Nicolás de Hidalgo (UMSNH in its Spanish acronym) the projects from which they derived (Gómez 2012; Zamora 2012), and the studies of the

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

725

vegetation made during the last two decades (Casanova-Lugo et al. 2014; Cué-Bär et al. 2006; Martínez-Cruz et al. 2013; Méndez-Toribio et al. 2014; Peña-Estrada et al. 2016). In 2019, specialists in different plant groups made an important effort to systematize the available information about the flora of the Balsas Depression physiographic region – which includes the Tierra Caliente region – and documented 1519 species of flowering plants, 23 species of conifers and cycads, 102 species of ferns, and seven moss species, all of which agreed there still were large gaps in the knowledge of the regional biodiversity (CONABIO 2019a). An even larger information gap exists in the documentation of the human-nature interactions in the region. Among the published ethnobotanical studies in the region, the more relevant are those of the medicinal flora of the Balsas River Basin by José Carmen Soto and Mario Sousa (Soto 1987; Soto-Nuñez and Sousa 1995), who during their fieldwork – mostly conducted during the 1980s – collected specimens and interviewed people from 13 municipalities, nine of which were considered to be part of the Tierra Caliente region. Soto (2010) and other authors (Beltrán-Rodríguez et al. 2017), stated that this region, as others in Mexico where tropical dry forest is predominant, the ethnobotanical studies are scarce, involving isolated efforts, mostly reported in dissertations or technical reports. Despite the lack of complete information, the available records allowed the inclusion of the Infiernillo zone within the country’s Priority Terrestrial Regions for biodiversity conservation (Arriaga et al. 2000). Since the beginning of the twenty-first century, projects for promoting biodiversity conservation and sustainable management were emphasized by the Mexican government (Toledo et al. 2006), mainly through the implementation of the COINBIO program carried out between 2002 and 2013 (Caldera-Cano 2014; Toledo et al. 2006). In 2007, the government created the ZIBR, implemented several projects for conservation and sustainable development activities in the municipalities of La Huacana, Arteaga, Churumuco, and Tumbiscatio (CONANP 2014). While COINBIO recognized the essential role of communities in the maintenance and preservation of ecosystems, the objectives of the ZIBR emphasize the rescue and dissemination of TEK, practices, and technologies (CONANP 2014). This chapter reviews the literature, reports, databases, and herbarium specimens to analyze the present state of the ethnobotanical information documented in the Tierra Caliente region and, through case studies, to identify the role of TEK in communitarian projects, identifying the needs for further ethnobotanical and related research.

Methods Delimitation of the Tierra Caliente Region in Michoaca´n Following the general description by González (2001), we included in the Tierra Caliente region all the municipalities of Michoacán having 50–100% of its surface within the Balsas River hydrographic basin, and the municipalities in which the tropical vegetation types (tropical dry forest, tropical sub-deciduous forest, palm forest, and xerophilous scrub) had the same or larger coverage of the temperate vegetation (oak forest or coniferous forests). The surfaces of the vegetation types

726

S. Rangel-Landa et al.

were determined based on vectorial and elevation data (García and CONABIO 1998; INEGI 2013, 2016, 2019; INEGI, INE, CONAGUA 2007), using the geographic information system QGIS (2020). In the case of the municipalities of Aguililla, Arteaga, and Tumbiscatio whose limits exceed the Balsas River Basin, we only considered the part included within the limits of the basin to estimate the area of each vegetation type (71.45%, 57%, and 52%, respectively). Applying the latter criteria, we included 21 municipalities within the study region (Table 1).

Review of Literature, Reports, Databases, and Herbarium Specimens We reviewed the available scientific papers, books, and book chapters in English and Spanish using the academic search engines Google Scholar, Redalyc, ResearchGate, and AcademiaEdu, as well as dissertations from the UMSNH, the Universidad Nacional Autónoma de México (UNAM in its Spanish acronym), the Universidad Autónoma Chapingo (UACH in its Spanish acronym), and the Instituto Politécnico Nacional (IPN in its Spanish acronym). We searched for the keywords ethnobotany, useful flora, useful plants, traditional ecological knowledge (in both English and Spanish), Tierra Caliente, Zicuirán-Infiernillo, and the names of the 21 municipalities we included within the Tierra Caliente region using the above-mentioned criteria. The contents of the literature identified in the search results were consulted to make a record of each species having information about their common names, uses, and management practices. We requested the records of voucher specimens from the 21 chosen municipalities in the databases of the herbaria of the Faculty of Biology of the UMSNH (EBUM), the National Herbarium of Mexico (MEXU) at UNAM (MEXU-IBUNAM 2019) and the herbarium of the Bajío Region at the Institute of Ecology (IEB). Through this review we identified the records with information about common names, uses, and management form that were systematized as for the species identified in the literature search. In the case of the specimens recorded in the MEXU database, the digitized images of the vouchers were examined to supplement the information. We also consulted the Mexican Plant Ethnobotanical Database of the Botanic Garden of UNAM (BADEPLAM in its Spanish acronym) (Caballero et al. 2020). The scientific names of the plant species mentioned in the compiled records were updated according to the Taxonomic Authorities Catalogue of the National Biodiversity Information System (SNIB in its Spanish acronym) (CONABIO 2019b). The number of species having ethnobotanical information was compared with the number of species recorded in the review and those reported for the municipalities in the SNIB (CONABIO 2020).

Case Studies To identify the role of local people’s TEK about plants and ecosystems in the development of conservation and sustainable use projects in the region, we reviewed

Altitude (m a.s.l.) 232–2371

142–2016

47–1772

211–1797 350–1606 166–1800

290–1577 175–1721

673–2174 824–2508 115–2003

128–725 455–1793 138–1395 220–1400 582–2749

Municipality Aguilillaa

Apatzingán

Arteagaa

Buenavista Carácuaro Churumuco

Gabriel Zamora Huetamo

Juárez Jungapeo La Huacana

Múgica Nocupétaro Parácuaro San Lucas Susupuato

Climate (A)C(w2), Aw1, Awo, BS1(h’)w, C(w2) (A)C(w1), (A)C(w2), Aw1, Awo, BS1(h’)w (A)C(w1), Aw1, Awo, BS1(h’)w, BSo(h’)w (A)C(w1), Aw1, Awo, BS1(h’)w (A)C(w1), Aw1, Awo, BS1(h’)w (A)C(w1), Awo, BS1(h’)w, BSo (h’)w Aw1, Awo, BS1(h’)w (A)C(w1), Awo, BS1(h’)w, BSo (h’)w (A)C(w1), Aw1 (A)C(w1), Aw1, C(w1), C(w2) (A)C(w1), Aw1, Awo, BS1(h’)w, BSo(h’)w, C(w2) BS1(h’)w (A)C(w1), Aw1, Awo, BS1(h’)w (A)C(w1), Awo, BS1(h’)w Awo, BS1(h’)w (A)C(w1), Aw1, Awo, C(w1) 42.3 14.4 1.9

63.3

20.6 13.2 4.6

33.1 2.2

0.8

43.4

22.9

0.5 8.4 8.3 20.0 30.3

21.3 9.8 11.2

9.6 16.5

11.8 9.5 10.3

1.6

9.9

3.2 0.2 1.8 1.5 0.7

1.4 1.1 0.5

1.4 0.7

1.2 0.2 0.3

0.2

1.8

13.5

1.8 1.2

3.5 15.1

3.0

2.6

0.3

11.5

23.2

12.5

9.9 8.3 4.9

10.1 2.6

1.1 4.1 9.0

5.4

2.7

24.0 65.4 34.7 53.7 26.2

19.1 42.6 69.5

37.2 71.7

36.8 74.1 66.9

81.5

45.5

0.4 0.4

1.1

15.6 1.1 4.9

0.2

6.7

2.3

0.4

Other

(continued)

7.9 11.7 11.3 9.9 4.1

8.5 8.7 4.5

5.7 6.0

3.1 12.0 6.1

8.7

5.2

Vegetation types and land use (percentage of the surface) IrrAgr TmpAgr HuSet CF OF TDF Gras 7.5 6.8 0.8 28.5 6.9 29.2 20.3

Table 1 Climates and the surface by type of vegetation and land use of 21 municipalities of the Tierra Caliente of Michoacán

Traditional Ecological Knowledge and Biodiversity Conservation in. . . 727

355–2402

418–2705

Turicato

Tuzantla

(A)C(w1), (A)C(w2), Aw2, Awo, BS1(h’)w, C(w2) (A)C(w1), Aw1, Awo, BS1(h’)w, C(w2) (A)C(w1), (A)C(w2), Aw1, Awo, C(w2)

Climate (A)C(w2), Aw1, Awo, BS1(h’)w (A)C(w1), Aw1, Awo, C(w1)

2.8

4.6

0.3

9.6

16.2

4.3

0.3

0.3

0.2

9.3

0.8

39.0

17.5

14.4

6.7

52.9

51.4

38.8

7.7

12.2

10.8

Vegetation types and land use (percentage of the surface) IrrAgr TmpAgr HuSet CF OF TDF Gras 40.4 12.2 1.4 5.1 0.3 32.6 8.0 0.4 6.8 0.1 8.4 17.2 56.0 10.9 0.2

Other

Abbreviations: Climate abbreviations: Awo and Aw1 warm subhumid types, (A)C(w1) and (A)C(w2) semi-warm subhumid types, (A)C(m) humid semi-warm type, BSo(h’)w warm arid; warm semi-arid type, BS1(h’)w warm semi-arid type, C(w1) and C(w2) temperate subhumid types. Vegetation types and land use abbreviations: IrrAgr irrigated agriculture, TmpAgr temporary agriculture, HuSet human settlement, CF coniferous forest, OF oak forest, TDF tropical dry forest, Gras grassland, Other include water bodies, subtropical forest, xerophilous shrubland, Typha stands “tular,” palm stands, mesophyll forest of mountain and sites without vegetation cover. Source: Data from INEGI (2013, 2016) and García and CONABIO (1998) a Altitude, climate, and vegetation data correspond only to the portion of the municipality that is found within the Balsas basin

248–2340

Altitude (m a.s.l.) 239–2017 340–2158

Municipality Tepalcatepec Tiquicheo de Nicolás Romero Tumbiscatioa

Table 1 (continued)

728 S. Rangel-Landa et al.

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

729

and documented the experiences of the COINBIO project, the Ejido Huacana, and the non-governmental association Guacamayas Calentanas AC formed by inhabitants of El Chocolate, located within the zone of influence of the ZIBR. It is pertinent to say that the authors participate and/or collaborate with these projects and organizations.

The Tierra Caliente of Michoaca´n Environmental Description The Tierra Caliente of Michoacán is part of the Balsas River basin, mainly the Bajo Balsas province and a small portion of the Middle Balsas province (Fig. 1). The altitudinal range in the 21 municipalities identified as part of Tierra Caliente ranges from 47 to 2749 m (Table 1) (INEGI 2013). The highest lands of the study area belong to two mountainous areas that exist in west-east orientation: the TransMexican Volcanic Belt in the north and the Sierra Madre del Sur at the south, and between these mountainous areas exists a long and wide stripe of lowlands with several vegetation types and land uses (INEGI 2016).

Fig. 1 Location of the Tierra Caliente of Michoacán region, municipalities considered part of the Tierra Caliente region, and the Zicuirán-Infiernillo Biosphere Reserve (ZIBR)

730

S. Rangel-Landa et al.

Table 2 Plant biodiversity of the Tierra Caliente of Michoacán region Municipality Aguililla Apatzingán Arteaga Buenavista Carácuaro Churumuco Gabriel Zamora Huetamo Juárez Jungapeo La Huacana Múgica Nocupétaro Parácuaro San Lucas Susupuato Tepalcatepec Tiquicheo Tumbiscatío Turicato Tuzantla Total

Families 107 87 96 60 90 88 74 105 86 101 124 72 57 44 83 38 48 94 75 73 93 184

Genus 380 296 354 183 282 309 226 401 281 317 482 258 161 121 264 82 138 296 183 169 300 941

Species 635 471 626 241 409 511 311 717 446 448 912 422 203 146 380 103 176 447 242 223 456 2634

Relative richnessa 0.46 0.29 0.18 0.26 0.45 0.46 0.85 0.35 3.18 1.70 0.47 1.12 0.37 0.29 0.81 0.39 0.22 0.30 0.12 0.14 0.45 0.11

a

Expressed as a ratio between the number of species and the area of the municipalities and the region

Semiarid and arid warm climates predominate in the region, although temperate types are also present in the high areas of nine municipalities (Table 1) (García and CONABIO 1998). The tropical dry forest (TDF) is the most common type of vegetation in the study area and covers an average of 48% of the surface of the 21 municipalities, which indicates the existence of an important extension of TDF in the region. Actually, 10 municipalities still conserve between 50 and 80% of their surface as TDF (Arteaga, Carácuaro, Churumuco, Huetamo, La Huacana, Nocupétaro, San Lucas, Tiquicheo, Turicato, and Tuzantla). TDF is the most extended vegetation type along low-lying hills and valleys of the Mexican Pacific slope and in the Balsas River Basin. It is composed by low trees (6–12 m) adapted to seasonal variation in rainfall (8 months of drought and 4 months of rain in summer) and trees, shrubs, and herbs drop their leaves to survive in this context. The TDF from southwest of Mexico harbors high levels of endemic plant and vertebrate species (CONABIO 2019a). Although temperate plant communities are present in the middle and upper parts of the mountainous areas of 19 municipalities, these are not representative of this region, only the municipalities of Aguililla, Juárez, Susupuato, Tiquicheo, Tumbiscatio, and Tuzantla harbor between 20% and 45% of their surface with coniferous forests, oak forests, or a mixture of both, while the rest of the

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

731

municipalities have an average of only 7% with these temperate communities. Conversely, there are municipalities such as Tepalcatepec, Buenavista, Parácuaro, and Múgica that have between 40% and 60% of their surface with irrigated agriculture due to their location in the lowest and flat parts of the basin, while Huetamo, Juárez, Susupuato, Tepalcatepec, and Turicato have 10 to 30% of their surface as temporary agriculture. Most of the municipalities have induced pasture between 0% and 10% of their surface, but Aguililla stands out with 20% of this type of cover.

Plant Diversity From our review of the records by the SNIB (CONABIO 2020), the herbaria and the literature, we recorded a total of 2634 species for the region (Table 2). This plant diversity is distributed in 11 mosses taxa (Bryopsida), 5 species of Selaginella (Lycopodiopsida), 33 ferns species (Polypodiopsida), 3 cycads species (Cycadopsida), 18 conifers species (Pinopsida), and 2564 seed plants species (Angiospermae). A total of 115 species were exotic. The 2634 species recorded belong to 184 families and 941 genera (Table 2). We also recorded 22 taxa at the genus level, which did not coincide with any other taxon recorded at the level of species. The above values of richness by taxonomic group (with the exception of Angiospermae) are very similar to those reported for the Balsas physiographic region, which includes the 21 municipalities of the Tierra Caliente region, but having a larger extent (CONABIO 2019a). In the case of angiosperms, in the 21 municipalities of the Tierra Caliente region we found over 1000 more taxa than those reported for the Balsas Basin region in Michoacán (CONABIO 2019a). The values of richness by municipality also verify that voids remain to exist in the knowledge of biodiversity in the region (Table 2). The relative richness tends to be higher in municipalities with small surfaces and lower in the more extensive ones, even for Arteaga and Tumbiscatío, municipalities for which a higher number of species would be expected because of their environmental complexity. This pattern suggests that the recorded number of species per area unit reflects the sampling effort more than the actual diversity in the territories of the municipalities.

The Population of the Tierra Caliente Region in Michoaca´n The population in the 21 municipalities comprising the Tierra Caliente region of Michoacán (representing 40% of the state’s surface) was 570,639 inhabitants in 2015, or 12.45% of the state’s total (Table 3) (CONABIO 2012a; INPI 2017). Over half of that population (57%) live in rural communities (CONABIO 2012a). The predominant economic activity in the region is agriculture (including crop and animal husbandry), which is the livelihood of 50% of the economically active population in 9 of the 21 municipalities, 30–50% in 8, and less than 30% of that population in only 4 of these municipalities (Table 3) (CONABIO 2012b). Commerce and service provision are other important economic activities in terms of the fraction of the population that depends on them, which is higher than the proportion

34,243 40,818 14,387 21,548 45,484 8195 26,789 17,845 8804 23,842 13,731 6947 31,849 15,383 570,639

2050 140 264 1944 376 543 501 466 267 794 1486 2055 1539 1013 23,187

48 73 75 62 21 58 49 60 100 34 76 65 70 83

Rural population (%) 46 18 52 35 60 67 41 88 116 57 86 345 33 273 99 504 114 26 14 1043 10 5861

Indigenous population 168 1429 78 1282 2 18 76 High High Towering Medium High High High High Medium Medium Towering Medium High High

Migratory intensity High Medium Low Medium High High Medium High Medium Medium Medium High High High High Medium Towering High High High Low

Marginalization INPI Low Medium Medium High High Medium High 27 53 57 35 17 61 39 39 83 22 63 46 53 67

Agricultural activities 54 11 40 45 52 45 35 38 34 40 49 21 63 36 42 85 17 72 67 58 61

Households that use firewood (%) 30 13 37 19 61 65 30 55 (5) 9 (1) 19 68 (1) 5 12 (1) 23 28 12 (1) 35 18 13 22 27 514

Ejidos and communities 16 49 (1) 22 (1) 41 (1) 10 24 (1) 6

Note: Numbers between parentheses correspond to the number of indigenous communities. Source: Data from CONABIO (2012a, b), INPI (2017), Uribe et al. (2012), and RAN (2017a)

Municipality Aguililla Apatzingán Arteaga Buenavista Carácuaro Churumuco Gabriel Zamora Huetamo Juárez Jungapeo La Huacana Múgica Nocupétaro Parácuaro San Lucas Susupuato Tepalcatepec Tiquicheo Tumbiscatío Turicato Tuzantla Total

Population 2015 15,241 128,250 22,138 47,498 9485 15,455 22,707

Area (km 2) 1391 1633 3425 918 914 1105 365

Table 3 Socio-economic and territorial characteristics of the municipalities of the Tierra Caliente of Michoacán

732 S. Rangel-Landa et al.

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

733

of the population depending on farming only in the municipalities of Apatzingán, Gabriel Zamora, Huetamo, Múgica, Parácuaro, and Tepalcatepec (CONABIO 2012b). Another economic activity with historical importance in the region is mining, which has experienced a fast growth due to recent authorization of exploitation of mineral deposits in the region. Mining provides employment options for local people, but mine operation also implies negative socio-environmental issues (CONANP 2014). The intensity of people migration is high to very high in 13 of the 21 municipalities in the region and intermediate in the remaining municipalities, except in Arteaga where it is low (Table 3) (Uribe et al. 2012). Migration in this region has been associated with high marginalization that is high to very high in 12 of the 21 municipalities, and recently, to insecurity due to presence of organized crime groups (Table 3) (INPI 2017). The indigenous population inhabiting in the region receives little recognition – unlike what is recognized in the Purhépecha Plateau, the Coast, and Eastern regions of Michoacán – as it can be seen in the governmental programs that only address the Pirinda communities (which are related to the Nhanñu or Otomí people) in the municipality of Huetamo (SPI 2015). However, the 2015 census of population recorded a total of 5861 indigenous inhabitants in the region (Table 3), 1960 being speakers of 23 of the indigenous languages of Mexico, four of which are spoken by peoples that have historically been settled in Michoacán (INPI 2017). At present, the region goes through a process of reappraisal of ethnic roots, revitalization of tradition, and struggle for recognition of territorial rights, which is reflected in the official inclusion into the National Agrarian Registry of 13 indigenous communities located in the region (RAN 2017). Land tenure in the region is predominantly social, including 501 ejidos and 13 indigenous communities covering 52% of the regional municipalities (Table 3) (RAN 2017a, b). Social land tenure is determinant for access to natural resources by the population, especially of families belonging to ejidos and indigenous communities. Communitarian use areas are essential for small-scale animal husbandry, which depends on free foraging in them, and to satisfy basic needs like extraction of firewood for cooking (Table 3) (INPI 2017). The proportion in the region of households using firewood is 34% and dependence on this fuel is closely related to degree of marginalization, that is, the higher the margination index the more important to obtain low-cost fuels. Over 50% of the households in seven municipalities use biofuels, which indicates not only that firewood is a basic natural resource but also that ample knowledge exists in the region about firewoodproviding species and their management.

The State of Ethnobotanical Documentation in the Tierra Caliente of Michoaca´n Of the 20,789 records of plants obtained in our review, 4766 included ethnobotanical information, 91% corresponded to the 901 taxa identified at the species level, 6% to taxa identified at the genus level, and 3% were unidentified. Our records of plants

734

S. Rangel-Landa et al.

having ethnobotanical information were mostly obtained from labels in herbarium specimens (66.5%), 6.2% from technical reports of a COINBIO project, and 27% from 30 academic works including books, scientific papers, and dissertations (Table 4). Of the latter works, 6 were published between 1984 and 1995, 9 in the first decade of the present century, 14 in the last 10 years, and the botanical calendar of Miguel Tena published in the nineteenth century (1892), which was recently reprinted. These data portray the scarcity of scientific works with ethnobotanical information, but its rate of increment is 50%, like that of the presentation of ethnobotanical works in Mexican botanical congresses from 1980 to 2010 (Camou-Guerrero et al. 2016). The 13 dissertations we identified and consulted are the main source of scientific information about ethnobotany, for both their number and relative contribution (52%) to the total records reported in the literature (Table 4). The largest number of ethnobotanical records for the region were contributed by the dissertation of SotoNúñez (1987), which was later published as a book. Following in number, we recorded ten published scientific papers, eight book chapters, books, or congress abstracts (Table 4). Most publications documented ethnobotanical information for one municipality (Table 4), mainly Huetamo (nine), Carácuaro, Churumuco, La Huacana, and Tuzantla (seven), six in Apatzingán, one in Juárez and Parácuaro, and none in Susupuato. We identified nine disciplines from which works were made (Table 4). The main objective of 23% of the works was to document the local knowledge and humanplant interactions, the central goal of ethnobotany. The remaining publications used ethnobotanical information as a tool and were approached from disciplines different to ethnobotany, like zootechnics (27% of the publications), mainly directed to identify grazing plants known by farmers. Agronomy was the main approach in 20% of the publication, aimed at documenting local knowledge about crops, notably a dissertation documenting weedy plants in crop fields (Vargas 1989). We recorded one publication from Medical Anthropology, with relevant contribution to understanding the role of plants in traditional medical practices in the region (Pérez 2013). Nearly 10% of the publications had an ecological approach, while the rest has taxonomic, urban design, plant propagation, historical biogeographic, and genetic resources morphogenetic characterization scopes (Table 4). The publications we recorded contributed to three of the technical areas of ethnobotany proposed by Camou-Guerrero et al. (2016) as follows. The 60% of the works are mainly descriptive ethnobotany listing plants, names, and use. Some publications followed an ecological approach, but do not establish the kind of relationships between TEK or management and their ecological findings, reason why they are not considered within ecological ethnobotany (Table 4). The 37% of the publications have economic ethnobotanical approach, focusing on the identification of plants, their nutritional composition as fodder, and their potential for animal husbandry. Only one of the publications we recorded can be included in the cultural ethnobotany approach because it documents basic aspects of the role of plants in traditional medicine systems (Pérez 2013).

Ethb Agro Agro Ethb Ethb

Zoot Zoot Bot Zoot Zoot

Book BTh BTh BTh BTh

BTh Book

Art Art Art

Book

Book

García (2006) García and Linares (2010)

González et al. (2006) González et al. (2007) Guevara-Fefer and Rzedowski (1980) Gutiérrez-Vázquez et al. (2018) Gutiérrez-Vázquez et al. (2018)

Agro Bot, Udzn

PBA

Discipline Agro Zoot

Art

Publication type BTh Art

Bayuelo-Jiménez et al. (2006) Blanco-García et al. (2017) Diaz (1995) Figueroa (2005) Gallegos (2015) Galván (2005)

Reference Angón (1993) Ávila-Ramírez et al. (2007)

E

E

D E D

E D

D E E D D

D

Ethnobotanical approach D E

CN, CmC, Use CN, Use

Data type CN, O CN, Use, CmC CN, Use, Gc, M CN, Use, M CN, Use CN, Use, O CN, Use CN, Use, M, O CN, M, O CN, M, Use, O CN, Use CN, CmC CN

FW

R, FW

FW FW R

FW U

R, FW FW FW FW FW

U

Data collection FW FW

Table 4 Publications having ethnobotanical information records identified in the review of the literature

UMSNH

UMSNH

UMSNH UMSNH UMSNH

UACh IPN

UMSNH UMSNH UACh UACh UNAM

UMSNH

Institution UNAM UMSNH

15

56

80 64 1

1 65

62 2 9 43 46

1

Records 1 40

13

1

6 6 1

1 10

1 1 1 1 1

1

(continued)

Municipalities 1 1

Traditional Ecological Knowledge and Biodiversity Conservation in. . . 735

Discipline Zoot Zoot Ecol Ecol Agro Ethb Bot MedAnth Ecol Ethb Ethb HB Ethb Agro Zoot

Publication type CM

Art MTh Art BTh

BTh Art Book BTh BTh, Book Book Art BTh BTh Art

D D C D D D D D D E

E E D E

Ethnobotanical approach D

CN, Use CN, Use CN, Use CN, O CN, Use CN, Use, O CN, M CN, Use CN, D CN, Use

CN, CmC CN, Use CN, Use, O CN, Use, O

Data type CN, Use

FW R, FW FW FW FW FW R FW FW FW

FW FW FW FW

Data collection FW

UMSNH UMSNH ColMich UNAM UNAM UMSNH UAEM UNAM UMSNH UMSNH

UMSNH UNAM UNAM UACh

Institution UMSNH

45 25 47 44 284 12 18 8 150 7

67 15 14 1

Records 94

1 1 1 1 8 7 1 6 1 1

3 1 1 3

Municipalities 1

Abbreviations: Publication type: Art article, BTh bachelor thesis, CM congress memory, MTh master thesis. Discipline: Agro agronomy, Bot botany, Ecol ecology, Ethb ethnobotany, HB historical biography, MedAnth medical anthropology, PBA plant breeding, Udzn urban design, Zoot Zootechny. Ethnobotanical approach: C cultural, D descriptive, E economic. Data type: CmC chemical composition, CN common name, D disservice, Gc genetic characterization, M management, O others. Data collection: FW field work, R review, U undefined

Reference Gutiérrez-Vázquez et al. (2018) López et al. (2017) Luna (2011) Luna-Nieves et al. (2017) Martínez and de la Trinidad (2007) Molina (2010) Peña-Estrada et al. (2016) Pérez (2013) Rodríguez (2018) Soto (1987) Tena (1892) Thiébaut and Aguirre (2011) Torres (1984) Vargas (1989) Villa-Méndez et al. (2008)

Table 4 (continued)

736 S. Rangel-Landa et al.

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

737

All the publications we found included common names of plant species, 63% also state the kind of use, and only 10% provide information about management. Regarding the source of information, the authors of 87% of the publications stated they obtained data from fieldwork, 17% of the authors obtained information from literature reviews, and, in two works, the authors do not include their sources of information (Table 4). The first authors of the works we recorded were affiliated to six institutions, 53% to UMSNH, 23% to UNAM, and 13% UACH (Table 4). The 56% of the works made by authors with institutional affiliation to UMSNH were made in the Institute of Agricultural and Forestry Research (IIAF in its Spanish acronym) and the filiated Faculty of Veterinary Medicine and Zootechnics (FMVZ in its Spanish acronym), the latter focused on the identification of pasture plants. Works related to agronomy and identification of agrestal plants were published by authors mainly affiliated to UACH and the Faculty of Agrobiology of UMSNH. The authors of publications aimed to describe human-plant interactions were affiliated to Biology faculties or Botany departments from UMSNH, UNAM, and UACH. The nineteenth century work of Miguel Tena, was the only publication we recorded from a medical school.

The Useful Plant Species of the Tierra Caliente of Michoaca´n Region As mentioned above, 910 of the 4766 records of ethnobotanical information we systematized were identified to species – representing 35% of all plant species reported for the region – 296 were identified to genus, and 155 do not include a taxonomic identity. The ethnobotanical data more frequently provided by the authors of these records was common names (in 876 species, or 33% of the regional flora) and use (616, 23%). Management is mentioned in 13% of the species (115 species) recorded (Table 5). Authors provided information about potential damages caused by plants (agrestal in monoculture, toxicity, or undesired effects) in 10% of the recorded species, most of these reported in a dissertation focused on identifying agrestal plants in maize, rice, and sorghum monocultures (Vargas 1989). The 616 recorded species belong to 105 plant families, but only 10 of these families included 52% of them: Fabaceae (116 species), Asteraceae (49), Malvaceae (28), Apocynaceae (24), Burseraceae (21), Euphorbiaceae (18), Lamiaceae (18), Poaceae (18), Cactaceae (15), and Rubiaceae (13). These plant families are those reported in other ethnofloristic inventories for the country (Caballero and Cortés 2001) and for regions like the Tehuacan Valley (Lira et al. 2009). Bursera was the genus with the highest number of species uses (21% or 44% of the species in the genus known for the region) (Villaseñor 2016). Following in the number of uses mentioned were the genera Lonchocarpus, Quercus, and Senna, each represented by ten species. If we compare the percentage of species used relative to the total species in the region (23% in our case), other regions have a much larger proportion of used plants, as it is the case of the Tehuacan Valley where local people use 61.2% of the species (Lira et al. 2009). This noticeable discrepancy is due to the lower research effort

738

S. Rangel-Landa et al.

Table 5 Species by type of ethnobotanical information documented and percentage of useful species by municipality of the Tierra Caliente of Michoacán region Municipality Aguililla Apatzingán Arteaga Buenavista Carácuaro Churumuco Gabriel Zamora Huetamo Juárez Jungapeo La Huacana Múgica Nocupétaro Parácuaro San Lucas Susupuato Tepalcatepec Tiquicheo Tumbiscatío Turicato Tuzantla Total

Common name 112 154 71 106 287 264 47

Use 92 78 49 89 237 146 39

Management practices 1 7 0 20 0 3 4

Noxious 9 68 4 2 0 2 1

Useful species (%) 14 17 8 37 58 29 13

370 213 218 313 64 91 37 279 16 55 270 90 190 256 876

299 196 218 220 55 78 31 239 7 50 241 83 188 239 616

32 10 40 6 1 0 2 10 0 0 10 0 0 11 115

4 1 0 7 1 0 1 1 0 0 2 0 0 3 93

42 44 49 24 13 38 21 63 7 28 54 34 84 52 23.4

made in the region compared with that invested in the Tehuacan Valley. For instance, in some communities of Tierra Caliente of Michoacán like the ejido Ichamio, in the municipality of La Huacana we inventoried 202 plant species, 55% of them were reported used by local people (Rangel-Landa et al. 2003). The more frequent use categories we registered were medicine, animal feed, food, and obtention of wood used for buildings, making furniture, and as fuel (Table 6), similarly as in other regions of Mexico (Caballero and Cortés 2001; Lira et al. 2009). The plant species in the region with most reports of uses known by local people were Enterolobium cyclocarpum with 11 uses, and Crescentia alata, Guazuma ulmifolia, Sideroxylon capiri, and Swietenia humilis with 9 uses each. All these species are used as animal fodder, firewood, medicine, fuel, and material for building and making furniture (Table 7). The management of these multiple-use species is little documented, only for one of them did we find information in that regard, that despite that local people commonly keep these plants in their crop fields and pastures, where in many cases, they form part of traditional agroecosystems.

Traditional Ecological Knowledge and Biodiversity Conservation in. . . Table 6 Categories of use of the useful plant species of the Tierra Caliente of Michoacán region

Use Medicinal Fodder Edible Wood and construction Firewood Ornamental Shade Living fence Domestic tools Other Handicrafts manufacture Veterinary medicine Ceremonial Melliferous Fibers Sources of saponins Insecticide Ludic Animal control Aromatizing Soil control Additive Poison Tanning Cosmetic Facilitator of other plants Dyes Glue Stimulants Hunting bait Water indicator

739 Number of species 356 186 160 116 110 94 61 54 48 29 27 27 18 12 9 8 6 5 5 4 4 3 3 3 2 2 1 1 1 1 1

The COINBIO Project and Valuing of Traditional Ecological Knowledge for Conservation The official biodiversity conservation model applied in Mexico from the early twentieth century is based on the creation of protected natural areas (ANP in its Spanish acronym) by the government, a model that reached its peak in the period 1994–2000 with the establishment of the National Commission of Protected Natural Areas (CONANP in its Spanish acronym) (Velázquez et al. 2005). However, the Mexican government decreed most ANPs without taking into consideration the desire, participation, or interest of local inhabitants and landowners, and they operate with scarce human and material resources insufficient for the optimal accomplishment of their goal. This scenario created the need to generate biodiversity

740

S. Rangel-Landa et al.

Table 7 Species of the Tierra Caliente of Michoacán region with the highest number of reported uses Species Enterolobium cyclocarpum (Jacq.) Griseb. (Fabaceae)

Common name Parota

Crescentia alata Kunth (Bignoniaceae)

Cirián

Guazuma ulmifolia Lam. (Malvaceae)

Caulote, cablote

Sideroxylon capiri (A. DC.) Pittier (Sapotaceae)

Capire

Swietenia humilis Zucc. (Meliaceae)

Cóbano

Gliricidia sepium (Jacq.) Kunth ex Walp. (Fabaceae)

Cacahuananche

Leucaena leucocephala (Lam.) de Wit (Fabaceae)

Guaje Blanco

Lysiloma divaricatum (Jacq.) J.F. Macbr. (Fabaceae) Pithecellobium dulce (Roxb.) Benth. (Fabaceae)

Cuitaz

Pseudobombax ellipticum (Kunth) Dugand (Malvaceae) Spondias purpurea L. (Anacardiaceae)

Escobetillo

Vachellia campechiana (Mill.) Seigler & Ebinger (Fabaceae)

Espino, huisache

Pinzán

Ciruela

Use and management Edible, firewood, fodder, handicrafts manufacture, living fence, medicinal, poison, shade, sources of saponins, tanning, wood and construction Domestic tools, edible, firewood, fodder, handicrafts manufacture, medicinal, ornamental, veterinary medicine, wood and construction Domestic tools, edible, firewood, fodder, medicinal, melliferous, shade, veterinary medicine, wood and construction Additive, domestic tools, edible, firewood, fodder, living fence, medicinal, shade, wood and construction; Protection Domestic tools, edible, firewood, fodder, medicinal, ornamental, shade, wood and construction, other Firewood, fodder, living fence, medicinal, ornamental, poison, shade, wood and construction Edible, firewood, fodder, living fence, medicinal, soil control, veterinary medicine, wood and construction; Tolerance, cultivated Domestic tools, edible, firewood, fodder, handicrafts manufacture, medicinal, wood and construction, other Edible, firewood, fodder, living fence, medicinal, shade, veterinary medicine, wood and construction; cultivated Domestic tools, firewood, fodder, handicrafts manufacture, hunting bait, medicinal, sources of saponins, wood and construction Domestic tools, glue, edible, firewood, fodder, living fence, medicinal, wood and construction; cultivated Firewood, fodder, living fence, medicinal, ornamental, shade, wood and construction, other

conservation alternatives based on the direct and informed participation of landowners of areas concentrating mostly primary vegetation and high biodiversity, which in the case of Mexico corresponds to the territories of ejidos and indigenous communities (Boege 2008). One of these alternatives was the COINBIO, which operated in the states of Oaxaca, Guerrero, and Michoacán, in its first stage coordinated by a National Committee from 2002 to 2007, and in the second stage from 2008 to 2013, under the coordination by the government of the state of Michoacán (Sosa et al. 2019).

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

741

In the state of Michoacán, four regions were initially defined for its implementation: Costa-Sierra de Coalcomán, Oriente, Sierra Purépecha, and Infiernillo, which share a similar biological richness, but had contrasting ecological and social characteristics due to differences in the occurrence of indigenous communities and ecosystem types. The Infiernillo region – where TDF is dominant – included the municipalities of Arteaga, Churumuco, La Huacana, Tumbiscatio, Nuevo Urecho, and Ario de Rosales, where ejidos (whose members do not belong to an indigenous group) cover over 60% of the territory, except for Arteaga and Tumbiscatio where that percentage is smaller (RAN 2017a, b). At the begenning of the COINBIO project, the first questions were if people of the region practiced traditional uses and management favorably for conserving biodiversity and if the communitarian cohesion and institutions of the ejidos were as solid as in indigenous communities. These apparent social weaknesses gradually dissipated with the implementation of the project consolidating the ejido Assembly as the highest decision.making instance use, management, and conservation of the natural resources within the territory. The ejidos defined the activities of the COINBIO project based on rural participatory evaluations made by the men, women, and youngsters of each ejido, the results of which were approved by the ejido Assemblies. Through that process, projects addressing the knowledge of the territory and its planned use (flora and fauna inventories and territorial planning), development and strengthening of local capabilities (training, exchange of experiences, and workshops to reinforce organization processes), and operation of actions aimed at the conservation and sustainable use of the ejido resources (implementation of productive projects and transference of ecotechnologies). During the first stage of COINBIO, 105 projects were carried out in 28 ejidos in the municipality of La Huacana, 51 in 16 ejidos and one indigenous community in the municipality of Arteaga, 51 in 12 ejidos and one indigenous community in the municipality of Churumuco, and nine in six ejidos in the municipality of Tumbiscatio (Caldera-Cano 2014). During the second stage of COINBIO, only 23 projects were carried out in 16 ejidos and communities in the municipalities of Juárez, Jungapeo, Susupuato, and Tuzantla (Caldera-Cano 2014). According to Caldera-Cano’s (2014) evaluation of the COINBIO projects in Michoacán, the Infiernillo region contributed most of the generated social capital compared to the other regions. This result is essentially explained by three factors: (1) The communities’ responsibility for accomplishing the goals of each project, a transparent use of the resources used, and the supervision of the ejido Assemblies; (2) The accompaniment of interdisciplinary groups of strong commitment professionals that advise and provide technical assistance to communities going beyond the accomplishment of agreed projects by focusing on reinforcing communitarian skills and capabilities and enabling the dialogue between TEK and scientific knowledge; (3) Valuing the TEK of local populations, which thanks to the COINBIO projects began to become the pride and identity of local communities and an essential premise for the development and execution of projects supported by other institutions. Communitarian areas for biodiversity conservation were established by many ejidos, with which they showed their capacity for preservation of ecosystems based on processes of reinforcement of TEK systems and local organization structures (Toledo et al. 2006).

742

S. Rangel-Landa et al.

One main result of the projects during the first stage of COINBIO was the decree of the Zicuirán-Infiernillo Biosphere Reserve (ZIBR) in 2007, and in a second stage until 2013, these projects contributed to the establishment and operation of the biosphere reserve. The decree of the ZIBR involved the active participation of the inhabitants of the ejidos and communities of the area, a process in part driven by CONANP projects (CONANP 2014; Sosa et al. 2019). The ZIBR has tried to give continuity to productive projects through support provided mainly by the Program for Conservation and Sustainable Development (PROCODES in its Spanish acronym). These projects are in charge of organizing groups of local inhabitants, who actively play a central role and apply their knowledge about the territory through a dialogue of knowledge with participating external actors like scientists and civil organizations. The establishment of fodder banks and production of organic hibiscus flower (Hibiscus sabdariffa) in several locations are examples of projects followed by the ZIBR staff and civil organizations. In these projects local knowledge about the species’ ecological aspects and traditional cultivation techniques have been essential for developing local methods to implement productive projects. Another example of a product of the knowledge dialogue between local inhabitants and the biosphere reserve staff members is the restoration of eroded soils in the ejido Toluquilla, where stone walls were built to using the columnar cactus Stenocereus fricii to stabilize the soil and increase fruit production for the market. The members of this ejido have traditionally cultivated S. fricii in crop fields, pastures, and home gardens, accumulating knowledge about the biology of this plant with high cultural importance in the region. The combination of traditional knowledge and knowledge about geological and hydrological processes generated through a dialogue with ZIBR technical staff has been the key both for the immediate results and the permanence of the project. Villa-Reyes (2018) found that the factors with most weight in abandoning the project by local people were related to: insecurity due to the operation of criminal organizations in the region; and factors that commonly discourage productive initiatives, as economic profits not meeting the expectancies of participant beneficiaries. However, local people realized that it is essential take care of environmental issues, especially water availability and soil fertility, and have recognized the value of their own knowledge and the productive opportunities offered by local biodiversity. These local awareness processes have fostered the emergence of conservation initiatives from within the communities (even from those outside the protected natural area), and the accomplishment of goals and permanence of communitarian initiatives.

The Trees of the Ejido Huacana: A Collaborative Exercise The remarkable biodiversity and the adverse social context of the region makes crucial the involvement of the academic sector in generation and popularization of scientific knowledge about its biological richness to help resources owners and managers to properly socialize that information among themselves. The book Trees of the Ejido Huacana, Michoacán: an illustrated guide (Blanco-García et al. 2017) arose from the interaction between scholars and members of the Ejido

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

743

Huacana since 2013, willingly to receive groups of students from the Faculty of Biology of UMSNH. The purpose was to teach them about subjects like communitarian organization, biodiversity management, or alternative productive projects, always with coordination and support from the ZIBR staff. In 2014, some members of the ejido Huacana expressed their desire to know in more detail the work of the biologists and other scientists working in the area to understand what the practical value of their fieldwork is and have access to the information to design better ways of managing their natural resources. With that purpose, the ejido and the Restoration Ecology Research Laboratory proposed a collaboration initially aimed at making an illustrated guide to the wild trees in the ejido and their local uses taking advantage of the TEK of key actors. The commitment was made to coordinate the assistance of students from the Faculty of Biology to participate in fieldwork, collect and identify herbarium specimens, and create digital records for the plants, while members of the ejido and the ZIBR staff would find funding for printing the resulting book. The guide contains information about 62 wild tree species occurring in the Ejido Huacana, and although it was not intended to provide a formal taxonomic list, it does represent an exercise that contributes to disseminate a small part of the biological richness in the ejido. Biologists and members of the ejido collected specimens of the listed trees and photographs of the entire tree, leaves, bark, flowers, and fruits, in field trips conducted in 2015. The herbarium specimens were then identified. Each page of the guide contains information about one of the tree species listed by local and scientific names, the uses reported locally or elsewhere, and the tree’s local conservation status assessed as local abundance or its category of risk according to the Mexican norms (SEMARNAT 2010). The book also includes a short botanical glossary. The Huacana ejido Directive Board received copies of the printed book, which was distributed among its members and the local public schools, the ZIBR distributed some copies to other communities. The ample recognition of this uncommon exercise of collaboration between local landowners, scholars, and public institutions led members of the ejido to plan a second edition of the book including more of the trees in the ejido land to cover the demand for the book, both from within and beyond the region.

The Guacamayas Calentanas A.C. Self-Management Initiative for Biocultural Conservation People of the community El Chocolate, a settlement within the ZIBR have committed their community to actions mainly focused on management of vegetation to conserve the guacamaya verde (the military macaw Ara militaris). Motivated by this purpose, they began to innovate the communities’ productive activities through their experience and by documenting and transmitting regional TEK and formulating a project whose implementation would satisfy human needs in the community in synchrony with preservation of local natural resources. The presence of the Ara militaris was familiar to all the inhabitants of El Chocolate, but the initiative of some people motivated by the bird’s beauty and the awareness of its risk to extinction, led them to find out more about their behavior, feeding habits, and movements in the territory. They started asking to people of the

744

S. Rangel-Landa et al.

community what they know, then they observed the birds, recorded the observations in a notebook, and took photographs, which attracted the interest of more members of the community and generated a process of transmission and generation of knowledge. The growing interest in the community for Ara militaris created the need to ask for the advice from the ZIBR technical staff members to implement the actions to conserve the guacamaya verde, an initiative to which scholars from the Faculty of Biology of UMSNH adhered as collaborators. The participation of scholars in the community’s initiative led to creation of two environmental awareness groups in the locality, the Grupo Ambiental Guacamaya Verde integrated by adults and the Grupo Ambiental Infantil Guacamaya Verde integrated by children, and the civil association Guacamayas Calentanas (Fig. 2). One of the actions for conservation of Ara militaris was the establishment of a communitarian monitoring system based on the experience accumulated by people’s observations and training in ecological monitoring, use of photographic, and geolocalization equipment provided by ZIBR staff and researchers from the UMSNH and UNAM. All the participants recognized that generating an impact on the conservation of the guacamaya verde required joining efforts throughout the region, which led them to form a network of regional monitoring. The network included participants from other five localities, whose training was mainly provided by members from the community El Chocolate, thus generating a rich intercommunitarian exchange of knowledge (Fig. 2). Motivated by the work of adults, the children in El Chocolate started to participate in the community’s activities. That fact led the adults to compare the children’s and their own environmental knowledge and recognize that a large part of that knowledge was not being transmitted. That reflection generated a conscious exercise of knowledge exchange between children and adult groups. The enthusiasm of the children also enhanced them to participate in groups in environmental education events in schools and science popularization activities throughout the region where people share the TEK acquired during their observations in the field and their conversations with experts external to the community (Fig. 2). Plant and vegetation management to conserve the guacamaya verde. The environmental awareness groups observed that Ara militaris remained in the locality only during some months, and that their abundance was correlated with the availability of food in the territory. That knowledge led them to cultivate in their homegardens the plants eaten by the macaws. That posed them the challenge of knowing by experimentation how to germinate the seeds of each plant after realizing that each had different environmental requirements for germination, some species presenting dormancy. That started another cycle of observation of natural processes, dialogue with experts, and experimentation, which resulted in the continued propagation of plant species that provide food to the macaws, are sources of animal feed, or are in risk of extinction like Celtis sp., Jacaratia mexicana, Senna atomaria, Coulteria platyloba, Heteroflorum sclerocarpum, Albizia tomentosa, Swietenia humilis, Dalbergia granadillo, and others. The propagated plants have been transplanted by the local inhabitants in TDF and the limits of field crops and pastures, reforestation actions which have enriched the habitat of Ara militaris and reinforced

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

745

Fig. 2 Guacamayas Calentanas is a civil association that was consolidated in 2017 to be the legal representation of the Grupo Ambiental Guacamaya Verde and Grupo Ambiental Infantil Guacamaya Verde environmental awareness groups in El Chocolate, Churumuco in Tierra Caliente region. Some of its activities are community monitoring of the guacamaya verde (military macaw, Ara militaris) and other birds, seed collection, and germination in a community nursery for native species, reforestation, recovery of endangered species, environmental education workshops, creation and presentation of a play for children about the military macaw and its habitat, and others. At present, people from over 15 localities in the Bajo Balsas region participate in the management and conservation activities carried out by the association, the documentation of plant knowledge in their localities, and in a process for developing productive systems incorporating conservation actions

746

S. Rangel-Landa et al.

the awareness of the preservation of traditional management of productive systems following the agroforestry approach. Productive activities and conservation. The members of the environmental awareness group in the community El Chocolate were challenged by the need to change their farming practices and stop hunting wild fauna – a deeply rooted activity in the region as a source of food – because these activities were incompatible with their new scope of conservation. The reflection by local people began the whole process of environmental awareness in the locality, is illustrative of how this shift of scope was driven in local farmers: . . . me acabo yo, se acaba mi tierra y todavía no me alcanza para vivir, había algo que no encajaba . . . (. . . I get wasted, my land gets wasted, and I still don’t have enough to make a living; something was not fitting . . .). With that in mind, local people analyzed the cost-benefit ratio of agricultural production based on their experience and estimated that producing maize for the market, instead of being profitable, resulted in economic losses. Producing 1 kg of maize cost MX 7 (USD 0.3) and buyers paid producers only MX 5 (USD 0.2), a balance that in recent years has made local farmers to limit their production of maize to self-consumption levels and to invest their efforts in other productive activities. A similar analysis applied to cattle raising, during the dry season, farmers had to invest MX 150 (USD 6.5) daily to feed each head of cattle, so after estimating the volume of feed present in their pastures, they decided to reduce their herds to make buying feed unnecessary, which made the activity more profitable and reduced the environmental cost for pastures (Fig. 3). In 2019, the civil association Guacamayas Calentanas participated in the recently initiated Federal government program Jóvenes Construyendo el Futuro (Youngsters Building the Future) by enrolling 71 trainees of ages between 18 and 29 from 15 localities in the municipalities of Churumuco, La Huacana, Turicato, and Arteaga. This initiative created a space for intergenerational, interlocality, and intersectorial knowledge exchange and hands-on learning through theoretical and practical activities. Such process has contributed to developing communities’ capabilities for

Fig. 3 Productive activities, change of management practices, and innovation in El Chocolate. (a) Recovery of cultivation of local maize varieties in the milpa traditional polyculture system, at present made organically. (b) Reduction of herd size, rotation of pasture areas, and planting of high value feed plants ensuring their availability to cattle and increasing forest cover in pastures. (c) Innovation initiative Arte para la conservación involving women and men from several localities who make jewelry and small handicraft objects using the seeds and branch wood of native species. (Photo by the authors)

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

747

conservation, restoration, and sustainable use of natural resources, rising the interest in other localities for working toward these goals. Also, it has motivated the reflection about the importance of interlocality collaboration to preserve the region’s scarce resources like water and primary vegetation areas that provide environmental benefits to local communities. With the goal of generating productive alternatives based on the sustainable use of local natural resources, the trainees hired through Guacamayas Calentanas in collaboration with a group of ethnobotanists from the UNAM carried out a workshop for making handicrafts from seeds and wood obtained in the region. With that, training for the elaboration of handicrafts was transmitted to the local inhabitants, but also, they were motivated to observe and identify plants of potential use and learn about their availability and propagation to ensure their supply (Fig. 3). The project is in its initial phase through the Art for conservation progra, That is the beginning of a process to gain experience in finding buyers and selling their products. This process in the future could build a commercial network that, besides providing income to local households, would represent a new channel for making public the work for conservation. The implementation of the association’s communitarian nursery has allowed increasing the number of plant species propagated to satisfy the needs of its members, but also aims at obtaining profit through the sale of plants. To materialize this initiative, the association has started the process of registration of a Management Unit for Wildlife Conservation (UMA in its Spanish acronym). Documentation of local traditional knowledge. The recognition by young members of the group of their low level of knowledge about the environment compared to that of adult members and elders in the community and that this is not being transmitted to their progeny, created the need for finding support to document local TEK. That need resulted in the making of an ethnozoology and an ethnobotany bachelor’s degree dissertations. Members of the young and adult members of the environmental awareness groups collaborate in documenting the ethnobotanical dissertation and work sessions were carried out with trainees of the Youngsters Building the Future program. The participants became interested in documenting the knowledge of elders from their localities, in particular regarding traditional medicine, so that they could learn directly from their parents and grandparents and making knowledge available to future generations. Until the present, a group of trainees has documented 68 traditional remedies with images of the plants used for that purpose, the methods for their preparation. The started to reutilize them, and are planning to publish a book to be distributed in the region. Challenges for the permanence of the initiative and consolidation of the integrated management project. The work of people of El Chocolate in their organizational process, valuing TEK, and acting for conservation of Ara militaris and its habitat is being recognized by outsiders. However, the permanence of the initiative requires that the participants face the challenges posed by the oppressive socioeconomic context in which they live. The struggle between locally operating organized crime has affected the activities of the association to the point that in early 2020 several of its members were forced to emigrate to preserve their integrity and that of their families. Members of the association are under constant pressure from

748

S. Rangel-Landa et al.

the uncertainty of access to income for supporting their families, which might have a negative impact on the organizational process and the achievements so far reached, as expressed during one of the work sessions by a member: “. . . I grow organic maize because I am paid [the Federal government subsidizes the management of native germplasm under an organic methodology], but I wouldn’t if there was no money; I need to eat, seeing the guacamaya is nice, caring for the environment is good for ourselves, but if I have to cut down a tree to eat, I will cut it down . . .” This statement makes evident the need to act to ensure the availability to local inhabitants of basic resources like water, increase the profit of their productive activities, and reinforce the recently implemented productive alternatives, all of which is essential for the continuity of this initiative. Several of these requirements seem unreachable, but as demonstrated by the history of El Chocolate, the interest and motivation of its inhabitants has allowed them to remain active in the monitoring of the guacamaya verde in six localities in the region, maintain the communitarian nursery, and make restoration work despite the lack of institutional funding. The persistence of these processes shows that reinforcement of local capabilities based on the recognition of local knowledge and reciprocal participation of scholars and governmental officers, like members of the ZIBR technical staff, is essential for projects’ continuation, regardless of the adverse conditions there could be. It also shows that ethnobotanical research can be of help to the region’s communities initiatives.

Reflections, Perspectives, and Concluding Remarks The result of our review shows the incipient status of ethnobotanical research in the Tierra Caliente of Michoacán region. Most of the records of ethnobotanical information that we found were associated with botanical collections and are unpublished. We found few scientific publications and dissertations, and most have a descriptive approach focused on one of the components of TEK. Some municipalities in the region lack in-depth documentation about some plant groups or topics. Academic groups specialized in disciplines different from ethnobotany, like zootechnics, have found that ethnobotanical documentation is an important research tool. These trends agree with reports for the initial development stages of ethnobotanical research in Mexico, and as stressed by Camou-Guerrero et al. (2016), there is a need for broadening the scope of research and to collaborate in interdisciplinary and transdisciplinary groups. The TEK of local inhabitants has been essential for developing regional programs, as illustrated for the Zicuirán-Infiernillo area; however, there is a clear need to document such knowledge, understand the factors that influence its preservation or loss, and the risks and potentials of present plant resource management. Ethnobotanical research might contribute to the local communities’ and institutional efforts to conserve and manage biodiversity in a sustainable way, both with the generation of information and through research processes made following the principles of respect and reciprocity. A continued dialogue needs to be established between scholars,

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

749

government officers, and local communities that allow the reflection and motivation of all participants, and the putting into practice of management actions. The identification of culturally important plants and the factors involved in their management might provide information for decision-making about which plant species to consider for productive and conservation projects, the systems in which they would be incorporated, and documentation can aid making decisions about which productive projects should be continued. Documentation of plant management practices and the systems in which they are managed is essential for evaluating the so far little recognized contributions of traditional agroforestry management for conservation, identifying its issues, and designing proposals to improve their functionality and productivity. The processes of TEK erosion acknowledged by communities themselves call for understanding this phenomenon and its driving factors, and documenting the components of TEK systems, which would contribute to reinforce the regulatory framework about access to resources in compliance with the norms, beliefs, and worldview involved in these systems. The research subjects that can contribute to planning and improving resources management in concrete cases include the identification of the importance and effects of actual or potential plant use, estimation of optimal harvesting rates, and identifying methods to diminish the environmental impact of extraction. Likewise, understanding the processes of plant and landscape domestication that originated and maintain this region’s biocultural heritage have conceptual importance, but it would also accelerate the desired process of general acknowledgement of local farmers’ traditional management and its relationship with the identity of people of the Tierra Caliente of Michoacán. Acknowledgments We thank the contributions of staff members of the Faculty of Biology, UMSNH, Instituto de Investigaciones en Ecosistemas, UNAM, Guacamayas Calentanas A.C., and RBZI – CONANP, especially to Miguel Ángel Tornez Álvarez and Hugo Zepeda Castro who gave us their support during this work and shared their work experience in the region. Acknowledgments are due to the authorities of the Ichamio ejido and its Tizatal annex and the staff of the EBUM, IEB, MEXU, and BADEPLAM herbaria, especially to Rosa Isabel Fuentes Chávez, Laura Cortés Zarraga, and Rosalinda Medina Lemos for their help in database queries and information curation. We also wish to thank the Consejo Nacional de Ciencia y Tecnología, México (CONACYT) – Project CONACYT A1-S-14306; Dirección General de Asuntos del Personal Académico UNAM – Project IN206520 and IN224023; and Consejo Nacional para el Uso y Conocimiento de la Biodiversidad (CONABIO/GEF/FAO, GEF project ID 9380 770) – Project RG023 for granting us funding for fieldwork, fellowships, and salaries.

References Altieri MA, Toledo VM. The agroecological revolution in Latin America: rescuing nature, ensuring food sovereignty and empowering peasants. J Peasant Stud. 2011;38:587–612. Álvarez E. El viaje a América de Alexander Von Humboldt y Aimé Bonpland y las relaciones científicas de ambos expedicionarios con los naturalistas españoles de su tiempo. An del Inst Botánico JA Cavanilles (An Jard Bot Madr). 1965;22:9–60. Angón MP. Espacio agrícola y uso de la fuerza de trabajo en el Valle del Tepalcatepec, Michoacán. Bachelor thesis, Universidad Nacional Autónoma de México; 1993.

750

S. Rangel-Landa et al.

Arriaga L, Espinoza JM, Aguilar C, Martínez E, Gómez L, Loa E, editors. Regiones terrestres prioritarias de México. Tlalpan: CONABIO; 2000. Aswani S, Lemahieu A, Sauer WHH. Global trends of local ecological knowledge and future implications. PLoS One. 2018;13:e0195440. Ávila-Ramírez NA, Ayala-Burgos A, Gutiérrez E, Herrera-Camacho J, Madrigal-Sánchez X, Ontiveros-Alvarado S. Taxonomía y composición química de la necromasa foliar de las especies arbóreas y arbustivas consumidas durante la época de sequía en la selva baja caducifolia en el municipio de La Huacana, Michoacán México. Livest Res Rural Dev. 2007;19(6):11. Bayuelo-Jiménez JS, Lozano JC, Ochoa IE. Caracterización morfológica de Byrsonima crassifolia (l.) Kunth nativa de Churumuco, Michoacán, México. Rev Fitotec Mex. 2006;29(2):316. Beltrán-Rodríguez L, Manzo-Ramos F, Maldonado-Almanza B, Martínez-Ballesté A, Blancas VJ. Wild medicinal species traded in the Balsas Basin, Mexico: risk analysis and recommendations for their conservation. J Ethnobiol. 2017;37:743–64. Berkes F. Sacred ecology: traditional ecological knowledge and resource management. Philadelphia/New York: Taylor & Francis; 1999. Berkes F, Folke C, Gadgil M. Traditional ecological knowledge, biodiversity, resilience and sustainability. In: Perrings CA, Mäler KG, Folke C, Holling CS, Jansson BO, editors. Biodiversity conservation. Dordrecht: Kluwer Academic Publishers; 1995. p. 281–99. Berkes F, Colding J, Folke C. Rediscovery of traditional ecological knowledge as adaptive management. Ecol Appl. 2000;10:1251–62. Blanco-García A, Farfán R, Tornez MÁ, Madrigal R, Guzmán G, Farías N, Linares M, CarbajalNavarro A, Manrique A, Ríos G, Ramos I, Mejía ME. Guía ilustrada: árboles del Ejido Huacana, Michoacán. 2nd ed. Morelia: UMSNH; 2017. Boege E. El patrimonio biocultural de los pueblos indígenas de México. México City: Instituto Nacional de Antropología e Historia & Comisión Nacional para el Desarrollo de los Pueblos Indígenas; 2008. Burgos A, Cuevas G, Carlón T, Ramírez V, Tinoco A, Solorio G, Pérez S. Determinación de cambios de uso del suelo como base del ordenamiento territorial de la Reserva de la Biosfera Zicuirán Infiernillo. Morelia: GTZ-CONANP; 2010. p. 47. Caballero J, Cortés L. Percepción, uso y manejo tradicional de los recursos vegetales en México. In: Rendón B, Rebollar S, Caballero J, Martínez MA, editors. Plantas, cultura y sociedad: estudio sobre la relación entre humanos y plantas en los albores del siglo XXI. Mexico City: Universidad Autónoma Metropolitana, SEMARNAP; 2001. p. 79–100. Caballero J, Martínez-Ballesté A, Cortés L. Base de Datos Etnobotánicos de México (BADEPLAM). Ciudad de México: Jardín Botánico del Instituto de Biología, UNAM; 2020. Caldera D. Proyecto de Conservación Comunitaria de la Biodiversidad (COINBIO): análisis de su contribución en la conservación de la biodiversidad en el Estado de Michoacán, México (MSc thesis). Morelia: UMSNH; 2014. Camou-Guerrero A, Casas A, Moreno-Calles AI, Aguilera-Lara J, Garrido-Rojas D, Rangel-LandaS, Torres I, Pérez-Negrón E, Solís L, Blancas J, Guillén S, Parra F, Rivera-Lozoya E. Ethnobotany in Mexico: history, development, and perspectives. In: Lira R, Casas A, Blancas J, editors. Ethnobotany of Mexico: interactions of people and plants in Mesoamerica. New York: Springer; 2016. p. 21–39. Casanova-Lugo F, Flores-Estrada MX, López-Santiago G, García-Gómez M. Estructura, composición y usos de los árboles de la selva baja caducifolia en Apatzingán, Michoacán. Trop Subtrop Agroecosyst. 2014;17:155–259. Casas A, Caballero J, Mapes C, Zárate S. Manejo de la vegetación, domesticación de plantas y origen de la agricultura en Mesoamérica. Bot Sci. 1997;61:31–47. Casas A, Otero-Arnaiz A, Pérez-Negrón E, Valiente-Banuet A. In situ management and domestication of plants in Mesoamerica. Ann Bot. 2007;100(5):1101–15. Casas A, Lira R, Torres I, Delgado A, Moreno-Calles AI, Rangel-Landa S, Blancas J, Larios C, Solís L, Pérez-Negrón E, Vallejo M, Parra F, Farfán-Heredia B, Arellanes Y, Campos N. Ethnobotany for sustainable ecosystem management: a regional perspective in the Tehuacán Valley. In: Lira R, Casas A, Blancas J, editors. Ethnobotany of Mexico: interactions of people and plants in Mesoamerica. New York: Springer; 2016a. p. 179–206.

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

751

Casas A, Moreno-Calles AI, Vallejo M, Parra F. Importancia cultural y potencial de los recursos genéticos. In: Casas A, Torres-Guevara J, Parra F, editors. Domesticación en el continente americano: manejo de la Biodiversidad y evolución dirigida por las culturas del Nuevo Mundo. UNALM, UNAM: Lima; 2016b. p. 51–74. CONABIO. Características sociodemográficas de México por municipio, 2010, scale: 1:250000. Comisión Nacional para el Conocimiento, Uso de la Biodiversidad, Instituto Nacional de Estadística y Geografía: Mexico City; 2012a. CONABIO. Actividades económicas en México por municipio, 2010. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad, Instituto Nacional de Estadística y Geografía: Mexico City; 2012b. CONABIO. La biodiversidad en Michoacán Estudio de Estado 2. México City: CONABIO; 2019a. CONABIO. Catálogo de autoridades taxonómicas de especies de flora y fauna con distribución en México. Base datos SNIB-CONABIO. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad: México City; 2019b. CONABIO. Sistema Nacional de Información sobre Biodiversidad. Registros de ejemplares de plantas. Ciudad de México: CONABIO; 2020. CONANP. Programa de manejo de la Reserva de la Biosfera Zicuirán-Infiernillo. Ciudad de México: SEMARNAT, CONANP; 2014. Cruse-Sanders JM, Friar EA, Parker KC, Huang DI, Mashayekhi S, Prince LM, Otero-Arnaiz A, Casas A. Managing diversity: domestication and gene flow in Stenocereus stellatus Riccob. in Mexico. Ecol Evol. 2013;3(5):1340–55. Cué-Bär EM, Villaseñor JL, Arredondo-Amezcua L, Cornejo-Tenorio G, Ibarra-Manríquez G. La flora arbórea de Michoacán, Mexico. Bot Sci. 2006;78:47–81. Delfín M. Los Pirindas de Michoacán: ¿inicio de un proceso de etnogénesis? Cuiculco. 2011;50: 145–58. Diaz JL. Conocimiento del manejo, comercialización y utilización de dos especies de pitaya (Stenocereus spp.) en el municipio de Huetamo, Michoacán. Bachelor thesis, Universidad Michoacana de San Nicolás de Hidalgo; 1995. Fernández R, Rodríguez C, Arreguín M d l L, Rodríguez A. Listado florístico de la cuenca del Río Balsas, Mexico. Polibotánica. 1998;9:1–151. Figueroa M. Estrategias productivas, comerciales y de vida de los productores de mango en Apatzigán, Michoacán, México. Morelia: Bachelor thesis, Universidad Autónoma de Chapingo; 2005. Gallegos LL. Caracterización de la flora de Tumbiscatío, Michoacán, México. Texcoco: Bachelor thesis, Universidad Autónoma de Chapingo; 2015. Galván JDA. Conocimiento de la dualidad alimenticio-medicinal de la flora de la localidad de las Anonas, municipio de Jungapeo de Juárez, Michoacán: un estudio etnobotánico. Los Reyes Iztacala: Bachelor thesis, Universidad Nacional Autónoma de México; 2005. García S. Caracterización de la zona productora de limón mexicano (Citrus aurantifolia Swingle) en el valle de Apatzingán, Michoacán. Bachelor thesis, Universidad Autónoma de Chapingo; 2006. García E, CONABIO. Climas catálogo de metadatos geográficos. Shapefile, scale 1:1000000. México City: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 1998. García I, Linares A. Árboles y arbustos de la cuenca del Río Tepalcatepec (Michoacán y Jalisco, México) para uso urbano. Zamora: El Colegio de Michoacán, Instituto Politécnico Nacional; 2010. Gómez M. El herbario de la Facultad de Biología de la Universidad Michoacana de San Nicolás de Hidalgo (EBUM). In: Chávez MA, Gómez M, editors. Memoria de las plantas, pasado y presente del Herbario de la Facultad de Biología. Morelia: UMSNH; 2012. p. 77–128. Gómez-Baggethun E, Corbera E, Reyes-García V. Traditional ecological knowledge and global environmental change: research findings and policy implications. Ecol Soc. 2013;18:72. González L. Introducción: La Tierra Caliente de Michoacán. In: Zárate JE, editor. La Tierra Caliente Michoacán. Zamora: El Colegio de Michoacán, Gobierno del Estado de Michoacán; 2001. p. 17–66.

752

S. Rangel-Landa et al.

González JC, Ayala A, Gutiérrez E. Composición química de especies arbóreas con potencial forrajero de la Región de Tierra Caliente, Michoacán, México. Rev Cuba Cienc Agrícola. 2007;41(1):8793. González J, Madrigal X, Ayala A, Juárez A, Gutiérrez E. Especies arbóreas de uso múltiple para la ganadería de Tierra Caliente del Estado de Michoacán, México. Livest Res Rural Dev. 2006;18(8). Guevara-Fefer F, Rzedowski J. Notas sobre el género Bursera (Burseraceae) en Michoacán (México) I. Tres especies nuevas de los alrededores de la presa del infiernillo, con algunos datos relativos de la región. Boletín la Soc. Botánica México. 1980;39:6381. Gutiérrez-Vázquez E, Ávila-Ramírez N, Madrigal X, López-Hernández N, Villalba-Sánchez CA, Juárez-Caratachea A, et al. Árboles forrajeros de la selva baja caducifolia en el Estado de Michoacán: avances y perspectivas. In: Herrera J, Chay AJ, Casanova F, Piñeiro ÁT, Márquez L, Santillán E, et al., editors. Av. la Investig. sobre Prod. Anim. y Segur. Aliment. en México. Universidad Michoacana de San Nicolás de Hidalgo; 2018. p. 25370. Hinton J, Rzedowski J. George B. Hinton, collector of plants in southwestern Mexico. J Arnold Arbor. 1972;53:141–81. Ihl T, Bautista F, Mendoza ME. Conservación e intensidad de uso de la tierra en la Reserva de la Biósfera Zicuirán-Infiernillo, Michoacán. Terra Digit. 2017;1:1–5. INEGI. Continuo de Elevaciones Mexicano 3.0 (CEM 3.0). Raster, 15 m. Aguascalientes: Instituto Nacional de Estadística y Geografía; 2013. INEGI. Uso de suelo y vegetación, serie VI. Shapefile, scale 1:250000 (continuo nacional). Aguascalientes: Instituto Nacional de Estadística y Geografía; 2016. INEGI. División política municipal. Shapefile 1:250000. Aguascalientes: Instituto Nacional de Estadística y Geografía; 2019. INEGI, INE, CONAGUA. Cuencas Hidrográficas de México, 2007. Shapefile, scale 1:250000, elaborated by Priego A G, Isunza E, Luna N, Pérez J L. México City: Instituto Nacional de Estadística Geografía e Informática, Instituto Nacional de Ecología, Comisión Nacional de Agua; 2007. INPI. Indicadores socioeconómicos de los pueblos indígenas de México, 2015. Instituto Nacional de los Pueblos Indígenas: Mexico City; 2017. Leavenworth WC. A preliminary study of the vegetation of the region between Cerro Tancitaro and the Rio Tepalcatepec, Michoacan, Mexico. Am Midl Nat. 1946;36:137–206. Léonard E. Una historia de vacas y golondrinas: ganaderos y campesinos temporeros del trópico seco mexicano. Mexico City: Fondo de Cultura Económica, El Colegio de Michoacán, ORSTOM; 1995. Lira R, Casas A, Rosas-López R, Paredes-Flores M, Pérez-Negrón E, Rangel-Landa S, et al. Traditional knowledge and useful plant richness in the Tehuacán–Cuicatlán Valley, Mexico. Econ Bot. 2009;63:271–87. López N, Gutiérrez E, Juárez A, Salas G, Ayala A, García A. Valor nutricional de nuevas especies arbóres nativas con potencial forrajero en el Estado de Michoacán. Engormix. 2017;6/6/2017. Luna AL. Identificación, selección y aprovechamiento de árboles semilleros en áreas de conservación comunitaria en el municipio de Churumuco, Michoacán, México. Morelia: MSc thesis, Universidad Nacional Autónoma de México; 2011. Luna-Nieves AL, Meave J, Cerdeira LP, Ibarra-Manríquez G. Reproductive phenology of useful seasonally dry tropical forest trees: guiding patterns for seed collection and plant propagation in nurseries. For Ecol Manage. 2017;393:5262. MacNeish RS. The origins of agriculture and settled life. Norman: University of Oklahoma Press; 1992. Martínez J, de la Trinidad NK. Costos y competitividad de la producción de mango (Magnifera indica L.) en la región de Tierra Caliente, Michoacán. Texcoco: Bachelor thesis, Universidad Autónoma de Chapingo; 2007. Martínez R, Cadena E, Sandoval EA. Sociodemografía en la Región de Tierra Caliente (Sur Estado de México, Este de Michoacán y Norte de Guerrero), 1990–2010. Geogr y Sist Inf Geogr. 2016;8:114–35.

Traditional Ecological Knowledge and Biodiversity Conservation in. . .

753

Martínez-Cruz J, Méndez-Toribio M, Cortés-Flores J, Coba-Pérez P, Cornejo-Tenorio G, IbarraManríquez G. Estructura y diversidad de los bosques estacionales desaparecidos por la construcción de la presa Gral. Francisco J. Múgica, en la Depresión del Balsas, Michoacán, México. Rev Mex Biodivers. 2013;84:1216–34. Méndez-Toribio M, Martínez-Cruz J, Cortés-Flores J, Rendón-Sandoval FJ, Ibarra-Manríquez G. Composición, estructura y diversidad de la comunidad arbórea del bosque tropical caducifolio en Tziritzícuaro, Depresión del Balsas, Michoacán, Mexico. Rev Mex Biodivers. 2014;85:1117–28. MEXU-IBUNAM. Plantas Vasculares. Portal de Datos Abiertos UNAM Colecciones Universitarias. México City: Departamento de Botánica, Herbario Nacional de México (MEXU), Instituto de Biología (IBUNAM), UNAM; 2019. Available from: https:// datosabiertos.unam.mx/. Accessed Oct 2019. Molina MA. Uso de las plantas en tres localidades de municipio de Churumuco. Morelia: Bachelor thesis, Universidad Michoacana de San Nicolás de Hidalgo; 2010. Moreno-Calles AI, Toledo VM, Casas A. Los sistemas agroforestales tradicionales de México: una aproximación biocultural. Bot Sci. 2013;91(4):375–98. Moreno-Calles AI, Casas A, Rivero-Romero AD, Romero-Bautista YA, Rangel-Landa S, FisherOrtíz RA, Alvarado-Ramos F, Vallejo-Ramos M, Santos-Fita D. Ethnoagroforestry: integration of biocultural diversity for food sovereignty in Mexico. J Ethnobiol Ethnomed. 2016;12:54. Oliveros JA, editor. Raíces culturales en la historia de la Tierra Caliente de Michoacán. México City: El Colegio de Michoacán; 2011. Orozco Q, Odenthal J. Lenguas prehispánicas México y Centro América. Shapefile 1:5000000. Morelia: Centro de Investigaciones en Geografía Ambiental, UNAM; 2016. Parra F, Casas A, Peñaloza-Ramírez J, Cortés-Palomec AC, V. Rocha-Ramírez y A. GonzálezRodríguez. Evolution under domestication: ongoing artificial selection and divergence of wild and managed Stenocereus pruinosus (Cactaceae) populations in the Tehuacán Valley, Mexico. Ann Bot. 2010;106:483–96. Peña-Estrada D, Robles J, Montero JC. Cactáceas de San Francisco de los Ranchos, Reserva de la Biosfera Zicuirán-Infiernillo, Michoacán, Mexico. Biológicas. 2016;18:29–35. Pérez S. Medicina doméstica en la Tierra Caliente Michoacana. Zamora: El Colegio de Michoacán; 2013. QGIS_Development Team. QGIS Geographic Information System 3.12.3. Open Source Geospatial Foundation Project; 2020. RAN. Catálogo de núcleos agrarios de la propiedad social. Registro Agrario Nacional. 2017a. Available from: https://datos.gob.mx/busca/dataset/catalogo-de-nucleos-agrarios-de-la-pro piedad-social. Accessed 9 Oct 2020. RAN. Perimetrales núcleos agrarios SHAPE Entidad Federativa Michoacán. 2017b. Available from: https://datos.gob.mx/busca/dataset/datos-geograficos-perimetrales-de-los-nucleosagrarios-certificados-por-estado. Accessed 9 Oct 2020. Rangel-Landa S, Armas-Bautista ME, Ortiz-Ávila T, Burgos A, Granados J, Carmona AR, et al. Inventario de flora y fauna en el Ejido Ichamio, La Huacana, Mich. Morelia: Ejido Ichamio, Grupo GESTA, COINBIO; 2003. Rodríguez A. Riqueza, estructura y carbono aéreo de la flora leñosa de un bosque tropical caducifolio en La Huacana, Michoacán. Bachelor thesis, Universidad Nacional Autónoma de México; 2018. Sánchez G, Chávez MA. Una aproximación a los estudios y las colectas botánicas en Michoacán. De la época colonial a la primera década del siglo XX. In: Chávez MA, Gómez M, editors. Memoria de las plantas, pasado y presente del Herbario de la Facultad de Biología. Morelia: UMSNH; 2012. p. 9–76. Saynes-Vásquez A, Caballero J, Meave JA, Chiang F. Global trends of local ecological knowledge and future implications knowledge among the Isthmus Zapotecs of Mexico. J Ethnobiol Ethnomed. 2013;9:40. SEMARNAT. Norma oficial mexicana NOM-059-SEMARNAT-2010, Protección ambientalEspecies nativas de México de flora y fauna silvestres-Categorías de riesgo y especificaciones

754

S. Rangel-Landa et al.

para su inclusión, exclusión o cambio-Lista de especies en riesgo. México City: SEMARNAT, Diario Oficial de la Federación; 2010. Sosa GN, Zarazúa R, Baltazar GB. Programa de conservación comunitaria de la biodiversidad: política pública con enfoque participativo. La biodiversidad en Michoacán estudio de Estado 2. México City: CONABIO; 2019. p. 225–30. Soto JC. Las plantas medicinales y su uso tradicional en la cuenca del Río Balsas; Estado de Michoacán y Guerrero, México (Bachelor thesis). Morelia: UNAM; 1987. Soto NJ. Plantas útiles de la Cuenca del Balsas. In: Ceballos G, Martínez L, García A, Espinoza E, Bezaury J, Dirzo J, editors. Diversidad, amenazas y áreas prioritarias para la conservación de las selvas secas del Pacífico Mexicano. México City: Fondo de Cultura Económica, CONABIO; 2010. p. 285–320. Soto-Nuñez JC, Sousa M. Plantas medicinales de la Cuenca del Río Balsas. Ciudad de México: Instituto de Biología, UNAM; 1995. SPI. Programa especial para los pueblos indígenas de Michoacán 2015–2021. Morelia: Secretaría de los Pueblos Indígenas, Gobierno del Estado de Michoacán; 2015. Tena M. Calendario botánico de Michoacán 1892. 3rd ed. Morelia: Centro de Investigación y Desarrollo del Estado de Michoacán, Secretaria de Urbanismo y Medio Ambiente, UMSNH; 1892. Thiebaut V. Movilidad de la población y evolución de los paisajes en la Tierra Caliente de Michoacán, Mexico. Rev Geogr Am Cent. 2011;2:1–15. Thiébaut V, Aguirre A. Procesos en los paisajes de la Tierra Caliente de Michoacán: asentamientos humanos y usos de suelo en Buenavista Tomatlán. Contrib desde Coatepec. 2011;(20):75100. Toledo VM, Barrera-Bassols N. La Memoria Biocultural: la importancia ecológica de las sabidurías tradicionales. Barcelona: Icaria Editorial; 2008. Toledo VM, Alarcón-Cháires P, Ortíz-Ávila T, Acosta-Moreno L. Atlas de experiencias comunitarias en el manejo sustentable de los recursos naturales en Michoacán. Morelia: Universidad Veracruzana; 2006. Torres J. Etnobotánica mexicana. Plantas tradicionalmente utilizadas para el tratamiento de enfermedades gastrointestinales en el Estado de Michoacán. Bachelor thesis, Universidad Nacional Autónoma de México; 1984. Uribe LM, Ramírez T, Labarthe R. El estado de la migración: índices de intensidad migratoria México-Estados Unidos 2010. México City: Consejo Nacional de Población; 2012. Urquijo PS. Humboldt y el Jorullo, historia de una exploración. México City: SEMARNAT, INE-SEMARNAT, UNAM, INE, CEDEM, ColMich, UMSNH, H. Ayunt. de La Huacana; 2010. Vargas E. Levantamiento ecológico de malezas en maíz, arroz y sorgo en el Valle de Apatzingán (Bachelor thesis). Uruapan: Facultad de Agrobiología, UMSNH; 1989. Velázquez M, Sosa G, Navarrete P, Torres G. Programa de conservación para el Estado de Michoacán 2003–2008. Bases para la conformación del sistema de áreas de conservación del Estado de Michoacán. Morelia: Secretaría de Urbanismo y Medio Ambiente del Gobierno del Estado de Michoacán, UNAM; 2005. Villa-Méndez CI, Tena MJ, Tzintzun R, Val D. Caracterización de los sistemas ganaderos en dos comunidades del municipio de Tuzantla de la región de Tierra Caliente, Michoacán. Av Investig Agropecu. 2008;12(2):45–58. Villa-Reyes. Análisis de los proyectos productivos del Programa de Conservación para el Desarrollo Sostenible (PROCODES) en la Reserva de la Biosfera Zicuirán-Infiernillo, Michoacán (2008–2015) (MRD thesis). Texcoco: Universidad Autónoma Chapingo; 2018. Villaseñor JL. Checklist of the native vascular plants of Mexico. Rev Mex Biodivers. 2016;87: 559–902. Villegas-Santibañez H. Políticas agrícolas, cambios productivos y migración internacional en Tierra Caliente, Michoacán. México City: El Colegio de la Frontera Norte; 2007. Zamora V. Los principales colectores del EBUM (1901–2010). In: Chávez MA, Gómez M, editors. Memoria de las plantas, pasado y presente del Herbario de la Facultad de Biología. Morelia: UMSNH; 2012. p. 129–44.

Ethnobotanical Science in the Clouds: Useful Plants of Northeastern Oaxaca, Mexico Beatriz Rendo´n-Aguilar, David Bravo-Avilez, Luis Alberto Bernal-Ramírez, Abisaí García-Mendoza, Adolfo Espejo-Serna, Ana Rosa Lo´pez-Ferrari, Carlos Dura´n-Espinosa, David S. Gernandt, Francisco Lorea-Herna´ndez, Guillermo Ibarra-Manríquez, Jaime Jime´nez-Ramírez, Jesu´s Ricardo de Santiago-Go´mez, Jorge Santana-Carrillo, Jose´ Luis Villasen˜or, Laura Ya´n˜ez-Espinosa, Lucio Lozada-Pe´rez, Marie-Ste´phanie Samain, Susana Valencia-A´valos, Rosa María Fonseca-Jua´rez, and Salvador Arias-Montes Introduction Oaxaca is one of the five Mexican states with the highest biological richness and one of the most diverse in ethnic composition. Knowledge and use of plant resources have been part of their culture and maintenance in territories with different B. Rendón-Aguilar (*) · D. Bravo-Avilez · L. A. Bernal-Ramírez · A. Espejo-Serna · A. R. López-Ferrari · J. Santana-Carrillo Departamento de Biología, Universidad Autónoma Metropolitana Iztapalapa, Ciudad de México, Mexico e-mail: [email protected]; [email protected]; [email protected] A. García-Mendoza · D. S. Gernandt · J. L. Villaseñor · S. Arias-Montes Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, México e-mail: [email protected]; [email protected]; [email protected]; [email protected] C. Durán-Espinosa · F. Lorea-Hernández Red de Biodiversidad y Sistemática, Instituto de Ecología, Mexico City, Mexico e-mail: [email protected]; [email protected] G. Ibarra-Manríquez Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de, Ciudad de México, México e-mail: [email protected] J. Jiménez-Ramírez · L. Lozada-Pérez1 · S. Valencia-Ávalos · R. M. Fonseca-Juárez Facultad de Ciencias, Universidad Nacional Autónoma de México, Ciudad de México, México e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_16

755

756

B. Rendo´n-Aguilar et al.

ecological, geological, and climatic conditions. Oaxaca contains 43.9% of the Mexican flora, with about 10,229 vascular plant species, distributed in 26 vegetation types. Ethnobotanical research in Oaxaca has been conducted during decades, focused on aspects like the integral traditional management, traditional classification systems, floristic composition and management of different ecosystems, agroecosystems. This chapter synthesizes a panorama of the ethno-floristic inventory in some preserved areas of northeastern Oaxaca, in three Priority Terrestrial Regions (PTR) with high bio-cultural diversity. The study area comprised 84 municipalities of Oaxaca, with eight ethnic groups in the Tehuacán-Cuicatlán Valley, the Sierras del Norte de Oaxaca-Mixe and the Selva Zoque-La Sepultura. Ethnobotanical data were recorded by participative observation, ethnobotanic fieldtrips, and free interviews with the local guides, for almost 4 years. Specimens of useful plants were collected and photographed the vouchers deposited in UAM-IZ, MEXU, and OAX. Field information and photographs of each specimen were integrated into the database BIÓTICA (CONABIO 2012) and the images bank of CONABIO (2018). More than 2,340 specimens were collected, a total of 139 families, 441 genera, 804 species, 7 subspecies, and 18 varieties were identified. The genus Quercus was the richest one with 34 species, followed by Tillandsia L. (22 species), and Pinus L. (14). Most of the useful species were found in four vegetation types: montane cloud forests (MCF), oak forests (OF), pine forests (PF), and tropical semi-evergreen forest (TSF). We recorded 11 use categories, the most important being medicinal, edible, and ornamental plants. Variation in number of useful families and species among ethnic groups was detected. Those with the greatest number of useful species and botanical families are the Northern Zapotec, Mixe, and Mazatec people, whereas the ethnic groups with the lowest values of useful species and families are also distributed in few municipalities, like Mixtec and Nahua. Results of this research reinforce the statement that people of ethnic groups are local safeguards of biodiversity. Patterns of regeneration processes indicate high rates of resilience of ecosystems, which needs to be analyzed to understand the kinds of management that people practice to recovering vegetation, since apparently, ethnic groups of Northeastern Oaxaca assist empirically to ecosystems restoration. The total number of useful species recorded in the present study increase in almost 10% these records. For many localities these are the first records of useful plants. Traditional knowledge must persist by its own right to persist, because we must be clear and sensitive that there are many explanations of world and life.

J. R. de Santiago-Gómez Facultad de Ciencias, Universidad Nacional Autónoma de México, Ciudad de México, México L. Yáñez-Espinosa Instituto de Investigación de Zonas Desérticas, Universidad Autónoma de San Luis Potosí, San Luis Potosí, México e-mail: [email protected] M.-S. Samain2 Red de Diversidad Biológica del Occidente, Instituto de Ecología, Xalapa, Veracruz, México

Ethnobotanical Science in the Clouds: Useful Plants of. . .

757

Mexico is one of the ten most important countries of the world in terms of high levels of biological and cultural diversity (CONABIO 2018). Based on archeological records, it is also known that use and management of natural resources have occurred for at least 10,000 years (Rindos 1984; Smith 1967), which implies cumulative processes of learning, trying, using, management, and eventually domestication of those resources that were selected for different purposes. Moreover, a philosophical context has evolved, which means that ancestral people developed a worldview to link that knowledge and practices with metaphysical concepts to give a sense to their surrounding world (Toledo and Barrera-Bassols 2008). Currently, ethnobotanical research has tried to understand the complex world referred to above, in order to explain patterns and processes involved in it. Consequently, different studies have been conducted throughout Mexico to record the importance of biodiversity for food, medicinal, ceremonial, and other human cultural uses, as well as traditional ecological knowledge (Aswani et al. 2018; LunaJosé and Rendón 2008). Oaxaca is considered one of the five Mexican states with the highest biological richness, as well as one of the most diverse in terms of ethnic composition (Caballero et al. 2004; García-Mendoza and Meave 2012). Local people belonging to different ethnic groups refer to histories about their village foundation, conflicts, social and political dynamics, as well as legends and myths of their surrounding ecological environment. In this context, knowledge and use of plant resources have been part of their own culture and maintenance in different territories, with different ecological, geological, and climatic conditions into this state. Oaxaca has been classified into different regions, based on cultural-ethnic and political division, the 590 municipalities of the state have been categorized in eight regions and 30 districts (Boege 2008; INAFED-SEGOB 2010; Ordóñez 2000). In addition, joining biodiversity and cultural approaches, the state harbors eight of the 151 Priority Terrestrial Regions proposed for Mexico (Arriaga et al. 2000). In terms of floristic records, Oaxaca contains 43.9% of the Mexican flora, conformed of about 10,229 vascular plant species (Villaseñor 2016), distributed in 26 vegetation types (Mickel and Beitel 1988; Torres-Colín 2004), representing 70% of all vegetation types present in Mexico (García-Mendoza et al. 2004). The most important plant communities are the temperate forests (i.e., pine, oak, and pine-oak forest), montane cloud forest (which has its highest distribution in Oaxaca), tropical evergreen/semi-evergreen forests, tropical dry deciduous forest, grasslands, and shrublands. Due to this complex bio-cultural diversity, ethnobotanical research in Oaxaca has been developed during decades, focused on different aspects, from the integral traditional management (Zizumbo and Colunga 1982), traditional classification systems (Hunn 2008; Messer 1978), floristic composition and management of different ecosystems (Ventura-Aquino et al. 2008), agroecosystems such as traditional coffee plantations, family orchards, and milpas (Cardoso 2004; Luna-José 2006; Luna-José and Rendón 2008; Manzanero et al. 2009; Montalvo 2006; Van der Wal 1996) of traditional uses, forms or application of medicinal plants, and specific taxonomic groups (Bernal-Ramírez et al. 2019; Martínez-Bautista et al. 2019; Rendón-Aguilar et al. 2017b). However, different areas remain poorly known,

758

B. Rendo´n-Aguilar et al.

such as the Mixteca, Sierra Madre del Sur, Istmo de Tehuantepec, Depresión del Balsas, Valles Centrales, Sierra Atravesada, Sierra Madre de Oaxaca (Sierra de Juárez, Chinantla, and Sierra Mixe), and Costa (Prance and Campbell 1988; Torres-Colín 2004). More recently, the CONABIO (National Commission for the Knowledge and Use of Biodiversity) has promoted floristic inventories in many areas of Mexico, including Oaxaca (e.g., Meave et al. 2017; Rendón-Aguilar et al. 2017a), but only 55 studies have been supported from 1999 to the present, and 55% were carried out in the Tehuacán-Cuicatlán Valley (e.g., Blancas et al. 2010; Casas et al. 2001; Lira et al. 2009), so many regions have been poorly studied. Every study undoubtedly extends ethnobotanical background about useful plants and increases our consciousness about traditional ecological knowledge. This information contributes to understand and confirm, the role of ethnic groups in preserving ecosystems and safeguarding nature (Bernal-Ramírez et al. 2019; Toledo and BarreraBassols 2008), and the importance to understand how they do this, which traditional practices are followed, and the limits or deficiencies of these practices. The present chapter represents a contribution to the understanding of this knowledge in some preserved areas of northeastern Oaxaca, specifically within three Priority Terrestrial Regions (PTR) (which also represent areas of high bio-cultural diversity) and is a synthesis of the ethno-floristic inventory carried out by Rendón-Aguilar et al. (2017a).

Methods Study Area The study area comprised 84 municipalities belonging to ten Political Districts, five Territorial Regions, and three Priority Terrestrial Regions (PTRs) of Oaxaca (Fig. 1). Based on our fieldwork, we found eight ethnic groups distributed as follows: PTR 121 (Valle de Tehuacán-Cuicatlán), occupied by the Cuicatec and Mazatec people; PTR 130 (Sierras del Norte de Oaxaca-Mixe) by the Chinantec, Mixe, Mixtec, Nahua, and Zapotec people; and PTR 132 (Selva Zoque-La Sepultura), inhabited by the Zoque people. This area also included four municipalities with mestizo populations (Natividad, San Pedro Teutila, Santiago Xiacuí in PTR 130, and San Pedro Tapanatepec in PTR 132). The number of municipalities inhabited by each ethnic group was highly variable because it depends on the PTR selected (Table 1) (Rendón-Aguilar et al. 2017a). This area comprises a wide mosaic of vegetation types, with a complex distribution, occurring at a wide elevation range.

Ethnobotanical Information The purposes of our research were explained to local authorities, who always were very helpful, and who authorized us to work into their territories, and to collect plant specimens. They also helped by assigning us local guides who knew the area, and who had extensive knowledge on the plants used in the past and at the present, as well as their Spanish and indigenous languages names. In almost all the cases, two or

Ethnobotanical Science in the Clouds: Useful Plants of. . .

759

Fig. 1 Study área, indicating the three PTRs of Oaxaca were study was followed: PTR 121: Valle de Tehuacán-Cuicatlán; PTR 130: Sierras del Norte de Oaxaca-Mixe; PTR 132: Selva Zoque-La Sepultura. Blue dots represent collecting sites of Magnoliophyta taxa, red triangles are the collecting sites of Coniferophyta taxa, and yellow stars indicate collecting sites of Lycophyta and Polypodiophyta taxa

three local guides were assigned in each municipal authority. Most of the guides were men from 20 to 65 years old; female guides were assigned only in four localities (Bernal-Ramírez et al. 2019; Martínez-Bautista et al. 2019; RendónAguilar et al. 2017b). The total of our ethnobotanical records was documented with the help of about 200 local people and collected from approximately 650 sites of 84 municipalities. Ethnobotanical data were recorded combining different methods: participative observation (Spradley 1980), ethnobotanic fieldtrips (Aguilar et al. 1994), and free interviews with the local guides (Albuquerque et al. 2014; Alexiades 1996). For almost 4 years (from January 2013 to July 2016), fieldtrips of 10–12 days every two months were conducted to collect useful plants within the selected municipalities (Rendón-Aguilar et al. 2017a, b). Each municipality was visited once. In many cases, local guides noted useful species that were not mature at the time of the fieldwork and were neither collected nor recorded, because of the difficulty for identifying them without reproductive structures. Specimens were collected only if local guides recognized them as useful plant (in the past or currently), and if they contained some reproductive structures (e.g., buds, flowers, fruits, sori, or cones). These specimens were photographed and collected, and ecological and geographical data were recorded, including date, municipality, locality (collection site), geographic reference, elevation, vegetation type (according to Torres-Colín 2004), visual soil characteristics, biological data of each species (i.e., habitat, size, growth

B. Rendo´n-Aguilar et al.

760

Table 1 List of municipalities visited during the present study. Geopolitical, biological, and ethnic classifications where they belong to, are indicated. Ethnic group(s) and useful plant families are indicated for each municipality

Municipalities Ayotzintepec

PTR 130

Culturalethnic region Papaloapam

San Felipe Usila

130

Papaloapam

Tuxtepec

Chinantecs

San José Chiltepec

130

Papaloapam

Tuxtepec

Chinantecs

San Juan Bautista Valle Nacional

130

Papaloapam

Tuxtepec

Chinantecs

San Juan Lalana

130

Papaloapam

Choápam

Chinantecs

Political district Tuxtepec

Ethnic groups Chinantecs

Useful plant families Araceae, Arecaceae, Asteraceae, Begoniaceae, Bromeliaceae, Cucurbitaceae, Euphorbiaceae, Gesneriaceae, Moraceae, Muntingiaceae, Plantaginaceae, Pteridaceae, Urticaceae, Violaceae, Zamiaceae Asteraceae, Begoniaceae, Piperaceae, Siparunaceae, Urticaceae Araceae, Asteraceae, Cecropiaceae, Commelinaceae, Euphorbiaceae, Iridaceae, Marantaceae, Melastomataceae, Moraceae, Piperaceae, Rubiaceae, Sapotaceae Apocynaceae, Asparagaceae, Asteraceae, Clusiaceae, Convolvulaceae, Euphorbiaceae, Malvaceae, Piperaceae, Primulaceae, Rubiaceae, Selaginellaceae, Siparunaceae, Thelypteridaceae Altingiaceae, Annonaceae, Arecaceae, Bixaceae, Cannabaceae, Cannaceae, Euphorbiaceae, Malvaceae, Melastomataceae, (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

761

Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

San Juan Quiotepec

130

Sierra Norte

Ixtlán

Chinantecs

San Pedro Sochiápam

130

Cañada

Cuicatlán

Chinantecs

San Pedro Yólox

130

Sierra Norte

Ixtlán

Chinantecs

Useful plant families Phytolaccaceae, Poaceae, Primulaceae, Rubiaceae, Zamiaceae Actinidiaceae, Alstroemeriaceae, Apiaceae, Araliaceae, Asteraceae, Bignoniaceae, Clethraceae, Cucurbitaceae, Ericaceae, Fagaceae, Lauraceae, Myricaceae, Myrtaceae, Orchidaceae, Pentaphylacaceae, Phytolaccaceae, Pinaceae, Piperaceae, Plantaginaceae, Rubiaceae, Solanaceae Actinidiaceae, Altingiaceae, Asteraceae, Bromeliaceae, Chloranthaceae, Ericaceae, Fabaceae, Malvaceae, Melastomataceae, Orchidaceae, Phytolaccaceae, Pinaceae, Piperaceae, Solanaceae Altingiaceae, Amaranthaceae, Apiaceae, Araceae, Arecaceae, Asteraceae, Cannaceae, Cecropiaceae, Clusiaceae, Cucurbitaceae, Ericaceae, Fabaceae, Lamiaceae, Lindsaeaceae, Lycopodiaceae, Malvaceae, Orchidaceae, Papaveraceae, (continued)

B. Rendo´n-Aguilar et al.

762 Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

Santa María Jacatepec

130

Papaloapam

Tuxtepec

Chinantecs

Santiago Comaltepec

130

Sierra Norte

Ixtlán

Chinantecs

Santiago Jocotepec San Andrés Teotilálpam

130

Papaloapam

Choápam

Chinantecs

130

Cañada

Cuicatlán

Cuicatecs

San Francisco Chapulapa

130

Cañada

Cuicatlán

Cuicatecs

Useful plant families Piperaceae, Polygalaceae, Rubiaceae, Selaginellaceae Apocynaceae, Arecaceae, Asteraceae, Caricaceae, Convolvulaceae, Euphorbiaceae, Fabaceae, Lauraceae, Malvaceae, Moraceae, Myrtaceae, Piperaceae, Rubiaceae, Sapotaceae, Solanaceae, Violaceae, Vitaceae Araceae, Arecaceae, Bromeliaceae, Costaceae, Cucurbitaceae, Cyatheaceae, Fabaceae, Hypericaceae, Juglandaceae, Lycopodiaceae, Melastomataceae, Piperaceae, Rubiaceae, Solanaceae, Zamiaceae Araceae Cactaceae, Fabaceae, Juglandaceae, Melastomataceae, Phytholacaceae, Piperaceae, Ticodendraceae Actinidiaceae, Adoxaceae, Altingiaceae, Annonaceae, Apiaceae, Asteraceae, Crassulaceae, Equisetaceae, Fabaceae, Fagaceae, Lauraceae, Malvaceae, Moraceae, Papaveraceae, Pinaceae, Piperaceae, (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

763

Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

Santa María Pápalo

121

Cañada

Cuicatlán

Cuicatecs

Santa María Tlalixtac

130

Cañada

Cuicatlán

Cuicatecs

Guevea de Humboldt

130

Istmo

Tehuantepec

Isthmus Zapotecs

Santa María Guienagati

130

Istmo

Tehuantepec

Isthmus Zapotecs

Useful plant families Plantaginaceae, Primulaceae, Rosaceae, Solanaceae, Vitaceae Asteraceae, Betulaceae, Cactaceae, Campanulaceae, Cleomaceae, Cucurbitaceae, Ericaceae, Fagaceae, Iridaceae, Lamiaceae, Pinaceae, Solanaceae Adoxaceae, Altingiaceae, Arecaceae, Asparagaceae, Asteraceae, Bromeliaceae, Caryophyllaceae, Cecropiaceae, Clusiaceae, Commelinaceae, Fabaceae, Fagaceae, Lamiaceae, Lauraceae, Lindsaeaceae, Melastomataceae, Myricaceae, Myrtaceae, Orchidaceae, Pinaceae, Polypodiaceae, Ranunculaceae, Rosaceae Altingiaceae, Araceae, Bromeliaceae, Cyclanthaceae, Fabaceae, Fagaceae, Lauraceae, Melastomataceae, Orchidaceae, Pinaceae, Piperaceae Araceae, Arecaceae, Asparagaceae, Asteraceae, Boraginaceae, Bromeliaceae, Calophyllaceae, (continued)

B. Rendo´n-Aguilar et al.

764 Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

Santiago Laollaga

130

Istmo

Tehuantepec

Isthmus Zapotecs

Santo Domingo Chihuitán

130

Istmo

Tehuantepec

Isthmus Zapotecs

Santo Domingo Petapa

130

Istmo

Juchitán

Isthmus Zapotecs

Useful plant families Clusiaceae, Cyatheaceae, Fagaceae, Lauraceae, Magnoliaceae, Melastomataceae, Orchidaceae, Pinaceae, Piperaceae, Plantaginaceae, Podocarpaceae, Sapotaceae, Zingiberaceae Anacardiaceae, Apocynaceae, Bromeliaceae, Burseraceae, Cactaceae, Iridaceae, Malvaceae, Meliaceae, Moraceae, Polygonaceae Anacardiaceae, Bixaceae, Cactaceae, Capparaceae, Carvicaceae, Ebenaceae, Euphorbiaceae, Fabaceae, Hernandiaceae, Malvaceae, Muntingiaceae, Primulaceae, Salicaceae, Solanaceae, Typhaceae, Zygophyllaceae Actinidiaceae, Altingiaceae, Araliaceae, Arecaceae, Aspleniaceae, Begoniaceae, Bromeliaceae, Commelinaceae, Fabaceae, Malvaceae, Melastomataceae, Pinaceae, Polypodiaceae, Primulaceae, Rubiaceae, Siparunaceae (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

765

Table 1 (continued)

Municipalities Eloxochitlán de Flores Magón

PTR 130

Culturalethnic region Cañada

Huautepec

130

Cañada

Teotitlán

Mazatecs

Huautla de Jiménez

130

Cañada

Teotitlán

Mazatecs

Political district Teotitlán

Ethnic groups Mazatecs

Useful plant families Adoxaceae, Altingiaceae, Annonaceae, Asteraceae, Bromeliaceae, Chloranthaceae, Ericaceae, Fagaceae, Hypericaceae, Lauraceae, Myrtaceae, Orchidaceae, Piperaceae, Plantaginaceae, Selaginellaceae, Solanaceae, Verbenaceae Altingiaceae, Asteraceae, Begoniaceae, Cupressaceae, Dicksoniaceae, Ericaceae, Orchidaceae, Phyllonomaceae, Pinaceae, Piperaceae, Pteridaceae, Rosaceae, Smilacaceae, Verbenaceae Actinidiaceae, Adoxaceae, Asteraceae, Begoniaceae, Betulaceae, Boraginaceae, Clethraceae, Ericaceae, Fagaceae, Gesneriaceae, Lauraceae, Malvaceae, Melastomataceae, Onagraceae, Orchidaceae, Papaveraceae, Piperaceae, Polypodiaceae (continued)

B. Rendo´n-Aguilar et al.

766 Table 1 (continued)

Municipalities Mazatlán Villa de Flores

PTR 130

Culturalethnic region Cañada

San Bartolomé Ayautla

130

Cañada

Teotitlán

Mazatecs

San Francisco Huehuetlán

0

Cañada

Teotitlán

Mazatecs

Political district Teotitlán

Ethnic groups Mazatecs

Useful plant families Amaranthaceae, Annonaceae, Apocynaceae, Asteraceae, Cupressaceae, Cyperaceae, Ebenaceae, Equisetaceae, Ericaceae, Euphorbiaceae, Fabaceae, Fagaceae, Martyniaceae, Meliaceae, Menyanthaceae, Muntingiaceae, Myrtaceae, Pinaceae, Piperaceae, Platanaceae, Poaceae, Rosaceae, Rubiaceae, Solanaceae Amaranthaceae, Apocynaceae, Araceae, Asteraceae, Cecropiaceae, Convolvulaceae, Cyatheaceae, Euphorbiaceae, Plantaginaceae, Solanaceae, Zamiaceae Altingiaceae, Asteraceae, Begoniaceae, Betulaceae, Campanulaceae, Caprifoliaceae, Dicksoniaceae, Dryopteridaceae, Equisetaceae, Ericaceae, Fabaceae, Fagaceae, Gesneriaceae, Lamiaceae, Lauraceae, Lycopodiaceae, Lythraceae, Orchidaceae, (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

767

Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

San José Tenango

130

Cañada

Teotitlán

Mazatecs

San Juan de los Cués

121

Cañada

Teotitlán

Mazatecs

San Lorenzo Cuanecuiltitla

0

Cañada

Teotitlán

Mazatecs

Useful plant families Pentaphylacaceae, Phytolaccaceae, Pinaceae, Piperaceae, Platanaceae, Polypodiaceae, Pteridaceae, Rosaceae, Rubiaceae, Verbenaceae Actinidiaceae, Apocynaceae, Araceae, Begoniaceae, Blechnaceae, Cecropiaceae, Chloranthaceae, Clusiaceae, Cupressaceae, Dryopteridaceae, Fagaceae, Heliconiaceae, Malvaceae, Marantaceae, Meliaceae, Myrtaceae, Orchidaceae, Papaveraceae, Piperaceae, Plantaginaceae, Polypodiaceae, Solanaceae, Urticaceae, Verbenaceae, Vitaceae Anacardiaceae, Burseraceae, Convolvulaceae, Euphorbiaceae, Fabaceae, Malvaceae, Rhamnaceae, Verbenaceae, Zamiaceae Adoxaceae, Altingiaceae, Asteraceae, Betulaceae, Dicksoniaceae, Equisetaceae, Ericaceae, Euphorbiaceae, Fagaceae, Lamiaceae, Lauraceae, (continued)

B. Rendo´n-Aguilar et al.

768 Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

San Lucas Zoquiápam

130

Cañada

Teotitlán

Mazatecs

San Mateo Yoloxochitlán

0

Cañada

Teotitlán

Mazatecs

Santa Ana Ateixtlahuaca

0

Cañada

Teotitlán

Mazatecs

Useful plant families Lentibulariaceae, Myrtaceae, Onagraceae, Papaveraceae, Pinaceae, Piperaceae, Rosaceae, Rubiaceae, Solanaceae Asparagaceae, Asteraceae, Betulaceae, Bromeliaceae, Cactaceae, Calceolariaceae, Clethraceae, Crassulaceae, Cucurbitaceae, Ericaceae, Euphorbiaceae, Fagaceae, Gesneriaceae, Lauraceae, Melastomataceae, Orchidaceae, Pentaphylacaceae, Rosaceae, Winteraceae Begoniaceae, Bromeliaceae, Cyatheaceae, Equisetaceae, Ericaceae, Fagaceae, and Orchidaceae Bromeliaceae, Clusiaceae, Crassulaceae, Ericaceae, Fagaceae, Lamiaceae, Orchidaceae, Polypodiaceae, Pteridaceae, Urticaceae (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

769

Table 1 (continued)

Municipalities Santa Cruz Acatepec

PTR 0

Culturalethnic region Cañada

Santa María Chilchotla Cuyamecalco Villa de Zaragoza

130

Cañada

Teotitlán

Mazatecs

130

Cañada

Cuicatlán

Mazatecs and Mixtecs

Natividad

130

Sierra Norte

Ixtlán

Mestizos

San Pedro Tapanatepec

132

Istmo

Juchitán

Mestizos

Political district Teotitlán

Ethnic groups Mazatecs

Useful plant families Adoxaceae, Altingiaceae, Asteraceae, Bromeliaceae, Dioscoreaceae, Fagaceae, Lamiaceae, Loranthaceae, Orchidaceae, Pinaceae, Piperaceae, Plantaginaceae, Santalaceae, Solanaceae Dracaenaceae Orchidaceae Amaranthaceae, Anacardiaceae, Arecaceae, Asparagaceae, Asteraceae, Betulaceae, Bromeliaceae, Burseraceae, Dicksoniaceae, Dryopteridaceae, Ericaceae, Fabaceae, Fagaceae, Lycopodiaceae, Malvaceae, Meliaceae, Myrtaceae, Pinaceae, Rutaceae, Sapindaceae, Sapotaceae, Scrophulariaceae, Verbenaceae, Winteraceae Asteraceae, Bromeliaceae, Equisetaceae, Ericaceae, Fagaceae, Pinaceae, Polypodiaceae Anacardiaceae, Bignoniaceae, Boraginaceae, Capparaceae, (continued)

B. Rendo´n-Aguilar et al.

770 Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

San Pedro Teutila

130

Cañada

Cuicatlán

Mestizos

Santiago Xiacuí

130

Sierra Norte

Ixtlán

Mestizos

Asunción Cacalotepec

130

Sierra Norte

Mixe

Mixes

Useful plant families Combretaceae, Fabaceae, Malvaceae, Phytolaccaceae, Polygonaceae Altingiaceae, Araceae, Asteraceae, Begoniaceae, Bromeliaceae, Convolvulaceae, Ebenaceae, Fabaceae, Fagaceae, Heliconiaceae, Lamiaceae, Lauraceae, Malvaceae, Melastomataceae, Menispermaceae, Moraceae, Orchidaceae, Papaveraceae, Pinaceae, Piperaceae, Sapindaceae, Siparunaceae, Solanaceae, Verbenaceae, Zamiaceae Asteraceae, Bromeliaceae, Equisetaceae, Ericaceae, Fabaceae, Fagaceae, Lamiaceae, Lauraceae, Onagraceae, Orchidaceae, Phyllonomaceae, Pinaceae, Piperaceae, Rosaceae, Solanaceae Asteraceae, Bromeliaceae, Fagaceae, Lentibulariaceae, Orchidaceae (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

771

Table 1 (continued)

Municipalities Mixistlán de la Reforma

PTR 130

Culturalethnic region Sierra Norte

San Juan Cotzocón

130

Sierra Norte

Mixe

Mixes

San Juan Guichicovi

130

Istmo

Juchitán

Mixes

San Juan Juquila Mixes

130

Sierra Sur

Yautepec

Mixes

Political district Mixe

Ethnic groups Mixes

Useful plant families Alstroemeriaceae, Asparagaceae, Asteraceae, Boraginaceae, Bromeliaceae, Burseraceae, Fabaceae, Fagaceae, Myrtaceae, Pinaceae, Pinaceae , Solanaceae, Vitaceae Annonaceae, Arecaceae, Bignoniaceae, Cyatheaceae, Euphorbiaceae, Fabaceae, Fagaceae, Heliconiaceae, Lauraceae, Malpighiaceae, Malvaceae, Melastomataceae, Myrtaceae, Pinaceae, Piperaceae, Sapotaceae, Siparunaceae, Vitaceae Annonaceae, Apocynaceae, Araceae, Arecaceae, Caricaceae, Costaceae, Fabaceae, Fagaceae, Malpighiaceae, Pinaceae, Polypodiaceae, Siparunaceae, Solanaceae Anacardiaceae, Asparagaceae, Asteraceae, Ericaceae, Fabaceae, Fagaceae, Lauraceae, Lentibulariaceae, Lythraceae, Myricaceae, Myrtaceae, Pinaceae, Rosaceae, Solanaceae (continued)

B. Rendo´n-Aguilar et al.

772 Table 1 (continued)

Municipalities San Juan Mazatlán

PTR 130

Culturalethnic region Sierra Norte

San Miguel Quetzaltepec

130

Sierra Norte

Mixe

Mixes

Santa María Alotepec

130

Sierra Norte

Mixe

Mixes

Santa María Tepantlali

130

Sierra Norte

Mixe

Mixes

Political district Mixe

Ethnic groups Mixes

Useful plant families Arecaceae, Bromeliaceae, Fagaceae, Orchidaceae, Pinaceae, Zingiberaceae Arecaceae, Bromeliaceae, Meliaceae, Orchidaceae, Piperaceae, Solanaceae, Zamiaceae Actinidiaceae, Adoxaceae, Altingiaceae, Asparagaceae, Begoniaceae, Betulaceae, Clethraceae, Ericaceae, Euphorbiaceae, Fabaceae, Fagaceae, Malvaceae, Melastomataceae, Pinaceae, Plantaginaceae, Polygonaceae, Polypodiaceae, Primulaceae, Pteridaceae, Selaginellaceae, Siparunaceae, Smilaceceae, Zamiaceae Actinidiaceae, Altingiaceae, Asteraceae, Betulaceae, Blechnaceae, Bromeliaceae, Cannaceae, Clethraceae, Dennstaedtiaceae, Equisetaceae, Euphorbiaceae, Fabaceae, Fagaceae, Lentibulariaceae, Lindsaeaceae, (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

773

Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

Santa María Tlahuitoltepec

130

Sierra Norte

Mixe

Mixes

Santiago Atitlán

130

Sierra Norte

Mixe

Mixes

Useful plant families Lycopodiaceae, Lythraceae, Malvaceae, Orchidaceae, Phyllonomaceae, Pinaceae, Polypodiaceae, Pteridaceae, Rosaceae, Solanaceae, Zamiaceae Adoxaceae, Anacardiaceae, Asteraceae, Betulaceae, Equisetaceae, Ericaceae, Fabaceae, Lamiaceae, Lythraceae, Malvaceae, Onagraceae, Oxalidaceae, Pentaphylacaceae, Phyllonomaceae, Pinaceae, Primulaceae, Pteridaceae, Rosaceae, Solanaceae Altingiaceae, Begoniaceae, Cannaceae, Clethraceae, Cucurbitaceae, Equisetaceae, Ericaceae, Fabaceae, Fagaceae, Malvaceae, Orchidaceae, Orobanchaceae, Polygalaceae, Rosaceae, Solanaceae, Verbenaceae, Vitaceae (continued)

B. Rendo´n-Aguilar et al.

774 Table 1 (continued)

Municipalities Santiago Ixcuintepec

PTR 130

Culturalethnic region Sierra Norte

Santiago Zacatepec

130

Sierra Norte

Mixe

Mixes

Santo Domingo Tepuxtepec

130

Sierra Norte

Mixe

Mixes

Political district Mixe

Ethnic groups Mixes

Useful plant families Apiaceae, Bignoniaceae, Cecropiaceae, Ebenaceae, Equisetaceae, Fabaceae, Fagaceae, Lauraceae, Malvaceae, Marantheaceae, Muntingiaceae, Piperaceae, Rubiaceae, Sapotaceae, Solanaceae Altingiaceae, Annonaceae, Cannabaceae, Cannaceae, Cucurbitaceae, Dennstaedtiaceae, Equisetaceae, Malvaceae, Moraceae, Piperaceae, Rosaceae, Rubiaceae Anacardiaceae, Apocynaceae, Asparagaceae, Asteraceae, Betulaceae, Bromeliaceae, Burseraceae, Dennstaedtiaceae, Equisetaceae, Ericaceae, Euphorbiaceae, Fagaceae, Lamiaceae, Lauraceae, Malvaceae, Myrtaceae, Orchidaceae, Pinaceae, Poaceae, Rhamnaceae, Rosaceae (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

775

Table 1 (continued)

Municipalities Tamazulápam del Espíritu Santo

PTR 130

Culturalethnic region Sierra Norte

Totontepec Villa de Morelos

130

Sierra Norte

Mixe

Mixes

San Juan Coatzóspam

130

Cañada

Teotitlán

Mixtecs

San Miguel Santa Flor

130

Cañada

Cuicatlán

Mixtecs

Political district Mixe

Ethnic groups Mixes

Useful plant families Anacardiaceae, Asteraceae, Betulaceae, Bromeliaceae, Campanulaceae, Crassulaceae, Cupressaceae, Ericaceae, Euphorbiaceae, Fagaceae, Iridaceae, Pinaceae, Poaceae, Polygalaceae, Rosaceae, Sapindaceae, Scrophulariaceae, Verbenaceae Altingiaceae, Apiaceae, Arecaceae, Cannabaceae, Cannaceae, Cyatheaceae, Dioscoreaceae, Equisetaceae, Ericaceae, Fagaceae, Gentianaceae, Lauraceae, Malvaceae, Pinaceae, Plantaginaceae, Siparunaceae, Solanaceae, Vitaceae Asparagaceae, Asteraceae, Cactaceae, Ebenaceae, Fagaceae, Orchidaceae, Pinaceae, Zamiaceae Adoxaceae, Apocynaceae, Asteraceae, Betulaceae, Bromeliaceae, Cactaceae, Crassulaceae, Dryopteridaceae, Ericaceae, Fagaceae, Iridaceae, Lamiaceae, (continued)

B. Rendo´n-Aguilar et al.

776 Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

Santa María Teopoxco

130

Cañada

Teotitlán

Nahuas

Santiago Texcalcingo

130

Cañada

Teotitlán

Nahuas

Useful plant families Onagraceae, Pinaceae, Podocarpaceae, Rosaceae, Scrophulariaceae, Smilacaceae Apiaceae, Asparagaceae, Betulaceae, Bromeliaceae, Clethraceae, Convolvulaceae, Dennstaedtiaceae, Ericaceae, Euphorbiaceae, Fabaceae, Fagaceae, Lauraceae, Malvaceae, Onagraceae, Orchidaceae, Pentaphylacaceae, Phytolaccaceae, Piperaceae, Rosaceae, Solanaceae Adoxaceae, Asteraceae, Betulaceae, Clethraceae, Convolvulaceae, Equisetaceae, Ericaceae, Fagaceae, Iridaceae, Lycopodiaceae, Melastomataceae, Onagraceae, Pentaphylacaceae, Phytolaccaceae, Piperaceae, Plantaginaceae, Platanaceae, Poaceae, Polygonaceae, Rosaceae, Rubiaceae (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

777

Table 1 (continued)

Municipalities Abejones

PTR 130

Culturalethnic region Sierra Norte

Nuevo Zoquiápam

130

Sierra Norte

Ixtlán

Northern Zapotecs

San Cristóbal Lachirioag

130

Sierra Norte

Villa Alta

Northern Zapotecs

Political district Ixtlán

Ethnic groups Northern Zapotecs

Useful plant families Asteraceae, Burseraceae, Ericaceae, Euphorbiaceae, Fabaceae, Fagaceae, Lamiaceae, Malvaceae, Pinaceae, Piperaceae, Rosaceae, Sapindaceae, Solanaceae Asteraceae, Berberidaceae, Brassicaceae, Bromeliaceae, Crassulaceae, Dryopteridaceae, Equisetaceae, Ericaceae, Fabaceae, Fagaceae, Lamiaceae, Lauraceae, Malvaceae, Orchidaceae, Pinaceae, Rosaceae Anacardiaceae, Asteraceae, Brassicaceae, Bromeliaceae, Cannabaceae, Clethraceae, Euphorbiaceae, Fabaceae, Fagaceae, Lamiaceae, Malvaceae, Melastomataceae, Myrtaceae, Pinaceae, Piperaceae, Polypodiaceae, Primulaceae, Pteridaceae, Rubiaceae, Rutaceae, Sapindaceae, Solanaceae, Verbenaceae (continued)

B. Rendo´n-Aguilar et al.

778 Table 1 (continued)

Municipalities San Francisco Cajonos

PTR 0

Culturalethnic region Sierra Norte

San Ildefonso Villa Alta

130

Sierra Norte

Villa Alta

Northern Zapotecs

San Juan Atepec

130

Sierra Norte

Ixtlán

Northern Zapotecs

San Juan Comaltepec

130

Papaloapam

Choápam

Northern Zapotecs

Political district Villa Alta

Ethnic groups Northern Zapotecs

Useful plant families Betulaceae, Campanulaceae, Ericaceae, Fagaceae, Lamiaceae, Lauraceae, Lycopodiaceae, Orchidaceae, Phytolaccaceae, Pinaceae, Rosaceae, Solanaceae Altingiaceae, Asparagaceae, Asteraceae, Betulaceae, Bromeliaceae, Convolvulaceae, Ericaceae, Euphorbiaceae, Fabaceae, Fagaceae, Loranthaceae, Lycopodiaceae, Lythraceae, Pinaceae, Piperaceae, Rosaceae Anacardiaceae, Asteraceae, Brassicaceae, Burseraceae, Cactaceae, Fabaceae, Fagaceae, Lauraceae, Myrtaceae, Pinaceae, Rosaceae, Sapindaceae Actinidiaceae, Araceae, Arecaceae, Asteraceae, Begoniaceae, Clethraceae, Convolvulaceae, Euphorbiaceae, Fabaceae, Gleicheniaceae, Lycopodiaceae, Malphigiaceae, Malvaceae, Melastomataceae, Pinaceae, Plantaginaceae, Poaceae, Primulaceae, Pteridaceae, (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

779

Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

San Juan Tabaá

130

Sierra Norte

Villa Alta

Northern Zapotecs

San Juan Yaeé

130

Sierra Norte

Villa Alta

Northern Zapotecs

San Miguel Aloápam

130

Sierra Norte

Ixtlán

Northern Zapotecs

San Miguel del Río

130

Sierra Norte

Ixtlán

Northern Zapotecs

Useful plant families Rhamnaceae, Sellaginellaceae, Solanaceae Asteraceae, Betulaceae, Bromeliaceae, Dryopteridaceae, Ericaceae, Fabaceae, Fagaceae, Lauraceae, Lycopodiaceae, Malvaceae, Meliaceae, Myrtaceae, Phytolaccaceae, Pinaceae, Poaceae, Rhamnaceae Altingiaceae, Asteraceae, Cactaceae, Cannabaceae, Caprifoliaceae, Clethraceae, Convolvulaceae, Ericaceae, Fagaceae, Gentianaceae, Melastomataceae, Moraceae, Myrtaceae, Onagraceae, Pinaceae, Piperaceae, Polypodiaceae, Pteridaceae, Rubiaceae Asteraceae, Betulaceae, Brassicaceae, Dryopteridaceae, Ericaceae, Fabaceae, Fagaceae, Lamiaceae, Lauraceae, Oxalidaceae, Pinaceae, Piperaceae Anacardiaceae, Asteraceae, Basellaceae, Bromeliaceae, Ericaceae, Fabaceae, Fagaceae, Orchidaceae, Pinaceae, Rosaceae, Sapindaceae (continued)

B. Rendo´n-Aguilar et al.

780 Table 1 (continued)

Municipalities San Pablo Macuiltianguis

PTR 130

Culturalethnic region Sierra Norte

San Pablo Yaganiza

0

Sierra Norte

Villa Alta

Northern Zapotecs

San Pedro Cajonos

130

Sierra Norte

Villa Alta

Northern Zapotecs

Santa Ana Yareni

130

Sierra Norte

Ixtlán

Northern Zapotecs

Santa María Jaltianguis

130

Sierra Norte

Ixtlán

Northern Zapotecs

Political district Ixtlán

Ethnic groups Northern Zapotecs

Useful plant families Apocynaceae, Asteraceae, Betulaceae, Burseraceae, Cucurbitaceae, Fabaceae, Fagaceae, Lauraceae, Onagraceae, Pinaceae, Rhamnaceae, Sapindaceae, Scrophulariaceae Asparagaceae, Asteraceae, Bromeliaceae, Burseraceae, Ericaceae, Fagaceae, Pinaceae, and Primulaceae Asteraceae, Betulaceae, Cactaceae, Crassulaceae, Ericaceae, Fagaceae, Gentianaceae, Lamiaceae, Lauraceae, Pinaceae, Solanaceae Anacardiaceae, Asparagaceae, Asteraceae, Burseraceae, Convolvulaceae, Cupressaceae, Ericaceae, Fabaceae, Fagaceae, Grossulariaceae, Papaveraceae, Pinaceae, Rubiaceae, Sapindaceae, Solanaceae, Zygophyllaceae Asteraceae, Dryopteridaceae, Equisetaceae, Ericaceae, Fagaceae, Lamiaceae, Lauraceae, Pinaceae, Piperaceae, Rosaceae (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

781

Table 1 (continued)

Municipalities Santa María Yalina

PTR 130

Culturalethnic region Sierra Norte

Santiago Camotlán

130

Sierra Norte

Villa Alta

Northern Zapotecs

Santiago Choápam

130

Papaloapam

Choápam

Northern Zapotecs

Santo Domingo Roayaga

130

Sierra Norte

Villa Alta

Northern Zapotecs

Political district Villa Alta

Ethnic groups Northern Zapotecs

Useful plant families Asparagaceae, Asteraceae, Betulaceae, Bromeliaceae, Crassulaceae, Cucurbitaceae, Ericaceae, Fabaceae, Fagaceae, Lamiaceae, Lauraceae, Lycopodiaceae, Melastomataceae, Myrtaceae, Onagraceae, Pinaceae, Rosaceae, Solanaceae Asteraceae, Calceolariaceae, Commelinaceae, Convolvulaceae, Dennstaedtiaceae, Fabaceae, Malvaceae, Melastomataceae, Myrtaceae, Phytolaccaceae, Pinaceae, Piperaceae, Plantaginaceae, Primulaceae, Pteridaceae, Rubiaceae, Solanaceae, Verbenaceae Altingiaceae, Araceae, Asteraceae, Cecropiaceae, Convolvulaceae, Cyatheaceae, Lindsaeaceae, Lycopodiaceae, Malvaceae, Melastomataceae, Myricaceae, Phytolaccaceae, Rubiaceae, Vitaceae Actinidiaceae, Altingiaceae, Asteraceae, Betulaceae, Cannabaceae, Dicksoniaceae, (continued)

B. Rendo´n-Aguilar et al.

782 Table 1 (continued)

Municipalities

PTR

Culturalethnic region

Political district

Ethnic groups

Santo Domingo Xagacía

130

Sierra Norte

Villa Alta

Northern Zapotecs

Teococuilco de Marcos Pérez

130

Sierra Norte

Ixtlán

Northern Zapotecs

Villa Talea de Castro

130

Sierra Norte

Villa Alta

Northern Zapotecs

Useful plant families Fabaceae, Fagaceae, Malphigiaceae, Myrtaceae, Phyllonomaceae, Pinaceae, Piperaceae, Polypodiaceae, Pteridaceae, Rosaceae, Rubiaceae, Solanaceae Asparagaceae, Betulaceae, Cactaceae, Cleomaceae, Ericaceae, Fagaceae, Lamiaceae, Lauraceae, Pinaceae Anacardiaceae, Asteraceae, Bignoneaceae, Boraginaceae, Brassicaceae, Bromeliaceae, Ericaceae, Fabaceae, Fagaceae, Loranthaceae, Papaveraceae, Phytolaccaceae, Pinaceae, Piperaceae, Pteridaceae, Rosaceae, Salicaceae, Scrophulariaceae Asteraceae, Brassicaceae, Cannabaceae, Cecropiaceae, Equisetaceae, Fabaceae, Malvaceae, Orobanchaceae, Oxalidaceae, Primulaceae, Sapindaceae, Solanaceae, Verbenaceae (continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . .

783

Table 1 (continued)

Municipalities San Miguel Chimalapa

PTR 132

Culturalethnic region Istmo

Santa María Chimalapa

132

Istmo

Political district Juchitán

Ethnic groups Zoques

Juchitán

Zoques

Useful plant families Altingiaceae, Apocynaceae, Araceae, Arecaceae, Asparagaceae, Asteraceae, Bignoniaceae, Bromeliaceae, Burseraceae, Cactaceae, Chloranthaceae, Clusiaceae, Cyatheaceae, Dryopteridaceae, Fabaceae, Fagaceae, Lycopodiaceae, Malpighiaceae, Myrtaceae, Piperaceae, Podocarpaceae, Polygonaceae, Sapindaceae, Solanaceae, Woodsiaceae, Zamiaceae Amaranthaceae, Apocynaceae, Araceae, Arecaceae, Asteraceae, Bromeliaceae, Burseraceae, Calophyllaceae, Cecropiaceae, Clusiaceae, Costaceae, Cyatheaceae, Cyclanthaceae, Elaeocarpaceae, Fabaceae, Malvaceae, Melastomataceae, Meliaceae, Moraceae, Orchidaceae, Pinaceae, Piperaceae, Rubiaceae, Vitaceae, Zamiaceae

784

B. Rendo´n-Aguilar et al.

form, and characteristics of reproductive structures), common name(s), use(s), plant structure(s) used, and descriptions of their use.

Species Determination of Plant Specimens Herbarium specimens were identified with the help of specialized literature and taxonomists working in Mexican herbaria (i.e., FCME, IEB, MEXU, UAMIZ, and XAL) (acronyms according to (Thiers 2016)). The main set of specimens was assembled and labeled. Specimens are being distributed and deposited in the Herbario Metropolitano (UAMIZ), Universidad Autónoma Metropolitana Iztapalapa, Ciudad de México (UAM-IZ). Duplicates are in process to be deposited in the Herbario Nacional de Mexico (MEXU) of the Universidad Nacional Autónoma de México UNAM and the herbarium OAX, at the Centro Interdisciplinario de Investigación para el Desarrollo Integral Regional Unidad Oaxaca (CIIDIR-Oaxaca). Nomenclature, as well as authors, where reviewed following the Royal Botanical Gardens, Kew database (PoWO 2019).

Data Analysis Field information and photographs of each specimen were integrated into the database BIÓTICA© 5.0.3 (CONABIO 2012) and the images bank of CONABIO (2018). We extracted and systematized this information to analyze it from the following perspectives: (a) Geographical and ecological aspects of ethnofloristic approach sampling. Elevational range of each specimen, vegetation type. (b) Taxonomic aspects of useful species. This includes the number of species, and the families to which they belong, best represented according to their uses; endemism, status in NOM-059-SEMARNAT-2010 (SEMARNAT 2010), and IUCN Red List of Threatened Species (IUCN 2019). (c) Traditional knowledge. Use categories; Number of uses; Name(s) in a local language (i.e., traditional nomenclature, (De Ávila 2004)/Spanish languages. (d) Distribution of traditional knowledge among ethnic groups. Total number of useful families and species was obtained for each ethnic group, considering number of municipalities inhabited by each one.

Results Geographical and Ecological Aspects of Ethnofloristic Approach Sampling Within the 84 municipalities selected, we collected specimens in 327 localities. Elevations where plants were collected range from 75 to 3,334 m. Lowlands were more common in the slope of Gulf of Mexico (e.g., the Chinantla region, and the

Ethnobotanical Science in the Clouds: Useful Plants of. . .

785

municipality of Ayotzintepec), and in the slope of Pac\ific Coast (e.g., the Isthmus region and the municipalities of Santo Domingo Chihuitán, and San Pedro Tapanatepec). Highland localities were the most widely distributed, and were found in the mountains in the Northeast of the state (e.g., the Sierra Norte region, municipalities of Abejones, Nuevo Zoquiápam, San Pedro Cajonos, Santa María Jaltianguis, and Santo Domingo Xagacía). Plant specimens were collected in 15 vegetation types and ecotones. Secondary vegetation was also sampled, which is very frequent throughout the forests, because different natural and/or human disturbances occurred decades ago. Distribution of vegetation exhibit very complex and dynamic patterns. Pine forest (PF) and oak forest (OF) exhibit a wide distribution, from warm areas at 164 m a.s.l., to temperate areas reaching elevations of 3,000 m. River-margin vegetation (RMV) also included warm and temperate areas, from 90 to 2,541 m a.s.l., but when this vegetation type is dominated by Platanus mexicana Moric., the elevation range was restricted between 988 and 2,541 m a.s.l. Some vegetation types have a very restricted distribution in low altitudes, like submersed aquatic vegetation (SAV), and tropical thorn forest (TTF) (under 500 m a.s.l.) or high altitudes (above 2,500 m a.s.l.), like Oak dominated chaparral (ODC) (Fig. 2). The PTR 130 (Sierras del Norte de Oaxaca-Mixe) recorded the greatest number of collected plants, which is related to the greatest area of its territory (19,382 km2), the

Fig. 2 Elevational range of each vegetation type recorded in the present study. Acronyms: TTF: tropical thorn forest; SRV: swamp-reed vegetation; SAV: submersed and floating aquatic vegetation; TEF: tropical evergreen forest; SHR: shrubland; TSF: tropical semi-evergreen forest; TDF: tropical dry deciduous forest; RMV: river-margin vegetation; ODC: oak-dominated chaparral; MCF: montane cloud forest; GRS: grassland; PF: pine forest; OF: oak forest; FF: fir forest

786

B. Rendo´n-Aguilar et al.

greatest number of municipalities belonging to it (165), and the number of ethnic groups that inhabit it (e.g., Chinantec, Mazatec, Mixe, Mixtec, Nahua, and Zapotec).

Taxonomic Aspects of Useful Plants Approximately more than 2,340 specimens were collected, distributed as follows: 118 Lycophyta and Polypodiophyta (15 families); 142 Coniferophyta (4 families), and 2,089 Magnoliophyta (120 families). A total of 139 families, 441 genera, 804 species, 7 subspecies, and 18 varieties were determined. Comparing with the total taxa reported for Oaxaca (García-Mendoza and Meave 2012), in the present study we collected 53% of families (total 261), 22.6% of genera (total 1952), and 9% of species (total 1952). The 45% of the useful plants (363 species) are included in 12 families: Asteraceae (76), Fabaceae (49), Bromeliaceae and Fagaceae (34 species each), Orchidaceae (31), Malvaceae (25), Solanaceae (24), Piperaceae (23), Melastomataceae (22), Pinaceae, Rosaceae, and Rubiaceae (15 species each) (Fig. 3). The genus Quercus was the richest one with 34 species, followed by Tillandsia L. (22 species), and Pinus L. (14). Most of the useful species were found in four vegetation types: montane cloud forests (MCF), oak forests (OF), pine forests (PF), and tropical semi-evergreen forest (TSF) (Fig. 4). One hundred and one species are catalogued in a risk category (Table 2). Particularly, 75 taxa are included in the IUCN Red List of Threatened Species (IUCN 2019), and 41 species in the NOM-059-SEMARNAT-2010 (SEMARNAT 2010).

Fig. 3 Plant families with highest number of useful species in northeastern Oaxaca

Ethnobotanical Science in the Clouds: Useful Plants of. . .

787

Fig. 4 Number of useful species recorded in each vegetation type in northeastern Oaxaca. Acronyms: MCF: montane cloud forest; OF: oak forest; PF: pine forest; TSF: tropical semievergreen forest; ECT: ecotones; TEF: tropical evergreen forest; TDF: tropical dry deciduous forest; RMV: river-margin vegetation; TTF: tropical thorn forest; SHR: shrubland; FF: fir forest; ODC: oak-dominated chaparral; SAV: submersed and floating aquatic vegetation; SRV: swampreed vegetation; GRS: grassland

Traditional Knowledge (a) Use categories. We recorded 11 use categories. The most important, based on the number of species reported, are medicinal, edible, and ornamental (Fig. 5). For every use category, there are families that are more important. For instance, Asteraceae has the largest number of medicinal species. For edible plants species in the family Fabaceae are the most significant. Most Ornamental taxa belong to Bromeliaceae and Orchidaceae. (b) Number of uses. Almost 60% of the useful species have a single use, considering as exclusive species, because they are used to satisfy a specific need. Contrarily, there are few species that have more than five uses but a variable pattern of uses was detected throughout the study area (Table 2). Such is the case of Alnus acuminata Kunth (“Aile”), a species with many uses throughout the study area (e.g., ceremonial, construction, edible, environmental, firewood, handcraft, medicinal, timber) but they differ among ethnic groups. Thus, the Cuicatec people use it to make handcrafts, whereas the Nahua people use it as firewood. Another example corresponds to Arbutus xalapensis Kunth, which is used for construction, edible, firewood, fodder, handcraft, medicinal, ornamental, only in the PTR 130, among the Mazatec, Mixe, and Northern Zapotec people; whereas

Cyatheaceae

Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Cyatheaceae

Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Cyatheaceae

Cyatheaceae

Cyatheaceae

Lycophyta and Polypodiophyta

Cyatheaceae

Blechnaceae

Blechnaceae

Family Aspleniaceae

Lycophyta and Polypodiophyta

Division Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Cyathea myosuroides (Liebm.) R.M. Tryon Cyathea schiedeana Domin Sphaeropteris horrida (Liebm.) R.M. Tryon

Scientific name Asplenium sessilifolium Desv. Blechnum appendiculatum Willd. Woodwardia spinulosa M. Martens & Galeotii Alsophila firma (Baker) D.S. Conant Cyathea divergens var. tuerckheimii (Maxon) R.M. Tryon Cyathea fulva Fée var. tuerckheimii

Infraspecies

MCF, TSF

MCF

PF

MCF, OF, TSF

OF

TEF, TSF

RMV

TSF

Vegetation types MCF

Maz, Mix, NZ

IZ, NZ

Mix

Chi, Maz, Zoq

Maz

Maz, Zoq

Mix

Maz

Ethnic groupS IZ

Construction, Medicinal, Ornamental

Construction

Construction, Handcraft, Medicinla, Ornamental Construction

Construction, Medicinal

Medicinal

Ceremonial

Medicinal

Use Categories Ornamental

- / Special protection - / Special protection

- / Special protection

- / Special protection - / Special protection

Risk status IUCN / SEMARNAT

Table 2 Useful species collected. The vegetation types where useful species are distributed, the ethnic groups that manage them, and the use categories and species catalogued in a risk category are shown. Vegetation types: GRS, grassland; MCF, montane cloud forest; OF, oak forest; PF, pine forest; RMV, rivermargin vegetation; TDF, tropical dry deciduous forest; TEF, tropical evergreen forest; TTF, thorn forest; TSF, tropical semi-evergreen forest; and ECT, ecotones. Ethnic groups: Chi, Chinantec; Cui, Cuicatec; IZ, Isthmus Zapotec; Maz, Mazatec; Mes, Mestizo; Mix, Mixe; Mxt, Mixtec; Nah, Nahua; NZ, Northern Zapotec; and Zoq, Zoque

788 B. Rendo´n-Aguilar et al.

Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta

Dennstaedtiaceae Pteridium aquilinum var. feei (W. Schaffn.) Maxon Dennstaedtiaceae Pteridium arachnoideum (Kaulf.) Maxon Dennstaedtiaceae Pteridium caudatum Maxon Dicksoniaceae Lophosoria quadripinnata (J.F. Gmel.) C. Chr. Dryopteridaceae Dryopteris wallichiana (Spreng.) Hyl. Dryopteridaceae Elaphoglossum petiolatum Urb. Dryopteridaceae Elaphoglossum sartorii (Liebm.) Mickel Dryopteridaceae Elaphoglossum vestitum (Schltdl. & Cham.) Schott Dryopteridaceae Polybotrya polybotryoides (Baker) Christ Equisetaceae Equisetum hyemale var. affine (Engelm.) A.A. Eaton Equisetaceae Equisetum myriochaetum Schltdl. & Cham. var. affine

var. feei

MCF, OF, PF, RMV,

OF

OF, TEF

Cui, Mas, Mes, Mix, Nah, NZ

Mes, Mix

Zoq

Maz

Maz

TSF

TSF

NZ

PF

Maz, Mxt, NZ

Maz, NZ

MCF, OF, PF MCF, PF

Mix

Mix, NZ

Mix, Nah

PF

OF, TEF

MCF, OF

Fodder, Medicinal, Veterinary medicine

Medicinal

Ornamental

Medicinal

Medicinal

Ceremonial

Ceremonial, Construction, Handcraft, Medicinal Handcraft, Ornamental

Handcraft

Handcraft, Medicinal

Handcraft, Medicinal

(continued)

Least concern / -

Ethnobotanical Science in the Clouds: Useful Plants of. . . 789

Lindsaeaceae

Lycopodiaceae

Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Gleicheniaceae

Lycophyta and Polypodiophyta

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Lycopodiaceae

Lycopodiaceae

Family

Division

Table 2 (continued)

Gleichenella pectinata (Willd.) Ching Odontosoria schlechtendalii (C. Presl) C. Chr. Lycopodiella cernua (L.) Pic. Serm. Lycopodium clavatum L. Lycopodium thyoides Willd. Campyloneurum angustifolium Fée Campyloneurum tenuipes Maxon Niphidium crassifolium (L.) Lellinger Pecluma alfredii (Rosenst.) M.G. Price Phlebodium areolatum (Willd.) J. Sm.

Scientific name

Infraspecies

MCF, OF, PF

MCF

TSF

TSF

TSF

MCF

MCF, PF

MCF, TSF

Cui, Maz, Mes, Mix, NZ

Mix

IZ

Maz

Maz

Maz, Mix, NZ, Zoq Maz, Nah

Chi, NZ

Chi, Cui, Mix, NZ

NZ

TDF, TEF, TSF MCF

MCF, OF

Ethnic groupS

Vegetation types

Medicinal, Ornamental

Ornamental

Ornamental

Medicinal

Medicinal

Handcraft, Ornamental

Handcraft, Ornamental

Ornamental

Ceremonial, Handcraft, Medicinal, Ornamental

Handcraft

Use Categories

Risk status IUCN / SEMARNAT

790 B. Rendo´n-Aguilar et al.

Lycophyta and Polypodiophyta

Pteridaceae

Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Pteridaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Polypodiaceae

Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Mildella intramarginalis (Kaulf. ex Link) Christenh. Pityrogramma calomelanos (L.) Link

Polypodium collinsii Maxon Polypodium lepidotrichum Maxon Polypodium martensii Mett. Polypodium plesiosorum Kunze Polypodium polypodioides (L.) Watt var. polypodioides Serpocaulon triseriale (Sw.) A.R. Sm. Adiantum andicola Liebm. Adiantum concinnum Humb. & Bonpl. ex Willd. Cheilanthes notholaenoides Maxon ex Weath. Hemionitis palmata L. var. polypodioides

OF, PF, TSF

PF

TEF

ECT (MCF-PF)

MCF, OF, PF PF

PF

TSF

MCF

OF

OF

PF

NZ

Mix

Chi

Maz

Mix

Maz, Mix

NZ

Maz, Mix

Maz

Mes

Mes

Mix

Ceremonial, Medicinal

Handcraft

Ceremonial

Ceremonial

Ceremonial, Handcraft, Medicinal, Ornamental Handcraft

Ornamental

Medicinal, Ornamental

Ornamental

Ornamental

Ornamental

Ornamental

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 791

Coniferophyta

Coniferophyta

Coniferophyta

Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta Coniferophyta

Division Lycophyta and Polypodiophyta Lycophyta and Polypodiophyta

Scientific name Pityrogramma ebenea (L.) Proctor Pteridaceae Pityrogramma trifoliata (L.) R.M. Tryon Selaginellaceae Selaginella illecebrosa Alston Selaginellaceae Selaginella pallescens (C. Presl) Spring Selaginellaceae Selaginella porphyrospora A. Br. Selaginellaceae Selaginella stellata Spring Thelypteridaceae Thelypteris imbricata (Liebm.) C.F. Reed Woodsiaceae Diplazium ternatum Liebm. Cupressaceae Cupressus benthamii (Endl.) Bartel Cupressaceae Cupressus lusitanica Mill. Cupressaceae Juniperus flaccida Schltdl. Cupressaceae Taxodium mucronatum Ten.

Family Pteridaceae

Table 2 (continued)

Infraspecies

Mix Chi, Mix, NZ Chi Zoq Maz

MCF MCF

TEF OF

PF

PF, OF

PF

Maz

Mix, NZ

Maz

Maz

TSF

TEF

Chi

NZ

SHR

TEF

Ethnic groupS Maz, NZ

Vegetation types OF, RMV

Ornamental

Construction, Timber

Timber

Construction, Timber

Ornamental

Medicinal

Medicinal, Ornamental

Ornamental

Medicinal

Ornamental

Medicinal

Use Categories Ceremonial, Medicinal

Least concern / Least concern / -

- / Danger of extinction

Risk status IUCN / SEMARNAT

792 B. Rendo´n-Aguilar et al.

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Pinaceae

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Pinus lawsonii Roezl ex Gordon var. lawsonii Pinus leiophylla Schiede ex Schltdl. & Cham. var. leiophylla Pinus maximinoi H.E. Moore Pinus montezumae Lamb. Pinus oocarpa Schiede ex Schltdl.

Pinus douglasiana Martínez Pinus hartwegii Lindl.

Pinus devoniana Lindl.

Pinus chiapensis (Martínez) Andresen

Pinus ayacahuite Ehrenb. ex Schltdl.

Abies hickelii Flous & Gaussen

var. leiophylla

var. lawsonii

MCF, OF, PF

ECT (MCF-OF) MCF

PF

OF

FF, PF

PF

PF

MCF, OF, PF

MCF, OF, PF

FF, MCF, PF

IZ, Maz, Mix, Mxt, NZ, Zoq

NZ

Mes

Mix, NZ

Chi, Mix, NZ, Zoq Mes, Mix, NZ

NZ

Chi, Cui, IZ, Mix, NZ Mix, NZ

Cui, Maz, Mxt, NZ

NZ

Construction, Firewood, Handcraft, Medicinal, Timber

Construction, Medicinal, Timber, Veterinary medicine Construction, Environmental, Firewood, Timber Ceremonial, Handcraft, Timber Firewood, Handcraft

Construction, Timber

Ceremonial, Construction, Firewood, Timber Timber

Construction, Environmental, Firewood, Medicinal, Ornamental, Timber Construction, Firewood, Timber

Construction, Ornamental, Timber

(continued)

Least concern /Least concern /Least concern /-

Least concern /-

Least concern /Least concern /Least concern /-

Endangered / Special protection Least concern / -

Endangered / Danger of extinction Least concern / -

Ethnobotanical Science in the Clouds: Useful Plants of. . . 793

Family Pinaceae

Pinaceae

Pinaceae

Pinaceae

Podocarpaceae

Podocarpaceae

Zamiaceae

Zamiaceae

Zamiaceae

Division Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Table 2 (continued)

Ceratozamia chimalapensis PérezFarr. & Vovides Ceratozamia mixeorum Chemnick, T.J. Greg. & SalasMor. Ceratozamia norstogii D.W. Stev.

Pinus teocote Schiede ex Schltdl. & Cham. Podocarpus aff. guatemalensis Standl. Podocarpus matudae Lundell

Pinus pseudostrobus (Lindl.) Shaw

Scientific name Pinus patula var. longipedunculata Loock ex Martínez Pinus pringlei Shaw var. apulcensis

Infraspecies var. longipedunculata

MCF

MCF, OF

MCF

MCF

MCF

PF

MCF, OF, PF

PF

Vegetation types MCF, OF, PF

Zoq

Mix

Zoq

Mxt

IZ, Zoq

NZ

Maz, Mes, Mix, NZ

NZ

Ethnic groupS Cui, Maz, Mes, NZ

Handcraft, Ornamental

Ceremonial, Edible

Environmental, Ornamental

Construction

Firewood, Timber

Ceremonial, Construction, Firewood, Medicinal, Timber Construction, Timber

Construction, Timber

Use Categories Construction, Medicinal, Timber

Endangered / Danger of extinction

Endangered / Danger of extinction

Least concern /Least concern /Vulnerable / Special protection - / Danger of extinction

Least concern /Least concern /-

Risk status IUCN / SEMARNAT Least concern /-

794 B. Rendo´n-Aguilar et al.

Zamiaceae

Zamiaceae

Zamiaceae

Zamiaceae

Zamiaceae

Zamiaceae

Acanthaceae

Acanthaceae

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Coniferophyta

Magnoliophyta

Magnoliophyta

TEF

Justicia fimbriata (Nees) V.A.W. Graham Odontonema callistachyum (Schltdl. & Cham.) Kuntze TEF

OF

TSF

TSF

OF

TSF

MCF, TEF

Zamia spartea A.DC.

Zamia purpurea Vovides, J.D. Rees & Vázq. Torres

Zamia paucijuga Wieland

Dioon purpusii Rose

Ceratozamia robusta Miq. Ceratozamia zoquorum Pérez-Farr., Vovides, Iglesias

Chi

Chi

Zoq

Chi

Mxt

Maz

Chi, Mes, Mix Maz

Ornamental

Ornamental

Ornamental

Medicinal

Ornamental

Ornamental

Environmental, Handcraft, Ornamental Handcraft

(continued)

Endangered / Threatened Critically endangered / Danger of extinction Vulnerable / Special protection Near threatened / Special protection Critically endangered / Special protection Critically endangered / Special protection

Ethnobotanical Science in the Clouds: Useful Plants of. . . 795

Actinidiaceae Adoxaceae

Adoxaceae

Adoxaceae

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Actinidiaceae

Magnoliophyta

Saurauia matudae Lundell Saurauia villosa DC. Sambucus canadensis L. Sambucus nigra L.

Scientific name Saurauia comitisrossei R.E. Schult.

Viburnum acutifolium Benth. Adoxaceae Viburnum hartwegii Benth. Adoxaceae Viburnum microcarpum Schltdl. & Cham. Alstroemeriaceae Bomarea edulis (Tussac) Herb. Altingiaceae Liquidambar styraciflua L.

Family Actinidiaceae

Division Magnoliophyta

Table 2 (continued)

Infraspecies

MCF, OF, PF, TEF, TSF

MCF

ECT (MCF-PF) TSF

TDF

MCF, OF

OF OF, TSF

MCF

Vegetation types MCF, OF, PF, TSF

Chi, Cui, IZ, Maz, Mes, Mix, Nz, Zoq

Chi, Mix

Maz

Maz

Maz, Mix, Nah Mxt

Maz Cui

Ethnic groupS Chi, Cui, IZ, Maz, Mix, NZ Chi

Ceremonial, Construction, Firewood, Medicinal, Ornamental, Timber

Medicinal, Ornamental

Firewood

Firewood, Handcraft

Handcraft, Ornamental

Handcraft, Medicinal

Edible, Firewood Firewood, Medicinal

Use Categories Edible, Firewood, Fodder, Handcraft, Medicinal Edible

Lower risk / -

Vulnerable / -

Risk status IUCN / SEMARNAT

796 B. Rendo´n-Aguilar et al.

Amaranthaceae

Amaranthaceae

Amaranthaceae

Amaranthaceae

Anacardiaceae

Anacardiaceae

Anacardiaceae

Anacardiaceae

Anacardiaceae Anacardiaceae

Anacardiaceae Anacardiaceae Anacardiaceae

Anacardiaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta Magnoliophyta Magnoliophyta

Magnoliophyta

Amaranthus hybridus L. Amaranthus spinosus L. Dysphania ambrosioides (L.) Mosyakin & Clemants Iresine diffusa Humb. & Bonpl. ex Willd. Amphipterygium adstringens (Schltdl.) Schiede ex Standl. Chrysobalanus icaco L. Cyrtocarpa procera Kunth Pistacia mexicana Kunth Rhus andrieuxii Engl. Rhus chondroloma Standl. Rhus galeottii Standl. Rhus oaxacana Loes. Rhus standleyi F.A. Barkley Rhus terebinthifolia Schltdl. & Cham. TSF

PF OF SHR, TDF

PF OF

TDF

TDF

TTF

TDF, TTF

MCF

TDF

TEF

OF

NZ

Mix Mix NZ

Mix NZ

NZ

Maz

Mes

IZ, Maz

Chi

Maz

Zoq

Maz

Firewood, Handcraft

Medicinal Medicinal Edible

Medicinal Edible, Medicinal

Firewood

Edible

Edible

Edible, Medicinal

Edible

Edible

Medicinal

Edible

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 797

Family Anacardiaceae

Anacardiaceae Annonaceae

Annonaceae

Annonaceae

Annonaceae

Apiaceae

Apiaceae

Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Division Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Ottoa oenanthoides Kunth Asclepias aff. rosea M.E. Jones Asclepias circinalis (Decne.) Woodson Asclepias curassavica L. Cascabela ovata (Cav.) Lippold

Spondias mombin L. Annona cherimola Mill. Cymbopetalum parviflorum N.A. Murray Rollinia rensoniana Standl. Xylopia frutescens Aubl. Eryngium foetidum L.

Scientific name Rhus virens Lindh. ex A. Gray

Infraspecies

PF, TDF, TEF, TSF TDF

PF

MCF

MCF, TEF, TSF ODC

TEF

OF, TSF

TDF MCF, TDF, TSF TEF

Vegetation types OF, TDF

Chi, Maz, NZ Maz, Mix

Maz

Mxt

Chi, Cui, Maz Nah

Chi

Mix

Mix

Maz, NZ Cui, Maz

Ethnic groupS Mix, NZ

Ceremonial, Medicinal, Ornamental Edible, Firewood, Medicinal

Environmental

Ceremonial

Fodder

Edible

Firewood, Timber

Edible

Firewood

Use Categories Edible, Firewood, Handcraft, Veterinary medicine Edible Edible

Risk status IUCN / SEMARNAT

798 B. Rendo´n-Aguilar et al.

Apocynaceae

Apocynaceae Apocynaceae

Apocynaceae

Apocynaceae

Apocynaceae

Araceae

Araceae

Araceae

Araceae

Araceae

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Philodendron hederaceum (Jacq.) Schott Spathiphyllum aff. brevirostre (Liebm.) Schott Spathiphyllum aff. montanum (R.A. Baker) Grayum

Cascabela thevetia (L.) Lippold Plumeria rubra L. Rauvolfia tetraphylla L. Stemmadenia oaxacana L.O. Alvarado Tabernaemontana chrysocarpa S.F. Blake Tabernaemontana odontadeniiflora A.O. Simões & M.E. Endress Anthurium scandens (Aubl.) Engl. Anthurium schlechtendalii Kunth

TSF

MCF

MCF

MCF, TDF, TEF, TSF

MCF

TSF

TEF

TDF, TSF

TDF TEF

TDF

Maz

Chi

Chi, Cui, Maz, Mes, NZ, Zoq Zoq

IZ

Zoq

Zoq

IZ, Mix

IZ Chi

NZ

Edible

Ornamental

Edible

Medicinal, Ornamental

Ornamental

Medicinal

Medicinal

Edible, Environmental

Ornamental Medicinal

Handcraft

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 799

Family Araceae

Araceae

Araceae

Araceae

Araliaceae

Araliaceae

Arecaceae

Arecaceae

Arecaceae

Arecaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Spathiphyllum blandum Schott Spathiphyllum cochlearispathum (Liebm.) Engl. Syngonium aff. Podophyllum Schott Xanthosoma robustum Schott Dendropanax aff. arboreus (L.) Decne. & Planch. Oreopanax echinops (Schltdl. & Cham.) Decne. & Planch. Acrocomia aculeata (Jacq.) Lodd. ex R. Keith Acrocomia mexicana Karw. ex Mart Astrocaryum mexicanum Liebm. ex Mart. Chamaedorea elatior Mart.

Infraspecies

MCF, TEF

TEF

MCF

OF

PF

MCF

MCF, TSF

TSF

MCF

Vegetation types MCF

Chi, IZ, Mix

Chi, Zoq

NZ

Maz

Chi

IZ, Mes, NZ IZ

Maz, Mix

Chi

Ethnic groupS Zoq

Construction, Handcraft

Construction, Edible, Handcraft

Edible

Edible

Edible

Edible, Handcraft, Medicinal Construction

Medicinal, Ornamental

Edible

Use Categories Ornamental

Vulnerable / -

Risk status IUCN / SEMARNAT

800 B. Rendo´n-Aguilar et al.

Arecaceae

Arecaceae

Arecaceae

Arecaceae

Arecaceae

Arecaceae

Arecaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Asparagaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Chamaedorea elegans Mart. Chamaedorea ernestiaugusti H. Wendl. Chamaedorea metallica O.F. Cook ex H.E. Moore Chamaedorea oblongata Mart. Chamaedorea tepejilote Liebm. Chamaedorea woodsoniana L.H. Bailey Geonoma interrupta (Ruiz & Pav.) Mart. Agave angustifolia Haw. Agave obscura Schiede ex Schltdl. Agave potatorum Zucc. Agave seemanniana Jacobi Beaucarnea aff. hiriartiae L. Hern. Beaucarnea sanctomariana L. Hern. TDF

TSF

PF, TEF

TDF

OF

MCF

TEF

MCF

TEF, TSF

MCF, TEF

MCF, TEF, TSF

ECT (OF-TEF) TEF, TSF

Zoq

Chi, Mix, NZ Cui

NZ

Nah

NZ

Zoq

Chi, Cui, Mix IZ

Mix, Zoq

Chi, Cui, Zoq

IZ, Zoq

Mix

Ornamental

Environmental

Edible, Medicinal

Edible, Medicinal

Edible

Edible

Ornamental

Edible

Edible, Ornamental

Edible, Ornamental

Handcraft, Ornamental

Ornamental

Edible

(continued)

-/ Threatened

-/ Threatened

Vulnerable / -

-/ Threatened - / Danger of extinction

Ethnobotanical Science in the Clouds: Useful Plants of. . . 801

Family Asparagaceae

Asparagaceae

Asparagaceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Dasylirion acrotrichum (Schiede) Zucc. Maianthemum paniculatum (M. Martens & Galeotti) La Frankie Nolina aff. longifolia (Karw. ex Schult. & Schult. f.) Hemsl. Ageratina ligustrina (DC.) R.M. King & H. Rob. Ageratina petiolaris (Moc. & Sessé ex DC.) R.M. King & H. Rob. Ageratum houstonianum Mill. Aldama dentata La Llave Alloispermum integrifolium (DC.) F. Rob.

Infraspecies

MCF

MCF, TEF, TSF OF

OF, PF

OF, PF

PF

MCF, OF

Vegetation types ECT (PF-MCF)

Chi

Chi, Cui, Maz, Mes Cui

NZ

Maz, Mix

NZ

IZ, Maz, Mix, Mxt

Ethnic groupS NZ

Medicinal

Fodder

Medicinal

Medicinal

Construction, Firewood, Ornamental

Construction

Ornamental

Use Categories Handcraft

Risk status IUCN / SEMARNAT

802 B. Rendo´n-Aguilar et al.

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Baccharis heterophylla Kunth Baccharis multiflora Kunth Baccharis salicifolia (Ruiz & Pav.) Pers. Baccharis trinervis Pers. Barkleyanthus salicifolius (Kunth) H. Rob. & Brettell Bartlettina tuerckheimii (Klatt) R.M. King & H. Rob. Bidens triplinervia Kunth Calea urticifolia (Mill.) DC. Chromolaena odorata (L.) R.M. King & H. Rob.

Aztecaster pyramidatus (B.L. Rob. & Greenm.) G.L. Nesom Baccharis conferta Kunth

var. urticifolia

subsp. salicifolia

var. multiflora

TEF

TSF

PF

MCF

PF, SHR, TDF MCF, PF, TDF, TEF MCF, PF

OF

OF, PF

FF, MCF, OF, PF, TDF

MCF, TDF

Chi

Cui, Mix

NZ

Chi, NZ

Chi, Maz, Mes, NZ Maz, NZ

Mix, NZ

Mix

Cui, Maz, Mes, Mxt, NZ NZ

NZ

Medicinal

Medicinal

Medicinal

Medicinal, Ornamental

Medicinal

Ceremonial, Medicinal

Medicinal

Medicinal

Handcraft

Ceremonial, Handcraft, Medicinal

Firewood, Handcraft

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 803

Family Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Cirsium lappoides (Less.) Sch. Bip. Cirsium pinetorum Greenm. Cirsium rhaphilepis Petr. Cirsium subcoriaceum (Less.) Sch. Bip. Critonia morifolia (Mill.) R.M. King & H. Rob. Dahlia australis (Scherff) P.D. Sørensen Dahlia tenuicaulis P.D. Sørensen Erigeron karvinskianus DC. Galinsoga parviflora Cav. Gamochaeta americana (Mill.) Wedd. var. australis

Infraspecies

ECT (MCF-PF) OF

MCF, OF

FF

MCF, PF

TEF

PF

OF

MCF, OF

Vegetation types TDF

Mix

NZ

Maz, Mix

NZ

NZ

Chi

NZ

Maz, Mix, NZ Maz

Ethnic groupS NZ

Medicinal

Fodder

Medicinal

Ornamental

Ornamental

Medicinal

Medicinal

Edible

Edible, Medicinal

Use Categories Medicinal

- / Special protection

Risk status IUCN / SEMARNAT

804 B. Rendo´n-Aguilar et al.

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Gymnosperma glutinosum Less. Lagascea helianthifolia Kunth Lasianthaea fruticosa (L.) K.M. Becker Lepidaploa tortuosa (L.) H. Rob. Leucanthemum lacustre (Brot.) Samp. Melampodium divaricatum (Rich.) DC. Mikania pyramidata Donn. Sm. Neomirandea araliifolia (Less.) R.M. King. & H. Rob. Neurolaena lobata (L.) Cass. Neurolaena macrocephala Sch. Bip. ex Hemsl. Oxylobus subglabrus R.M. King & H. Rob. Pinaropappus roseus (Less.) Less. Piqueria trinervia Cav. var. fruticosa

OF

PF

MCF, OF

MCF, TEF, TSF TEF, TSF

TEF

MCF

TSF

PF

TEF

TSF

OF

TDF

Mix

NZ

NZ

Chi

Chi, Maz

Chi

Chi

Mes

Maz

Chi

NZ

Mix

NZ

Veterinary medicine

Medicinal

Medicinal

Medicinal, Veterinary medicine

Medicinal, Ornamental

Medicinal

Medicinal

Fodder

Ornamental

Ceremonial

Firewood

Fodder

Medicinal

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 805

Family Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Podachaenium eminens (Lag.) Sch. Bip. Podachaenium pachyphyllum (Klotzsch) R.K. Jansen, N.A. Jansen, N.A. Harriman & Urbatsch Pseudognaphalium chartaceum (Greenm.) Anderb. Pseudognaphalium elegans Kartesz Pseudognaphalium roseum (Kunth) Anderb. Pseudognaphalium viscosum (Kunth) Anderb. Roldana angulifolia (DC.) H. Rob. & Brettell Roldana oaxacana (Hemsl.) H. Rob & Brettell

Infraspecies

MCF, OF

PF

RMV

OF, PF, TSF

OF

OF

MCF

Vegetation types TEF

Cui, Mix

Mix

Nah

Mix, NZ

NZ

Maz

Chi

Ethnic groupS Chi

Medicinal

Handcraft

Medicinal

Medicinal

Medicinal

Medicinal

Ornamental

Use Categories Firewood

Risk status IUCN / SEMARNAT

806 B. Rendo´n-Aguilar et al.

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae Asteraceae Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Tagetes lucida Cav.

MCF, OF OF, PF

OF

Tagetes filifolia Lag.

OF, PF, RMV PF

MCF, OF, PF, TDF

OF, TDF

var. oaxacensis

var. ovata

var. oaxacana

MCF OF TDF

PF

TDF

PF

MCF

Stevia polycephala Grashoff Stevia seemannii Sch. Bip. Tagetes erecta L.

Stevia ovata Willd.

Rumfordia floribunda DC. Sanvitalia fruticosa Hemsl. Sanvitalia procumbens Lam. Senecio callosus Sch. Bip. Sonchus oleraceus L. Stevia hirsuta DC. Stevia lucida (DC.) Grashoff Stevia microchaeta Sch. Bip.

Cui, Maz, NZ IZ, Mes, NZ

Maz, Mix

Mes

Chi, Maz, Mes, Mxt, Nah Mes, Mix, NZ Mix

Mes Mes NZ

NZ

Maz, NZ

Mix

NZ

Ceremonial, Edible, Medicinal, Ornamental

Ceremonial, Handcraft, Medicinal, Ornamental Edible, Medicinal

Ornamental

Ornamental

Medicinal, Ornamental

Ceremonial, Handcraft, Medicinal, Ornamental

Edible Ceremonial Medicinal

Ornamental

Medicinal, Ornamental

Medicinal

Ornamental

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 807

Family Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Asteraceae Asteraceae

Asteraceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Tagetes micrantha Cav. Tagetes tenuifolia Cav. Tanacetum parthenium (L.) Sch. Bip. Telanthophora grandifolia (Less.) H. Rob. & Brettell Tetrachyron manicatum Schltdl. Tithonia diversifolia (Hemsl.) A. Gray Tithonia longiradiata (Bertol.) S.F. Blake Tithonia rotundifolia (Mill.) S.F. Blake Tridax procumbens L. Trigonospermum melampodioides DC. Verbesina hypoglauca Sch. Bip. ex Klatt var. grandifolia

Infraspecies

PF

TSF MCF, PF

OF

MCF

TDF, TEF

MCF

MCF

PF

PF

Vegetation types MCF

Mix

NZ Maz, Mix

Zoq

Chi

Mix, Zoq

Maz

Chi

NZ

Mix

Ethnic groupS Cui

Fodder

Edible Fodder

Medicinal

Ornamental

Medicinal

Fodder

Medicinal, Ornamental

Medicinal

Ceremonial

Use Categories Edible

Risk status IUCN / SEMARNAT

808 B. Rendo´n-Aguilar et al.

Asteraceae

Asteraceae

Asteraceae

Asteraceae

Basellaceae

Begoniaceae Begoniaceae

Begoniaceae

Begoniaceae

Begoniaceae

Begoniaceae

Begoniaceae

Begoniaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Begonia heracleifolia Schltdl. & Cham. Begonia hispidivillosa Ziesenh. Begonia manicata Brongn. Begonia oaxacana A. DC. Begonia pinetorum A. DC. Begonia rhodochlamys L.B. Sm. & B.G. Schub.

Verbesina neriifolia Hemsl. Verbesina perymenioides Sch. Bip. ex Kaltt Verbesina turbacensis Kunth Vernonanthura patens (Kunth) H. Rob. Anredera vesicaria (Lam.) C.F. Gaertn. Begonia fusca Liebm. Begonia glabra Aubl.

TSF

TSF

OF

MCF

OF, TEF, TSF TEF

OF, PF, TSF MCF, TSF

TDF

MCF, TDF

MCF, OF

OF

OF

Maz

Maz

Maz

Mix

Maz, Mix Chi, IZ, Mes Chi, Maz, NZ Chi

NZ

Chi, Mxt

Cui, Mes

Mes

Mix

Medicinal

Edible, Medicinal

Edible, Ornamental

Ornamental

Fodder, Medicinal, Ornamental Medicinal

Ornamental Medicinal, Ornamental

Environmental, Firewood Medicinal

Ceremonial, Medicinal

Ceremonial, Firewood

Medicinal

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 809

Family Berberidaceae

Betulaceae

Bignoneaceae

Bignoniaceae Bignoniaceae

Bignoniaceae

Bignoniaceae

Bixaceae

Boraginaceae

Boraginaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Parmentiera aculeata (Kunth) Seem. Crescentia cujete L. Handroanthus chrysanthus (Jacq.) S.O. Grose Paragonia aff. pyramidata (Rich.) Bureau Pithecoctenium crucigerum (L.) A.H. Gentry Cochlospermum vitifolium (Willd.) Spreng. Bourreria laevis G. Don Cordia alliodora (Ruiz & Pav.) Oken

Scientific name Berberis moranensis Schult. & Schult. f. Alnus acuminata Kunth

Infraspecies

ECT (MCF-OF) TSF, TTF

TEF, TTF

TTF

OF

OF, TSF PF

SHR

MCF, OF, PF, RMV, TDF

Vegetation types PF

IZ, Mes

Maz

Chi, IZ

Mes

Zoq

Mix Chi

NZ

Cui, Maz, Mix, Mxt, Nah, NZ

Ethnic groupS NZ

Timber

Medicinal

Ceremonial, Construction

Handcraft

Timber

Handcraft Edible

Ceremonial, Construction, Edible, Environmental, Firewood, Handcraft, Medicinal, Timber Edible

Use Categories Ornamental

Least concern / -

Least concern / -

Risk status IUCN / SEMARNAT

810 B. Rendo´n-Aguilar et al.

Boraginaceae

Boraginaceae

Brassicaceae Brassicaceae

Brassicaceae

Brassicaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Heliotropium macrostachyum (DC.) Hemsl. Tournefortia hirsutissima L. Brassica rapa L. Cardamine flexuosa With. Lepidium virginicum L. Nasturtium officinale W.T. Aiton Ananas comosus (L.) Merr. Bromelia palmeri Mez Catopsis compacta Mez Catopsis paniculata É. Morren Catopsis sessiliflora (Ruiz & Pav.) Mez Greigia oaxacana L.B. Sm. Guzmania nicaraguensis Mez & C.F. Baker Hechtia aff. caudata L.B. Sm. TDF

MCF

MCF

MCF, OF

MCF

OF

TDF

OF

PF

SHR, TSF

PF TSF

SHR

TDF

IZ

Mix, Zoq

Cui, Maz, Nah Chi

Nah

Mix

IZ

NZ

NZ

NZ

NZ NZ

NZ

Mix

Edible, Ornamental

Ornamental

Edible

Handcraft, Ornamental

Handcraft

Ornamental

Edible, Ornamental

Edible

Edible

Edible, Medicinal

Edible Edible

Medicinal

Handcraft

(continued)

Least concern / -

Ethnobotanical Science in the Clouds: Useful Plants of. . . 811

Family Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Hechtia rosea É. Morren ex Baker Pitcairnia heterophylla (Lindl.) Beer Pitcairnia imbricata (Brongn.) Regel Tillandsia achyrostachys É. Morren ex Baker Tillandsia deppeana Steud. Tillandsia filifolia Schltdl. & Cham. Tillandsia gymnobotrya Baker Tillandsia imperialis É. Morren ex Roezl Tillandsia ionantha Planch. Tillandsia lampropoda L.B. Sm. Tillandsia leiboldiana Schltdl.

Infraspecies

MCF

PF

MCF, OF, PF MCF, OF, PF TDF

TEF

OF

OF, SHR, TDF

MCF, PF

MCF

Vegetation types TDF

IZ

NZ

Chi, Maz, Mxt IZ

Maz, NZ

Zoq

Maz

Mix, NZ

IZ, Mix

IZ

Ethnic groupS IZ

Ornamental

Ornamental

Ornamental

Ceremonial, Fodder, Ornamental Ornamental

Ornamental

Ornamental

Ornamental

Handcraft, Ornamental

Environmental

Use Categories Edible

-/ Threatened Least concern / -/ Threatened

Risk status IUCN / SEMARNAT

812 B. Rendo´n-Aguilar et al.

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Bromeliaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Tillandsia limbata Schltdl. Tillandsia macdougallii L.B. Sm. Tillandsia macrochlamys Baker Tillandsia makoyana Baker Tillandsia multicaulis Steud. Tillandsia oaxacana L.B. Sm. Tillandsia polystachia (L.) L. Tillandsia prodigiosa (Lem.) Baker Tillandsia punctulata Schltdl. & Cham. Tillandsia schiedeana Steud. Tillandsia seleriana Mez Tillandsia tricolor Schltdl. & Cham. Tillandsia usneoides (L.) L. Tillandsia zoquensis Ehlers MCF

OF, PF

TSF

MCF

TEF

MCF, OF

PF

OF, TSF

PF

MCF, OF

TDF

PF

PF

TSF

Maz, Mes, Mix, NZ NZ

Mes

IZ

Cui, Maz, NZ Chi

NZ

Cui, Mes

Chi, Maz, Mix Mix

NZ

Mix

NZ

Cui

Ceremonial, Ornamental Ornamental

Ornamental

Ornamental

Ceremonial

Ornamental

Ornamental

Ornamental

Ornamental

Ornamental

Ornamental

Handcraft, Ornamental

Ceremonial, Ornamental Ornamental

(continued)

-/ Threatened -/ Threatened

Least concern / -

Ethnobotanical Science in the Clouds: Useful Plants of. . . 813

Family Bromeliaceae

Burseraceae

Burseraceae

Burseraceae

Burseraceae

Burseraceae

Burseraceae

Burseraceae

Burseraceae

Burseraceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Bursera glabrifolia (Kunth) Engl. Bursera linanoe (La Llave) Rzed., Calderón & Medina Bursera simaruba (L.) Sarg. Bursera submoniliformis Engl. Protium copal (Schltdl. & Cham.) Engl.

Scientific name Werauhia werckleana (Mez) J.R. Grant Bursera aloexylon (Schiede ex Schltdl.) Engl. Bursera bipinnata (Moc. & Sessé ex DC.) Engl. Bursera excelsa (Kunth) Engl. Bursera fagaroides (Kunth) Engl.

Infraspecies

ECT (TDF-SHR) TDF

TEF, TSF

ECT (PF-TDF) TDF

TDF

TDF

TDF

TDF

Vegetation types OF

Maz

Maz

Zoq

Maz

NZ

Mix, NZ

IZ

Maz, Mix, NZ

Maz

Ethnic groupS Mix

Ceremonial

Construction, Medicinal Ceremonial

Handcraft

Construction, Environmental, Medicinal Ceremonial

Medicinal

Ceremonial, Handcraft, Medicinal

Ceremonial

Use Categories Ornamental

Risk status IUCN / SEMARNAT

814 B. Rendo´n-Aguilar et al.

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Cactaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Aporocactus martianus (Zucc. ex Pfeiff.) Britton & Rose Disocactus ackermannii (Haw.) Ralf Bauer Epiphyllum oxypetalum (DC.) Haw. Mammillaria deherdtiana (Bravo) D.R. Hunt Mammillaria voburnensis (Britton & Rose) D.R. Hunt Opuntia lasiacantha Pfeiff. Pachycereus pectenaboriginum (Engelm. ex S. Watson) Britton & Rose Pereskia lychnidiflora DC. Pilosocereus leucocephalus (Poselg.) Byles & G.D. Rowley subsp. collinsii

subsp. dodsonii

TDF

TTF

TTF

TDF, TTF

TDF

FF, PF

MCF

MCF, OF

MCF, OF

Zoq

IZ

IZ, NZ, Zoq IZ

IZ

NZ

Cui

Cui, Maz, Mxt

Maz, Mxt

Edible

Environmental

Handcraft

Construction, Edible

Edible

Medicinal, Ornamental

Edible

Edible, Ornamental

Medicinal, Ornamental

(continued)

Least concern / Least concern / -

Least concern / Least concern / -

Vulnerable / Threatened

Least concern / -

Least concern / -

Ethnobotanical Science in the Clouds: Useful Plants of. . . 815

Family Cactaceae

Cactaceae

Calceolariaceae

Calceolariaceae

Calophyllaceae

Campanulaceae

Cannabaceae Cannabaceae

Cannaceae

Capparaceae

Capparaceae

Caprifoliaceae Caprifoliaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Table 2 (continued)

Quadrella incana (Kunth) Iltis & Cornejo Quadrella indica (L.) Iltis & Cornejo Valeriana scandens L. Valeriana sorbifolia Kunth

Scientific name Rhipsalis baccifera (J.S. Muell.) Stearn Strophocactus testudo (Karw. ex Zucc.) Ralf Bauer Calceolaria mexicana Benth. Calceolaria tripartita Ruiz & Pav. Calophyllum brasiliense Cambess. Lobelia laxiflora Kunth Celtis caudata Planch. Trema micrantha (L.) Blume Canna indica L.

Infraspecies

MCF MCF

TTF

MCF, OF, PF TSF MCF, OF, TEF, TSF MCF, OF, RMV, TEF TTF

MCF, TEF

RMV

MCF

Vegetation types ECT (MCF-PF) TDF

NZ Maz

Mes

IZ

Cui, Maz, Mix, NZ NZ Chi, Mix, NZ Chi, Mix

IZ, Zoq

NZ

Maz

Zoq

Ethnic groupS NZ

Medicinal Ornamental

Edible

Edible, Fodder, Medicinal, Ornamental Handcraft Edible, Firewood, Fodder, Handcraft Edible, Handcraft, Medicinal, Ornamental Firewood, Edible

Construction, Timber

Edible

Medicinal

Edible

Use Categories Environmental

-/ Threatened

Risk status IUCN / SEMARNAT Least concern / Least concern / -

816 B. Rendo´n-Aguilar et al.

Caricaceae Caricaceae

Caryophyllaceae

Cecropiaceae

Chloranthaceae

Cleomaceae

Cleomaceae Clethraceae

Clethraceae

Clethraceae

Clethraceae

Clethraceae

Clethraceae

Clethraceae

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Hedyosmum mexicanum C. Cordem. Cleome magnifica Briq. Cleome viridiflora L. Clethra galeottiana Briq. Clethra hartwegii Britton Clethra hondurensis Britton Clethra kenoyeri Lundell Clethra lanata M. Martens & Galeotti Clethra pringlei S. Watson Clethra suaveolens Turcz.

Carica papaya L. Jacaratia mexicana A. DC. Stellaria media (L.) Vill. Cecropia obtusifolia Bertol.

MCF

OF, PF

PF

MCF, OF, PF OF

MCF ECT (ODC-MCF) OF

MCF

MCF, TSF

MCF, TEF, TSF

OF

TSF, TEF TTF

Nah

Chi

Mix

Maz, Mix, NZ Maz

Mix

Cui Nah

NZ

Chi, Cui, Maz, Mix, NZ, Zoq Chi, Maz, Zoq

Cui

Chi, Mix IZ

Firewood

Construction

Firewood

Construction, Edible, Firewood, Ornamental Construction, Timber

Construction

Edible Construction

Edible

Construction, Edible, Firewood, Medicinal

Construction, Edible, Firewood, Medicinal

Ceremonial

Edible, Medicinal Edible

(continued)

Vulnerable / -

Lower risk / -

Ethnobotanical Science in the Clouds: Useful Plants of. . . 817

Family Clusiaceae

Clusiaceae

Clusiaceae

Clusiaceae Clusiaceae

Clusiaceae

Combretaceae

Commelinaceae

Commelinaceae

Commelinaceae

Convolvulaceae

Convolvulaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Clusia lundellii Standl. Clusia quadrangula Bartlett Clusia salvinii Donn. Sm. Garcinia edulis Exell Garcinia macrophylla Mart. Rheedia edulis (Seem.) Planch. & Triana Combretum farinosum Kunth Tradescantia zanonia (L.) Sw. Tradescantia zebrina Bosse Tripogandra serrulata (Vahl) Handlos Cuscuta corymbosa Ruiz & Pav. Cuscuta jalapensis Schltdl.

Infraspecies

MCF

OF

OF

TEF

MCF, TSF

TTF

TEF

MCF, TEF, TSF MCF MCF

Vegetation types ECT (MCF-PF) TSF

Nah

Nah

NZ

Chi

Cui, IZ

Mes

Chi

Cui, Mz, Zoq IZ Zoq

Maz

Ethnic groupS Chi

Fodder

Medicinal

Medicinal

Medicinal

Edible, Medicinal

Handcraft

Firewood

Edible Construction

Construction, Edible

Edible, Firewood

Use Categories Edible, Handcraft

Risk status IUCN / SEMARNAT

818 B. Rendo´n-Aguilar et al.

Convolvulaceae

Convolvulaceae

Convolvulaceae

Convolvulaceae

Convolvulaceae Convolvulaceae

Convolvulaceae

Costaceae Costaceae

Crassulaceae

Crassulaceae

Crassulaceae

Crassulaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Ipomoea chenopodiifolia (M. Martens & Galeotti) Hemsl. Ipomoea dumosa (Benth.) L.O. Williams Ipomoea hederifolia L. Ipomoea ignava House Ipomoea nil (L.) Roth Ipomoea pauciflora M. Martens & Galeotti Turbina corymbosa (L.) Raf. Costus pictus D. Don. Costus pulverulentus C. Presl Echeveria montana Rose Echeveria rosea Lindl. Sedum praealtum A. DC. Villadia guatemalensis Rose MCF, OF

FF, PF

MCF

OF

TEF TEF, TSF

TSF

MCF TDF

PF

OF

TEF, TSF

MCF, TSF

Cui, Mix

NZ

Mxt, NZ

Zoq Chi, Mix, Zoq Maz

Maz

NZ Maz

NZ

NZ

Chi, Mes, NZ

NZ

Ornamental

Ornamental

Medicinal

Ornamental

Ornamental Medicinal, Ornamental

Medicinal

Fodder Medicinal

Handcraft

Edible

Edible, Fodder

Handcraft

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 819

Family Cucurbitaceae

Cucurbitaceae

Cucurbitaceae

Cucurbitaceae

Cucurbitaceae

Cyclanthaceae

Cyperaceae

Cyperaceae

Dioscoreaceae

Dracaenaceae

Ebenaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Asplundia aff. Utilis (Oerst.) Harling Eleocharis elegans (Kunth) Roem. & Schult. Eleocharis mutata (L.) Roem. & Schult. Dioscorea composita Hemsl. Dracaena americana Donn. Sm. Diospyros digyna Jacq.

Scientific name Cyclanthera ribiflora (Schltdl.) Cogn. Microsechium helleri (Peyr.) Cogn. Microsechium palmatum (Ser.) Cogn. Momordica charantia L. Sicyos galeottii Cogn.

Infraspecies

MCF, TDF, TSF

TSF

OF

SAV

SAV

MCF, TEF

MCF

TEF

MCF, OF, TEF, TSF ECT (PF-MCF)

Vegetation types OF

Maz, Mes, Mix, Mxt

Maz

Maz, Mix

Maz

Maz

IZ, Zoq

Chi

Chi

Maz

Mix, NZ

Ethnic groupS Cui

Edible, Firewood, Medicinal, Veterinary medicine

Edible, Handcraft, Medicinal Environmental

Fodder

Fodder, Handcraft, Veterinary medicine Construction, Environmental Fodder

Medicinal

Handcraft, Medicinal

Fodder, Handcraft

Use Categories Handcraft

Risk status IUCN / SEMARNAT

820 B. Rendo´n-Aguilar et al.

Ebenaceae

Elaeocarpaceae

Ericaceae

Ericaceae

Ericaceae

Ericaceae

Ericaceae

Ericaceae

Ericaceae

Ericaceae

Ericaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Gaultheria erecta Vent.

Arctostaphylos pungens Kunth Bejaria aestuans Mutis Cavendishia bracteata (Ruiz & Pav. ex J. St.-Hil.) Hoerold Chimaphila maculata (L.) Pursh Comarostaphylis arbutoides Lindl. Comarostaphylis discolor (Hook.) Diggs Gaultheria acuminata Schltdl. & Cham.

Diospyros oaxacana Standl. Sloanea tuerckheimii Dnn. Sm. Arbutus xalapensis Kunth

MCF, OF, PF

MCF, OF, PF, TSF

MCF, PF

OF, SHR

PF

OF

MCF, OF

MCF, OF, PF, RMV, TDF, TEF, TSF OF, PF

TEF

TTF

Chi, Maz, Mix, Mxt, Nah Chi, Cui, Mix, Nah, NZ

Chi, Nah, NZ

NZ

Maz

Maz

Cui, Mes, NZ Mix, NZ

Maz, Mix, NZ

Zoq

IZ

Lower risk / -

Ceremonial, Edible, Firewood, Handcraft, Medicinal, Ornamental Ceremonial, Edible, Medicinal, Ornamental (continued)

Edible, Firewood, Handcraft, Timber Ceremonial, - / Special Construction, Firewood protection

Environmental

Ceremonial, Edible, Medicinal

Edible, Firewood, Medicinal Firewood, Ornamental

Construction, Edible, Firewood, Handcraft, Medicinal, Ornamental

Fodder

Edible

Ethnobotanical Science in the Clouds: Useful Plants of. . . 821

Family Ericaceae

Ericaceae

Ericaceae

Ericaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Croton ramillatus Croizat

Adelia barbinervis Schltdl. & Cham. Alchornea latifolia Sw. Cnidoscolus tubulosus (Müll. Arg.) I.M. Johnst. Cnidoscolus urens (L.) Arthur Croton draco Schltdl. & Cham. Croton niveus Jacq.

Scientific name Macleania insignis M. Martens & Galeotti Orthaea stipitata (Luteyn) Luteyn Pernettya prostrata (Cav.) DC. Vaccinium leucanthum Schltdl.

Infraspecies

TDF

MCF, OF, TEF, TSF TTF

TDF

MCF, TSF

OF

TEF

MCF, OF, PF

ECT (MCF-PF) ODC

Vegetation types OF

Maz

Chi, Mix, NZ IZ

Maz

NZ

Maz

Chi, Maz, Mix, Nah, NZ Chi

Nah

Chi

Ethnic groupS Maz

Handcraft

Environmental, Firewood, Medicinal Medicinal

Handcraft

Handcraft, Medicinal

Timber

Firewood

Edible, Firewood, Fodder, Handcraft

Edible, Medicinal

Ornamental

Use Categories Edible

Least concern / -

Risk status IUCN / SEMARNAT

822 B. Rendo´n-Aguilar et al.

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Euphorbiaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Acacia centralis (Britton & Rose) Lundell Acacia cochliacantha Humb. & Bonpl. ex Willd. Acacia farnesiana (L.) Willd. Acacia pennatula (Schltdl. & Cham.) Benth.

Euphorbia nutans Lag. Euphorbia pulcherrima Will. ex Klotzsch Euphorbia schiedeana (Klotzsch & Garcke) Mayfield ex C. Nelson Euphorbia tehuacana (Brandegee) V.W. Steinm. Jatropha gossypiifolia L. Ricinus communis L.

OF, PF, TDF

TDF, TTF

ECT (TDF-SHR)

MCF, OF, PF, TSF OF

RMV

SHR

PF

ECT (TDF-PF) ECT (TDF-PF)

Mix, NZ, Zoq

IZ, NZ

Maz

Maz, Mix, Nah Mes

IZ

Maz

Mix

Mix

NZ

Ceremonial, Fodder, Medicinal Construction, Firewood, Handcraft, Medicinal, Timber

Fodder

Firewood, Handcraftr, Medicinal Timber

Fodder

Medicinal

Medicinal

Ornamental

Medicinal

(continued)

-/ Threatened

Ethnobotanical Science in the Clouds: Useful Plants of. . . 823

Family Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Acacia pringlei (Brandegee) Y.S. Lee, D. Seigler & Ebinger Acaciella angustissima (Mill.) Britton & Rose Acosmium panamense (Benth.) Yakovlev Bauhinia divaricata L. Caesalpinia pulcherrima (L.) Sw. Calliandra houstoniana (Mill.) Standl. Canavalia villosa Benth. Cojoba arborea (L.) Britton & Rose Crotalaria longirostrata Hook. & Arn. Crotalaria pumila Ortega var. arborea

Infraspecies subsp. californica

ECT (PF-TDF)

TDF

MCF, TEF

OF

OF, PF

TDF

TEF

TSF

OF

Vegetation types TTF

NZ

Maz

IZ, Zoq

Mix

Mix

Maz

Chi

Mix

Mix, Nah, NZ

Ethnic groupS IZ

Edible

Edible

Timber

Construction

Edible, Firewood, Handcraft

Firewood

Firewood

Medicinal

Edible, Handcraft

Use Categories Firewood, Handcraft

Least concern / -

-/ Threatened

Risk status IUCN / SEMARNAT

824 B. Rendo´n-Aguilar et al.

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Hymenaea courbaril L.

Crotalaria purshii DC. Crotalaria quercetorum Brandegee Crotalaria sagittalis L. Diphysa aff. americana (Mill.) M. Sousa Diphysa carthagenensis Jacq. Diphysa floribunda Peyr. Enterolobium cyclocarpum (Jacq.) Griseb. Erythrina aff. hondurensis Standl. Erythrina americana Mill. Erythrina caribaea Krukoff & Barneby Gliricidia sepium (Jacq.) Steud. ECT (MCF-TDF)

OF, TTF

MCF

MCF

TSF

OF, TDF, TSF ECT (MCF-TSF) TDF, TSF

TSF

MCF, OF

ECT (OF-MCF) TSF

IZ

IZ, Mes, Zoq

Maz

Cui

Mix

Maz, NZ

Mix

NZ

Mix

Chi, Mix

Mes

Mix

Construction, Handcraft, Medicinla, Timber Ceremonial, Handcraft

Edible

Edible

Construction, Edible, Environmental, Firewood, Timber Construction

Construction, Edible, Medicinal Construction, Firewood

Construction, Medicinal

Ceremonial, Edible

Edible, Handcraft, Medicinal

Medicinal

(continued)

Least concern / -

Least concern / -

Ethnobotanical Science in the Clouds: Useful Plants of. . . 825

Family Fabaceae

Fabaceae Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Division Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Mimosa polyantha Benth.

Leucaena esculenta (Moc. & Sessé ex DC.) Benth. Leucaena leucocephala (Lam.) de Wit Leucaena macrophylla Benth Lupinus aff. Montanus Kunth Lysiloma acapulcense (Kunth) Benth. Mimosa albida Humb. & Bonpl. ex Willd.

Scientific name Inga aff. laurina (Sw.) Willd. Inga eriocarpa Benth. Inga oerstediana Benth Inga vera (Benth.) J. León

var. albida

subsp. eriocarpa

Infraspecies

ECT (TDF-SHR)

PF, RMV, TDF MCF, SHR, TDF, TSF

ECT (PF-TDF) PF

SHR, TSF

OF, TDF

MCF, PF, TEF

TSF MCF, RMV

Vegetation types MCF

IZ, Mix, NZ Chi, Cui, Maz, Mes, NZ Maz

NZ

NZ

NZ

Mix IZ, Maz, NZ Chi, Cui, Mes, Mix, NZ, Zoq Mix, NZ

Ethnic groupS Cui

Fodder

Construction, Edib, Firewood, Fodder Ceremonial, Medicinal

Ornamental

Edible

Edible, Medicinal

Edible

Edible, Environmental, Firewood, Handcraft

Environmental Edible, Firewood

Use Categories Edible

Least concern / Least concern / -

Endangered / -

Risk status IUCN / SEMARNAT

826 B. Rendo´n-Aguilar et al.

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Fabaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Mimosa tenuiflora (Willd.) Poir. Mucuna urens (L.) Medik. Phaseolus coccineus L. Prosopis laevigata (Humb. & Bonpl. ex Willd.) M.C. Johnst. Pterocarpus orbiculatus DC. Rhynchosia longeracemosa M. Martens & Galeotti Senna alata (L.) Roxb. Senna atomaria (L.) H.S. Irwin & Barneby Senna foetidissima (Ruiz & Pav. ex G. Don) H.S. Irwin & Barneby Senna fruticosa (Mill.) H.S. Irwin & Barneby Senna racemosa (Mill.) H.S. Irwin & Barneby OF

TSF

ECT (TDF-SHR) MCF

RMV

TDF

MCF

MCF, OF, PF TTF

ECT (MCF-TDF) MCF, TSF

Mix

Mix

Mix

Maz

IZ

Maz

NZ

IZ

Cui, NZ

Mix

IZ

Construction

Edible

Ornamental

Firewood, Fodder

Fodder

Medicinal

Firewood, Handcraft

Medicinal

Edible, Firewood

Handcraft, Medicinal

Medicinal

(continued)

Lower risk / -

Ethnobotanical Science in the Clouds: Useful Plants of. . . 827

Family Fabaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Quercus conzattii Trel. Quercus corrugata Hook. Quercus crassifolia Bonpl.

Quercus conspersa Benth.

Quercus castanea Née

Quercus acutifolia Née Quercus aff. Rubramenta Trel. Quercus affinis Scheidw. Quercus candicans Née

Scientific name Zornia thymifolia Kunth Quercus acatenangensis Trel.

Infraspecies

OF, PF

OF, PF

OF, PF

OF, PF

OF, PF

MCF, OF, PF

MCF, OF

PF

OF

MCF, PF

Vegetation types MCF

Chi, Mix, NZ

Mes, Mix, NZ Maz

Mes, Mix, MXt, NZ Maz, Mix, NZ

Cui, Maz, Mix, NZ

Mxt, Nah

Maz

NZ

Mix, NZ

Ethnic groupS Cui

Construction, Firewood, Handcraft, Timber Firewood, Handcraftr, Medicinal Construction, Firewood, Timber Construction, Environmental, Firewood, Handcraft

Construction, Edible, Firewood, Handcraft, Medicinal, Timber Firewood, Timber

Firewood, Handcraft

Firewood

Construction, Firewood, Handcraft, Ornamental Firewood, Medicinal

Use Categories Medicinal

Least concern / -

Lower risk / -

Vulnerable

Risk status IUCN / SEMARNAT

828 B. Rendo´n-Aguilar et al.

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Quercus liebmannii Oerst. ex Trel. Quercus magnoliifolia Née Quercus martinezii C.H. Mull.

Quercus lancifolia Schltdl. & Cham. Quercus laurina Bonpl.

Quercus glaucescens Bonpl. Quercus glaucoides M. Martens & Galeotti Quercus greggii (A. DC.) Trel. Quercus laeta Liebm.

Quercus glabrescens Benth.

Quercus depressa Bonpl. Quercus elliptica Née

PF

ECT (TDF-OF) OF, PF

MCF, OF, PF

MCF

OF

OF

MCF, OF, PF, TEF OF, TSF

MCF, OF, PF

MCF, OF, PF

ODC

Mix

Mix

Chi, Maz, Mes, Nah, NZ NZ

Cui

Maz

NZ

NZ

IZ, Mix

Chi, Cui, Maz, Mix, Nz, Zoq Mes, Mxt, NZ

Nah

Firewood, Handcraft

Construction, Firewood

Construction, Firewood, Handcraft, Medicinal Firewood

Construction, Firewood, Timber Firewood, Timber

Firewood

Construction, Firewood, Handcraft, Timber Construction, Firewood, Medicinal Firewood

Construction, Firewood, Medicinal

Firewood

(continued)

Lower risk / -

Lower risk / -

Lower risk / -

Ethnobotanical Science in the Clouds: Useful Plants of. . . 829

Family Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Fagaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Quercus sapotifolia Liebm. Quercus sartorii Liebm. Quercus scytophylla Liebm. Quercus segoviensis Liebm. Quercus skinneri Benth. Quercus trinitatis Trel.

Scientific name Quercus nixoniana S. Valencia & LozadaPérez Quercus oleoides Schltdl. & Cham. Quercus peduncularis Née Quercus pinnativenulosa C.H. Mull. Quercus polymorpha Schltdl. & Cham. Quercus rugosa Née

Infraspecies

MCF, OF

MCF, OF, PF MCF

MCF, OF, PF OF, PF

MCF, OF, PF OF

OF, PF

MCF, OF, PF MCF, PF

OF

Vegetation types MCF

IZ, Mix, Mxt, Zoq Chi, Maz, Mix, Mxt

Maz, Mix, NZ Maz, NZ

Cui, Maz

Cui, Mix, NZ Mix

Nah, NZ

Maz, Mes, NZ

Cui, Mix

Mix, Zoq

Ethnic groupS Cui

Construction, Firewood, Timber

Construction, Firewood, Timber Construction, Firewood, Medicinal Construction, Firewood, Timber Firewood, Timber

Construction, Firewood

Firewood, Handcraft

Construction, Firewood, Medicinal Construction, Firewood, Fodder, Medicinal Firewood, Medicinal

Construction, Firewood

Use Categories Firewood

Vulnerable / -

Risk status IUCN / SEMARNAT

830 B. Rendo´n-Aguilar et al.

Gentianaceae

Gentianaceae

Gesneriaceae

Gesneriaceae

Gesneriaceae

Gesneriaceae

Grossulariaceae

Heliconiaceae

Heliconiaceae

Hernandiaceae

Hypericaceae

Iridaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Chelonanthus alatus (Aubl.) Pulle Gentiana ovatiloba Kusn. Diastema aff. racemiferum Benth. Moussonia deppeana (Schltdl. & Cham.) Hanst. Moussonia elegans Decne. Moussonia larryskogii Ram.-Roa Ribes ciliatum Humb. & Bonpl. ex Roem. & Schult. Heliconia bourgaeana Petersen Heliconia collinsiana Griggs Gyrocarpus mocinoi Espejo Vismia baccifera (L.) Triana & Planch. Alophia drummondii (Graham) R.C. Foster TDF

MCF, TSF

TTF

TEF

TSF

PF

MCF

MCF

ECT (MCF-OF)

TSF

PF

MCF

IZ

Chi, Maz

IZ

Mix

Maz, Mes

NZ

Maz

Maz

Maz

Chi

NZ

Mix, NZ

Ornamental

Construction, Firewood

Firewood, Handcraft

Construction

Medicinal, Ornamental

Edible

Medicinal

Ornamental

Firewood, Medicinal

Medicinal

Ceremonial, Medicinal, Ornamental Ornamental

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 831

Family Iridaceae

Iridaceae

Iridaceae

Iridaceae

Juglandaceae

Juglandaceae

Lamiaceae

Lamiaceae

Lamiaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Neomarica variegata (M. Martens & Galeotti) Henrich & Goldblatt Rigidella inusitata Cruden Sisyrinchium angustissimum (B.L. Rob. & Greenm.) Greenm. & C.H. Thomps. Tritonia x crocosmiiflora G. Nicholson Carya illinoinensis (Wangenh.) K. Korch Oreomunnea mexicana (Standl.) J.F. Leroy Clinopodium macrostemum (Moc. & Sessé ex Benth.) Kuntze Cunila lythrifolia Benth. Hyptis mutabilis (Rich.) Briq.

Infraspecies

OF

OF

MCF, PF

MCF

MCF

RMV, PF

MCF

OF

Vegetation types TEF

Maz, NZ

Maz

Cui, Mix, Mxt, NZ

Chi

Cui

Mix, Nah

Mxt

Cui

Ethnic groupS Chi

Medicinal

Medicinal

Edible, Medicinal

Firewood

Timber

Ornamental

Ornamental

Edible

Use Categories Ceremonial

Risk status IUCN / SEMARNAT

832 B. Rendo´n-Aguilar et al.

Lamiaceae

Lamiaceae

Lamiaceae

Lamiaceae Lamiaceae

Lamiaceae

Lamiaceae

Lauraceae

Lauraceae

Lauraceae

Lauraceae

Lauraceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Litsea guatemalensis Mez Litsea neesiana (S. Schauer) Hemsl.

Beilschmiedia ovalioides Sach. Nishida Cinnamomum areolatum (Lundell) Kosterm. Litsea glaucescens Kunth

Ocimum campechianum Mill. Ocimum micranthum Willd. Salvia albocaerulea Linden Salvia circinnata Cav. Salvia lavanduloides Kunth Salvia polystachia Cav. Salvia purpurea Cav.

PF

PF

GRS, MCF, OF, PF, TSF

MCF

MCF

MCF, PF

MCF

OF OF

MCF

MCF

MCF

Mix

Chi, NZ

Cui, IZ, Maz, Mes, Mix, Nah, NZ

IZ

Mes

Cui, Mes

Maz

Mes Mix

Maz

Chi

Mes

Ceremonial, Construction, Edibleble, Firewood, Handcraft, Medicinal, Ornamental Ceremonial, Edible, Medicinal, Ornamental Firewood

Environmental

Ceremonial, Ornamental Edible

Medicinal

Medicinal Medicinal

Ornamental

Edible

Medicinal

(continued)

- / Danger of extinction

Ethnobotanical Science in the Clouds: Useful Plants of. . . 833

Family Lauraceae

Lauraceae

Lauraceae

Lauraceae

Lauraceae

Lauraceae

Lentibulariaceae

Lentibulariaceae

Loranthaceae

Loranthaceae

Loranthaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Ocotea acuminatissima (Lundell) Rohwer Ocotea effusa (Meisn.) Hemsl. Ocotea sarcodes Lorea-Hern. Persea americana Mill. Persea pallescens (Mez) Lorea-Hern. Persea schiedeana Nees Pinguicula lilacina Schltdl. & Cham. Pinguicula moranensis Kunth Psittacanthus calyculatus (DC.) G. Don Struthanthus hartwegii (Benth.) Standl. Struthanthus quercicola (Schltdl. & Cham.) Blume

Infraspecies

OF

MCF, PF

SHR

OF, PF

PF

RMV

MCF

TEF, TSF

OF

TEF

Vegetation types MCF

Maz

NZ

NZ

Maz, Mix

Mix

Mix

Chi, Mes, Mix Maz

Maz

Mix

Ethnic groupS Cui

Medicinal

Handcraft

Environmental, Medicinal Medicinal

Medicinal

Edible

Fodder

Edible, Firewood

Timber

Firewood

Use Categories Firewood, Fodder

Vulnerable / -

Risk status IUCN / SEMARNAT

834 B. Rendo´n-Aguilar et al.

Lythraceae

Lythraceae

Magnoliaceae

Malphigiaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Helicteres baruensis Jacq. Helicteres guazumifolia Kunth Heliocarpus aff. mexicanus (Turcz.) Sprague

Chiranthodendron pentadactylon Larreat. Guazuma ulmifolia Lam.

Cuphea aequipetala Cav. Cuphea hyssopifolia Kunth Magnolia aff. schiedeana Schltdl. Byrsonima crassifolia (L.) Kunth Anoda cristata (L.) Schltdl. Ceiba aesculifolia (Kunth) Britten & Baker f. Ceiba parvifolia Rose

OF

OF

PF

TDF, TEF, TSF, TTF

PF

TDF, TTF

MCF, OF, PF OF, RMV, TSF TDF

MCF

MCF

OF, PF

NZ

Zoq

Mix

IZ, Maz, Mes, Mix, NZ, Zoq

NZ

IZ, Maz

IZ

Mix, NZ, Zoq Mix, NZ

IZ

Maz

Mix, NZ

Environmental

Medicinal

Environmental, Firewood, Fodder, Handcraft, Medicinal, Veterinary medicine Handcraft

Edible, Firewood, Handcraft Medicinal

Edible, Firewood, Medicinal Medicinal, Veterinary medicine Construction

Medicinal, Ornamental

Medicinal

Medicinal

(continued)

-/ Threatened

Vulnerable / Threatened

Ethnobotanical Science in the Clouds: Useful Plants of. . . 835

Family Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae Malvaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Table 2 (continued)

Hibiscus costatus A. Rich. Luehea candida (DC.) Mart. Luehea speciosa Willd. Malva parviflora L. Malvaviscus aff. lanceolatus Rose Malvaviscus arboreus Cav. Ochroma pyramidale (Cav. ex Lam.) Urb. Robinsonella mirandae Gómez Pompa Sida acuta Burm. f. Sida haenkeana C. Presl

Scientific name Heliocarpus americanus L. Heliocarpus appendiculatus Turcz.

Infraspecies

TSF OF, TDF, TEF

TEF

TEF

TEF

PF TEF

OF

TEF, TSF

Vegetation types MCF, PF, TSF MCF, OF, RMV, TDF, TEF, TSF PF

NZ Chi, Mix, NZ

Zoq

Zoq

Chi, Zoq

NZ Zoq

Maz, Mix

Chi, Maz

Ethnic groupS Mes, Mix, NZ Chi, IZ, Mes, Mix, NZ Mix

Handcraft Handcraft, Medicinal

Medicinal

Handcraft

Medicinal

Medicinal Handcraft

Construction, Firewood, Timber Construction

Use Categories Construction, Firewood, Handcraft Construction, Firewood, Handcraft, Medicinal Ornamental

Vulnerable / -

Risk status IUCN / SEMARNAT

836 B. Rendo´n-Aguilar et al.

Malvaceae

Malvaceae

Malvaceae

Malvaceae

Malvaceae Marantaceae

Marantaceae

Martyniaceae Melastomataceae

Melastomataceae

Melastomataceae

Melastomataceae

Melastomataceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Trichospermum grewiifolium (A. Rich.) Kosterm. Triumfetta bogotensis DC. Triumfetta semitriloba Jacq. Waltheria indica L. Calathea aff. crotalifera S. Watson Calathea lutea (Aubl.) Schult. Martynia annua L. Arthrostemma ciliatum Pav. ex D. Don Arthrostemma primaevum Almeda Centradenia inaequilateralis (Schltdl. & Cham.) G. Don Clidemia sericea D. Don Clidemia setosa (Triana) Gleason

Sida rhombifolia L.

TSF

MCF, PF

MCF

MCF

TDF OF

TEF, TSF

TDF TSF

TDF

MCF

OF

MCF, OF, RMV, TSF

Chi

Mix, NZ

IZ

Chi

Maz Cui

Chi, Mix

Maz Maz

Mix

Mix

Chi, Cui, Mix, Nah, NZ Mix

Ceremonial

Edible

Ornamental

Edible, Medicinal

Medicinal Medicinal

Handcraft

Medicinal Edible

Medicinal

Handcraft

Handcraft

Handcraft, Medicinal

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 837

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Division Magnoliophyta

Family Scientific name Melastomataceae Conostegia icosandra (Sw. ex Wikstr.) Urb. Melastomataceae Conostegia xalapensis (Bonpl.) D. Don ex DC. Melastomataceae Heterocentron axillare Naudin Melastomataceae Heterocentron macrostachyum Naudin Melastomataceae Heterocentron subtriplinervium (Link & Otto) A. Braun & C.D. Bouché Melastomataceae Leandra subseriata (Naudin) Cogn. Melastomataceae Miconia aff. Oligotricha (DC.) Naudin Melastomataceae Miconia affinis DC. Melastomataceae Miconia argentea (Sw.) DC. Melastomataceae Miconia glaberrima (Schltdl.) Naudin

Table 2 (continued)

Infraspecies

MCF

MCF TEF

OF

MCF

MCF

RMV, MCF

MCF

MCF, OF, TSF

Vegetation types MCF, PF

NZ

Cui Zoq

NZ

Maz

IZ

NZ

Mix

Chi, Mes, NZ

Ethnic groupS Chi, Mix

Timber

Handcraft Construction

Firewood

Medicinal

Edible, Medicinal

Edible

Ceremonial, Medicinal

Edible, Firewood

Use Categories Firewood

Risk status IUCN / SEMARNAT

838 B. Rendo´n-Aguilar et al.

Menispermaceae

Menyanthaceae

Moraceae

Moraceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Cissampelos pareira L. Nymphoides fallax Ornduff Brosimum alicastrum Sw. Dorstenia drakena L.

Swietenia humilis Zucc.

Meliaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Melastomataceae Miconia ibaguensis (Bonpl.) Triana Melastomataceae Miconia impetiolaris (Sw.) D. Don ex DC. Melastomataceae Miconia liebmannii Cogn. Melastomataceae Miconia mexicana (Bonpl.) Naudin Melastomataceae Miconia minutiflora (Bonpl.) DC. Melastomataceae Miconia serrulata (DC.) Naudin Melastomataceae Tibouchina scabriuscula (Schltdl.) Cogn. Meliaceae Cedrela odorata L.

Magnoliophyta

OF, TEF, TSF TDF

SAV

MCF

TDF

MCF, TDF, TEF, TSF

OF

TEF

ECT (MCF-OF) MCF

MCF

TEF

PF

IZ

Chi, Mix

Maz

Mes

Maz, NZ

IZ, Maz, Mix, Zoq

Nah

Zoq

NZ

Maz

Chi, Cui

Chi

Mix

Edible, Firewood, Timber Handcraft

Fodder

Construction, Firewood, Handcraft, Timber Medicinal

Construction, Timber

Medicinal

Edible, Handcraft

Handcraft

Edible

Ceremonial, Handcraft

Ornamental

Firewood

(continued)

Vulnerable / Special protection Vulnerable / -

Ethnobotanical Science in the Clouds: Useful Plants of. . . 839

Family Moraceae Moraceae

Moraceae Moraceae

Moraceae

Moraceae

Muntingiaceae

Myricaceae

Myricaceae Myrtaceae

Myrtaceae

Myrtaceae

Myrtaceae

Division Magnoliophyta Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Ficus aurea Nutt. Ficus lapathifolia (Liebm.) Miq. Ficus pertusa L. f. Poulsenia armata (Miq.) Standl. Pseudolmedia oxyphyllaria Donn. Sm. Trophis mexicana (Liebm.) Bureau Muntingia calabura L. Morella cerifera (L.) Small Myrica cerifera L. Myrcia splendens (Sw.) DC. Myrcianthes fragrans (Sw.) McVaugh Pimenta dioica (L.) Merr. Psidium friedrichsthalianum (O. Berg) Nied.

Infraspecies

ECT (TDF-OF)

MCF

TDF

OF MCF

TDF, TEF, TTF MCF, PF

TEF

TEF

TSF TEF

Vegetation types PF TEF

Maz

Zoq

NZ

Chi, IZ, Maz, Mix Chi, Mix, NZ Cui NZ

Chi

Mix

Mes, NZ Zoq

Ethnic groupS Cui Zoq

Edible

Medicinal

Handcraft

Construction, Firewood, Handcraft Firewood Handcraft

Edible, Firewood

Edible

Firewood

Firewood, Medicinal Fodder

Use Categories Construction, Edible Medicinal

Vulnerable / -

Risk status IUCN / SEMARNAT

840 B. Rendo´n-Aguilar et al.

Myrtaceae

Myrtaceae

Myrtaceae

Myrtaceae

Myrtaceae

Onagraceae

Onagraceae

Onagraceae

Onagraceae

Orchidaceae

Orchidaceae

Orchidaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Syzygium jambos (L.) Alston Ugni myricoides (Kunth) O. Berg Fuchsia microphylla Kunth Fuchsia paniculata Lindl. Lopezia racemosa Cav. Oenothera rosea L’Hér. ex Aiton Bletia greenmaniana L.O. Williams Calanthe calanthoides (A. Rich. & Galeotti) Hamer & Garay Dichaea glauca (Sw.) Lindl.

Psidium guineense Sw. Psidium sartorianum (O. Berg) Nied.

Psidium guajava L.

MCF, OF, PF

MCF, TSF

OF

MCF, TDF

MCF, OF

PF

MCF

PF

OF

MCF, TDF, TSF PF, TDF, TSF

MCF, OF, PF , TEF

Cui, Maz, NZ

Cui, Maz

Mix

Mix, Nah, NZ Maz, Mxt

Mes

Nah

Chi

Mix

Mix, NZ

Chi, Cui, Maz, Mix, NZ Maz

Medicinal, Ornamental

Ceremonial, Environmental Ornamental

Fodder, Medicinal, Ornamental Medicinal

Ornamental

Medicinal

Edible, Medicinal

Construction, Edible, Firewood, Handcraft, Medicinal Edible

Edible, Medicinal

Edible, Firewood, Medicinal

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 841

Family Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Epidendrum longipetalum A. Rich. & Galeotti Epidendrum nitens Rchb. f. Epidendrum pseudoramosum Schltr. Epidendrum radicans Pav. ex Lindl. Epidendrum tuxtlense Hágsater, García-Cruz & L. Sánchez

Scientific name Dichaea graminoides (Sw.) Lindl. Dichaea neglecta Schltr. Dichaea trichocarpa (Sw.) Lindl. Elleanthus cynarocephalus (Rchb. f.) Rchb. f. Epidendrum camposii Hágsater Epidendrum ciliare L.

Infraspecies

MCF, OF, PF OF

OF

OF

ECT (TDF-PF) OF

OF

MCF

MCF

MCF

Vegetation types MCF

Maz

Chi, Mix

Maz

Maz

Maz

Mix

Maz

IZ

Maz

IZ

Ethnic groupS Mix

Ornamental

Handcraft, Ornamental

Ornamental

Ornamental

Ornamental

Handcraft

Ornamental

Ornamental

Ornamental

Ornamental

Use Categories Environmental

Risk status IUCN / SEMARNAT

842 B. Rendo´n-Aguilar et al.

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Epidendrum verrucosum Sw. Gongora galeata (Lindl. ex Lodd.) Rchb f. Goodyera cf. striata Rchb. f. Jacquiniella teretifolia (Sw.) Britton & P. Wilson Maxillaria variabilis Bateman ex Lindl. Prosthechea brassavolae (Rchb. f.) W.E. Higgins Prosthechea cochleata (L.9 W.E. Higgins Prosthechea concolor (Lex.) W.E. Higgins Prosthechea ochracea (Lindl.) W.E. Higgins Prosthechea rhynchophora (A. Rich. & Galeotti) W.E. Higgins Prosthechea varicosa (Bateman ex Lindl.) W.E. Higgins MCF, OF, PF

PF

OF, RMV

PF

MCF, TSF

MCF

MCF, TSF

PF

MCF

MCF

MCF

Maz, Mes, Mxt

Mix

Maz, Mix

NZ

Maz, Mes

IZ

Maz, Mix

Mix

Chi

Maz

IZ

Handcraft, Ornamental

Environmental

Ornamental

Edible, Ornamental

Ornamental

Environmental, Ornamental Ornamental

Ornamental

Ornamental

Handcraft

Environmental

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 843

Family Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orchidaceae

Orobanchaceae

Orobanchaceae

Oxalidaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Prosthechea vitellina (Lindl.) W.E. Higgins Rhynchostele cervantesii (Lindl.) Soto Arenas & Salazar Sobralia macrantha Lindl. Trichocentrum cosymbephorum (C. Morren) R. Jiménez & Carnevali Trichocentrum pachyphyllum (Hook.) R. Jiménez & Carnevali Vanilla planifolia Andrews Lamourouxia viscosa Kunth Melasma physalodes (D. Don) Melch. Oxalis corniculata DC. subsp. membranacea

Infraspecies

PF, TSF

TSF

OF

MCF, TEF

MCF

MCF, OF, TSF TSF

PF

Vegetation types MCF

NZ

NZ

Mix

Chi, Zoq

Maz

IZ, Maz, Mix Mix

NZ

Ethnic groupS Nah

Edible

Medicinal

Medicinal

Edible, Handcraft

Ornamental

Ornamental

Ornamental

Ornamental

Use Categories Ornamental

- / Special protection

Risk status IUCN / SEMARNAT - / Special protection -/ Threatened

844 B. Rendo´n-Aguilar et al.

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Oxalis intermedia A. Rich. Papaveraceae Argemone mexicana L. Papaveraceae Argemone platyceras Link & Otto Papaveraceae Bocconia arborea S. Watson Papaveraceae Bocconia frutescens L. Pentaphylacaceae Ternstroemia aff. sylvatica Schltdl. & Cham. Pentaphylacaceae Ternstroemia lineata DC. Pentaphylacaceae Ternstroemia tepezapote Schltdl. & Cham. Phyllonomaceae Phyllonoma laticuspis (Turcz.) Engl. Phytolaccaceae Petiveria alliacea L. Phytolaccaceae Phytolacca americana L. Phytolaccaceae Phytolacca icosandra L. Phytolaccaceae Phytolacca rivinoides Kunth & C.D. Bouché

Oxalidaceae

MCF, OF, PF, SHR MCF, TEF

MCF, OF, PF TTF MCF

MCF

MCF, OF

OF

ECT (PF-TDF) MCF, PF, TSF MCF, TSF

SHR

OF

Chi, Cui, NZ

Nah, NZ

Maz, Mes, Mix, NZ Mes Maz

Maz, Mix, Nah Maz

Chi

Chi, Cui, Maz, Mes Maz

NZ

NZ

Mix

Edible, Handcraft

Edible, Handcraft

Firewood, Handcraftr, Medicinal Medicinal Edible

Medicinal

Medicinal

Medicinal

Medicinal

Medicinal

Medicinal

Medicinal

Ceremonial

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 845

Family Phytolaccaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Phytolacca rugosa A. Braun & C.D. Bouché Peperomia aff. glabella (Sw.) A. Dietr. Peperomia aff. glabra C. DC. Peperomia alpina (Sw.) A. Dietr. Peperomia blanda (Jacq.) Kunth Peperomia clavigera Standl. & Steyerm. Peperomia hernandiifolia (Vahl) A. Dietr. Peperomia leptophylla Miq. Peperomia maculosa (L.) Hook. Peperomia obtusifolia (L.) A. Dietr. Peperomia pecuniifolia Trel. & Standl.

Infraspecies

PF

MCF

MCF, OF

MCF, OF

MCF

TSF

MCF

TEF

MCF

TEF

Vegetation types OF, PF, RMV

NZ

IZ

IZ, Mix

Chi, Maz

Chi, Mes, Mix

Chi, Cui, Maz Chi

Mix

Cui

Chi

Ethnic groupS Chi, Nan

Edible

Ornamental

Edible, Ornamental

Ceremonial, Medicinal

Edible

Medicinal

Ceremonial, Medicinal

Medicinal

Ornamental

Medicinal

Use Categories Edible, Fodder

Risk status IUCN / SEMARNAT

846 B. Rendo´n-Aguilar et al.

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae

Piperaceae Piperaceae

Piperaceae

Piperaceae Piperaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Piper chinantlense M. Martens & Galeotti Piper peltatum L. Piper schiedeanum Steud.

Piper aff. uhdei C. DC. Piper amalago L. Piper auritum Kunth

Peperomia peltilimba C. DC. ex Trel. Peperomia pseudoalpina Trel. Peperomia quadrangularis (J.V. Thomps.) A. Dietr. Peperomia quadrifolia (L.) Kunth Peperomia rotundifolia (L.) Kunth Piper aduncum L.

TEF TEF

TSF MCF, OF, PF, RMV, TDF, TEF, TSF TEF

TEF

MCF, OF, TEF, TSF

MCF

MCF

OF, PF, TSF

MCF

TEF

Chi Chi

Mix Cui, Maz, Mes, Mix, Nah, NZ, Zoq Chi

Zoq

Chi, NZ

Cui

Maz

Maz, Mes, NZ

IZ

Chi

Veterinary medicine Medicinal

Firewood, Medicinal

Ceremonial Edible, Medicinal

Construction, Environmental, Medicinal Firewood

Ornamental

Ornamental

Ceremonial, Edible

Ornamental

Medicinal

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 847

Family Piperaceae

Plantaginaceae

Plantaginaceae

Plantaginaceae

Plantaginaceae

Plantaginaceae Plantaginaceae

Platanaceae

Poaceae

Poaceae

Poaceae

Poaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Melinis repens (Willd.) Zizka Muhlenbergia versicolor Swallen

Russelia coccinea (L.) Wettst. Russelia sarmentosa Jacq. Scoparia dulcis L. Stemodia verticillata (Mill.) Hassl. Platanus mexicana Moric. Andropogon bicornis L. Arundo donax L.

Plantago australis Lam. Plantago major L.

Scientific name Piper umbellatum L.

Infraspecies

PF

RMV, TDF, TSF TSF

TEF, TSF

RMV

MCF TSF

MCF

MCF, TEF

MCF, TSF

OF

Vegetation types MCF, TEF

Mix

Maz, Mix, Nah, NZ NZ

Chi, NZ

Maz, Nah

NZ Maz

IZ

Chi, Cui, Maz, Mix, Nah Chi, Mix

Ethnic groupS Chi, Maz, NZ Maz, NZ

Construction

Construction, Fodder, Handcraft, Medicinal Fodder

Construction, Firewood, Timber Handcraft

Medicinal Ceremonial

Medicinal

Medicinal

Medicinal

Medicinal

Use Categories Medicinal

Least concern / -

Least concern /-

Risk status IUCN / SEMARNAT

848 B. Rendo´n-Aguilar et al.

Poaceae

Polygalaceae

Polygalaceae

Polygalaceae

Polygonaceae

Polygonaceae

Polygonaceae Polygonaceae Polygonaceae

Primulaceae

Primulaceae

Primulaceae

Primulaceae

Primulaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Ardisia compressa Kunth Jacquinia macrocarpa Cav. Myrsine coriacea (Sw.) R. Br. ex Roem. & Schult. Parathesis rekoi Standl.

Pennisetum purpureum Schumach. Monnina xalapensis Kunth Polygala floribunda Benth. Polygala paniculata L. Coccoloba liebmannii Lindau Muehlenbeckia tamnifolia (Kunth) Meisn. Rumex crispus L. Rumex obtusifolius L. Ruprechtia fusca Fernald Anagallis arvensis L.

MCF

MCF, OF

MCF, OF, TSF MCF, PF, TEF TTF

MCF OF TTF

OF

TDF, TSF

MCF

OF

PF

RMV, TSF

Mix

NZ

Chi, IZ, NZ IZ

Cui, Mix

Mix Nah Mes

Nah

IZ, Zoq

Chi

Mix

Mix

Nah, NZ

Construction, Firewood, Handcraft, Timber Firewood

Handcraft

Edible, Medicinal

Fodder Edible Construction, Medicinal Ceremonial, Medicinal

Handcraft

Edible

Ceremonial, Handcraft, Medicinal Medicinal

Fodder

Fodder

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 849

Family Primulaceae

Ranunculaceae

Rhamnaceae

Rhamnaceae

Rhamnaceae

Rhamnaceae

Rhamnaceae

Rhamnaceae

Rosaceae

Rosaceae

Rosaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Samolus floribundus Kunth Clematis grossa Benth Adolphia infesta (Kunth) Meisn. Ceanothus caeruleus Lag. Colubrina elliptica (Sw.) Brizicky & W.L. Stern Gouania stipularis DC. Rhamnus longistyla C.B. Wolf Ziziphus amole (Sessé & Moc.) M.C. Johnst. Alchemilla pectinata Kunth Amelanchier denticulata (Kunth) K. Koch Cercocarpus macrophyllus C.K. Schneid.

Infraspecies

OF, PF, TDF

OF, PF

ODC

ECT (MCF-PF) TDF

TSF

TSF

RMV

TDF

MCF

Vegetation types TSF

Mix, NZ

NZ

Nah

Maz

NZ

NZ

NZ

Mix

Maz

Cui

Ethnic groupS NZ

Construction, Firewood, Fodder, Handcraft, Timber

Handcraft

Fodder

Handcraft, Medicinal

Firewood

Medicinal

Firewood

Medicinal

Medicinal

Veterinary medicine

Use Categories Edible

Risk status IUCN / SEMARNAT

850 B. Rendo´n-Aguilar et al.

Rosaceae

Rosaceae

Rosaceae Rosaceae

Rosaceae

Rosaceae

Rosaceae

Rosaceae Rosaceae

Rosaceae Rosaceae

Rosaceae

Rubiaceae

Rubiaceae

Rubiaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Rubus aff. cymosus Rydb. Rubus eriocarpus Liebm. Rubus nelsonii Rydb. Rubus philyrophyllus Rydb. Rubus pringlei Rydb. Rubus sapidus Schltdl. Rubus urticifolius Poir. Arachnothryx scabra (Hemsl.) Borhidi Borreria remota (Lam.) Bacigalupo & E.L. Cabral Bouvardia ternifolia (Cav.) Schltdl.

Crataegus mexicana DC. Eriobotrya japonica (Thunb.) Lindl. Fragaria vesca L. Prunus brachybotrya Zucc. Rubus adenotrichos Schltdl.

PF

RMV

TEF

MCF

OF, PF OF, PF

MCF OF, PF, TSF

TEF

TDF

MCF, OF, PF, RMV

MCF, PF PF

TDF

PF

NZ

NZ

Chi

Maz

Maz, NZ Cui, Mix, Nah Maz, NZ Mes, Mix

Mix

Cui, Maz, Mix, Nah, NZ Mxt

Cui, NZ NZ

Maz

NZ

Medicinal

Medicinal

Firewood

Edible

Edible Edible

Edible Edible

Edible

Edible

Edible

Edible Handcraft

Edible

Edible

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 851

Family Rubiaceae

Rubiaceae

Rubiaceae Rubiaceae

Rubiaceae

Rubiaceae

Rubiaceae

Rubiaceae

Rubiaceae

Rubiaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Palicourea tetragona (Donm. Sm.) C.M. Taylor & Lorence Psychotria elata (Sw.) Hammel Psychotria poeppigiana Müll. Arg.

Scientific name Coccocypselum hirsutum Bartl. ex DC: Crusea calocephala DC. Crusea coccinea DC. Deppea grandiflora Schltdl. Diodella sarmentosa (Sw.) Bacigalupo & E, L. Cabral Galium hypocarpium (L.) Endl. ex Griseb. Hamelia patens Jacq.

Infraspecies

MCF, TEF, TSF TSF

TDF, TEF, TSF TEF

MCF

PF

MCF MCF

TEF

Vegetation types MCF, OF, TEF

Chi, IZ, NZ Chi

Chi, Maz, Mix, Zoq Chi

Maz, Nah

NZ

Chi Maz

Zoq

Ethnic groupS Chi, Mix, NZ

Medicinal

Ornamental

Firewood

Firewood, Medicinal

Medicinal, Ornamental

Medicinal

Ornamental Ornamental

Edible

Use Categories Ceremonial, Medicinal

Risk status IUCN / SEMARNAT

852 B. Rendo´n-Aguilar et al.

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Psychotria tomentosa (Oerst.) Hemsl. Rubiaceae Randia aff. oaxacana Standl. Rutaceae Zanthoxylum fagara (L.) Sarg. Salicaceae Salix humboldtiana Willd. Santalaceae Phoradendron falcatum Eichler Sapindaceae Dodonaea viscosa Jacq. Sapindaceae Sapindus saponaria L. Sapindaceae Serjania triquetra Radlk. Sapotaceae Manilkara zapota (L.) P. Royen Sapotaceae Pouteria sapota (Jacq.) H.E. Moore & Stearn Sapotaceae Sideroxylon capiri (A. DC.) Pittier Scrophulariaceae Buddleja americana L. Scrophulariaceae Buddleja cordata Kunth Scrophulariaceae Buddleja sessiliflora Kunth

Rubiaceae

subsp. cordata

SHR

PF

MCF

TDF

OF, TEF, TSF

TEF

TDF

OF, TSF

PF, TDF

OF

RMV, SHR

TDF, TSF

TEF

TEF

NZ

Mix, NZ

Maz, Mxt

Maz

Chi, IZ, Mix

Chi

Mes, Mix, NZ, Zoq Maz

NZ

Maz

IZ, NZ

Maz, NZ

Zoq

Chi

Handcraft

Construction, Firewood

Medicinal

Edible, Firewood

Construction, Edible, Medicinal

Edible, Firewood

Medicinal

Ceremonial, Firewood, Medicinal Handcraft, Medicinal

Firewood, Handcraftr, Timber Firewood, Medicinal, Ornamental Medicinal

Edible

Ornamental

(continued)

-/ Threatened

Ethnobotanical Science in the Clouds: Useful Plants of. . . 853

Family Siparunaceae

Siparunaceae

Siparunaceae

Siparunaceae

Smilacaceae

Smilacaceae

Smilaceceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Cyphomandra betacea (Cav.) Miers Datura stramonium L.

Scientific name Siparuna andina (Tul.) A. DC. Siparuna austromexicana Lorence Siparuna gesnerioides (Kunth) A. DC. Siparuna thecaphora (Poepp. & Endl.) A. DC. Smilax jalapensis Schltdl. Smilax moranensis M. Martens & Galeotti Smilax regelii Killip & C.V. Morton Brugmansia x candida Pers. Cestrum nocturnum L.

Infraspecies

RMV

MCF, OF, PF, TEF, TSF TSF

TSF

MCF

OF

MCF

MCF, TSF

TEF

MCF, TEF

Vegetation types TSF

IZ

Mix

Chi, Cui, Mix, NZ

Mix

Mix

Maz

Mxt

IZ, Mix

Mix

Chi, Mes, Mix

Ethnic groupS Chi

Ceremonial, Medicinal

Edible

Edible, Medicinal, Ornamental

Fodder

Handcraft

Construction

Construction

Ceremonial, Eddible, Medicinal

Medicinal

Ceremonial, Firewood, Medicinal

Use Categories Medicinal

Risk status IUCN / SEMARNAT

854 B. Rendo´n-Aguilar et al.

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Solanum appendiculatum Dunal Solanum chrysotrichum Schltdl.

Jaltomata procumbens (Cav.) J.L. Gentry Lycianthes heteroclita (Sendtn.) Bitter Lycopersicon esculentum Mill. Nicandra physalodes (L.) Gaertn. Nicotiana glauca Graham Nicotiana tabacum L. Physalis angustiphysa Waterf. Physalis lagascae Roem. & Schult. Physalis orizabae Dunal Physalis vel. aff. stapelioides (Regel) Bitter Solanum americanum Mill.

PF, RMV

MCF, OF, PF, RMV, TEF, TSF PF

PF

TSF

OF

ECT (PF-TDF) TSF PF

PF

TSF

TSF

TSF

Mix

Chi, Cui, Maz, Mes, Mix, NZ NZ

NZ

Cui

Mix

Maz Cui

NZ

Mix

NZ

Mix

Mix

Medicinal

Edible

Edible, Medicinal

Edible

Edible

Edible

Ceremonial, Medicinal Edible

Ceremonial

Veterinary medicine

Edible

Edible

Edible

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 855

Family Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Solanaceae

Ticodendraceae

Typhaceae

Urticaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Solanum lanceolatum Cav. Solanum lycopersicum L. Solanum nigrescens M. Martens & Galeotti Solanum nigricans M. Martens & Galeotti Solanum pubigerum Dunal Solanum rudepannum Dunal Witheringia solanacea L’Hér. Ticodendron incognitum GómezLaur. & L.D. Gómez Typha domingensis Pers. Pilea microphylla (L.) Liebm.

Infraspecies

TSF

SRV

MCF

TSF

MCF, TSF

OF

OF

OF, PF, TDF, TSF

TDF

Vegetation types MCF, OF

Chi, Maz

IZ

Cui

Cui, Maz, Zoq Maz

Mix

Nah

Chi, Maz, NZ

Mix

Ethnic groupS Maz

Medicinal

Ornamental

Firewood

Ceremonial, Handcraft, Medicinal Edible

Edible

Firewood

Edible, Medicinal

Edible

Use Categories Ceremonial, Medicinal

Least concern / -

Vulnerable / -

Risk status IUCN / SEMARNAT

856 B. Rendo´n-Aguilar et al.

Urticaceae

Verbenaceae

Verbenaceae

Verbenaceae

Verbenaceae

Verbenaceae

Verbenaceae

Verbenaceae

Violaceae

Vitaceae

Vitaceae

Vitaceae

Winteraceae

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Magnoliophyta

Verbena litoralis Kunth Rinorea hummelii Sprague Vitis bourgaeana Planch. Vitis popenoei J.L. Fennell Vitis tiliifolia Humb. & Bonpl. ex Schult. Drimys granadensis L. f.

Verbena carolina L.

Lippia berlandieri Schauer Lippia umbellata Cav.

Pilea pansamalana Donn. Sm. Citharexylum mocinnoi D. Don. Lantana aff. Velutina M. Martens & Galeotti Lantana camara L.

MCF, PF, TSF MCF, OF

MCF

TEF, TSF

TEF

OF, RMV, TSF MCF, RMV, TSF OF

TDF

OF, TSF

TDF

MCF, OF

TEF

Cui, Mix, NZ Maz

Chi, Maz, Mix, Zoq Mix

Chi

Maz, Mes, NZ Maz

Mix, NZ

Maz, Mix, NZ Maz

Maz

Maz

Chi

Edible, Handcraft, Medicinal Medicinal, Timber

Medicinal

Edible, Medicinal

Construction, Firewood

Medicinal

Medicinal

Firewood, Medicinal

Handcraft, Medicinal, Ornamental Edible, Medicinal

Medicinal

Handcraft

Medicinal

(continued)

Ethnobotanical Science in the Clouds: Useful Plants of. . . 857

Family Zingiberaceae

Zygophyllaceae

Zygophyllaceae

Division Magnoliophyta

Magnoliophyta

Magnoliophyta

Table 2 (continued)

Scientific name Renealmia mexicana Klotzsch ex Petersen Guaiacum coulteri A. Gray Tribulus cistoides L.

Infraspecies

ECT (PF-TDF)

TTF

Vegetation types MCF

NZ

IZ

Ethnic groupS IZ, Mix

Firewood, Handcraftr, Ornamental Medicinal

Use Categories Medicinal, Ornamental

Lower risk / Threatened

Risk status IUCN / SEMARNAT

858 B. Rendo´n-Aguilar et al.

Ethnobotanical Science in the Clouds: Useful Plants of. . .

859

Fig. 5 Use categories with highest number of useful species in northeastern Oaxaca

Liquidambar styraciflua L. is used by almost all ethnic groups (e.g., ceremonial, construction, firewood, medicinal, ornamental, timber) even when both species are widely distributed (Table 2). (c) Common names in Spanish and indigenous languages. Regarding the 804 species here reported (with their respective subspecies and varieties), 149 (18.5%) have a name solely in an indigenous language; 443 (55%) have a name in Spanish and only in one indigenous language. This means that 592 species (73.6%) are named in at least one of the indigenous languages; 129 species (16%) have only a name in Spanish and 83 species have not a common name. Of the 592 taxa that are named in one or more languages, 419 species (70.8%) are named only in one; 108 (18.2%) are named in two; 39 species (6.6%) have names in three; 14 (2.4%) are named in four; six species (1%) are named in five, and only four species (0.7%) have names in six. It should be mentioned that the latter correspond to widely used species throughout the study area, such as Alnus acuminata, Piper auritum Kunth, Litsea glaucescens Kunth, and Quercus laurina Raf.

Distribution of Traditional Knowledge Among Ethnic Groups Variation in number of useful families and species among ethnic groups was detected. Those with the greatest number of useful species and botanical families are the Northern Zapotec, Mixe, and Mazatec people (Fig. 6). A proportional relation

B. Rendo´n-Aguilar et al.

860

between the number of useful species and families and the number of municipalities inhabited by each ethnic group is observed, but only in the case of ethnic groups with greater number of useful species. On the other hand, ethnic groups with the lowest values of useful species and families are also distributed in few municipalities, like Mixtec and Nahua people (Fig. 7). Some ethnic groups, like Chinantec or Mazatec NZ Mix Maz Chi IZ Mes Cui Mxt Zoq Nah 0

50

100 Number of families

150

200

250

Number of species

Fig. 6 Number of useful species distributed among Oaxacan ethnic groups. Acronyms: Chi: Chinantec; Cui: Cuicatec; IZ: Isthmus Zapotec; Maz: Mazatec; Mes: Mestizo; Mix: Mixe; Mxt: Mixtec; Nah: Nahua; NZ: Northern Zapotec; and Zoq: Zoque

Fig. 7 Number of species and families recorded among Oaxacan ethnic groups. Number of municipalities occupied by each ethnic group is also included

Ethnobotanical Science in the Clouds: Useful Plants of. . .

861

people, have a notorious contribution in terms of number of useful species and families, even when this information came from few municipalities. It should also be noted that the area potentially used by the Zoque people is the greatest of all the ethnic groups. When comparing the useful families of the species registered there, they are almost the same that the recorded for the rest of ethnic groups, such as Altingiaceae, Asparagaceae, Asteraceae, Fabaceae, and Piperaceae. Conversely, some species, and families, are used only by one ethnic group, and in one municipality, such as Oreomunnea mexicana (Standl.) J.-F. Leroy “caudillo, (leader)” (Juglandaceae), which is distributed and used exclusively in the Chinantla region (Santiago Comaltepec), and Magnolia schiedeana Schltdl. “flor del corazón, (heart flower)” (Magnoliaceae) in the Isthmus region (Santa María Guienagati), which, although people mention that it was present in several municipalities, was always referred to as occurring in remote areas near mountain peaks.

Discussion Geographical and Ecological Aspects of the Ethno-Floristic Sampling Approach The geological and physiognomic complexity of the state of Oaxaca undoubtedly have contributed to its floristic richness, and in this historic scenario many indigenous communities have appropriated, used, and transformed landscapes and local floras. During almost 4 years, localities with representation of 15 of the 26 vegetation types reported by Torres-Colín (2004) were explored, and the useful plants growing there were collected. The data obtained are significant in terms of useful species recorded, because only 14.7% of the total area of Oaxaca was visited (84 of 570 municipalities). Previous data suggest that a much larger sampling effort is needed to complete the ethnofloristic scene of Oaxaca. Results of this research also reinforce the statement that people of ethnic groups are local safeguards of biodiversity (Bernal-Ramírez et al. 2019; Boege 2008; Toledo and Barrera-Bassols 2008). Many areas visited were conceived as preserving areas, but most of them were modified 30 or 40 years ago, and currently are under natural regeneration process. During this time, people have maintained low levels of ecosystem management and use in these sites (commonly named “tierra en descanso, (land at rest)”). From these places, local people obtain firewood, medicinal plants, and sometimes hunt animals such as birds, lizards, and small mammals (BernalRamírez et al. 2019; Martínez-Bautista et al. 2019; Rendón-Aguilar et al. 2017b; Pérez-Nicolás et al. 2017), which is a very common activity throughout the study area. These patterns in regeneration processes indicate high rates of resilience of ecosystems, which needs to be analyzed profoundly in order to understand the kinds of management that people practice to recovering vegetation. Apparently, ethnic groups of Northeastern Oaxaca assist empirically to ecosystems restoration (GascaZamora et al. 2010; Pérez-Nicolás et al. 2017). Part of these processes are associated with migration phenomena and farm abandonment (Saynes-Vásquez et al. 2013), but

862

B. Rendo´n-Aguilar et al.

also with the indigenous initiatives for voluntary conservation of preserving areas (Monroy et al. 2015), the local conflicts to recover their own territories, the local policies related to territorial ordering, and the forest management programs, among others (Gasca-Zamora 2014; Gasca-Zamora et al. 2010).

Taxonomic Aspects of Useful Plants The most recent inventory of the flora of Oaxaca includes 10,229 vascular plant species (Villaseñor 2016). The present contribution served the purpose of completing this inventory including data about the distribution of the species and new records for the flora of Oaxaca, and also to register new uses of previously reported species, increasing ethnobotanical information, which is fundamental given the great cultural richness of this state. The number of families registered in the present study is ca. 50% of those reported by García-Mendoza and Meave (2012) (261 families without considering the mosses). This means that a high number of families have some use and are represented by at least one useful species. A similar situation happens at the genus level. If we consider the total number of records, ca. 20% have species with some utility. There are approximately 7,000 useful species in Mexico and in the Ethnobotanical Database of Mexican Plants (BADEPLAM by its Spanish acronym, Caballero and Cortés 2001); there are 3,500 species of vascular plants registered. The total number of useful species recorded in the present study increase in almost 10% of these records, considering the fact that our study took place in a single Mexican state, and that for many localities these are the first records of useful plants. The 12 best represented families, with the largest number of useful species is concordant with 50% of those reported by Caballero and Cortés (2001) for the country, and with 58% of the floristic inventory of Oaxaca (García-Mendoza and Meave 2012). In fact, species in many families are used only in the state of Oaxaca, among them Melastomataceae and Piperaceae, of which the information was substantially increased (Martínez-Bautista et al. 2019). Moreover, the number of useful species of Asteraceae also increased substantially in relation with those previously reported for Oaxaca (Caballero and Cortés 2001). Something similar happens with Fabaceae (Leguminosae). Interestingly, 34 (65.4%) of the 52 oak species known from Oaxaca (Valencia and Nixon 2004) were registered as useful plants, a situation that is not observed in the BADEPLAM (Caballero and Cortés 2001). Mexico is the country with greatest pine diversity in the world (Farjon and Styles 1997; Gernandt and Pérez-de la Rosa 2014), and Oaxaca is one of the three Mexican states with more species (Bernal-Ramírez et al. 2019) and 15 of the 24 native species (62.5%) were recorded as useful, like one of the two species of Abies Mill. (i.e., A. hickelii Flous & Gaussen), and fourteen species of Pinus (Bernal-Ramírez et al. 2019). In summary, the data of the number of useful species of gymnosperms recorded in this study exceeds the numbers registered in previous contributions.

Ethnobotanical Science in the Clouds: Useful Plants of. . .

863

Traditional Knowledge (a) Use categories. The importance of medicinal plants, as the most significant category, indicates that they are functional and effective. Although allopathic medicine is permeating culture of ethnic groups, people continue using plants to cure pain, illness, and psychosomatic ailments (García-Hernández et al. 2015; Pérez-Nicolás et al. 2017). Currently, different studies still continue demonstrating that medicinal plants are of daily basic use by ethnic groups (MagañaAlejandro et al. 2010; Pérez-Nicolás et al. 2017; Hernández-Pasteur et al. 2019). The role played by this group of plants goes beyond politic borders, to the extent that migrants seek professional help from traditional healers. The same goes for edible plants, the second most important use category recorded here. As a basic aspect of life, of plant components which provide different nutriments to people (e.g., proteins, vitamins, sugars), such as the edible fruits of Arctostaphylos pungens Kunth, Conostegia spp., Persea spp., Psidium spp., Rubus spp., Saurauia spp., and Spondias spp.; leaves of Cestrum nocturnum L., Cleome spp., Piper auritum Kunth, Peperomia spp.; and floral buds of Astrocaryum mexicanum Liebm. ex Mart., Cirsium pinetorum Greenm., Spathiphyllum cochlearispathum (Liebm.) Engl., Phaseolus coccineus L., Macleania insignis M. Martens & Galeotti, Agave spp., and Chamaedorea spp. The importance of these two categories is well known. Our results agree with previous studies that points to these uses as the most frequent in the ethnofloristic inventories (Blancas et al. 2010; Caballero et al. 2004; Lira et al. 2009; Luna-José and Rendón 2008; Martínez 1959; Martínez-Bautista et al. 2019; Rendón-Aguilar et al. 2017b; Vibrans 2016). Use categories represent tangible data of complex processes coming from field observations, and probing nature, so the concept of utility is not anthropocentric. In several occasions, local guides mentioned plants that were consumed by different animals, so people deduce whether a plant can be useful at some time of food shortage, or whether they are rather toxic plants, through biological interactions and observations. For example, Lopezia racemosa Cav., Poulsenia armata (Miq.) Standl., and Lysiloma spp., are taxa considered with high potential for food, whereas Serjania spp. are considered toxic plants. IbarraManríquez et al. (1997) report that the fruit of Poulsenia armata is edible in Veracruz, Mexico; this information supports that the knowledge acquisition occurs through nature observations by Oaxacan ethnic groups. This knowledge also reflects the holistic view behind the use of plant resources. For example, Quercus glaucoides M. Martens & Galeotti (Fagaceae) species located in the category of plants with environmental use is named “palo de agua, (water stick)” and people recognize it as a moisture indicator of the ground. People like the presence in the woods of bromeliads and orchids; from their own perspective, these plants are elements that accentuate the beauty of surrounding ecosystems. (b) Number of uses (exclusive species vs. multifunctional species). The record of a high percentage of species with a single use indicates that knowledge about plants is very fine and precise. This means that only a certain species satisfies a

864

B. Rendo´n-Aguilar et al.

particular need. In this context, several examples are found: Coccocypselum hirsutum Bartl. ex DC. (Rubiaceae) “bolita de miones, (pisser-ball)” is used only to cease the need to urinate at night in children without urinary sphincter control. In some cases, there are interesting associations equivalent to the doctrine of signatures, where physical characteristics of plants reveal their therapeutic value (Bennett 2007). For instance, the fern Hemionitis palmata L. (Pteridaceae) known as “susto de perro (dog scare)” is used exclusively when people get scared by a dog´s barking and curiously the frond form is like the footprint of a dog paw. Additionally, people associated odors with certain function of the plants. For example, Gaultheria acuminata Schldtl. & Cham. (Ericaceae), which is called “bengue” (a borrowed name), is used to alleviate muscle aches, and it smells and functions in a similar way as commercial ointments. Clidemia setosa (Triana) Gleason (Melastomataceae) called “remedio de mujer (woman’s remedy)” is used when men get frightened after they are discovered in extramarital sexual relations. In this case, people associate the tip part of petiole with the vagina shape. For these reasons, inventories of useful plants must go beyond the registering and trying to understand the philosophical and theoretical bases of this ancestral knowledge. (c) Common names. The persistence of maternal language in plant classifications is another indicator of the important role that the nature has for these communities. This knowledge is still shared, at least among the elders, although in some communities there is an interest in starting again to educate children within the ethnic values and knowledge, which includes language. The names from many of the specimens collected have no translation into Spanish, which indicates that there is still an interest in preserving these names, and that they correspond to species that have very specific uses within each community or ethnic group. For example, several species of the genus Rhus (R. andrieuxii Engl., R. chondroloma Standl., R. galeotti Standl., R. oaxacana Loes., R. standleyi F.A. Barkley, and R. terebinthifolia Schltdl. & Cham.) were recorded only in Mixe municipalities, or several Begonia L. species were recorded only among the Mazatec sites (B. heracleifolia Schltdl. & Cham., B. oaxacana A. DC., B pinetorum A. DC., and B. rhodochlamys L.B. Sm. & B.G. Schub.). In the process of collecting data about the names of plants, we detected an ethnotaxonomic structure in the way to name them, an aspect that Berlin (1992) has addressed in his studies and has been analyzed extensively by De Ávila (2004) for the diverse languages used in Oaxaca. The dominant species of any of the vegetation types visited always had a name in the indigenous language. Some of them were currently used, and others were not, but they always have a name. It is not accidental that some of the species present in montane cloud forests, which are the vegetation with the highest number of species registered in the present study, receive names in several languages (e.g., Saurauia comitis-rossei R.E. Schult., Litsea glaucescens, and Liquidambar styraciflua). The syntax of the terms used to classify the different biological forms (tree, shrub, grass, vine, orchid, etc.) also corresponds to the theoretical aspects that have been analyzed by other authors (Berlin 1992; Hunn 1988, 2008) and has been studied by De

Ethnobotanical Science in the Clouds: Useful Plants of. . .

865

Ávila (2004) for ethnic groups of Oaxaca. It is important highlight the fact that names in the indigenous language prevail, and it is necessary to rescue them, especially because many schools insist that children speak only in Spanish, an attitude that extends to their homes, where sometimes it is no longer possible to have interlocutors, because they do not know the words in their mother´s language. (d) Distribution of useful botanical families by ethnic groups in the PTRs. Some taxa are used preferentially by some ethnic groups. Polypodiophyta and Lycophyta are used mainly by the Mazatec, Zapotec, and Mixe people (Rendón-Aguilar et al. 2017a), the Mixe and Mazatec people used more the gymnosperm species (Bernal-Ramírez et al. 2019). This is evident in the well-represented families. For example, Piperaceae are more used among the Chinantec and Zapotec people (Martínez-Bautista et al. 2019). A possible explanation rests on the area of distribution of each ethnic group, and the different ecosystems they depend on. For example, in terms of useful families, Northern Zapotec people recorded the higher number of useful species and families and they are distributed in 22 municipalities. Nevertheless, some families were collected only in one municipality, and different families were added as more municipalities were visited. Even when ecological limitations were present throughout the project, and many species were not collected in some municipalities because of their phenology state, it is evident that intracultural variation exists, and gradual addition of families used in each municipality gives a total high value. Final aspects of this research are related with an old discussion about reversion of traditional knowledge (Barrera-Marín 1979; Gómez-Pompa 1982; Toledo 1982). Maybe the base of these studies most rest on the need to give it a scientific support, but not all traditional knowledge can be subject to inquiry, because it goes further than only a material world, and some uses function only under specific circumstances. For example, some medicinal plants function in a cultural context, with particular cultural values associated like of “envy,” or “courage”; the use category “ornament” has a cultural perception of “beauty.” Even food plants are preferred by some ethnic groups, but not by others. In this context, traditional knowledge must persist by its own right to persist, because we must be clear and sensitive that there are many explanations of world and life.

References Aguilar A, Camacho JR, Chino S, Jácquez P, López ME, Información etnobotánica. Herbario medicinal del Instituto Mexicano del Seguro Social (IMSS). México: EDITORIAL; 1994. Albuquerque UP, de Lucena P, RF, Neto EM de F. Selection of Research Participants. In: Albuquerque UP, Cruz LVF, de Lucena RF P, Alves RRN, editors. Methods and techniques in ethnobiology and ethnoecology. New York: Springer; 2014. Alexiades MN. Selected guidelines for ethnobotanical research: a field manual. New York: The New York Botanical Garden; 1996.

866

B. Rendo´n-Aguilar et al.

Arriaga L, Espinoza JM, Aguilar C, Martínez E, Gómez L, Loa E. Regiones Terrestres Prioritarias de México. México: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2000. Aswani S, Lemahieu A, Sauer WHH. Global trends of local ecological knowledge and future implications. PLoS One. 2018;13:e0195440. https://doi.org/10.1371/journal.pone.0195440. Bennett BC. Doctrine of signatures: an explanation of medicinal plant discovery or dissemination of knowledge? Econ Bot. 2007;61:246–255. https://doi.org/10.1663/0013-0001(2007)61[246: DOSAEO]2.0.CO;2 Berlin B. Ethnobiological classification: principles of categorization of plants and animals in traditional societies. Princeton: Princeton University Press; 1992. Bernal-Ramírez LA, Bravo-Avilez D, Fonseca-Juárez RM, Yáñez-Espinosa L, Gernandt DS, Rendón-Aguilar B. Usos y conocimiento tradicional de las gimnospermas en el noreste de Oaxaca, México. Acta Bot Mex. 2019;126:e1471. https://doi.org/10.221829/abm126.2019.1471 Blancas J, Casas A, Rangel-Landa S, Moreno-Calles A, Torres I, Pérez-Negrón E, Solís L, DelgadoLemus A, Parra F, Arellanes Y, Caballero J, Cortés L, Lira R, Dávila P. Plant management in the Tehuacán-Cuicatlán Valley, Mexico. Econ Bot. 2010;64:287–302. https://doi.org/10.1007/ s12231-010-9133-0. Boege SE. El patrimonio biocultural de los pueblos indígenas de México. Hacia la conservación in situ de la biodiversidad y agrodiversidad en los territorios indígenas. México: Instituto Nacional de Antropología e Historia: Comisión Nacional para el Desarrollo de los Pueblos Indígenas; 2008. Brown CH, Chase PK. Animal classification in Juchitán Zapotec. J Anthropol Res. 1981;37:61–70. https://doi.org/10.1086/jar.37.1.3629515. Caballero J, Cortés L. Percepción, uso y manejo tradicional de los recursos vegetales en México. In: Rendón B, Caballero J, Martínez-Alfaro MA, editors. Plantas Cultura y Sociedad. México: Universidad Autónoma Metropolitana, SEMARNAP; 2001. Caballero J, Casas A, Cortés L, Mapes C. Patrones en el conocimiento, uso y manejo de plantas en pueblos indígenas de México. Estudios Atacameños. 1998;16:181–95. Caballero J, Cortés L, Martínez-Alfaro MA, Lira-Saade R. Uso y manejo tradicional de la diversidad vegetal. In: García-Mendoza AJ, Ordóñez MJ, Briones-Salas M, editors. Biodiversidad de Oaxaca. México: Ed Redacta-Instituto de Biología, Universidad Nacional Autónoma de México-Fondo Oaxaqueño para la Conservación de la Biodiversidad-World Wildlife Fund; 2004. Cardoso J. Estudio etnobotánico del agroecosistema milpa en el municipio de Candelaria Loxicha, Oaxaca. México: Servicio Social. Universidad Autónoma Metropolitana-Xochimilco; 2004. Casas A, Valiente-Banuet A, Viveros JL, Caballero J, Cortés L, Dávila P, Lira R, Rodríguez I. Plant Resources of the Tehuacán-Cuicatlán Valley. Mexico Econ Bot. 2001;55:129–66. https://doi. org/10.1007/BF02864551. CONABIO. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad. Sistema de Información Biótica. Manual de usuario. Versión 5.0.3. México: CONABIO; 2012. CONABIO. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad. 2018. https:// www.biodiversidad.gob.mx/pais/quees.html. Accessed online on April, 2020. Cruz S, Cruz R. Zoología zapoteca de Santiago Xanica sierra sur de Oaxaca. In: González A, Vásquez MA, coordinators. Etnias, desarrollo, recursos y tecnologías en Oaxaca. Oaxaca: Centro de Investigaciones y Estudios Superiores en Antropología Social CIESAS, Gobierno del Estado de Oaxaca; 1992. De Ávila A. Clasificación de las lenguas de Oaxaca. In: García-Mendoza AJ, Ordóñez MJ, BrionesSalas M, editors. Biodiversidad de Oaxaca. México: Ed Redacta-Instituto de Biología Universidad Nacional Autónoma de México-Fondo Oaxaqueño para la conservación de la naturaleza-World Wildlife Fund; 2004. De Ávila A. La diversidad lingüística y el conocimiento etnobiológico. In: Soberón J, Halffter G, Llorente J, compilers. Capital Natural de México, vol. I: Conocimiento actual de la biodiversidad. México: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2008.

Ethnobotanical Science in the Clouds: Useful Plants of. . .

867

Delgado-Altamirano ML, Piñón-Tápanes A, Vibrans H, Rivero-Cruz JF, Ibarra-Alvarado C, RojasMolina A. In vitro antileishmanial activity of Mexican medicinal plants. Heliyon. 2017;3: e00394. https://doi.org/10.1016/j.heliyon.2017.e00394. Farjon A, Filer D. An atlas of the world’s conifers: an analysis of their distribution, biogeography, diversity and conservation status. Boston: Brill; 2013. Farjon A, Styles TB. Flora Neotropica, Pinus (Pinaceae). New York: The New York Botanical Garden; 1997. García-Hernández KY, Vibrans H, Rivas-Guevara M, Aguilar-Contreras A. This plant treats that illness? The hot-cold system and therapeutic procedures mediate medicinal plant use in San Miguel Tulancingo, Oaxaca, Mexico. J Ethnopharmacol. 2015;163:12–30. https://doi.org/10. 1016/j.jep.2015.01.001. García-Mendoza AJ, Meave JA, editors. Diversidad florística de Oaxaca: de musgos a angiospermas (colecciones y lista de especies). 2ª ed. México: Universidad Nacional Autónoma de México-Comisión Nacional para el Conocimiento y Uso de la Biodiversidad-Instituto Estatal de Ecología y Desarrollo Sustentable; 2012. García-Mendoza AJ, Ordóñez MJ, Briones-Salas M. Introducción. In: García-Mendoza AJ, Ordóñez MJ, Briones-Salas M, editors. Biodiversidad de Oaxaca. México: Ed Redacta-Instituto de Biología. Universidad Nacional Autónoma de México-Fondo Oaxaqueño para la conservación de la naturaleza-World Wildlife Fund; 2004. Gasca-Zamora J. Gobernanza y gestión comunitaria de recursos naturales en la Sierra Norte de Oaxaca. Región y sociedad. 2014. Gasca-Zamora J, López-Pardo G, Palomino-Villavicencio B, Mathus-Alonso M. La Gestión comunitaria de recursos naturales y ecoturísticos en la Sierra Norte de Oaxaca, México. México: UNAM, Instituto de Investigaciones Económicas: Academia Mexicana de Investigación Turística; 2010. Gernandt DS, Pérez-de la Rosa JA. Biodiversidad de Pinophyta (coníferas) en México. Rev Mex Biodiv. 2014;85(1):126–33. https://doi.org/10.7550/rmb.32195. Hernández-Pasteur G, Silva-Bermúdez PS, Reyes-Chilpa R, Vibrans H, Soto-Hernández M. Evaluación in vitro de la actividad cicatrizante y antibacteriana de extractos de Buddlejja cordata Kunth y Vismia baccifera (L.) Triana & Planch. Rev Fitotec Mex. 2019;42:93–9. Hoagland SJ. Integrating traditional ecological knowledge with western science for optimal natural resource management. IK: Other Ways Knowing. 2017;3:1–15. https://doi.org/10.18113/ P8ik359744. Hunn E. Mixtepec Zapotec ethnobiological classification: a preliminary sketch and a theoretical commentary. Antropológica. 1988;1:35–48. Hunn E. A Zapotec natural history. Trees, herbs, and flowers, birds, beasts and bugs in the life of SanJuan Gbëë. Tucson: The University of Arizona Press; 2008. Hunn ES, Acuca D. La etnobiología en el vocabulario en lengua zapoteca del fray Juan de Córdova, comparación con la actual de san Juan Mixtepec. Cuadernos del Sur. 2001;7:21–32. Ibarra-Manríquez G, Ricker M, Ángeles G, Sinaca Colín S, Sinaca Colín MA. Useful plants of the Los Tuxtlas rainforest (Veracruz, México): considerations on their market potential. Econ Bot. 1997;51:362–76. https://doi.org/10.1007/BF02861046. INAFED-SEGOB. Enciclopedia de los municipios y delegaciones de México. Estado de Oaxaca; 2010. http://www.inafed.gob.mx/work/enciclopedia/EMM20oaxaca/presentacion.html. Accessed online on April 2020. IUCN The Red List of Threatened Species. Version 2019-2. 2019. https://www.iucnredlist.org La B-MA. Etnobotánica. In: Barrera A, editor. La Etnobotánica: tres puntos de vista y una perspectiva, Cuadernos de Divulgación de INIRE, vol. 5; 1979. p. 19–24. La G-PA. Etnobotánica en México. Biotica. 1982;7(2):151–61. Lira R, Casas A, Rosas-López R, Paredes-Flores M, Pérez-Negrón E, Rangel-Landa S, Solís L, Torres I, Dávila P. Traditional knowledge and useful plant richness in the Tehuacán-Cuicatlán Valley, Mexico. Econ Bot. 2009;63:271–87. https://doi.org/10.1007/s12231-009-9075-6.

868

B. Rendo´n-Aguilar et al.

Luna-José AL. Cuantificación de los recursos vegetales recolectados en tres municipios Zapotecos de la Sierra Madre del Sur, Oaxaca. México: Tesis de Maestría. Colegio de Posgraduados, Montecillos. México; 2006. Luna-José A, Rendón B. Recursos vegetales recolectados en diez comunidades de la Sierra Madre del Sur, Oaxaca. Polibotánica. 2008;26:196–242. Luna-Morales C, Aguirre R. Clasificación tradicional, aprovechamiento y distribución ecológica de la pitaya mixteca en México. Interciencia. 2001;26:18–24. Magaña-Alejandro MA, Gama-Campillo LM, Mariaca-Méndez R. El uso de las plantas medicinales en las comunidades Maya-Chontales de Nacajuca, Tabasco, México. Polibotánica. 2010;29: 213–62. Manzanero G, Flores A, Hunn ES. Los huertos familiares zapotecos de San Miguel Talea de Castro, Sierra Norte de Oaxaca, México. Etnobiología. 2009;7:9–29. Martin GJ. Ecological classification among the Chinantec and Mixe of Oaxaca, México. Etnoecológica. 1993;1. http://etnoecologia.uv.mx/Etnoecologica/Etnoecologica_vol1_n2/art_ martin.htm Martínez M. Plantas útiles de la flora mexicana. México: Ediciones Botas; 1959. Martínez-Bautista BG, Bernal-Ramírez LA, Bravo-Avilez D, Samain M-S, Ramírez-Amezcua JM, Rendón-Aguilar B. Traditional uses of the family Piperaceae in Oaxaca, Mexico. Trop Conserv Sci. 2019;12:1–22. https://doi.org/10.1177/1940082919879315. Meave JA, Rincón-Gutiérrez A, Ibarra-Manríquez G, Gallardo-Hernández C, Romero-Romero MA. Checklist of the vascular fora of a portion of the hyper-humid region of La Chinantla, Northern Oaxaca Range, Mexico. Bot Sci. 2017;95:722–59. https://doi.org/10.17129/ botsci.1812. Messer E. Zapotec plant knowledge: classification, uses and communication about plants in Mitla, Oaxaca, México. Memoirs of the Museum of Anthropology, University of Michigan. 10. México. Etnoecológica. 1978;1:17–33. Mickel JT, Beitel JM. Pteridophyte flora of Oaxaca, Mexico. Mem N Y Bot Gard. 1988;46:1–568. Monroy AG, Sánchez-Cordero V, Briones-Salas M, Lira-Saade R, Maass JM. Representatividad de los tipos de vegetación en distintas iniciativas de conservación en Oaxaca, México. Bosque. 2015;36:199–210. https://doi.org/10.4067/S0717-92002015000200006. Montalvo L. Manejo y estructura de la vegetación leñosa de los cafetales en la Sierra Madre del Sur, Oaxaca. México: Tesis de Maestría. Colegio de Posgraduados, Montecillos. México; 2006. Ordóñez MJ. El territorio del estado de Oaxaca: una revisión histórica. Investig Geogr. 2000;42: 67–86. Pérez-Nicolás M, Vibrans H, Romero-Manzanares A, Saynes-Vásquez A, Luna-Cavazos M, Flores-Cruz M, Lira-Saade R. Patterns of knowledge and use of medicinal plants in Santiago Camotlán, Oaxaca, Mexico. Econ Bot. 2017;71:209–23. https://doi.org/10.1007/s12231-0179384-0. Pérez-Ortega G, Gonzalez-Trujano ME, Ángeles-Lopez GE, Brindis F, Vibrans H, Reyes-Chilpa R. Tagetes lucida Cav.: ethnobotany, phytochemestry and pharmacology of its tranquilizing properties. J Ethnopharmacol. 2016;181:221–8. https://doi.org/10.1016/j.jep.2016.01.040. PoWO. Plants of the World Online. Facilited by the Royal Botanical Gardens, Kew. 2019. http:// plantsoftheworldonline.org Prance GT, Campbell DG. The present state of tropical floristic. Taxon. 1988;37:519–48. https:// doi.org/10.2307/1221097. Rendón-Aguilar B, Bernal-Ramírez LA, Bravo-Avilez D. Inventario etnoflorístico en regiones oaxaqueñas con gran biodiversidad. Universidad Autónoma Metropolitana, unidad Iztapalapa. Informe final Sistema Nacional de Información sobre Biodiversidad-Comisión Nacional para el Conocimiento y Uso de la Biodiversidad proyecto No. JF102. México: CONABIO; 2017a. http://www.conabio.gob.mx/institucion/proyectos/resultados/InfJF102.pdf Rendón-Aguilar B, Bernal-Ramírez LA, Bravo-Avilez D, Mendoza-Ruiz A. Ethnobotany of lycophyta and polypodiophyta in priority terrestrial regions of Oaxaca, México. Am Fern J. 2017b;107:200–18. https://doi.org/10.1640/0002-8444-107.4.200.

Ethnobotanical Science in the Clouds: Useful Plants of. . .

869

Rindos D. The origins of cultivated plants. Cambridge: Harvard University Press; 1984. Sánchez-González A. Una visión actual de la diversidad y distribución de los pinos de México. Madera y Bosques 2008;14:107–120. 10.21829/myb.2008.1411222 Saynes-Vásquez A, Caballero J, Meave JA, Chiang F. Cultural change and loss of ethnoecological knowledge among the Isthmus Zapotecs of Mexico. J Ethnobiol Ethnomed. 2013;9:40. https:// doi.org/10.1186/1746-4269-9-40. SEMARNAT. Modificación del Anexo Normativo III, Lista de especies en riesgo de la Norma Oficial Mexicana NOM-059-SEMARNAT-2010, Protección ambiental-Especies nativas de México de flora y fauna silvestres-Categorías de riesgo y especificaciones para su inclusión, exclusión o cambio-Lista de especies en riesgo, publicada el 30 de diciembre de 2010; 2019. http://dof.gob.mx/nota_detalle.php?codigo¼5578808&fecha¼14/11/2019 Smith C. Plant remains. In: Byers D, editor. The prehistory of Tehuacán Valley. Austin: University of Texas Press; 1967. Solís RL. Etnoecología Cuicateca en San Lorenzo Pápalo, Oaxaca. Tesis de maestría. Centro de Investigaciones en Ecosistemas, Universidad Nacional Autónoma de México, Morelia; 2006. Spradley JP. Participant observation. New York: Holt, Rinehart and Winston; 1980. Thiers B. Index Herbariorum: a global directory of public herbaria and associated staff. New York Botanical Garden’s Virtual Herbarium. 2016. http://sweetgum.nybg.org/science/hi/ Toledo VM. La Etnobotánica hoy. Reversión del conocimiento, lucha indígena y proyecto nacional. Biotica. 1982;7:141–50. Toledo VM, Barrera-Bassols N. La memoria biocultural. La importancia ecológica de las sabidurías tradicionales. Barcelona: In-caria editorial; 2008. Torres-Colín R. Tipos de vegetación. In: García-Mendoza AJ, Ordóñez MJ, Briones-Salas M, editors. Biodiversidad de Oaxaca. México: Ed Redacta-Instituto de Biología. Universidad Nacional Autónoma de México-Fondo Oaxaqueño para la conservación de la naturalezaWorld Wildlife Fund; 2004. Valencia S, Nixon KC. Encinos. In: García-Mendoza A, Ordóñez MJ, Briones-Salas M, editors. Biodiversidad de Oaxaca. México: Ed Redacta-Instituto de Biología, Universidad Nacional Autónoma de México-Fondo Oaxaqueño para la conservación de la naturaleza-World Wildlife Fund; 2004. Van der Wal H. Modificación de la vegetación y el suelo por los chinantecos en Santiago Tlatepusco, Oaxaca, México. Etnoecológica. 1996;3:37–57. Ventura-Aquino Y, Rendón B, Rebollar S, Hernández G. Use and conservation of forest resources in the municipality of San Agustin Loxicha, Sierra Madre del Sur, Oaxaca, Mexico. Agrofor Syst. 2008;74:167–80. Vibrans H. Ethnobotany of Mexican weeds. In: Lira R, Casas A, Blanco J, editors. Ethnobotany of Mexico – interactions of people and plants in Mesoamerica. New York: Springer; 2016. Villaseñor JL. Checklist of the native vascular plants of Mexico. Rev Mex Biodiv. 2016;87: 559–902. https://doi.org/10.1016/j.rmb.2016.06.017. Zizumbo D, Colunga P. Aspectos etnobotánicos entre los Huaves de San Mateo del Mar, Oaxaca, México. Biótica. 1982;7:223–70.

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca Marco Antonio Va´squez-Da´vila, Gladys I. Manzanero-Medina, Adonicam Santiago-Martínez, and Sunem Pascual-Mendoza

Abstract

Life in the mountains of Oaxaca (southern Mexico) is characterized by high biological and cultural diversities; both are outstanding in both magnitude and importance and play a major role in the dynamics of the biocultural landscape. This chapter employs the transdisciplinary approach of biocultural ethnobotany to describe and interweave these diversities in mountainous Zapotec regions of Oaxaca. This is exemplified by two case studies in the Sierra Norte and one in the Sierra Sur; those in the Sierra Norte focus on milpa, home garden, and coffee agroecosystems and food plants, while the one in the Sierra Sur deals with medicinal plants. The ethnobotanical knowledge and management of mountain people, as well as the ecosystem services provided by the mountains, are key elements in the design of biocultural conservation policies.

Introduction The high biocultural diversity characterizes life in the mountains of Oaxaca (southern Mexico). Ecological, cultural, and linguistic diversities are outstanding in magnitude and importance and play a major role in the dynamics of the biocultural landscape. The present chapter deals with floristic diversity and Zapotec ethnobotany in mountainous regions of Oaxaca. These aspects are exemplified with three case studies in two geopolitical regions of Oaxaca: Sierra Norte and Sierra Sur.

M. A. Vásquez-Dávila (*) Instituto Tecnológico del Valle de Oaxaca, Oaxaca, Mexico G. I. Manzanero-Medina · A. Santiago-Martínez · S. Pascual-Mendoza Instituto Politécnico Nacional, Centro Interdisciplinario de Investigación para el Desarrollo Integral Regional, Unidad Oaxaca, Oaxaca, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_23

871

872

M. A. Va´squez-Da´vila et al.

For addressing the complex interrelationship of the human species with the plant kingdom, we rely on biocultural ethnobotany, defined here as “the study of the cosmovision, knowledge and wisdom, as well as the use and management that social groups carry out on the biocultural diversity present in an agroecosystem, vegetation type, region or landscape, in a specific biological, cultural and historical context” (Cfr. Vásquez-Dávila and Manzanero 2019).

Ethnobotanical Background for Oaxaca, Sierra Norte, and Sierra Sur Ethnobotanical research in Oaxaca includes publications on plant resources in general (Campos-Villanueva et al. 1992; Caballero et al. 2004); homegardens (Manzanero et al. 2018); studies on particular plant groups like the Piperaceae family (Martínez-Bautista et al. 2019); medicinal plants in general (Cruz-Pérez et al. 2021); and plant species and cultures interacting in traditional markets (Vásquez-Dávila and Manzanero 2021). Ethnobotanical works carried out in the Zapotec localities of the Sierra Norte of Oaxaca deal with the following topics: (1) Plant resources at the community or regional level, (2) medicinal plants, (3) food plants, (4) ornamental plants, (5) agroecosystems, and (6) markets. Some studies on plant resources are developed at the community or regional level (Aguilar-Santelisis 2007; Bautista et al. 2014; López-Santiago 2015; Bernal-Ramírez et al. 2019). Some studies have focused their attention on documenting use and management of medicinal plants (Vásquez-Cortez 2016; Domínguez-Yescas and Vázquez-García 2019; Pérez-Nicolás et al. 2018). Also, there have been several theses on medicinal plants from the Sierra Norte integrated to the Biology bachelor degree program at the Instituto Tecnológico del Valle de Oaxaca (see for instance the dissertations by Gómez-Hernández 2010; Pascacio-González 2011; Sosa-Pérez 2013; García-Serrano and López-Hernández 2016; Ramírez-Martínez 2017). There are additional references on food plants (Vázquez-García et al. 2009; Pascual-Mendoza et al. 2021), while ornamental plants have been more studied (Rees 1976; Mondragón 2008; Mondragón and VillaGuzmán 2008; Martínez-López et al. 2016; Hernández-Rodríguez and DelgadilloMoya 2021; Hernández-Rodríguez and López-Santiago 2022). Agricultural or agroecosystem-related ethnobotany has been also carried out (González 2001; Manzanero et al. 2009; Gómez-Luna et al. 2017; López-Mendoza 2020; PascualMendoza et al. 2020), but studies on traditional markets in the Sierra Norte de Oaxaca are still scarce (Berg 1975; Cruz-Arenas and Cruz-Hernández 2021). Fewer ethnobotanical studies have been conducted in the Sierra Sur than in the Sierra Norte; the studies available have mainly been related to (1) plant resources at the regional or community level, (2) medicinal plants, (3) food plants, (4) agroecosystems, and (5) markets. A number of articles and books have been published on plant resources of the regions (Delgadillo-Moya 2000; Hunn 2008; Ventura-Aquino et al. 2008; Luna-José and Rendón-Aguilar 2008, 2012). Works on medicinal plants are more abundant, although many of them remain as unpublished theses (Weiss 1995; Zurita-Vásquez et al. 2012; Cruz-Espinoza and Guendulain-Hernández 2013;

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

873

Castro-Juárez et al. 2014; Ordaz-Santana and Ordaz-Santana 2018; Tello-Ortega et al. 2020). Two studies on edible plant species have been addressed (Ramírez-Aragón 2016; Pérez-José and García-Morales 2016), ethnobotanical aspects of home gardens and coffee plantations have been described (Aguilar-Støen et al. 2009, 2011; ZuritaVásquez et al. 2020), and finally, there is a work on the Miahuatlán market (VenegasRamírez 2009).

Floristic Diversity in the Mountains of Oaxaca Oaxaca’s geography is diverse: it includes wide coastal strips and an extensive and intricate mountain complex (see Fig. 1). The mountains of Oaxaca (and of Mexico, in general) create ecological islands on summits and in isolated ravines and valleys, determining a marked influence on the biological diversity and biogeography of the area (Morrone 2019; Gómez-Mendoza et al. 2006). In some cases, they have functioned as barriers to species dispersal, especially of tropical fauna and flora; in other cases, mountains have been corridors through which species disperse, as it is the case of neotropical plants of Andean affinity, found mostly in mesophilic cloud forests of moderate altitude regions, whose typical example are tree ferns of the Cyatheaceae family (Martínez-Salas and Ramos 2014). There are three well-defined mountain ranges in Oaxaca: The Sierra Norte de Oaxaca, the Sierra Madre del Sur, and the Sierra Atravesada. With an abrupt and rugged relief and elevations of more than 3000 m, the Sierra Norte de Oaxaca is composed of both marine and continental sediments. The rivers that cross it pour their waters into the Atlantic Ocean (see Fig. 2). The Sierra Madre del Sur, on the Pacific Ocean slope, has an average elevation of 2000 m and is crossed by the Balsas and

Fig. 1 Sierra in Oaxaca, México. (Photo by: Elí García Padilla)

874

M. A. Va´squez-Da´vila et al.

Fig. 2 Sierra Norte of Oaxaca, México. (Photo by: Carlos Alberto Masés García)

Verde River systems. The presence of marine sediments and Cretaceous lavas suggests the tectonic collision of an oceanic island arc with the continent (Morrone 2019). The Sierra Atravesada (located in the Isthmus region) is part of the Sierra Madre de Chiapas and influences the Pacific and Atlantic slopes. It is the oldest formation in the country and is made up of igneous rocks of Paleozoic to Precambrian age. In general, the mountains of the Pacific coast are closer to the sea than those of the Gulf of Mexico slope. The above has facilitated the isolation of biotic communities and their species; therefore, it has produced a high number of endemism (Suárez-Mota et al. 2018). Oaxaca is the Mexican state with the highest richness of vascular plants, with 8903 species (García-Mendoza and Meave 2011; Villaseñor 2016). The four main vegetation types associated with the mountains of Oaxaca are: temperate forest, cloud or mesophyll forest, humid forest, and sub-humid forest. Figure 3 shows the vegetation types and highlights the Sierras of the State of Oaxaca. Table 1 shows areas of well-preserved forests and jungles in mountainous areas of five regions of Oaxaca (Sierra Norte, Sierra Sur, Isthmus, Papaloapan, and Cañada). Temperate forests – also known as coniferous, pine, fir or oyamel, táscate, oak and mixed oak-pine forests – prevail in the mountainous systems of the Oaxacan territory (Rzedowski 2006). The coniferous forest is dominated by species of Pinus, Juniperus, and Abies. Pines are widely distributed in the mountain systems, while firs are restricted to isolated patches on hills, slopes, and ravines. The Abies forest is located in elevated and humid parts of the Sierra Norte, Sierra Sur, and Sierra de Chiapas; the outstanding species are: Abies guatemalensis, A. hickelii, A. oaxacana, and Cupressus lindleyi. Juniperus forest is also found in various parts of the state, with Juniperus deppeana (táscate) and J. flaccida being the most important species; these species have medicinal, edible, and artisanal uses (Luna-José and RendónAguilar 2008).

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

875

Fig. 3 Vegetation, Sierras and Ethnic groups of Oaxaca, México. (Map drawn by: Mario Lavariega Nolasco)

The following Pinus species are common in the Sierra Norte and Sierra Sur: P. chiapensis, P. lawsoni, P. michoacana, P. oaxacana, P. patula, P. pseudostrobus, and P. rudis, accompanied by species of Quercus acutifolia, Q. crassifolia, Q. glaucescens, Q. laurina, Q. magnoliifolia, and Q. rugosa, and other species such as Arbutus xalapensis (madroño in Spanish), Alnus arguta (palo de águila), and Clethra mexicana (Flores and Manzanero 1999). In the pine forest of the Sierra Norte, the most abundant trees are Pinus hartwegii Lindl., P. patula Schltdl. & Cham., P. pseudostrobus Gordon, and P. teocote Schltdl. & Cham.; Abies hickelii Flous & Gaussen is found in smaller quantities (Zacarías-Eslava and del Castillo 2010). The Pinus-Quercus forest is characterized by being formed by different species of Pinus and Quercus, with dominance of the former, being placed as a separate category due to the large areas it occupies in all the mountainous systems of the entity. On the Pacific slope it is present from an altitude of 1000 m on average, although it descends to 400 m in the Sierra de Miahuatlán. The most outstanding species are Pinus douglassiana, P. michoacana, P. montezumae, P. oocarpa, Quercus elliptica, Q. glaucescens, Q. obtusata, and Q. rugosa (Flores and Manzanero 1999). The pine-oak and oak-pine forests of the Sierra Norte feature (in addition to the pine species already mentioned) Quercus castanea Née, Q. conzattii Trel. & Bonpl., Q. depressa Humb. & Bonpl., Q. glabrescens Benth., Q. glaucoides M. Martens & Galeotti, and Q. laurina Humb. & Bonpl (ZacaríasEslava and del Castillo 2010). All these species are used in different ways by local people (Luna-José and Rendón-Aguilar 2008).

M. A. Va´squez-Da´vila et al.

876

Table 1 Temperate and Humid Forests of the Sierras of Oaxaca, Mexico Region Sierra Norte

Temperate forest Sierra Juárez San Bernardino Cerro Rabón Sierra Monteflor Retumbadero Cuasimulco

Sierra Sur

Istmo

Yotao Cerro Negro Sierra de Juquila Sierra Loxicha Nube Flandes Cerro Encantado Cerro El Tigre Cerro Tlacuache Sierra Atravesada

Mesophilic forest Cerro San Felipe Comaltepec Talea de Castro El Rincón

Humid rainforest Sierra Veinte Cerros Cuasimulco Retumbadero

Cañada

Sierra Monteflor

Zempoaltépetl Huautla de Jiménez

Juquila

Loxicha

Cerro El Tigre Cerro Encantado Loxicha Miahuatlán Cerro Piedra Larga Sierra de Niltepec

Chimalapas

Sierra Atravesada Santiago Lachiguirí Papaloapan

Sub-humid rainforest Cerro San Felipe

Cerro Guiengola Chimalapas

Usila Cerro Rabón Pápalos

Cuicatlán

The Quercus or oak forest is found in all the mountainous systems of the state. The diversity of oak species in the state is very high. In the Sierra Norte, in the windward zone, Quercus chinantlensis, Q. laurina, Q. liebmannii, Q. sororia, Oreopanax xalapensis, and Ternstroemia pringlei are common. In its leeward portion Quercus acutifolia, Q. glaucoides, and Q. magnoliifolia are common. In the southern Sierra, Q. elliptica, Q. glaucescens, Q. glaucoides, and Q. magnoliifolia predominate, sometimes accompanied by species of Pinus, Juniperus, Clethra, Byrsonima, and Rhus (Flores and Manzanero 1999; Rzedowski 2006). In Oaxaca, the mountain cloud mesophyll cloud forest is distributed discontinuously in small portions of the Sierra Sur, Sierra Norte, and Sierra Atravesada. Table 1 gives examples of places where this type of vegetation exists. On the Gulf slope, the cloud forest is mainly composed of Quercus candicans Née, Q. corrugata Hook, Q. glaucescens Bonpl., Q. laurina Bonpl., Q. magnoliifolia Née, Q. mexicana Bonpl., Q. skinneri Benth., and Q. sororia Liebm. Other species are part of this

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

877

forest, for example: Alnus acuminata Kunth, Cestrum spp., Drimys winteri J.R. Forst. & G. Forst., Litsea glaucescens Kunth, Liquidambar styraciflua L., Podocarpus matudae Lundell, Saurauia spp., and Ternstroemia lineata DC., In the low altitude portions are found Brosimum alicastrum Sw., Manilkara zapota (L.) P. Royen, Attalea rostrata Oerst., Handroanthus chrysanthus (Jacq.), and Terminalia amazonia (J.G. Gmel.) Exell (Flores and Manzanero 1999; Rzedowski 2006). It should be noted that most species are used for fuel, construction, and shade (Carbajal-Esquivel 2008). As an example of the complexity of the distribution of vegetation in Oaxaca, we mention that in the middle part of the Sierra Norte (on the Gulf of Mexico slope) there are areas of high evergreen forest, mesophill mountain cloud forest, pine forest, pine-oak forest, oak-pine forest, and oak forest. The rainforest (also known as tropical evergreen forest or high evergreen forest) extends in a continuous strip in the lowlands of the Sierra Norte and Sierra Sur. It is dominated by Brosimum alicastrum Sw. (ramón) and Manilkara zapota (L.) P. Royen (chicozapote), which are considered multiple-use species (Flores and Manzanero 1999; Domínguez-Lagunes 2012). The sub-humid forest (or sub-deciduous tropical forest) occupies large portions of the Sierra Norte, Sierra Sur, Isthmus and Cañada (see Table 1). The characteristic species are: Calycophyllum candidissimum (Vahl) DC., Aphananthe monoica (Hemsl.) J.F. Leroy, Pterocarpus acapulcensis Rose, and species whose fruits and/or seeds are edible, such as Brosimum alicastrum Sw., Sideroxylon persimile (Hemsl.) T.D. Penn, Enterolobium cyclocarpum (Jacq.) Griseb, Manilkara zapota (L.) P. Royen, Vitex mollis Kunth, Psidium sartorianum (O. Berg) Nied., Ceiba pentandra (L.) Gaertn., Cordia alliodora (Ruiz & Pav.) Oken, Hymenaea courbaril L., Morisonia americana L., Parmentiera aculeata (Kunth) Seem. Other companion species are: Andira inermis (W. Wright) Kunth ex DC., Albizia niopoides (Spruce ex Benth.) Burkart, A guachapele (Kunth) Dugand, A. tomentosa (Micheli) Standl., Libidibia coriaria (Jacq.) Schltdl., Caesalpinia velutina (Britton & Rose) Standl., Godmania aesculifolia (Kunth) Standl, Hura polyandra Baill, Ficus aurea Nutt, Lafoensia punicifolia DC, Licania arborea Seem, Lysiloma acapulcense (Kunth) Benth, Attalea guacuyule (Liebm. ex Mart.) Zona, Swietenia humilis L., and Trichilia havanensis Jacq (Flores and Manzanero 1999; Rzedowski 2006). The mountains of Oaxaca play a very important role in the state, national, and international biodiversity scenarios. The environmental heterogeneity of Oaxaca has favored the evolution of numerous species of fungi, plants, and animals. During the last 10,000 years, humans have contributed to this evolution by interacting with the ecosystems and landscapes. Biodiversity has allowed at the same time, the flourishing and development of the plurality of human groups, resulting in the presence, since pre-Hispanic times, of 15 different ethnolinguistic groups in the state and a surprising cultural richness (CONABIO 2018). This is why the Zapotec mountains contain biocultural richness and provide diverse ecosystem services. These two characteristics are tools to address the socio-environmental crisis “through the development of ecosystem-based actions, the contribution of different knowledge systems including indigenous and local,

878

M. A. Va´squez-Da´vila et al.

community participation, ecosystem resilience and ecological integration for adaptive transformation in territorial dimensions” (CONABIO 2018).

Outline of Oaxaca’s Cultural Diversity Oaxaca is the Mexican state with the largest number of ethnic groups, differentiated by speaking a native language, having their own culture and having been settled in the territory of this state since before the European invasion of the sixteenth century. Some of the groups that inhabit the mountains or near them are the Amuzgo, Chatino, Chinantec, Chocholtec, Chontal, Cuicatec, Ixcatec, Mazatec, Mixe, Mixtec, Nahua, Triqui, Zapotec, and Zoque. The Tzotzil, originally from Chiapas, have recently settled in the Chimalapas region. Figure 3 shows the distribution of the ethnic groups of Oaxaca. In Oaxaca there are 1,221,555 people over 3 years of age who speak an indigenous language. Most indigenous languages spoken in this entity are: Zapotec (420,324 speakers), Mixtec (267,221), Mazatec (170,155), and Mixe (118,882) (INEGI 2020). There are 179 linguistic variants in the state of Oaxaca (INALI 2008), which provide an extraordinary cultural and socio-environmental diversity and imply a millenary knowledge, use, management and conservation of nature developed in situ (Aswani et al. 2018; Fletcher et al. 2021). The Sierra Norte Region is divided into 68 municipalities grouped in the districts of Ixtlán, Villa Alta, and Mixes. It has an area of 8972 km2 and is inhabited by Zapotec, Chinantes, and Mixe. The Sierra Sur region covers an area of 14,753.26 km2, with 70 municipalities grouped in four districts: Putla, Sola de Vega, Miahuatlán, and Yautepec. This region is inhabited by Zapotec, Mixtec, Chatino, Chontal, Amuzgo, and Triqui peoples. In the highlands of Oaxaca, small towns prevail; however, there are highly populated mountain municipalities whose municipal capitals are cities, such as San Pedro Mixtepec (with 49,780 inhabitants), San Pedro Pochutla (n ¼ 48,204), Acatlán de Pérez Figueroa (45,167), Huautla de Jiménez (31,710), and San Juan Guichicovi (29,802), among others (INEGI 2020). Ecology, culture, economy, and politics are factors that shape the structure and functioning of Oaxaca’s ethnic territories. Ecologically, all indigenous settlements are located in ecotones, that is, at the intersection of two or more ecosystems. Ethnic territories share characteristics of complementarity, that is, the use of natural resources in the manner most conducive to productive endeavor (Bélisle et al. 2018; Ban et al. 2018). Regional markets are a tangible example of this fact (Vásquez-Dávila and Manzanero 2021). At the political level, mountainous territories are regions of refuge (Aguirre-Beltrán 2009), that is, places where the original inhabitants took refuge from the pressure of the European invasion in the sixteenth century. Similarly, ethnic territories developed a philosophy of community organization, communality (Sánchez-Antonio 2021), with their own style of governance (known as the “system of uses and customs”) and strategies for defending the territory (Mraz-Bartra 2020). The correlation between biological and linguistic diversities of Oaxaca may be due, among others, to the following causes: (a) its large territorial extension with a

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

879

Table 2 Ecological zones in Zapotec territories of Oaxaca Ecological zones Tropical warm humid Tropical warm sub-humid Temperate humid Temperate sub-humid

Ecosystems Humid rainforest Sub-humid forest Mesophilic forest Temperate forest

Ecological levels High Medium Low

geographical, ecosystem, specific (biological species sensu stricto) and genetic variety; (b) the existence of geophysical barriers or large distances that historically allowed linguistic specialization; (c) the diversity of temperate and tropical climates (humid, sub-humid, and dry) that favor the high number and density of species (and hence the need to symbolize and name them); (d) the co-evolution of small human groups with their local ecosystems. Table 2 shows the ecological zones, ecosystems, and ecological levels of the Zapotec territories of Oaxaca.

Biocultural Ethnobotany of the Zapotec Mountains The Sierras of Oaxaca are regions of great biocultural density due to their biodiversity, the fact that they are centers of origin and diversification of species, the vitality of agroecosystems with domesticated native agrobiodiversity, and, of course, the cultural plurality that conceives and sustains them. The Sierra Norte of Oaxaca has been classified as Priority Terrestrial Region (RTP) number 130 (Arriaga-Cabrera et al. 2009) and Priority Biocultural Region (RBP) 17 (Arriaga-Cabrera et al. 2009; Boege 2008). The Sierra Sur de Oaxaca is RTP number 129 and RBP 11 (Boege 2008). The first two authors of this chapter have developed their fieldwork in Zapotec communities in these two Oaxacan sierras since the nineties of the last century. More recently, the third and fourth authors have worked in their own localities of origin as part of their master’s and doctoral theses. Below, we break down the case studies in the Sierra Norte and Sierra Sur. Figure 4 provides information on the three study sites we chose to develop the biocultural ethnobotany theme.

Zapotec Ethno-Agroecology: Milpas, Home Gardens, and Coffee Plantations in Zoogochí, Sierra Norte Here we report an ethnographic overview of the agriculture of the Zapotec of Santa María Zoogochí based on the authors’ fieldwork. Zoogochí is a small town in the municipality of Ixtlán de Juárez with 850 inhabitants. Its territory is 7.8 km2 and is located between 600 and 2200 m; therefore, it has three climates: semi-warm humid, sub-humid, and temperate humid, which correspond to what the local inhabitants call warm land, temperate land and cold land, respectively. The vegetation is composed of pine forest, pine-oak forest, mesophyll cloud forest, gallery vegetation,

880

M. A. Va´squez-Da´vila et al.

Fig. 4 Study sites. (Map drawn by: Mario Lavariega Nolasco)

secondary vegetation, and cultivated areas. The predominant agriculture is rainfed (based on rainfall). The most common crops are native: corn (Zea mays L.), beans (Phaseolus vulgaris L.), squash (Cucurbita spp.), and chile (Capsicum annuum var. annuum L.) and introduced crops like sugar cane (Saccharum officinarum L.) and coffee (Coffea arabica L.). The main product obtained from sugar cane is panela, a kind of round hard bread made from sugar cane molasses. Table 3 shows part of the ethnobotanical diversity of the Zapotec of Zoogochí.

The Milpa The milpa is the crop par excellence in Mesoamerica. It consists of a multiple temporal coexistence of maize, beans, squashes, other plant species, trees on the edges of the land, and a great associated biodiversity: spontaneous plants and animals that come to the land. This ethno-agroecosystem has a temporal sequence that includes the cultivation stage and fallow periods (an agroecological succession) where other species emerge that are also used by peasant families. Most of Zoogochí’s farmland has very steep slopes. Maize is planted on the slopes of the hills, on rocky and steep terrain. Land preparation, planting, cultivation, and harvesting are strictly manual tasks. For this reason, corn cultivation is extremely tiring and exhausting.

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

881

Table 3 Ethnobotanical diversity of the Zapotecs of Zoogochí, Sierra Norte of Oaxaca Family Anacardiaceae Annonaceae Apiaceae Amaranthaceae Araceae Asteraceae Boraginaceae Burseraceae Cactaceae Caricaceae Cucurbitaceae Euphorbiaceae Fagaceae Lamiaceae Lauraceae Leguminosae

Magnoliaceae Malvaceae Moraceae Musaceae Myrtaceae Pinaceae Poaceae Rosaceae Rubiaceae Rutaceae

Sapotaceae Salicaceae

Scientific name Mangifera indica L. Annona cherimola Mill. Annona squamosa L. Coriandrum sativum L. Dysphania ambrosioides (L.) Mosyakin & Clemants Monstera deliciosa Liebm. Bellis perennis L. Tagetes erecta L. Cordia alliodora (Ruiz & Pav.) Oken Bursera simaruba (L.) Sarg. Opuntia ficus-indica (L.) Mill. Opuntia spp. Carica papaya L. Cucurbita spp. Euphorbia pulcherrima Willd. ex Klotzsch. Jatropha curcas L. Quercus spp. Mentha spicata L. Persea americana Mill. Acacia cornigera (L.) Willd.) Diphysa americana (Mill.) M. Sousa Enterolobium cyclocarpum (Jacq.) Griseb. Phaseolus vulgaris L. Pithecellobium dulce Mart. Magnolia yajlachhi A. Vázquez & Domínguez-Yescas Hibiscus rosa-sinensis L. Ficus carica L. Musa spp. Psidium guajava L. Pinus teocote Schiede ex Schltdl. Pinus spp. Saccharum officinarum L. Zea mays L. Rhaphiolepis bibas (Lour.) Galasso & Banfi Coffea arabica L. Citrus  sinensis (L.) Osbeck Citrus  aurantium L. Citrus  limon (L.) Osbeck Citrus x paradisi Macfad. Citrus  aurantiifolia (Christm.) Swingle Manilkara chicle (Pittier) Gilly Pouteria sapota (Jacq.) H.E. Moore & Stearn. Homalium senarium Sessé & Moc

Common name Mango Anona Anona Cilantro Epazote Mimbre Margarita Cempasúchil Rabo de iguana Mulato Nopal tunero Nopal Papaya Calabaza Nochebuena Piñón Encino Hierbabuena Aguacate Cornezuelo Cuachepil Guanacastle Frijol Guamúchil Flor de corazón Tulipán Higo Plátano Guayabo Ocote Pino Caña de azúcar Maíz Níspero Café Naranja dulce Naranja agria Limón Toronja Lima Zapote Mamey palo de piedra (continued)

M. A. Va´squez-Da´vila et al.

882 Table 3 (continued) Family Solanaceae

Scientific name Capsicum annuum var. annuum L. Cestrum nocturnum L. Solanum americanum Mill.

Common name Chile mirasol Huele de noche Bishate

In the agrohabitats – determined by altitude – three varieties of corn are grown: (1) hill corn is typical of cold lands, (2) yellow or early corn (tempranero in Spanish), and (3) warm land corn. The hill variety has a vegetative cycle of 7 months and has a very strong ear. The yellow corn is harvested after 3 months and the warm land corn, which has the largest fruit size, has a 5-month cycle. The cultivation system is slash and burn, occupying the land for 2 years and leaving it to rest for 6 years. The cultivation of these varieties is carried out as descrived below. The clearing begins in February, using a machete. In March, the trees and bushes are cut with an axe and are chopped up so that they dry uniformly for 15 days. After this time, the guardarraya is made around the land as a previous step to the burning. Guardarraya is a strip of land surrounding agricultural plots that is left free of fuel material to prevent the passage of fire between the plots. In Zoogochí, this work consists of making a lane 2 to 3 m wide with the machete and a resistant branch (clayuda in local Spanish) to sweep away the leaf litter in order to prevent the passage of fire to the surrounding land. Subsequently, on a cloudless day, the burning is done starting on the highest side of the ground. Planting begins in the first days of May, when the moon is full. A cornezuelo stake (Acacia cornigera (L.) Willd.) is used. Four seeds are sown at each step. The distances between plants are: 2 m for hill corn, 1 m (or less) for yellow corn, and a shorter distance for warm land corn. The planting starts “up the hill,” since the rows go in the direction of the slope (a system that unfortunately causes a greater dragging of the soil due to rain). The seed sown is traditionally measured in bushels. Two bushels (8 kg) are used to sow one hectare of hill corn. Two weeding operations are carried out to eliminate species that compete with corn. Weeding is done 6 and 12 weeks after planting. Another agricultural operation (zacateo in Spanish) consists of cutting the leaves of the corn plant or the whole plant – in case it does not have an ear or has been eaten by a wild animal. The forage obtained is used to feed donkeys (Equus africanus asinus L.). In the case of only cutting leaves, these are cut by hand or with a machete on plants that have already passed the tender corn stage and that the farmer considers will not affect their further development. Similarly, plants whose fruit (elote in Spanish) has already been harvested, those that did not produce corncob, and those that were knocked down by the wind or eaten by animals are used as fodder grass. Two months after the second milpa cleaning, the first harvest of some of the tender fruits (elotes) of the hill corn is carried out to make atole and tamales (atole is a hot, thick corn drink and tamales are a meal made with steamed corn wrapped in the husks of the corn cob). The well-dried corn is harvested a little later. Both harvests are carried out on full moon days. On the day of the harvest, the whole

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

883

family goes to the plot of land to do this work in one attempt. In order to transport less bulky bundles, it is customary to remove the husks from the cob on the field. The corn cobs are stored in the house, on the ground, or in a covered shed. The shelled corn can be stored in large metal jars. The average yield in Zoogochí is 1 ton per hectare, both in cold and warm land. Corn production is sufficient to cover the family’s annual needs, but in some cases, corn is purchased from local stores. In this community, the farmers never abandoned maize cultivation for coffee cultivation; they now recognize their wisdom in not depending on external maize stocks. The corn plant is attacked by different animals at different stages of its development. During the planting season, zanates (Quiscalus mexicanus Gmelin) cut the seedlings. These birds can be scared away with a stone thrower or a scarecrow. With the first rain (which is the signal to start planting) the black ants (Formicidae) come out and attack the corn; but if the rains are continuous, these insects disappear. Fall armyworm (Spodoptera frugiperda J.E. Smith) appears sporadically when the plant begins to glean. To eradicate these two pests, farmers use various chemicals, mostly in powder. As the harvest approaches, a little mammal called tejón (Nasua narica L.) is a danger; to drive it away or hunt it, it is customary to bring trained dogs (Canis lupus familiaris L.) into the field to chase it up trees or kill it with a shotgun or rifle. The meat of this procyonid is edible; another omnivorous mammal is the raccoon (Procyon lotor L.), whose meat is also edible. Some years, during the months of July to August, a microscopic fungus, called chahuistle (Puccinia sorghi Schwein) attacks the plant, yellowing it. The tools used for milpa farming are machetes, axes, stakes, sacks, brooms for the guardarraya and mecapal (an elongated textile made of hard maguey fibers used to carry objects and placed on the front and back of the loader). Some people started using fertilizers a few years ago. Fertilizer is placed next to the plant when it is about 70 cm in height. Several varieties of beans (Phaseolus vulgaris L.) are cultivated. Creeper beans are planted in association with corn and squash. The red bean is planted in monoculture; during weeding, stakes or trellises are placed for the plant to climb. Other bush varieties of beans that are planted as a monoculture are: 1) zaa btó (garrote bean) or, by another name, zaa lit ( frijol de secas) and 2) frijol tripa, which are planted in warm land in May. The soil for planting beans is very fluffy because its cultivation, say the farmers, gives vigor to the soil. The pests that attack it are the coleopteran called mayagüila (Diabrotica sp.) and another insect described as “woolly, whitish, that scrapes the underside of the leaf.”

Homegardens The homegarden is an agroforestry system with a high biocultural diversity in a reduced area; it represents a space for socialization and interaction among family and community members. The species and their uses indicate the degree of association that social groups have with the plants (Vásquez-Dávila and Manzanero 2015).

884

M. A. Va´squez-Da´vila et al.

Most of the houses in Zoogochí have a vegetable garden whose size varies according to the dimensions of the living terraces, as well as the additional terraces existing in each property. In this sense, we find the conjunction of living terraces and agricultural terraces in the same space. The terrace platforms are built with cut stone walls and aligned to model the topography of the site. The builder decides the height of the walls and their location according to their experience and the precise conditions of the site. Once this wall is erected, the side next to the hill is filled with earth and tamped, thus forming the terrace. To level and widen it, the other end of the terrace is sometimes dug out. Each house in Zoogochí has at least one construction of this type, and may have others, at different levels, for annex constructions: laundry rooms, cement platforms for drying coffee and the corral. Homegarden is a system of intensive cultivation of plants (edible, medicinal, and ornamental, among others) for self-consumption. The labor force is contributed primarily by the woman and young children (Vásquez-Dávila and Manzanero 2015). In Zoogochí, fruit trees are scattered among coffee plantations, which serve as latrines. In addition to human manure, biodegradable garbage and organic matter produced by the different crops are added to the soil. It is common to transplant wild plants to the garden, mainly flowers and those used for medicinal purposes. Some of the plants present in Zoogochí’s home gardens are the following: piñón (Jatropha curcas L., whose seeds are edible in small quantities); cuachipil (Diphysa americana (Mill.) M. Sousa); nochebuena (Euphorbia pulcherrima Willd. ex Klotzsch), margarita (Bellis perennis L.), huele de noche (Cestrum nocturnum L.), marigold (Tagetes erecta L.), tulipán (Hibiscus rosasinensis L.), coffee (Coffea arabica L.), sugar cane (Saccharum officinarum L.), chile mirasol (Capsicum annuum var. annuum L.), prickly pear cactus (Opuntia ficus-indica (L.) Mill.). Among the fruit trees are: banana (Musa sp., with four varieties: ratán, cuatro lomos, enano and Campeche); papaya (Carica papaya L.), mamey (Pouteria sapota (Jacq.) H.E. Moore & Stearn.), mango (Mangifera indica L.), avocado (Persea americana Mill.), fig (Ficus carica L.), anona (Annona cherimola Mill.), orange (Citrus  aurantium L.), lemon (Citrus  limon (L.) Osbeck), lime (Citrus aurantiifolia (Christm.) Swingle), grapefruit (Citrus x paradisi Macfad.), sapote (Manilkara chicle (Pittier) Gilly), guava (Psidium guajava L.), and loquat (Rhaphiolepis bibas (Lour.) Galasso & Banfi).

The Coffee Plantation The natural conditions of El Rincón are favorable for the cultivation of coffee, which allowed the Zapotec communities to have a source of economic income. At the end of the 1940s and the beginnings of the 1950s, there was a coffee plantation boom. New towns were founded in the area, as Santa María La Luz, founded by immigrants from the municipality of San Juan Atepec, which also belongs to the district of Ixtlán. The existing coffee plantations in the area were formed from the process of clearing the mountain cloud forest or emerged in the resting lands or in the homegardens.

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

885

In relation to ethnoedaphology, the people know that the most adequate soils for planting coffee are “those with fertilized soil that is neither sad nor dead, where marshmallow (Malvaceae) grows, with a yellowish color and sandy texture.” In order to establishing, planting, taking care of, and harvesting a coffee plantation, the farmers of Zoogochí carried out the process shown in Table 4, which takes an average of 4 years. In a more detailed manner, the complete process consists of the following steps and activities: For the case of a “virgin land,” the land is cleared by eliminating the herbaceous and shrub strata, leaving the most suitable trees to provide shade for the coffee trees. For the local farmers, a good shade tree is one that provides shade in a regulated manner, since “there are trees such as the mulato (Bursera simaruba (L.) Sarg.), which in April sheds but then loses its leaves: this is not good. The suitable ones are those that change their leaves, but not completely, like the guamúchil (Pithecellobium dulce Mart.) or the guanacastle (Enterolobium cyclocarpum (Jacq.) Griseb.) that “fertilize the soil and give strength to the plants that grow next to them.” They also consider to be beneficial the cola de iguana (Cordia alliodora (Ruiz et Pavón) Oken) and the palo de piedra (Homalium senarium Sessé & Moc). Once the land has been cleared, stakes are used, which are obtained on the spot and distributed on the land. They are used to mark the rows where the coffee plants will be placed. Subsequently, in each place marked by a stake a hole of 50 cm in diameter and equal depth is made using a spade or shovel. It is then filled with compost, which is “rotten soil, not leaf litter.” The planter then places the coffee plant in the place of the stake. The seedlings can come from the farmers’ own land. The plants live for 3 years in the seedbed; at the moment of transplanting “they are just pulled out of the soil.” The plants used are those that already have “three crosses,” that is, three whorls or branches, one per year. Weeding is carried out twice a year, in May and October, with a curved machete. The weeds depend on the age of the coffee plantation: the older the plantation, the less weeds. Two years after planting, the coffee bush is bent so that lateral branches sprout from the stem, leaving five shoots. To support the bent stem, a stick hook is buried in the ground. In preparation for this work, the enclosed area is cleared of weeds, the plant is bent, and the tip is cut off so that only the five shoots remain. This process is a type of formation pruning. Production begins 4 years after transplanting. For the coffee harvest, the cherries are cut by hand and placed in small wicker baskets (made with Monstera deliciosa Table 4 Sequence of activities in coffee cultivation in Zoogochi, Sierra Norte of Oaxaca Year First Second Third Fourth

Main activity Preparing the soil Sowing Cultivation work Cultivation work Cultivation work

Agricultural work Sweeping, staking, making crates, filling in crates Planting three-year seedlings To incline the coffee plant, weeding Weeding Weeding, harvesting

886

M. A. Va´squez-Da´vila et al.

Liebm.). The average yield is 2.5 quintals per hectare (one quintal is equivalent to 45 kg). After harvesting, the coffee beans are pulped, fermented, washed, and roasted in the patios or asoleaderos of the houses. The pulping is generally done in manual pulping machines. The process ends with the packing of the beans and their transportation to the city of Oaxaca, where they are sold.

Gathering and Food Consumption Plant (and animal) gathering is an activity also practiced by the Zapotec of Zoogochí. They collect quelites (edible greens), firewood, wild fruits, culinary, and medicinal herbs. Among the wild plants used as medicine are palo mulato (Bursera simaruba (L.) Sarg,) and bishate (Solanum americanum Mill.). As for firewood, “firewood can be cut from any tree, but the best are oak (Quercus spp.) and pine (Pinus spp.).” The trees are cut down and, in order for them to lose moisture, they are cut up and sunbathed in the patios. The ocote (Pinus teocote Schiede ex Schltdl.) was formerly used for lighting, but is no longer used since electric lights were installed in the town. The most commonly consumed food is corn (Zea mays L.), in the form of tortillas. Nixtamal (corn cooked in water with lime) is ground on a metate. A metate is a ground stone tool used for processing grain (like corn) and seed (like cacao). The tortillas, 30 cm in diameter, are molded with the palm of the hand on the metate. An adult consumes six to seven tortillas per day. Next in importance are beans (Phaseolus spp.) and chile (Capsicum annuum var. annuum), squashes (Cucurbita spp.), and nopales (Opuntia spp.). Chicken eggs are frequently consumed and are prepared in sauce or batten. Vegetables (exogenous) are consumed in small quantities (generally by families with greater economic resources). Epazote (Dysphania ambrosioides (L.) Mosyakin & Clemants.), coriander (Coriandrum sativum L.), and mint (Mentha spicata L.) are used as condiments. Coffee sweetened with panela is currently the most common beverage. To prepare it, the beans are roasted on a comal and then ground in a metate. Comal is a flat, clay griddle used to cook tortillas or other foods. Two tablespoons of ground coffee are used to prepare one liter of coffee beverage. Regarding the eating schedule, breakfast is eaten between 8 and 9 o’clock in the morning. Breakfast consists of coffee and tortillas, sometimes with chilmole (chile sauce) or salt. Lunch is between noon and one o’clock in the afternoon. Occasionally they eat eggs, fowl or game meat, fish, or rice. Dinner is at about 7 o’clock at night, they drink coffee and tortillas. Sometimes, people drink tepache – made from fermented sugar cane (Saccharum officinarum L.) – and chínguere, a liquor made by distilling the fermented tepache. A lot of mezcal is drunk, which is bought in Oaxaca. The change in production (from staple crops to coffee) created new food patterns. With the income from coffee sales, it was easier to buy food than to grow it. Around 1940, stores appeared or increased. Table 5 shows some traditional elements and the change in diet. The following section discusses the role of the three agroecosystems described so far and provides data from another locality in the Sierra Norte.

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca Table 5 Traditional and modern elements of the Zapotec diet in the Sierra Norte of Oaxaca

Traditional Corn tortilla Corn tortilla Corn tortilla Corn tortilla; pumpkin seeds Quelites (greens) Panela; honey Aguas frescas (natural drinks) Tepache and aguardiente Fresh fish Fresh or dried chili Mother’s milk Creole egg

887 Modern Wheat flour Wheat bread Crackers Pasta soup Instant soup Refined sugar Soft drinks Beer Canned fish Canned chili Powdered milk Commercial egg

Complementarity of Mountain Agroecosystems: Las Delicias, Sierra Norte Las Delicias is a small community in the municipality of San Juan Juquila Vijanos with a 62 km2 territory located between 900 m and 2400 m above sea level. It has two climates: semi-humid and humid temperate (INEGI 2005). Rainfall is common throughout the year but is more abundant in summer (Pérez-García and del Castillo 2016) and temperature ranges from 16  C to 22  C. The landscape consists of pine forest, pine-oak forest, cloud forest, riparian vegetation, secondary vegetation, and cultivated areas (del Castillo and Blanco-Macías 2007). The slash-and-burn system has led to a mosaic of natural vegetation interspersed with cultivated areas (Pérez-García and del Castillo 2016). Broadly speaking, 78% is forest, 20% is agricultural land, and 2% is the town (INEGI 2005). Land ownership is communal (González 2001). Following communal rules, villagers can usufruct land in both parts of the territory, classified locally as tierra fría (cold land) and tierra caliente (warm land).

Agroecosystems The cornfields or milpas are assemblages of populations or agro-communities of corn (Zea mays), beans, squash, other crops, weeds, animals, fungi, and the soil microbiota. There are two types of cornfields or milpas: those located in the lowlands or “warm land” and those in the highlands or “cold land.” In the warm land, the yunta is used to remove the soil, the planting spaces are itinerant. Yunta is a pair of animals that work together, by means of a yoke, in order to carry out agricultural work. In the cold land, slash and burn is used on terraced permanent agricultural land. Around the land there are trees with ornamental value or for shade, such as Magnolia yajlachhi (Domínguez-Yescas and Vázquez-García 2019) and of food

888

M. A. Va´squez-Da´vila et al.

interest such as Mangifera indica. As a result, there is a greater richness of useful plants in cold-land agricultural areas. Four types of maize are sown by color: white, red, black, and yellow. Varieties of squashes (Cucurbita spp.), beans (Phaseolus spp.), chayotes (Sechium), bananas (Musa), and chiles (Capsicum spp.) are also grown. Examples of edible weeds are Galinsoga parviflora and Phytolacca icosandra. Self-consumption is the main destination of the milpa plants. Vegetables such as Allium neapolitanum and legumes such as Pisum sativum and Vicia faba are sold. The species that are exchanged are beans (Phaseolus spp.) and squashes (Cucurbita spp.). The Zapotec of Las Delicias conserve 238 species of useful plants in their homegardens, coffee plantations, and cornfields. In these spaces, a complementarity is developed that allows them to satisfy the needs of the inhabitants and conserve biodiversity. A total of 159 species are grown in homegardens, 148 in coffee plantations and 41 in milpas (Pascual-Mendoza et al. 2020). Homegardens occupy the space in front of and behind the houses. Each family establishes its arrangement based on available space and particular needs. The fences are spaces where vegetables are grown, delimited with a mesh to prevent the passage of chickens and sheep. The work in the family gardens is mainly conducted by women. Men remove the soil for planting, weeding, and irrigating. In the gardens, women decide to introduce and experiment with ornamental and edible species (Vásquez-Dávila and Manzanero 2015). Homegardens have edible plants such as Allium neapolitanum, Dysphania ambrosoides, Coriandrum sativum, Capsicum spp., Psidium guajava, Prunus persica, and Citrus. In homegardens there is a high representation of ornamental plants, mainly introduced species. Among the most frequent are agapanthus (Agapanthus africanus), gladiolas (Gladiolus sp.), alcatraces (Zantedeschia aethiopica), Rosa spp., and Lilium spp. The coffee multiple gardens are planted in areas with secondary vegetation, and are accompanied by diverse crops. In some cases, they are planted in areas of pine forest, with P. chiapensis being the trees that shade the crop. Therefore, the diversity associated with this agroecosystem is mainly made up of shade trees, such as Liquidambar straciflua, Alnus acuminata, Clethra Mexicana, and Inga jinicuil. Other uses of the trees in the coffee plantations are firewood, edible, construction, live fences, and medicine. Table 6 shows the multiple uses of 11 coffee trees. The coffee plantation also contains some medicinal herbaceous species such as Tithonia diversifolia, Cestrum nocturnum, and Gnaphalium viscosum. Each family has at least one coffee plot whose harvest is used for trade and self-consumption. Most of the family’s income comes from shade-grown coffee.

Categories of Plant Use The agroecosystems of Las Delicias provide diverse services such as food, medicines, and ornamental, among others. In the locality, there are 113 edible species, 79 ornamental, 50 fuel, 48 used as shade, 47 medicinal, 31 live fences, 13 for construction, 13 forage, and 9 for domestic use (Pascual-Mendoza et al. 2020).

Species Diphysa americana (Mill.) M. Sousa Prunus persica (L.) Batsch. Psidium guajava L. Alnus acuminata Kunth. Citrus medica L. Erythrina americana Mill. Inga jinicuil Schltdl. Inga edulis Mart. Mangifera indica L. Manilkara chicle (Pittier) Gilly. Persea schiedeana Nees Total

* * * * * * * 10

Edible * * *

Shade * * * * * * * * * * 10

Firewood * * * * * * * * * * 10

7

4

*

*

* * * * * *

Living fences * *

Construction *

Table 6 Multiple use of ten trees in the coffee plantations of Las Delicias, Sierra Norte of Oaxaca

* 3

* *

Medicinal

Total 5 5 4 4 4 4 4 4 4 4 2 44

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca 889

890

M. A. Va´squez-Da´vila et al.

Among the edible herbaceous plants are: Amaranthus hybridus, Crotalaria longirostrata, Portulaca oleracea, Sechium edule, and Cucurbita spp. Tree species such as oranges, guavas, peach, and bananas provide fruits. Accessibility and free of charge are two desirable characteristics of local medicinal plants. The most commonly used plants to alleviate stomach ailments are Mentha spicata, Chamaemelum nobile, citrus, and arnica. In another Zapotec community in the Sierra Norte (Camotlán), 90 medicinal species are used (Pérez-Nicolás et al. 2018). Inga spp., Cecropia obtusifolia, Heliocarpus donnellsmithii, and H. appendiculatus are species traditionally used by villagers as fuel or firewood. In the 1970s, INMECAFE promoted the use of the Citrus genus as shade. Live fences are used to delimiting, protecting and establishing boundaries for crops and livestock, in addition to providing food, medicine, ornament, and fuel. These plants provide scenic beauty, firewood for fuel, and produce edible fruits such as bananas. Plants used for construction include species of the genera Liquidambar, Alnus, Inga, Pinus, and Quercus. Forage species such as Cyperus esculentus and Setaria parviflora are herbaceous plants that are used as fodder for cattle and chickens. Examples of species with domestic use are: Lagenaria siceraria which is used to transport water or to serve food or coffee; Saccharum officinarum, whose dried leaves are used to protect panela; and Sida rhombhifolia with which brooms are made to sweep patios or the interior of houses.

Mountains Are Living Pharmacies: Coatla´n, Sierra Sur San Sebastián Coatlán is a municipality located in the Sierra Sur of Oaxaca, at an elevation of 1980 m. The total area of the municipality is 190 km2. It has three types of climates: warm sub-humid, semi-warm humid, and temperate sub-humid. Temperate forest predominates (75%), followed by pastureland (12%), sub-humid forest (6%), agricultural land (6%), and urban area (1%). Land tenure is communal, and local people (2678 inhabitants) are governed through the uses and customs regime (INEGI 2015). Most of the population is engaged in corn and bean agriculture. The Zapotec of Coatlán use 117 medicinal plant species (herbaceous, trees, shrubs, and climbers) to attend illnesses of all body systems including culturally affiliated ailments (Santiago-Martínez 2018). Homegardens in San Andrés Paxtlán, Miahuatlán, provide 150 medicinal plants (Zurita-Vásquez et al. 2012). The Zapotec of San Juan Mixtepec, in the Sierra Sur, have known most of the plants in their environment since the age of 10 years old and use 209 medicinal species (Hunn 2008). These plants are obtained from the mesophyll cloud forest, high evergreen forest and medium sub evergreen forest. Table 7 shows the number of families, genera, and medicinal species used in Coatlán. The botanical families with the highest number of species are Asteraceae with 16 species, Lamiaceae (n ¼ 9), Solanaceae (n ¼ 9), Leguminosae (n ¼ 6), and Rutaceae (n ¼ 5). In the Pochutla District, located on the coastal slope of the Sierra Sur, the Leguminosae, Asteraceae, and Solanaceae families predominate among the different categories of useful plants (Luna-José and Rendón-Aguilar 2008). The best

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

891

Table 7 Families, genera, and species of medicinal plants used in San Sebastián Coatlán, Sierra Sur of Oaxaca Group Angiosperms Gymnosperms Pteridophytes

Family 48 1 3

Genus 98 1 3

Species 113 1 3

represented genera are Citrus and Solanum, with four species each. About 70 species are native plants (either from the country, the region, or the whole Americas) and 47 are introduced species. Almost half of the species of medicinal plants (50 species) are found in homegardens, 23 occur along roadsides, 13% are in the forest or bush, 13 species can only be found in the market in Miahuatlán, 9 are riparian, and 7 are found in cultivated fields, either as cultivated plants or as weeds (Santiago-Martínez 2018). In Coatlán, the aerial part (38%) and leaves (29%) are the main plants used. To a lesser extent, the bark (9%), fruit (6%), and flower (4%) are used. Less used are stems, roots, and seeds (3%), bulbs (2%), and resin, tubers, and latex (1%). According to the degree of management of medicinal plants, the highest percentage corresponds to cultivated plants, followed by wild plants, and in decreasing order, plants that are tolerated, domesticated, promoted, and protected (Santiago-Martínez 2018). In San Sebastián Coatlán, Asteraceae is the family with the most medicinal species used. Asteraceae is one of the largest families of vascular plants in Mexico and the most diverse, representing approximately 13.5% of the total richness of the country (Rzedowski and Rzedowski 2005; Villaseñor 2018). This result reflects the floristic diversity of the state of Oaxaca, where the Asteraceae family is represented by 858 species, 502 of which are endemic to Mexico, and 113 are endemic to the state (Villaseñor 2018). Among the native species, Amphipterygium adstringens, Selaginella lepidophylla, and Hintonia latiflora can be mentioned as plants that are part of the medicinal ethnobotanical knowledge held by people in the community. Another case is that of species with a special protection status such as Litsea glaucescens, which is classified as an endangered plant in the Mexican Official Standard NOM-059SEMARNAT-2010. The medicinal importance attributed to it in the community is to remove fever, stomach pains, and energy, in all cases the leaves are used in infusions. Ethnobotanical research has made it possible to characterize the social factors related to the traditional use of this and other species and to understand the subsequent implications in terms of conservation. The existing record of the multiple uses of species in traditional medicine also indicates a long history of use and interaction of local cultures with plants (Jiménez-Pérez et al. 2011). The appropriation of introduced plants by community members evidences some characteristics of traditional knowledge such as its dynamism, collective transmission, and continuous expansion (Pochettino et al. 2008). In Coatlán, medicinal plant management practices that strengthen resilience were documented. These include:

892

M. A. Va´squez-Da´vila et al.

(1) flexibility in the use of different plant species to treat common diseases; (2) use of more abundant and more accessible species; (3) diversity of collection sites; and (4) knowledge of the different parts of the plants used. With these strategies of use and management, the pressures on habitats, species or plant parts collected decreases (Santiago-Martínez 2018). Women and older adults have a greater knowledge of local medicinal flora (Vázquez-Medina et al. 2011; Vázquez-García and Ortega 2016; Pérez-Nicolás et al. 2018), which can be explained in terms of their important role as primary healthcare providers in their families (Albuquerque et al. 2011) and to also to their multiple activities related to agriculture and the cultivation of useful plants in homegardens (Manzanero et al. 2009).

Final Comments The cultural diversity of the State of Oaxaca is a consequence of the cultural adaptation to the ecological diversity of its spaces, so that nature-culture form an inseparable unit. In fact, to a great extent, the ecological variability of Oaxaca influenced both the cultural peculiarities of the different zones and the cultural unity of the region by promoting an intense exchange of the products of each area. A policy that seeks the conservation of the biological and ecological heritage of the mountains cannot be implemented apart from one that seeks to support the cultural heritage that inhabits and symbolizes them. It is not possible in rugged Oaxaca to adopt a merely biological approach to mountain nature. To conceive of the sierras as de-socialized, dehumanized, or desacralized means adopting a reductionist and scientistic position, mistaken from a historical, social and political point of view. In the mountains of Oaxaca, each species of plant, fungus or animal, type of soil and subunit of the landscape almost always corresponds to a linguistic expression, a category of knowledge, a practical use, a religious content, or an individual or collective experience. In Oaxaca, where culture gives meaning to nature and nature gives meaning to indigenous societies, it is very important to transcend the scientistic conception of nature. Ecosystem services and ethnobotanical aspects performed by mountain people should be considered as key elements in the design of biocultural conservation policies. Hence, the policy for biodiversity conservation in the mountains of Oaxaca should integrate cultural aspects with nature protection in at least two ways: one historical and the other currently occurring. In the first case, it is desirable to create ecological archaeological spaces or zones such as in Yagul, Monte Alban, Guiengola, or Monte Negro, just to mention some of the vestiges of the pre-Hispanic Zapotec and Mixtec, but the same should be thought of for the other mountain cultures with historical antecedents in the territory of present-day Oaxaca. Similarly, each of the ethnic groups that inhabit the mountains of Oaxaca possess spaces or areas of ethno-biodiversity, which current conservationists could support (from the scientific discipline and managing both their legal protection and economic and institutional resources) to benefit from the declarations of priority biocultural

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

893

territories (Boege 2008), such as the Omora Ethnobotanical Park, located in Chile, which protects the southernmost forest on the planet (Rozzi et al. 2006). As a landscape or as a resource, mountains must be conserved and utilized, which inescapably implies the creation of its own design in the field of biocultural conservation in Oaxaca (Vásquez-Dávila and Manzanero 2019). To achieve this will require more research and action on mountain ecology and anthropology, people better prepared in inter- and transdisciplinary work (ethnobiologists, taxonomists, ecologists, geographers, anthropologists), strong regional centers of biological conservation, and a new type of political relations with indigenous people that prioritizes a new style of development that is more participatory, more conservationist, but, above all, more humane (CONABIO 2018). In the participatory design of conservation plans for the biocultural diversity of the mountains of Oaxaca, the importance (economic, social, and political) that this implies must be considered. It is vitally important to demonstrate that the conservation and rational use of natural resources is the true basis for sustainable development. In other words, sustainable development is impossible without effective conservation of biodiversity and respect, revaluation and support for cultural plurality. Are we willing to collaborate with the millenary inhabitants of the mountains in the study of biodiversity management in their ancestral territories to achieve its sustainable use and conservation? Are we capable of respecting the ideas and beliefs of others to build a common future, or is our interest in the biodiversity and culture of the owners and guardians of the most biodiverse territories on the planet only rhetorical?

Dedication To those who opened the gap of Ethnobotany in Mexico: Efraím Hernández X., Miguel Ángel Martínez Alfaro, and Javier Caballero, dear teachers and guides. Acknowledgments To the Zapotec families of the communities we worked with; to Mario Lavariega Nolasco for the elaboration of the maps; to Elí García Padilla and Carlos Alberto Masés García for allowing the use of their valuable photographic testimonies. GIMM thanks the Instituto Politécnico Nacional (IPN) for its support of the project “Ethnoecology of chile (Capsicum spp.) in indigenous localities of Oaxaca” (SIP number 20211931) and the COFFA grant. SPM and ASM thank to Consejo Nacional de Ciencia y Tecnología (CONACYT) and IPN for the scholarships to study the Master’s program at CIIDIR-Oaxaca of the IPN.

References Aguilar-Santelises MR. Etnobotánica cuantitativa en una región de bosque de niebla de la Sierra Norte, Oaxaca. MSc. Thesis. Oaxaca: CIIDIR; 2007. Aguilar-Støen M, Moe SR, Camargo-Ricalde SL. Home gardens sustain crop diversity and improve farm resilience in Candelaria Loxicha, Oaxaca, Mexico. Hum Ecol. 2009;37(1):55–77. Aguilar-Støen M, Angelsen A, Stølen KA, Moe SR. The emergence, persistence, and current challenges of coffee forest gardens: a case study from Candelaria Loxicha, Oaxaca, Mexico. Soc Nat Resour. 2011;24(12):1235–51.

894

M. A. Va´squez-Da´vila et al.

Aguirre-Beltrán G. Regiones de refugio: El desarrollo de la comunidad y el proceso dominical en Mestizoamérica. Xalpa: Universidad Veracruzana; 2009. Albuquerque UP, Soldati GT, Sieber SS, Ramos MA, de Sá JC, de Souza LC. The use of plants in the medical system of the Fulni-ô people (NE Brazil): a perspective on age and gender. J Ethnopharmacol. 2011;133(2):866–73. Arriaga-Cabrera L, Aguilar V, Espinoza JM. Regiones prioritarias y planeación para la conservación de la biodiversidad. In: Dirzo R, González R, March IJ, editors. Capital Natural de México, volumen II: Estado de Conservación y Tendencias de Cambio. México: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2009. Aswani S, Lemahieu A, Sauer WH. Global trends of local ecological knowledge and future implications. PLoS One. 2018;13(4):e0195440. Ban NC, Frid A, Reid M, Edgar B, Shaw D, Siwallace P. Incorporate indigenous perspectives for impactful research and effective management. Nature Ecol Evol. 2018;2(11):1680–3. Bautista LJ, Damon A, Ochoa-Gaona S, Tapia RC. Impact of silvicultural methods on vascular epiphytes (ferns, bromeliads and orchids) in a temperate forest in Oaxaca, Mexico. For Ecol Manag. 2014;329:10–20. Bélisle AC, Asselin H, LeBlanc P, Gauthier S. Local knowledge in ecological modeling. Ecol Soc. 2018;23(2) Berg RL. El sistema de plaza Zoogocho en el distrito de la sierra zapoteca de Villa Alta. In: Diskin M, Cook S, editors. Mercados de Oaxaca. México: Instituto Nacional Indigenista; 1975. p. 116–43. Bernal-Ramírez LA, Bravo-Avilez D, Fonseca-Juárez RM, Yáñez-Espinosa L, Gernandt DS, Rendón-Aguilar B. Usos y conocimiento tradicional de las gimnospermas en el noreste de Oaxaca, México. Acta Bot Mex. 2019;126(e1471):1–24. https://doi.org/10.21829/abm126. 2019.1471. Boege E. El patrimonio biocultural de los pueblos indígenas de México. Hacia la conservación in situ de la biodiversidad y agrodiversidad en los territorios indígenas. México: Instituto Nacional de Antropología e Historia, Comisión Nacional para el Desarrollo de los Pueblos Indígenas; 2008. Caballero J, Cortés L, Martínez-Alfaro MA, Lira SR. Uso y manejo tradicional de la diversidad vegetal. In: García-Mendoza AJ, Ordóñez-Díaz MJ, Briones-Salas M, editors. Biodiversidad de Oaxaca. Ciudad de México: Instituto de Biología-UNAM/Fondo Oaxaqueño para la Conservación de la Naturaleza/WWF; 2004. p. 541–64. Campos-Villanueva A, Cortés-Arriaga L, Dávila-Aranda P, García-Mendoza A, Reyes-Santiago J, Toris-Acosta G, Torres-Colín L, Torres-Colín R. Plantas y Flores de Oaxaca. Instituto de Biología, Universidad Nacional Autónoma de México; 1992. Carbajal-Esquivel H. Importancia de las plantas en la cultura alimentaria de la comunidad Xi’oi Las Guapas, Rayón, San Luis Potosí. MSc. Thesis. México: Universidad Autónoma de San Luis Potosí; 2008. Castro-Juárez CJ, Villa-Ruano N, Ramírez-García SA, Mosso-González C. Uso medicinal de plantas antidiabéticas en el legado etnobotánico oaxaqueño. Rev Cuba Plantas Med. 2014;19(1):101–20. CONABIO (Comisión Nacional para el Conocimiento y Uso de la Biodiversidad). Estrategia para la conservación y el uso sustentable de la biodiversidad del estado de Oaxaca. México: CONABIO; 2018. Cruz-Arenas RA, Cruz-Hernández AM. Mercados, hierbas y comerciantes: el intercambio de herbáceas en el tianguis de Ixtlán de Juárez, Oaxaca. Etnobiología. 2021;19(2):79–96. Cruz-Espinoza AD, Guendulain-Hernández N. Conocimiento tradicional de plantas medicinales de la comunidad de Santa Catarina Quioquitani, Yautepec, Oaxaca. Residencia Profesional. Xoxocotlán: Instituto Tecnológico del Valle de Oaxaca; 2013. Cruz-Pérez AL, Barrera-Ramos J, Bernal-Ramírez LA, Bravo-Avilez D, Rendón-Aguilar B. Actualized inventory of medicinal plants used in traditional medicine in Oaxaca, Mexico. J Ethnobiol Ethnomed. 2021;17(1):1–15.

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

895

Del Castillo RF, Blanco-Macías A. Secondary succession under slash and burn regime in a tropical montane cloud forest: soil and vegetation characteristics. In: Newton AC, editor. Biodiversity loss and conservation in fragmented forest landscapes: evidence from the forests of Montane Mexico and Temperate South America. Wallingford: CAB International; 2007. https://doi.org/ 10.1079/9781845932619.0158. Delgadillo-Moya C. Native moss names in Mexico. Evansia. 2000;17:84–6. Domínguez-Lagunes MA. Conocimiento local y crecimiento inicial del ojite (Brosimum alicastrum Swartz) bajo distintas condiciones de manejo en la zona de Lomeríos del municipio Paso de Ovejas, Veracruz, México. MSc. Thesis. Colegio de Posgraduados Campus Veracruz: Veracruz; 2012. Domínguez-Yescas R, Vázquez-García JA. Flower of the heart, Magnolia yajlachhi (subsect. Talauma, Magnoliaceae), a new species of ceremonial, medicinal, conservation and nurse tree relevance in the Zapotec culture, Sierra Norte de Oaxaca, Mexico. Phytotaxa. 2019;393(1): 21–34. https://doi.org/10.11646/phytotaxa.393.1.2. Fletcher MS, Hamilton R, Dressler W, Palmer L. Indigenous knowledge and the shackles of wilderness. Proc Natl Acad Sci. 2021;118(40):e2022218118. Flores Martínez A, Manzanero Medina GI. Tipos de vegetación del Estado de Oaxaca. In: Vásquez Dávila MA, editor. Vegetación y Flora. Sociedad y Naturaleza en Oaxaca, vol. 3. México: Instituto Tecnológico Agropecuario de Oaxaca; 1999. p. 7–48. García-Mendoza AJ, Meave JA, editors. Diversidad florística de Oaxaca: de musgos a angiospermas (colecciones y lista de especies). México: Universidad Nacional Autónoma de México y Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2011. García-Serrano S, López-Hernández TD. Cocimiento tradicional de las plantas medicinales del municipio de Santiago Comaltepec, Oaxaca. Bachelor’s Thesis. Xoxocotlán: Instituto Tecnológico del Valle de Oaxaca; 2016. Gómez-Hernández F. Sistematización del conocimiento local de la medicina tradicional en Capulalpam de Méndez, Oaxaca. Residencia Profesional. Xoxocotlán: Instituto Tecnológico del Valle de Oaxaca; 2010. Gómez-Luna RE, Manzanero-Medina GI, Vásquez-Dávila MA. Florística y aspectos sociales de huertas zapotecas en Lachatao, Sierra Norte de Oaxaca, México. Revista Bio Ciencias. 2017;4(4):15. Gómez-Mendoza L, Vega-Peña E, Ramírez MI, Palacio-Prieto JL, Galicia L. Projecting land-use change processes in the sierra Norte of Oaxaca, Mexico. Appl Geogr. 2006;26(3–4):276–90. https://doi.org/10.1016/j.apgeog.2006.09.003. González RJ. Zapotec science: farming and food in the northern sierra of Oaxaca. Texas: University of Texas Press; 2001. Hernández-Rodríguez E, Delgadillo-Moya C. The ethnobotany of bryophytes in Mexico. Bot Sci. 2021;99(1):13–27. https://doi.org/10.17129/botsci.268. Hernández-Rodríguez E, López-Santiago J. Uses and traditional knowledge of Dendropogonella rufescens (Bryophyta: Cryphaeaceae) in a Zapotec community of southeastern Mexico. Bot Sci. 2022;100(1):153–68. Hunn ES. A Zapotec natural history: trees, herbs, and flowers, birds, beasts, and bugs in the life of San Juan Gbee‘. Tucson: University of Arizona Press; 2008. INALI (Instituto Nacional de Lenguas Indígenas). Catálogo de las Lenguas Indígenas Nacionales. Variantes lingüísticas de México con sus autodenominaciones y referencias geoestadísticas. México: INALI; 2008. INEGI (Instituto Nacional de Estadística y Geografía). Prontuario de información geográfica municipal de los Estados Unidos Mexicanos. San Juan Juquila Vijanos, Oaxaca. Clave geoestadística 202001. INEGI: México; 2005. INEGI (Instituto Nacional de Estadística y Geografía). Principales resultados de la Encuesta Intercensal 2015 Oaxaca. México: INEGI; 2015. INEGI (Instituto Nacional de Estadística y Geografía). Censo de Población y Vivienda 2020. México: INEGI; 2020.

896

M. A. Va´squez-Da´vila et al.

Jiménez-Pérez NC, Lorea-Hernández FG, Jankowski CK, Reyes-Chilpa R. Essential Oils in Mexican Bays (Litsea spp., Lauraceae): Taxonomic Assortment and Ethnobotanical Implications. Econ Bot. 2011;65(2):178–89. López-Mendoza LM. Los huertos familiares del municipio de Santa Catarina Lachatao, Ixtlán, Oaxaca. Bachelor’s Thesis. Universidad de la Sierra Juárez; 2020. López-Santiago J. Plantas útiles del municipio de San Pablo Macuiltianguis, Oaxaca. Bachelor’s Thesis. Ingeniería Forestal. Oaxaca: Universidad de la Sierra Juárez; 2015. Luna-José AL, Rendón-Aguilar B. Recursos vegetales útiles en diez comunidades de la sierra madre del sur, Oaxaca, México. Polibotánica. 2008;26:193–242. Luna-José AL, Rendón-Aguilar B. Traditional knowledge among Zapotecs of Sierra Madre Del Sur, Oaxaca. Does it represent a base for plant resources management and conservation? J Ethnobiol Ethnomed. 2012;8(1):1–13. https://doi.org/10.1186/1746-4269-8-24. Manzanero Medina GI, Flores Martínez A, Hunn E. Los huertos familiares zapotecos de San Miguel Talea de Castro, Sierra Norte de Oaxaca, México. Etnobiología. 2009;7(1):9–29. Manzanero Medina GI, Vásquez-Dávila MA, Lustre Sánchez H, Gómez-Luna RE. Estado del arte sobre los estudios etnobotánicos de los huertos familiares de Oaxaca. In: Ordóñez Díaz MJ, editor. Atlas biocultural de huertos familiares en México: Chiapas, Hidalgo, Oaxaca, Veracruz y Península de Yucatán. CRIM-UNAM; 2018. https://doi.org/10.22201/crim.9786073007405e. Martínez-Bautista BG, Bernal-Ramírez LA, Bravo-Avilez D, Samain MS, Amezcua JMR, RendónAguilar B. Traditional uses of the family Piperaceae in Oaxaca, Mexico. Trop Conserv Sci. 2019;12 https://doi.org/10.1177/1940082919879315. Martínez-López J, Acosta-Ramos A, Martínez y Ojeda E, Manzano-Méndez F. Recursos forestales no maderables en dos comunidades zapotecas de la Sierra Juárez de Oaxaca. Rev Mex Cienc Forestales. 2016;7(35):37–52. Martínez-Salas E, Ramos CH. Biodiversidad de Pteridophyta en México. Rev Mex Biodivers. 2014;85:110–3. Mondragón D. La comercialización navideña de bromelias epífitas en la ciudad de Oaxaca, México. Etnobiología. 2008;6(1):24–8. Mondragón D, Villa-Guzmán DM. Estudio etnobotánico de las bromelias epífitas en la comunidad de Santa Catarina Ixtepeji, Oaxaca, México. Polibotánica. 2008;26:175–91. Morrone JJ. Regionalización biogeográfica y evolución biótica de México: encrucijada de la biodiversidad del Nuevo Mundo. Rev Mex Biodivers. 2019;90:e902980. https://doi.org/10. 22201/ib.20078706e.2019.90.2980. Mraz-Bartra AL. Lucha y resistencia en comunalidad. Ecofronteras. 2020:30–2. Ordaz-Santana EC, Ordaz-Santana LM. Conocimiento tradicional sobre plantas medicinales en Santa María Velató, Monjas, Miahuatlán, Oaxaca. Residencia Profesional. Instituto Tecnológico del Valle de Oaxaca: Xoxocotlán; 2018. Pascacio-González MM. Catálogo de las plantas medicinales de Santa Catarina Lachatao, Ixtlán, Oaxaca México. Residencia Profesional. Xoxocotlán: Instituto Tecnológico del Valle de Oaxaca; 2011. Pascual-Mendoza S, Manzanero Medina GI, Saynes-Vásquez A, Vásquez-Dávila MA. Agroforestry systems of a Zapotec community in the Northern Sierra of Oaxaca, Mexico. Bot Sci. 2020;98(1):128–44. Pascual-Mendoza S, Saynes-Vásquez A, Pérez-Herrera A. Traditional knowledge of edible plants in an indigenous community in the Sierra Norte of Oaxaca, Mexico. Plant Biosyst. 2021:1–13. Pérez-García O, del Castillo RF. The decline of the itinerant milpa and the maintenance of traditional agrobiodiversity: crops and weeds coexistence in a tropical cloud forest area in Oaxaca, Mexico. Agric Ecosyst Environ. 2016;228:30–7. https://doi.org/10.1016/j.agee.2016. 05.002. Pérez-José C, García-Morales PA. Estudio etnobotánico y contenido nutrimental de Salmea scandens en Buena Vista Loxicha, Pochutla, Oaxaca. Bachelor’s Thesis. Instituto Tecnológico del Valle de Oaxaca: Xoxocotlán; 2016.

Biocultural Ethnobotany of the Zapotec Mountains of Oaxaca

897

Pérez-Nicolás M, Vibrans H, Romero-Manzanares A. Can the use of medicinal plants motivate forest conservation in the humid mountains of northern Oaxaca, Mexico? Bot Sci. 2018;96(2): 267–85. Pochettino ML, Lema V, Capparelli A. ¿Aprendices de shaman o piratas de la naturaleza?: apropiación del conocimiento botánico tradicional y ética etnobotánica. In: Archila S, Giovannetti M, Lema V, editors, Arqueobotánica y teoría arqueológica: discusiones desde Suramérica. Colombia: Universidad de los Andes, Centro de Estudios Socioculturales e Internacionales; 2008. p. 253–71. Ramírez-Aragón AC. Estudio etnobotánico y bromatológico de Cryosophilla nana en Candelaria Loxicha, Pochutla, Oaxaca. Bachelor’s Thesis. Instituto Tecnológico del Valle de Oaxaca: Xoxocotlán; 2016. Ramírez-Martínez F. Conocimiento local sobre plantas medicinales en la comunidad de San Juan Atepec, Ixtlán, Oaxaca. Residencia Profesional. Xoxocotlán: Instituto Tecnológico del Valle de Oaxaca; 2017. Rees J. The Oaxaca Christmas plant market. J Bromeliad Soc. 1976;26(6):223–32. Rozzi R, Massardo F, Anderson CB, Heidinger K, Silander JA. Ten principles for biocultural conservation at the southern tip of the Americas: the approach of the Omora Ethnobotanical Park. Ecol Soc. 2006;11(1) Rzedowski J. Vegetación de México. México: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2006. Rzedowski GC, Rzedowski J. Flora fanerogámica del Valle de México. Pátzcuaro: Instituto de Ecología y Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2005. Sánchez-Antonio JC. Genealogía de la comunalidad indígena: Descolonialidad, transmodernidad y diálogos inter-civilizatorios. Lat Am Res Rev. 2021;56(3):696–710. Santiago-Martínez A. Conocimiento tradicional y valor de uso de plantas medicinales de San Sebastián Coatlán, Miahuatlán, Oaxaca. México: Instituto Politécnico Nacional. Centro Interdisciplinario de Investigación para el Desarrollo Integral Regional Unidad Oaxaca; 2018. Sosa-Pérez A. Plantas medicinales en el municipio de San Juan Evangelista Analco, Ixtlán, Oaxaca. Residencia Profesional. Xoxocotlán: Instituto Tecnológico del Valle de Oaxaca; 2013. Suárez-Mota ME, Villaseñor JL, Ramírez-Aguirre MB. Priority sites for the conservation of floristic richness and endemism of the Northern Sierra of Oaxaca, Mexico. Acta Bot Mex. 2018;124 https://doi.org/10.21829/abm124.2018.1296. Tello-Ortega KE, Hernández Santiago E, Rodríguez OG. Medicina alternativa complementaria en el tratamiento de enfermedades crónicas en el sur de Oaxaca, México. Ciencia Ergo Sum. 2020;27(2):134–45. Vásquez-Cortez VF. Estructura poblacional de Litsea glaucescens Kunth en bosques bajo manejo de la sierra norte de Oaxaca. MSc Thesis. Texcoco: Colegio de Posgraduados, Montecillo; 2016. Vásquez-Dávila MA, Manzanero Medina GI. Campesinidad y socialización en las huertas familiares de las mujeres zapotecas de la sierra norte de Oaxaca, México. Negocios & Desarrollo. 2015;2(2):92–117. Vásquez-Dávila MA, Manzanero Medina GI. Ecología biocultural: tejiendo la biodiversidad con el bienestar de Oaxaca: Reflexión metodológica. Tercer Encuentro Nacional sobre Estrategias de Biodiversidad. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad. Ciudad de México, 11 y 12 de julio. 2019. https://doi.org/10.13140/RG.2.2.13015.80800. Vásquez-Dávila MA, Manzanero Medina GI. Mercados regionales de Oaxaca como nodos bioculturales complejos. In: Moctezuma-Pérez S, Sandoval Genovez D, editors. Mercados y tianguis en el siglo XXI. Repensando sus problemáticas. Toluca: Universidad Autónoma del Estado de México; 2021. p. 241–63. Vázquez-García V, Ortega TO. Gender, local governance and non-timber forest products. The use and management of Satureja macrostema in Oaxaca's central valleys, Mexico. Women's Stud Int Forum. 2016;65:47–52.

898

M. A. Va´squez-Da´vila et al.

Vázquez-García V, Madamombe-Manduna I, Vibrans H. Género y conocimientos etnobotánicos en México y Zimbabwe. Un estudio comparativo. Sociedades Rurales, Producción y Medio Ambiente. 2009;18:21–48. Vázquez-Medina B, Martínez Corona B, Aliphat Fernández MM, Aguilar CA. Uso y conocimiento de plantas medicinales por hombres y mujeres en dos localidades indígenas en Coyomeapan, Puebla, México. Interciencia. 2011;36(7) Venegas-Ramírez Y. Plantas y hongos comestibles comercializados en el mercado de Miahuatlán, Oaxaca, México. Bachelor’s Thesis. México: Universidad Autónoma Benito Juárez de Oaxaca; 2009. Ventura-Aquino Y, Rendón B, Rebollar S, Hernández G. Use and conservation of forest resources in the municipality of San Agustin Loxicha, Sierra Madre del Sur, Oaxaca, Mexico. Agrofor Syst. 2008;73(3):167–80. Villaseñor JL. Catálogo de las plantas vasculares nativas de México. Rev Mex Biodivers. 2016;87(3):559–902. Villaseñor JL. Diversidad y distribución de la familia Asteraceae en México. Bot Sci. 2018;96(2): 332–58. https://doi.org/10.17129/botsci.1872. Weiss J. Comparación entre la medicina chatina y zapoteca de Oaxaca (México) y la medicina tradicional china. In: Vásquez-Dávila MA, editor, Recursos vegetales de Oaxaca. Sociedad y Naturaleza en Oaxaca 2. México: CONACYT, Instituto Tecnológico Agropecuario de Oaxaca; 1995. p. 33–52. Zacarías-Eslava Y, del Castillo RF. Comunidades vegetales templadas de la Sierra Juárez, Oaxaca: pisos altitudinales y sus posibles implicaciones ante el cambio climático. Bol Soc Bot Méx. 2010;87:13–28. Zurita-Vásquez GG, Manzanero Medina GI, Flores-Martínez A. Flora con uso medicinal de los huertos familiares de la Sierra Sur de Oaxaca. In: Vásquez-Dávila MA, Lope-Alzina D, editors. Aves y huertos de México. México: Carteles Editores-CONACYT; 2012. p. 118–9. Zurita-Vásquez GG, Manzanero Medina GI, Vásquez-Dávila MA, Lustre-Sánchez H. Mujer, huerta familiar zapoteca y seguridad alimentaria en San Andrés Paxtlán, Sierra Sur de Oaxaca, México. In: Moreno-Calles AI, et al., editors. Los Sistemas Agroforestales de México: Avances, experiencias, acciones y temas emergentes en México. UNAM - Red Temática de Sistemas Agroforestales de México del CONACYT; 2020. p. 209–23.

Nopal de Monte: Cacti Named and Used by a Mixtec Community in Mountainous Oaxaca Luis E. Ortiz-Martínez, Gladys I. Manzanero-Medina, Jordan Golubov, Marco Antonio Va´squez-Da´vila, and María C. Mandujano

Abstract

Indigenous communities are significant reservoirs of traditional ethnobotanical knowledge and represent a central source of information about use and management of wild plants. Cacti are among the most representative species in Neotropical deciduous forests. Since pre-Hispanic times, numerous species of Cactaceae are used by Indigenous communities of mountain regions as sources of food, medicine, dyes, fodder, and materials for construction. We documented cacti species used in a community located in the Mixtec highlands of Oaxaca: San Sebastián del Monte. This small town, located at the top of a mountain, is inhabited by proficient speakers of Mixtec. We obtained information regarding the native cacti traditional nomenclature and uses. Additionally, we calculated a use value index for each species. Twelve cacti were registered; all of them are known by its Mixtec and Spanish names and had some kind of traditional use. L. E. Ortiz-Martínez Instituto Politécnico Nacional, Centro Interdisciplinario de Investigación para el Desarrollo Integral Regional, Unidad Oaxaca, Mexico G. I. Manzanero-Medina Instituto Politécnico Nacional, Centro Interdisciplinario de Investigación para el Desarrollo Integral Regional, Unidad Oaxaca, Oaxaca, Mexico e-mail: [email protected] J. Golubov Universidad Autónoma Metropolitana Xochimilco, Ciudad de México, Mexico e-mail: [email protected] M. A. Vásquez-Dávila (*) Instituto Tecnológico del Valle de Oaxaca, Oaxaca, Mexico Tecnológico Nacional de México, Campus Valle de Oaxaca, Xoxocotlán, Oaxaca, Mexico e-mail: [email protected] M. C. Mandujano (*) Instituto de Ecología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_56

899

900

L. E. Ortiz-Martínez et al.

A remarkable name is nopal de monte, which refers to Opuntia velutina and its characteristic mountainous habitat. Food is the main local use of cacti (fruits of ten species are consumed). Since every structure of Ferocactus recurvus is used, the species has the highest use value index. Even when Opuntia pubescens is one of the species with lowest use value index, the traditional use of its fruits (prickly pears) to dye corn tortillas with an intense pink or purple color is a feature with great nutraceutical, culinary, and tourist potential. The stems of three species are used as medicine; two species are used as dyes, one as fodder and another as fuel. As in other communities from semiarid environments, cacti are an important source of food and raw materials in the mountain town of San Sebastián del Monte. Local knowledge on use and management is relevant for their conservation as biocultural heritage. Every season, when the nopal criollo (Opuntia decumbens Salm-Dyck) bears its fruits, several Mixtec families in San Sebastián del Monte prepare and consume very special corn tortillas. These tortillas are unique because of their beautiful deep pink color (provided by the pulp of the nopal criollo fruit). These tortillas are consumed at home and are also taken to sell in the municipal capital, Santo Domingo Tonalá. Although the local people do not attribute a ritual or sacred value to the pink tortillas, they do recognize them as something special and appreciate them for their color as well as their flavor. The academic interest in discovering the process of preparing the pink tortillas led us to conduct the ethnobotanical study of the Cactaceae family in San Sebastián de Monte presented in this chapter.

Introduction Tropical deciduous forests offer a rich variety of wild and cultivated useful plants (Pimienta-Barrios 1994; Ramírez-Rodríguez et al. 2020). Cacti (Cactaceae) are among the most representative groups of species in these environments. Species such as Hylocereus spp., Stenocereus spp., and Opuntia spp. are extensively cultivated and are valuable as crops (Ramírez-Rodríguez et al. 2020). Additionally, many wild cacti species are harvested and used by rural communities (Pérez-Negrón et al. 2014). The flowers, fruits, seeds, stems, and roots of many wild cacti species are used as food, medicine, dyes, fodder, and construction material (Bravo-Hollis 1978; Cota-Sánchez 2016). Indigenous communities are the main reservoirs of traditional ethnobotanical knowledge and represent an important source of information about the potential use of wild plants (Casas et al. 2014). Such traditional knowledge can be useful to identify new crops, diversified food production systems, to providing economic alternatives for local populations and developing strategies for their conservation and sustainable use (Pío-León et al. 2017). Rural Indigenous communities are under constant transformation, sometimes leading to the gradual loss of the knowledge concerning the use of wild plants (Shanley and Rosa 2004). In this chapter, we analyze the use of cacti species by an Indigenous community at the Mixteca region in Oaxaca state, Mexico. Through ethnobotanical interviews, we aim to answer the following questions: Which are the local names and uses of cacti in the community? And what are the use value of wild cacti species?

Nopal de Monte: Cacti Named and Used by a Mixtec Community in Mountainous . . .

901

Materials and Methods Study area. San Sebastián del Monte is a Mixtec community located at 17 230 N; 98 020 W, at the Sierra Madre del Sur in the southern Mexican state of Oaxaca. The town is located on the top of a mountain at an altitude of 1430 m above the sea level. The climate is semiarid, with annual precipitation averaging 440 mm, annual mean temperature ranging from 20 to 25  C, and the vegetation being dominated by tropical deciduous forest and xerophytic scrub (Katz 2008). The current population of San Sebastián del Monte is around 1500 inhabitants, most of them speakers of Mixtec, whose main economic activity is agriculture (INEGI 2010). Traditional knowledge and use value. In accordance with the Code of Ethics of the International Society of Ethnobiology for data collection, we showed our research projects and its objectives to the local traditional authorities (Commissariat of Communal Property) before beginning our study. With the help of a Spanish to Mixtec translator, we obtained Free Prior and Informed Consent from all the local collaborators. Two local guides conducted us through different plots which are used to collect plants. During these explorations, we identified all the cacti species present in the area. Semistructured interviews were held with 38 local collaborators from August to September 2017. In each interview, we asked collaborators about their age, main economic activity, and years of schooling. We also showed them photographs of the species or in front of the specimens during the fieldwork and asked the interviewees about their use of the species, the names of the plants in Spanish and Mixtec, and the structures they use and a brief description of the preparation methods. All the interviews were registered in notebooks with oral consent of the local collaborators. Interviews lasted 15–20 min on average. Use Value Index developed by Phillips and Gentry (Phillips and Gentry 1993) modified by Rossato et al. (1999) was employed to measure the relative importance of native cacti. This index was estimated as the sum of the uses recognized by each local collaborator for each species divided by the total number of collaborators according to the following formula: UVs ¼ (Uis)/N, where Uis is the number of uses mentioned by collaborator i for species s, and N is the total interviewees for each species.

Results Socioeconomic Characteristics of the Collaborators The ethnobotanical survey included 29 female and 9 male collaborators. Age of the collaborators averaged 50 years old (range 18–81 years). Total 76% of the interviewed people were bilingual (Mixtec and Spanish), and 24% had Mixtec as single language. The main economic activity of 63% of the collaborators was agriculture, and 37% were occupied in secondary activities (construction and manufacturing). The mean years of schooling was 3 years, and 39% of the collaborators never received formal schooling.

902

L. E. Ortiz-Martínez et al.

Traditional Knowledge Twelve cacti species were registered in the locality. These included five members of the Pachycereeae tribe: Escontria chiotilla (F.A.C. Weber ex Schum.) Rose, Cephalocereus mezcalaensis (Bravo) Backeb., Pachycereus weberi (J.M. Coult.) Backeb., Stenocereus griseus (Haw.) Buxb., and S. pruinosus (Otto ex Pfeiff.) Buxb. (Fig. 1); four from the Cacteae: Coryphantha retusa Britton & Rose, Ferocactus recurvus Britton & Rose, Mammillaria polyedra Mart., and M. dixanthocentron Backeb. (Fig. 2); and three Opuntioideae: Opuntia decumbens Salm-Dyck, O. pubescens H. Wendl. ex Pfeiff., and O. velutina F.A.C. Weber (Fig. 3). Every species was identified by its Mixtec and Spanish names (Table 1).

Uses and Use Value Index The use citations were classified into six categories: human food, fodder, medicinal, fuel, tool, and ornamental (Table 2). The most common use of the cacti in the community is food; all species except Opuntia pubescens are used for this purpose. Fruits are the most utilized structure, mainly as fresh fruit. The fruits of O. decumbens are used as pigment; fruits are grounded and mixed with corn to

Fig. 1 Cacti from the Pachycereeae tribe native to San Sebastián del Monte, Oaxaca, México: (a) Cephalocereus mezcalaensis; (b) Pachycereus weberi; (c) Escontria chiotilla; and (d) Senocereus griseus. (Photos by the authors)

Nopal de Monte: Cacti Named and Used by a Mixtec Community in Mountainous . . .

903

Fig. 2 Cacti from the Cacteae tribe native to San Sebastián del Monte, Oaxaca: (a) Mammillaria dixanthocentron; (b) Mammillaria polyedra; (c) Coryphantha retusa; and (d) Ferocactus recurvus. (Photos by the authors)

Fig. 3 Cacti from the Opuntioideae native to San Sebastián del Monte, Oaxaca: (a) Opuntia decumbens Salm-Dyck; (b) Opuntia velutina. (Photo by the authors)

prepare very special pink tortillas, tortillas are one of the staples of people’s diet, and this way of cooking them is a seasonal practice (Fig. 4). Only the fruits of O. velutina and O. pubescens are not considered edible. The flowers of C. mezcalaensis are consumed as vegetable. The stems are the second most utilized structure. Tender

904

L. E. Ortiz-Martínez et al.

Table 1 Nomenclature of the wild cacti species of San Sebastián del Monte, Oaxaca, Mexico Species Coryphantha retusa Escontria chiotilla

Spanish names Biznaga

Ferocactus recurvus Mammillaria polyedra Mammillaria dixanthocentron Cephalocereus mezcalaensis Opuntia decumbens Opuntia pubescens

Biznaga, Piña, Uña de gato

Opuntia velutina Pachycereus weberi Stenocereus griseus Stenocereus pruinosus

Mixtec names Tucushini, Ticushini ia, Shitala la Sholacate, Tochichi ya, Tochichi none, Tochichi ne Ticushini, Ticushi no

Jiotilla, pitayita

Chilitos, Biznaguita blanca

Ticushinini izo, Ticushini shinano, Ticushini ia Ticushinini izo, Ticushini shinano, Ticushini ia Tochichi be

Chilitos, Biznaguita blanca Pitayo, Órgano Nopal criollo, Tuna Cholla, Espina de burro, Nopal de burro, Espina de perro Nopal de campo, Nopal de monte

Chiqui iaie, Tomin dia, Tomin dia cana chiqui iaie Iñoji, Tomin dia

Chiquito, Órgano, Pitaya

Tomin cua, Tomin dia, Tomin azuco Tochiquito

Pitaya de agosto, Órgano

Tochichi caia, Tochichi tina,

Pitaya de mayo

Tochichi cua, Tochichi, Tochichi mayo

Table 2 Use, structure used, and number of mentions of useful cacti in San Sebastian del Monte, Oaxaca, Mexico Species Coryphantha retusa Escontria chiotilla Ferocactus recurvus Mammillaria polyedra Mammillaria dixanthocentron Cephalocereus mezcalaensis Opuntia decumbens Opuntia pubescens Opuntia velutina Pachycereus weberi Stenocereus griseus Stenocereus pruinosus

Use category Medicinal Human food Medicinal, human food, tool Ornamental, human food Ornamental, human food Human food, pigment Human food, pigment Medicinal Fodder, human food Human food, fuel Human food Human food

Number of mentions 29 37 35

Structure used Stem Fruit Stem, fruit, and spine Whole plant

31

Whole plant

29

Fruit, stem

33

Fruit Stem Stem Stem, fruit Fruit Fruit

37 27 30 38 35 38

Nopal de Monte: Cacti Named and Used by a Mixtec Community in Mountainous . . .

905

Fig. 4 Use value index of the cacti species of San Sebastián del Monte, Oaxaca, Mexico. (Data obtain upon open interviews from 38 local collaborators (From June to August 2017)) Fig. 5 Fruits of Opuntia decumbens used as eatable pigment to make a unique traditional tortilla that is pink in color at San Sebastián del Monte, Oaxaca, Mexico. (Photo by the authors)

pads called nopalitos (i.e., racket-shaped stems or cladodes) of O. velutina are consumed as vegetables; mature pads called nopales are used as fodder. Dried portions of the stem of P. weberi are used as firewood. The stems of three species have medicinal uses; C. retusa and F. recurvus are boiled and used as an infusion to treat kidney stones. The pads of O. pubescens are used in an infusion to improve the memory, and sections of the stem C. mezcalaensis are grounded and applied to the hair as gray hair dye. Only three plants are used as a hole, and the two Mammillaria species are used as ornamental plants; people recognized them as a single species.

906

L. E. Ortiz-Martínez et al.

Ferocactus recurvus was the species with the highest use value index (VU) (Fig. 5), every structure of the plant has at least one use, its stem is used to make a candy known in Mexico as acitrón, and its spines are used as hooks for the elaboration of hats and crafts made with the leaves of Brahea dulcis (Kunth) Mart. (Arecaceae).

Discussion The 12 cacti species registered in the study area have wide distribution ranges and currently are not included in any list of threatened species. Except for Stenocereus griseus, all the species are endemic to Mexico. Three of the species, Escontria chiotilla, Mammillaria dixanthocentron, and Ferocactus recurvus, are endemic to the states of Oaxaca and Puebla. Ethnobotanical studies have documented that in the Mixteca region, as well as in the Tehuacán-Cuicatlán Valley, columnar cacti species such as S. pruinosus, S. stellatus, E. chiotilla, Polaskia chichipe, and Myrtillocactus geometrizans are used to complement the diet of rural people (Casas et al. 1993; Mercado and Granados 1999), and are also exchanged for other crops in traditional markets (Arellanes et al. 2013). Fruits of columnar cacti are highly valuable as crops in Mexico and have been pointed out as a potential source of income of rural communities. Pérez-Negrón et al. (2014) estimated the total value of columnar cactus fruits to be $315.07 US dollars per hectare in the Tehuacán-Cuicatlán Valley. We did not measure the rates of extraction; however, extraction rates reported in other works appear not to be detrimental for the conservation of any of the species (Casas et al. 1993; Pérez-Negrón et al. 2014). Opuntia decumbens has the same use value (UV) as E. chiotilla and S. pruinosus. Local people use O. decumbens fruits as edible pigment that adds aesthetic and nutritional value to tortillas, a basic element of their diet. The use of natural pigments to dye tortillas and embellish them for religious rites is known in other parts of Mexico, including other localities from the Mixteca region, in which Hibiscus bifurcatus Cav. (Malvaceae) is used to dye a tortilla, and the species is known as flor de tortilla (tortilla flower); this traditional tortilla is only prepared in special occasions, such as weddings or altars for special festivities (Katz 2018). The addition of fruits and seeds of cacti to tortillas is a common practice in some localities of southern Mexico; it has been documented the seeds of Pachycereus weberi and S. pruinosus are added to tortillas to increase their nutritional value (Luna-Morales and Aguirre 2001). The commercial value of O. decumbens has not been explored, and the species is not included in any traditional management systems. Remarkably, the pigments of O. decumbens did not change their hue or intensity when exposed to heat in tortilla manufacture. Betalains are the pigments responsible for the color of Opuntia fruits (Castellar et al. 2003). Several studies have proved that Opuntia fruits are excellent sources of betalains since they have important nutritional and chemical properties with application in diverse industries (Butera et al. 2012; Slimen et al. 2017). However, betalains are known to suffer thermal degradation (color loss or browning)

Nopal de Monte: Cacti Named and Used by a Mixtec Community in Mountainous . . .

907

when they are exposed to high temperatures during prolonged heating periods (Slimen et al. 2017); perhaps there are other compounds present in O. decumbens that prevent color degradation, but this has to be studied yet. Some of the uses of the cacti cannot be considered sustainable. This is the particular case of the Ferocactus recurvus used to prepare acitrón candy. Acitrón is a crystallized candy made with the stems of several barrel cactus species belonging to the Echinocactus and Ferocactus genera (Vásquez-Dávila and Manzanero Medina 2019). The decline of many barrel cactus species due to the harvest of mature plant for the elaboration of this candy has forced the Mexican Government to prohibit the fabrication and trade of acitrón. Nevertheless, many people still manufacture, sell, and consume acitrón, as it is an important ingredient of several traditional dishes. The most common threats for cacti species in the studied area and in other arid regions of Mexico are ecosystem transformation into agricultural fields, overgrazing, and illegal collection of plants for the ornamental plant market (Goettsch et al. 2015). The effect of these activities could be lowered with the design of agroecological systems based on the experience of local communities that guarantee cacti conservation as well as income generation (Pérez-Negrón et al. 2014). People in rural communities have a long history of practice of in situ and ex situ traditional management of wild plants; ex situ practices include the tolerance, propagation, dispersal, and protection of wild plants in their natural habitats and within agricultural lands, and ex situ activities include in the propagation of useful plants in homegardens and other agroforestry systems (Casas et al. 2014).

Perspectives The studied site at the Mixteca Sierra Madre del Sur Mountain range experiences high levels of disturbance, mainly by the transformation of ecosystems into temporal agriculture plots. Further research is needed to assess the health of the cacti populations. The use of the fruits of Opuntia decumbens could be extended to other areas and become a source of income for the communities of the Mixteca region. However, more research is needed to identify the chemical properties of the pigments found in the fruit of this prickly pear and its potential uses.

References Arellanes Y, Casas A, Arellanes-Meixueiro A, Vega E, Blancas J, Vallejo M, Torres I, Solís L, Pérez-Negrón E. Influence of traditional markets and interchange on plant management in the Tehuacán Valley. J Ethnobiol Enthomed. 2013;9:38. Bravo-Hollis H. Las cactáceas de México ed. 2, vol. 1. Ciudad de México: Universidad Nacional Autónoma de México; 1978. Butera D, Tesosiere L, Di Gaudio F, Bongiorno A, Allegra D, Pintaudí A, Kohen R, Livrea M. Antioxidant activities of Sicilian prickly pear (Opuntia ficus-indica) fruit extracts and

908

L. E. Ortiz-Martínez et al.

reducing properties of its betalains: betanin and indicaxanthin. J Agric Food Chem. 2012;50: 6895–901. Casas A, Caballero J, Valiente-Banuet A. Use, management and domestication of columnar cacti in south-central Mexico: a historical perspective. J Ethnobiol. 1993;19:71–95. Casas A, Camou A, Rangel-Landa S, Solís L, Torres I, Delgado-Lemus A, Moreno AI, Vallejo M, Guillén S, Blancas J, Parra F, Aguirre X, Farfán-Heredia B, Arellanes Y, Pérez-Negrón E. Manejo tradicional de biodiversidad y ecosistemas en Mesoamérica: el valle de Tehuacán. Investigación Ambiental, Ciencia y Política Pública. 2014;6:23–44. Castellar M, Obón J, Alacid M, Fernández-López J. Color properties and stability of betacyanins from Opuntia fruits. J Agric Food Chem. 2003;51:2772–6. Cota-Sánchez JH. Nutritional composition of the prickly pear (Opuntia ficus- indica) fruit. In: Simmonds MSJ, Preedy VR, editors. Nutritional composition of fruit cultivars. San Diego: Academic Press; 2016. p. 691–712. Goettsch A, Hilton-Taylor C, Cruz-Piñon G, Gaston KJ, et al. High proportion of cactus species threatened with extinction. Nat Plants. 2015;1:15142. INEGI. Censo de Población y vivienda. Principales resultados por localidad (ITER). 2010. Katz E. Vapor, aves y serpientes: meteorología en la “Tierra de la Lluvia” (Mixteca Alta, Oaxaca). In: Lammel A, Goloubinoff M, Katz E, editors. Aires y lluvias: antropología del clima en México. CIESAS; 2008. Katz E. El chile en la Mixteca alta de Oaxaca: de la comida ritual. In: Aguilar-Meléndez A, Vásquez-Dávila MA, Katz E, Hernández MR, editors. Los chiles que dan sabor al mundo. Veracruz: Universidad Veracruzana; 2018. p. 177–212. Luna-Morales C, Aguirre R. Clasificación tradicional, aprovechamiento y distribución ecológica de la pitaya mixteca en México. Interciencia. 2001;26:18–24. Mercado A, Granados D. La pitaya: biología, Ecología, fisiología, sistemática, Etnobotánica. Mexico: Universidad Autónoma de Chapingo; 1999. Pérez-Negrón E, Dávila P, Casas A. Use of columnar cacti in the Tehuacán Valley, Mexico: perspectives for sustainable management of non-timber forest products. J Ethnobiol Ethnomed. 2014;10:79. Phillips O, Gentry A. The useful plants of Tambopata, Peru: statistical hypotheses tests with a new quantitative technique. Econ Bot. 1993;47:15–32. Pimienta-Barrios BE. Prickly pear (Opuntia spp.): a valuable fruit crop for the semi-arid lands of Mexico. J Arid Environ. 1994;28:1–12. Pío-León J, Delgado-Argas F, Murillo-Amador B, León de la Luz J, Vega-Aviña R, Nieto-GaribayA, Cordoba-Matson M, Ortega-Rubio A. Environmental traditional knowledge in a natural protected area as the basis for management and conservation policies. Journal of. Environ Manag. 2017;201:63–71. Ramírez-Rodríguez Y, Martínez-Huélamo M, Pedraza-Chaverri J, Ramirez V, Martínez-Tagüeña N, Trujillo J. Ethnobotanical, nutritional and medicinal properties of Mexican drylands Cactaceae fruits: recent findings and research opportunities. Food Chem. 2020;312:126073. Rossato SC, Leitao-Filho HF, Begossi A. Ethnobotany of Caicareas of the Alantic fores coast (Brazil). Econ Bot. 1999;53:387–95. Shanley P, Rosa NA. Eroding knowledge: an ethnobotanical inventory in eastern Amazonia’s logging frontier. Econ Bot. 2004;58:135–60. Slimen IB, Najar T, Abderrabba M. Chemical and antioxidant properties of betalains. J Agric Food Chem. 2017;65:675–89. Vásquez-Dávila MA, Manzanero Medina GI. Conservación biocultural de las cactáceas empleadas para elaborar el acitrón en México. Cact Suc Mex. 2019:58–61.

Floristic Diversity on Rubber Plantations and Their Importance for Subsistence at Foothill Landscapes of Mexico Juan Carlos Lo´pez-Acosta, Emmanuel Ismael Pantoja-Aparicio, Jorge Antonio Go´mez-Díaz, Maite Lascurain-Rangel, and Ina Falfa´n

Abstract

Human activities in the tropics of Veracruz, Mexico, have transformed its original vegetation into anthropized landscapes. This contemporary configuration of the territory raises questions about the capacity of these anthropized landscapes to maintain plant diversity. However, there are few detailed studies on the contribution of anthropized landscapes in maintaining floristic diversity. An ideal scenario to study the tropical landscapes within such perspective are the rubber plantations of the Uxpanapa Valley, Veracruz, because of its economic and coverage importance. The Uxpanapa Valley was once considered one of the most prominent centers of plant diversity in Mexico; however, 80% of its vegetation has disappeared due to the severe deforestation carried out in the last 50 years. Currently, there are plantations of Hevea brasiliensis in areas of low elevation (100–300 m) in the zone. In lower areas there are other important crops, such as orange, and large areas of grassland for livestock, setting a heterogeneous landscape in the Uxpanapa Valley. In this study, we identified structural plant elements associated with rubber plantations which have functions of subsistence and provision of income for owners who manage them under sustainability principles. We analyzed the contribution of the rubber plantations in the retention of woody plant diversity and characterized the rubber plantations management. We sampled five independent rubber plantations using 10 transects of 50  2 m in each plantation (0.5 ha total). In each of the sampling sites, we registered individuals J. C. López-Acosta · E. I. Pantoja-Aparicio · J. A. Gómez-Díaz Centro de Investigaciones Tropicales, Universidad Veracruzana, Xalapa, Veracruz, Mexico e-mail: [email protected]; [email protected] M. Lascurain-Rangel (*) Red Ambiente y Sustentabilidad, Instituto de Ecología, A.C, Xalapa, Veracruz, Mexico e-mail: [email protected] I. Falfán Red Ambiente y Sustentabilidad, Instituto de Ecología, A.C, Xalapa, Mexico © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_15

909

910

J. C. Lo´pez-Acosta et al.

with DBH  1 cm and those 30 cm in height (seedlings). We also conducted semi-structured interviews with 32 rubber producers. With these data, we quantified diversity metrics and characterized the type of management given to rubber plantations in the region. We found a high plant diversity associated with these systems, 197 species, many of them typical of primary forests (native trees) which could play a role in restoration programs and the promotion of the landscape connectivity of the Uxpanapa Valley. The heterogeneity of the plant richness reflects the management practices carried out by the rubber producers and the period of abandonment of the plantations. The interviews showed that the management of rubber plantations is varied, and selective plating prevails, particularly of useful species, which explains the high levels of species richness. Also, the results revealed the ability of rubber plantations to sustain the arrival and establishment of a high plant diversity. It is essential to encourage selective management that avoids eliminating useful species, both timber and non-timber, and promote the diversification of rubber plantations with species typical of tropical rainforest, as a strategy to increase their value and their potential as biological corridors to protect regional flora and fauna. The recognition of rubber plantations as elements of the landscape and their interaction with the human component is an indispensable step for the design of restoration and conservation strategies of the rural landscape in the Uxpanapa Valley.

Introduction Several studies refer to the broad disciplinary and conceptual approaches leading to analyze specific interactions between humans and nature (Ellen and Fukui 1996). These include questions in relation to how societies have been organized in the process of domestication of plant species and their environments, as well as the causes and effects articulated to socio-environmental, economic, cultural, and political factors related to such interactions at the level of species, ecosystems, and landscapes (Ellen and Fukui 1996; Wiersum 2008). Humans have intervened in the structure of natural forests for millennia, building a network between nature and culture, where people live together and depend on plant species “by living in and utilizing ‘nature’ and by assimilating it into ‘culture’” (Ellen and Fukui 1996; Levis et al. 2018). The complex and diverse practices of small producers, with its own socio-environmental and cultural conditions, have originated different stages of crop management, incorporating changes in the properties of species and production systems, building a process of coevolution (Wiersum 2008). This is of great relevance in the context of the several factors that reduce the diversity of important species for the subsistence of local communities (Lascurain-Rangel et al. 2019). There is an urgent need for finding solutions that achieve both forest resources provision for food security and conservation of the systems diversity to maintain ecosystems services and undertake restoration strategies. Millions of people depend on plant resources for food, health, shelter, and income, and the demand is growing by both present and future generations. Nonetheless, there is an evident modification of the vegetation cover, loss of species of flora and fauna,

Floristic Diversity on Rubber Plantations and Their Importance. . .

911

disturbance of ecological processes, and erosion of traditional knowledge (SánchezColón et al. 2009) affecting the strategies for food, medicine, and income security.

The Rubber Plantations in the Uxpanapa Valley, Veracruz, Mexico The Uxpanapa Valley in Veracruz, like the rest of Mexico, has undergone a profound anthropogenic impact, which significantly increased in the last 40 years (Ewell and Poleman 1982; Velasco-Toro 2011). In 1972 began the construction of the Cerro de Oro dam in the state of Oaxaca, which involved the displacement of indigenous population, mainly Mazatec and Chinantec people, to resettle in the Uxpanapa Valley. This phenomenon caused the erosion of the original culture of their new inhabitants and promoted the deforestation of the Valley (Velasco-Toro 2011; Velasco-Toro and Vargas-Montero 1990). This process was due to changes in land use, accelerated population growth, technological/industrial capacity, and the advance of the agricultural frontier (López-Acosta et al. 2011; Revel-Mouroz 1972). These anthropogenic changes transformed the tropical landscape of the Uxpanapa Valley into fragmented and deforested ecosystems immersed in an agricultural matrix (Gascon et al. 2004; Guevara et al. 2006). In 1990, the change in land cover and its use became more notorious with the introduction of citrus cultivation and rubber plantation, both with 16% occupancy of the Uxpanapa Valley, decreasing the forest cover from 58% to less than 16% in 1976, and increasing pastures in 28.3% (Hernández-Gómez 2014). The current configuration of the Uxpanapa Valley landscape consists in primary rainforests, secondary forest (abandoned vegetation in different states of successional stages and development time), riparian vegetation, and plantations of timber, perennial crops, live fences, and pastures for livestock (PRONATURA 2007; Zamora-Pedraza 2012) (Fig. 1). The situation of the Uxpanapa Valley landscape raises questions about the results of these changes in the medium and long term, as well as on its potential for maintaining the biological diversity and structure of management systems in benefit of the local people, and about what make them stable, predictable, and functional in one of the most transformed and ecologically, socially, culturally, and biologically affected areas of Mexico. Among the most conspicuous elements of the current landscape of the Uxpanapa Valley are the rubber plantations (Ortiz-Hernández 2011; PRONATURA 2007; Rojo-Martínez et al. 2011). Rubber plantations are systems formed by trees of Hevea brasiliensis, native to the Amazon rainforest (de Souza et al. 2012) (Fig. 2). Its planting was part of a technological package implemented by the Mexican government to promote the regional development which replaced natural forest for this crop (Ewell and Poleman 1982). Over time, rubber plantations have prevailed by resistance to local conditions and by having an economic niche facilitated by the locals, which has even grown in recent years (Monroy-Rivera and NávarChaidez 2004; Rojo-Martínez et al. 2011). Rubber plantations are an important element of the landscape in the region (Rojo-Martínez et al. 2011), representing 54% of the tree cover in the municipalities of Tezonapa, Las Choapas, and Uxpanapa (Monroy-Rivera and Návar-Chaidez 2004). The typical management

912

J. C. Lo´pez-Acosta et al.

Fig. 1 Overview of the Valley of Uxpanapa. (Photo by the authors)

Fig. 2 Hevea brasiliensis showing the scratching marks for rubber recollection. (Photo by the authors)

characteristics of these crops are the excessive use of chemical inputs and mechanical management methods (Rojo-Martínez et al. 2011). This chapter focuses on recognizing the main structural plant elements associated with rubber plantations, which have functions of subsistence and income for the owners who managed them sustainably (Hernández-Gómez et al. 2011; Wendt and Van der Werff 1987). Also, we present the retention values of woody floristic diversity in rubber plantations and its contribution to the well-being of the communities. The potential for biodiversity retention is the capacity of the different

Floristic Diversity on Rubber Plantations and Their Importance. . .

913

landscape elements to provide successful conditions for the recruitment, establishment, and survival of flora and fauna (Dirzo et al. 2009). This process is maintained due to the intrinsic characteristics of the system, which offer microclimatic conditions, specific habitats, shelter, food, and nutrients for survival. Therefore, we expected that the richness and diversity associated with these landscape elements would depend on the quality of these conditions, as well as the connectivity and intensity of the anthropogenic disturbances.

Methods Study Area The Uxpanapa Valley is located at the southeast of the state of Veracruz (Fig. 3). It belongs to the Neotropical belt of forests and rainforests in the southeast of Mexico (Espinosa-Organista et al. 2008) and is the only natural bridge between the sub-humid and humid tropical forests located on the Mexican coasts of the Pacific Ocean and the Gulf of Mexico (Halffter et al. 2008). It has an area of 1912 km2 and limits to the north with the municipalities of Jesús Carranza, Hidalgotitlán, Minatitlán, and Las Choapas and to the south with the state of Oaxaca. Most of the area is slightly rugged within an altitudinal range of 60 to 250 m (INEGI 2010). The climate is warm-humid with abundant rains in summer, reaching values of precipitation higher than 3000 mm per year in the highest altitudinal zones, the average annual temperature ranges between 24 and 26 C (INEGI 2010).

Vegetation Sampling In this study we carried out a sampling of vegetation from July 2010 to September 2011, in rubber systems. We used the methodology proposed by Gentry (1982, 1988) with some modifications (López and Dirzo 2007). The objective of this methodology is to analyze the vegetation richness, structure, and composition. In five representative plots of rubber plantations, we sampled 10 transects of 50  2 m (100 m2), covering a total of 0.1 ha per plot (0.5 total). We put transects randomly to avoid overlapping and possible edge effects of the plots. Within each transect, we identified and registered all individuals with a diameter at breast height (DBH)  1 cm, considering that such species constitute the current structure of rubber plantations. Additionally, in each transect we identified and registered all individuals whose DBH was 1 cm, but with a height  30 cm. We considered this group of individuals/species (seedlings) as the vegetation in regeneration which demonstrates the potential for species turnover for each of the sampled sites. We calculated the importance value index (I.V.I.) of each species (Lamprecht 1990). This function was obtained by summing the relative density (RDe), relative frequency (RF), and relative dominance (RDo) of each species. Where RDe ¼

914

J. C. Lo´pez-Acosta et al.

Fig. 3 Location of the Uxpanapa Valley, Veracruz, Mexico, and the rubber plantations in the Valley

number of individuals / 1,000 m2  100; RF ¼ number of transects in which the species appears / total number of transects  100; RDo ¼ % basal area of all individuals of the species / % basal area of the entire community  100. Consequently, the maximum value for the species larger than 1 cm of DBH was 300%. Likewise, we analyzed the species regeneration by adding up the RDe and the RF; RDe and RF have the same meaning as in I.V.I., but in this regeneration importance value, only the accumulated value for all species at this stage is estimated (Lamprecht 1990). Consequently, the maximum I.V.I. of species lower than 1 cm of DBH and taller than 30 cm was 200%. The result of this descriptive analysis is a hierarchical arrangement of the I.V.I. of the species which is a reliable parameter of their relevance in a plant community.

Floristic Diversity on Rubber Plantations and Their Importance. . .

915

Characterization of Rubber Plantation Management We carried out the characterization of the rubber plantations management from July to December 2013 through semi-structured interviews. We asked 32 rubber smallholders about the general characteristics of their rubber plantations (period of implementation, existing trees, trees in production, diversity), their management (management type, cleaning periods, fumigation periods, types of used herbicides), and the use they give to the species.

Results Vegetation and Structure in Rubber Plantations of Uxpanapa We quantified a total of 197 species in 0.5 ha of rubber plantations, having an average species richness of 19.2 ( 14.5). We registered a total of 55 species of adult individuals in the rubber plantations (Table 1), with H. brasiliensis having the highest I.V.I. of 133.0%, due mainly to its basal area, which was equivalent to more than a third of the total I.V.I., followed by Citrus sinensis with 20.4% of total I.V.I. and Cecropia obtusifolia with 14.7% of importance, both in terms of frequency, Cupania glabra with 12.1% and Miconia trinervia with 11.2% of importance, mainly in terms of abundance, and Dendropanax arboreus with 5.9%, Parathesis lenticellata with 5.2%, and Inga punctata with 5.2% of importance in terms of its frequency. Together, all these species comprised more than two thirds (207.8%) of the total I.V.I. (Fig. 4). Most of these species are typical pioneers of secondary vegetation, common in the area due to their rapid ability to regenerate, biomass gain, and tolerance by the owners of the rubber plots. Special mention deserves the record of Casearia tremula, C. commersoniana, Ocotea uxpanapana, Pterocarpus rohrii, and Talauma mexicana in the adult stratum, which together summed up 10% of the total I.V.I. for the rubber plantation. These species are typical of mature rainforest which show their ability to develop within the rubber plantation. When we analyzed the regeneration potential of these sites, we found 188 species (Table 1), with H. brasiliensis having 6.3% of the total I.V.I. followed by Piper hispidum, Cupania glabra, Tabernaemontana alba, and Conostegia xalapensis with 5.1, 5.0, 4.1, and 3.8% of the importance values, respectively, all in terms of abundance (Fig. 5). Also, we recorded species of successive systems such as Stemmadenia donnel-smithii, Vernonia pattens, Psychotria chiapensis, Piper nitidum, and Piper lapathifolium, which represented 57.5% of the total I.V.I. for the area. We recorded 165 species with few individuals and I.V.I < 2.5%, most of Table 1 Metrics (total values) for vegetation in adult and regeneration stages from rubber plantations

Metric Families Genera Species Individuals

Adult 27 26 55 476

Regeneration 52 88 188 983

916

J. C. Lo´pez-Acosta et al.

Fig. 4 Adult species (1 cm DBH) in rubber plantations

them typical of mature rainforest such as Nectandra sp., Poulsenia armata, Terminalia amazonia, Ficus sp., and Pseudomedia oxyfilaria, as well as some species of the shrub stratum as Psychotria sp. All of them with the ability to reach and settle in the rubber plantations, albeit the intensity and type of manual cleaning will determine their survival. The rubber plantations also protect species with some conservation status such as Zamia loddigesii and Talauma mexicana, both species under conservation status according with the Mexican Official Regulation (NOM-059-ECOL-2010) (SEMARNAT 2010) regarding environmental protection for native species of wild flora and fauna considering risk categories and specifications for inclusion, exclusion, or change and a list of species at risk.

Useful Species in Rubber Plantations In the adult stage, we recorded seven useful species. In addition to the rubber tree, there are orange trees (Citrus sinensis), the only exotic species that we recorded in

Floristic Diversity on Rubber Plantations and Their Importance. . .

917

Fig. 5 Regeneration species ( 30 cm height) in rubber plantations

the rubber plantations, which occupies the second place in ecological and economic importance within them. Then, the guarumo (Cecropia obtusifolia) and tepesin (Cupania glabra) followed in ecological importance. Also, there were timber species such as the water stick (Dendropanax arboreus), a medicinal plant whose wood is used to make handicrafts and as firewood, the capulin (Parathesis lenticellata) whose fruit is highly appreciated in the area, and finally Miconia trinervia, a tree that is not known to have any use in the area, but that is often used as firewood in other parts of the country. Regarding regeneration of vegetation in the rubber plantations, it is important to mention the recovering of populations of the first seven useful species after the rubber. In order of ecological importance these species were the comb (Piper hispidum) for medicinal use, followed by tepesin (C. glabra), the deer egg (Tabernaemontana glabra) whose latex is used to treat skin diseases, horse egg (Stemmadenia donnell-smithii) for timber, and finally, Vernonia patens which have no reported use.

918

J. C. Lo´pez-Acosta et al.

General Characteristics of the Rubber Plantations We interviewed a total of 32 rubber producers (30 from the Uxpanapa Valley and two from Las Choapas; Fig. 6). We found three types of rubber plantations, classified according to their period of establishing: (a) plantations from the 1980s (n ¼ 21), promoted by the colonization of the Valley, established more than 31 years ago; (b) plantations from 1995 (n ¼ 4), which coincide with the declaration of the local government, having an establishment period from 16–30 years; and (c) plantations from the year 2000 (n ¼ 7), of relatively recent establishment (0–15 years), encouraged by the increase in the price of rubber. Many of the plots in production were of 2 ha; each having from 201 to 400 trees depending on the survival of the plants (e.g., predation by gopher or cattle) and on the intensity of planting by the producers. The varieties of H. brasiliensis that have been tested in the Valley are RRIM600, IAN 710, IAN 873, IAN 754, RRIM 527, found in 15, 9, 5, 2, and 1 rubber plantations, respectively. The criteria of selection are based on the clonal production potential and susceptibility.

Characteristics of Management Practices We found two main types of management aimed at controlling weeds in rubber plantations: cleaning and fumigation (Table 2). The first consists of manual or mechanical (brush cutter-tractor) removal of weeds and the second involves the use of herbicides applied using sprinkler pumps. These techniques are not mutually exclusive and are done in all rubber plots. The difference lies in the period in which they are performed or in the extent of land covered within the rubber plantation in production (e.g., streets, only around the rubber trees, and lines). Regarding the periodicity of weed control, the interviews showed that this was carried out every 4, 6, and 12 months, with 6 months being the most reported period to work in the maintenance of the rubber plantations (Table 2), which is associated with the winter and rainy season in the area, as well as with the economic possibilities of the owners. Fig. 6 Latex extraction in the Uxpanapa Valley. (Photo by the authors)

Floristic Diversity on Rubber Plantations and Their Importance. . .

919

Table 2 Type and periodicity of management in rubber plantation Type of control Ground cleaning

Fumigation

Period in months 4 6 12 4 6 12

Number of rubber plantations 13 17 2 5 12 7

Fig. 7 Management of rubber plantations (n ¼ 32) in the Uxpanapa Valley

For rubber plantation cleaning, there are several subcategories (a) manual non-selective, in which every plant is cut, including H. brasiliensis seedlings; (b) manual selective, which tries to keep useful species (mainly woody) usually reported as “coming from the mount” (c) mechanical (both brush cutter and tractor) in which it is not possible to discriminate the species and their impact is greater on the understory of the rubber plantation (Fig. 7). These three variants of practices are carried out at three levels within the plantations: (a) streets: spaces of 6 m wide between line and line of trees; (b) lines: the spaces between tree and tree across the plantation, and (c) trees: where the cleaning is made exclusively around 2 m of each rubber tree. The interviews showed that the most common management variant was the selective manual, either at street, line, or tree level (Fig. 8). In the interviews, we detected that the cleaning of the of rubber plantation is highly variable, and that is not always preferred to maintain them clean

920

J. C. Lo´pez-Acosta et al.

Fig. 8 Rubber plantation with total cleaning. (Photo by the authors)

Fig. 9 Rubber plantation with no frequent cleaning. (Photo by the authors)

(Fig. 9), mainly due to the high costs of chemical inputs or the payment of wages for the control of weeds. Therefore, it is preferred to clean only around trees or lines or streets which reduces the work and fulfills the function of facilitating the scratching of the rubber (Fig. 6). On the contrary, there is also the practice of total cleaning, especially directed by producers that have more than 15 ha of land and with some level of technification, using machinery with no selective cut of the species (Fig. 8). Here is where we found an area of opportunity for the management of rubber plantations since, as reservoirs of seedlings from the tropical rainforest, they can be sites for conservation of this germplasm, especially of useful plant species. It is essential to encourage selective cleaning that avoids cutting useful species, both timber and non-timber, as well as to promote the diversification of rubber plantations with typical tropical rainforest species. This can be a strategy to increase the value of this productive system and their potential as biological corridors and shelters of flora and fauna of the region (Gadow et al. 2004; Parroquín-Pérez 2004). Also, for

Floristic Diversity on Rubber Plantations and Their Importance. . .

921

promoting evergreen species which would contrast with the typical deciduous pattern of H. brasiliensis, to protect soil and nutrient cycles (Ningsih et al. 2011). This proposal is viable since many of the producers recognize the importance of having more species within the rubber plantations, not only for their direct use but also for a local belief that “to a greater amount of forest, the fungi (Phytophthora palmivora, Diplodia sp., Brotryodiplodia theobromae, and Colletotrichum gloeosporiodes) that attack the rubber plantations, appears with less intensity.” This observation may be subject of subsequent research and would justify the retention of diversity if it is negatively related to the incidence of pathogens. Punctual observations indicated a high presence of epiphytic flora in the rubber plantations, mainly orchids, ferns, and bromeliads. All these established above the “scratched” area. This may represent another issue of enrichment of the rubber plantations, being places of protection of useful flora (ornamental), thus opening the possibility to constitute a production unit for this type of plants, giving added value to the rubber plantation.

Discussion The need to know and inventorying richness in anthropized systems is essential to recognize their retention capacity of plant diversity (Dirzo et al. 2009; Montagnini 2008). Under this context, we present a series of reflections based on this case study valuing systems that are not commonly considered as reservoirs of diversity or important elements of landscape conservation. As annotated above, rubber plantation are systems based on monoculture with high management inputs (like many other production systems in the tropics). However, our data show high levels of plant diversity occurring in these systems as also reported in other countries. For example, Bumrungsri et al. (2006) reported 52 species of seedlings and shrubs and 49 in the adult stage in rubber plantations of Thailand. Likewise, Ningsih et al. (2011) in Indonesia, reported seven species in regeneration and 35 adults for rubber plantations with 26 years of abandonment. Both studies report that these plantations were intensively managed, and the landscape matrix lacks sources of propagules that allow the recolonization of these sites. Along with findings of Bumrungsri et al. (2006) and Ningsih et al. (2011), our results indicate that rubber plantations have the potential for plant retention in different geographical areas, possibly higher in the Mexican neotropics, as shown in this study. Management practices have an important influence on the regeneration and maintenance of diversity. For example, Beukema et al. (2007) compared the floristic diversity of a mixed rubber (known as “jungle rubber gardens” or “rubber gardens”) against a technified rubber plantation. Showing that plant diversity differs over species richness patterns, which was higher for rubber gardens: with 145 tree species, 38 terrestrial pteridophytes, and 18 epiphytic pteridophytes. However, it is noted that the presence of these species is defined by management practices and by a logging and burning cycle for replanting rubber.

922

J. C. Lo´pez-Acosta et al.

In Mexico, studies in rubber plantations are aimed at developing technical knowledge about (1) timber production processes (Ordóñez-Candelaria and Martínez-Castillo 1997) such as biomass estimates for their management and payment for environmental services (Monroy-Rivera and Návar-Chaidez 2004), (2) technical and economic analysis of production process of the guild (RojoMartínez et al. 2005), (3) quality of the treatment of postharvest and its derived products (Obeso-Granados and Pérez-Méndez 2008), and (4) diagnoses on the vulnerability of farming families under unstable market scenarios (AguilarRomán et al. 2012). In the Uxpanapa Valley, our work is a pioneer in the characterization of vegetation associated with rubber plantations. However, there are some precedent studies on plant diversity in other elements of this agro-landscape, for example, Zamora (2012) focused on the orange crop (active and abandoned), finding (with a similar methodology) a total of 191 woody species, belonging to 60 families and 131 genera. Likewise, they made the differentiation between adult and regenerating plants, reporting 60 species in the adult group and 191 in regeneration; these values are barely higher than those reported by us (55 adults and 188 regeneration). Zamora (2012) argues that much of the species richness is given by zoochoric syndromes, promoted by orange trees (Citrus sinensis), due to the proximity of these crops to patches and remnants of well-preserved vegetation. Another element of the landscape described for the area are the living fences, which are plant component (mainly trees) set in a linear arrangement, used to delineate crop fields, pastures, and farms boundaries (Ruiz-Guerra et al. 2014; Harvey et al. 2005). Ruiz-Guerra et al. (2014) found 36 woody species as adults and 29 were registered as regenerating plants, arguing that the presence of remnants of the original vegetation is the key to the enrichment of seedlings within living fences. The composition and diversity of understory species in plantations are influenced by the rubber tree and management practices (which on a small scale tend to be less intense). Thus, showing the potential to provide buffer zones and stable habitats in benefit of species conservation. The rubber plantations show the ability to maintain connectivity between different elements of the landscape, mainly connecting preserved mature forest fragments with other disturbed fragments. In a study on seeds dispersion by chiropters in rubber plantations of the Uxpanapa area, a total of 6084 seeds were obtained from seven families and ten genera, with Piper auritum being the best represented species. It was found that the richness and diversity of bat species are greater in rubber plantations far from the rainforest remnant than in nearby ones. Moreover, rubber plantations are considered potentially useful ecosystems for natural restoration due to the diversity and quantity of seeds that are dispersed by bats in these environments, both near and far from the forest fragments (Santiago del Valle 2014). All the above mentioned demonstrates the importance of rubber plantations as biological corridors, especially in an area of high diversity such as the Uxpanapa Valley. The rubber plantations consist mainly of high-rise trees (H. brasiliensis), which structurally serve to maintain canopy cover and as a shelter for other species. Therefore, rubber plantations have a great value from an ecological perspective: they

Floristic Diversity on Rubber Plantations and Their Importance. . .

923

maintain biodiversity, influence animal movement patterns, function as biological corridors, and increase species persistence (Chacón et al. 2007; Harvey et al. 2005). Also, the rubber plantations help to increase the productivity and diversification of products in cattle farms, serve as a source of fodder, firewood, wood, and fruits (Harvey 2003; Tobar and Ibrahim 2010) as also reported for plantations of C. sinensis (Dirzo et al. 2009; Holl 2002; López-Acosta et al. 2018). This information demonstrates that rubber plantations can harbor a high diversity of plant species and their mutualists, depending on the management intensity that operates on them. Floristic richness increased with the time of abandonment has been also reported in other plantation systems as Citrus sp. and Chamaedorea elegans palm crop (Lascurain-Rangel et al. 2019; López-Acosta et al. 2018). For the palm crop a low recruitment of seedling and diversity in the understory was reported, which suggests a simplification favored by palm producers, which resulted in the anthropogenic interruption of the successional processes; however, their canopy cover provide better environmental services than those provided by previously existing vegetation (Lascurain-Rangel et al. 2019). This is evidence that the management strategies practiced by local people can have several consequences and that studying their trajectories helps to understand the retention of diversity at the landscape level. Our results show that rubber plantations are sites that promote the regeneration of native species. The commercial plantations have been developed in various parts of the world. In Mexico, the appearance, structure, and composition of rubber plantations, as well as live fences and other crops (C. elegans, Citrus sp.), are configured according to the region and its dynamics within the landscape, as well as the preferences of the producers, the availability of the seeds, the history of land use, production systems, farm size, and pasture management (agricultural mechanization and herbicide use) (Ibrahim et al. 2007; Lascurain-Rangel et al. 2019; López-Acosta et al. 2018). The rubber plantations can diversify their production by using non-timber forest products, which is part of the strategy that contributes to the welfare of populations near forest areas. An alternative could be to introduce the camedor palm (Chamaedorea spp.) into the understory, as several species are commercially important because of their attractive foliage used in flower arrangements. The introduction of this crop in rubber plantations could have a positive effect to prevent the loss of forest cover, which could be promoted by the transformation of rubber plantations to other land use.

Perspectives The rubber plantations of the Uxpanapa Valley can be managed integrally with live fences, orange crops, isolated trees, gallery vegetation, and karst domes with vegetation. Together, they can be elements that help to preserve the biodiversity of this area; the largest remnant of tropical forests in the State of Veracruz (Gómez-Pompa 1979; Wendt 1989).

924

J. C. Lo´pez-Acosta et al.

The recognition of elements of the landscape as the rubber plantations and their interaction with the human component is an indispensable step for the design of restoration strategies for the region. This modifies the classical view of considering the altered areas as unimportant habitats or systems with species of little biological importance, recognizing them as emerging plant associations with high capacity to retain biodiversity and ecological processes necessary to recover the natural capital of the tropics (Bridgewater et al. 2011). In summary, this work reports the ability of rubber plantations to support the arrival and establishment of a high plant diversity, similar to the richness of tropical rainforests and secondary forest, this provided that management practices are applied in a non-harmful manner and managed to maintain fragments of rainforest as sources of diversity and benefits for people. Therefore, these sites can be places of shelter and connectivity, since they occupy around 5,000 km2 within the Uxpanapa Valley, making it a landscape-level management opportunity for the area. The permanence of these systems would depend on proper management, which combines the primary interests of the inhabitants (rubber extraction) with a germplasm diversification and conservation approach. A meeting point can be the propagation in these sites of useful trees of the mid-stratum that support the deciduous conditions of the rubber plantations and provide extra resources to the owners of the plots (López-Acosta et al. 2018). The state of Veracruz has a high level of anthropic intervention, with tropical areas being one focus of attention due to its great heterogeneity and anthropogenic pressure. In addition to the useful species mentioned above, we found native species providing fodder, timber, poles, fruits, fuel, and medicine, such as Carica papaya, Persea americana, Coffea arabica, and Smilax sp. reported by Caballero et al. (1978). Also, we registered some seedlings: Capsicum sp., Smilax sp., Passiflora sp., Tabebuia sp., Gmelina sp., Cedrela sp., and Psidium guajava, among others. Many of these species survive (by selective plating) until they reach the adult stage. The historical and contemporary anthropogenic impact on biological and productive systems has changed the original vegetation, transforming it into a set of landscape elements with an intervention gradient, for the obtention of goods and ecological services. This new configuration of the territory raises questions about the capacity of these systems to maintain plant diversity and how their management can help conserve biodiversity. However, there are few detailed studies on the contribution in terms of maintaining floristic diversity of these systems. Besides, it is important to know the strategies in this type of systems for food, medicine, and income security. Acknowledgments The authors thank Reyna Paula Zárate Morales for managing cites and references, Gina Gallo Cadena for the figures editing, and hevea-producers of the Valley Uxpanapa, especially to Pantoja Aparicio family. We also thank the Regional “Unión Regional de Productores y Cultivadores de Hule Natural del Valle de Uxpanapa” for providing the facilities to obtain the field data and information of their records.

Floristic Diversity on Rubber Plantations and Their Importance. . .

925

References Aguilar-Román J, Galmiche-Tejeda Á, Domínguez M. El contexto de vulnerabilidad de familias productoras de hule (heveacultoras): estudios de caso en el municipio de Huimanguillo, Tabasco. Estud Soc. 2012;20(39):208–58. Beukema H, Danielsen F, Vincent G, Hardiwinoto S, Van Andel J. Plant and bird diversity in rubber agroforests in the lowlands of Sumatra, Indonesia. Agrofor Syst. 2007;70(3):217–42. Bridgewater P, Higgs ES, Hobbs RJ, Jackson ST. Engaging with novel ecosystems. Front Ecol Environ. 2011;9(8):423. Bumrungsri S, Sripao-raya E, Leelatiwong C. A quantitative analysis of plant community structure in an abandoned rubber plantations on Kho-Hong Hill, southern Thailand. Songklanakarin J Sci Technol. 2006;28(3):479–91. Caballero J, Toledo V, Argueta A, Aguirre E, Rojas P, Viccon J. Estudio botánico y ecológico de la Región del Río Uxpanapa, Veracruz, México. Flora útil o el uso tradicional de las plantas. Biótica 1978;3(2):103–44. Chacón LM, Harvey CA, Delgado D. Diversidad arbórea y almacenamiento de carbono en un paisaje fragmentado del bosque húmedo de la zona atlántica de Costa Rica. Recur Nat y Ambient. 2007;51–52:19–32. de Souza AD, de Oliveira RS, Furtado EL, Kageyama, PY, Freitas RGS, Ferraz PA. Árbol del caucho, siringa: Hevea brasiliensis (Willd. ex A. Juss.) Müll. Arg. In: Shanley P, Cymerys M, Serra M, Medina G, editor. Frutales y plantas útiles en la vida amazónica. Brasil: FAO, CIFOR, PPI; 2012. p. 39–46. Dirzo R, Aguirre A, López JC. Diversidad florística de las selvas húmedas en paisajes antropizados. Investig Ambient. 2009;1:19–22. Ellen RF, Fukui K. Redefining nature: ecology, culture, and domestication. Oxford: Berg Publishers; 1996. Espinosa-Organista D, Ocegueda-Cruz S, Aguilar-Zúñiga C, Flores-Villela Ó, Llorente-B J. El conocimiento biogeográfico de las especies y su regionalización natural. In: Sarukhán J, editor. Capital natural de México. Veracruz: CONABIO; 2008. p. 33–65. Ewell P, Poleman T. Uxpanapa: the new land. In: Ewell P, Poleman T, editors. Uxpanapa resettlement and agricultural development in the Mexican tropics. New York: AE ResearchDepartment of Agricultural Economics, Cornell University Agricultural Experiment Station; 1982. p. 93–110. von Gadow K, Sánchez-Orois S, Aguirre-Calderón OA. Manejo forestal con bases científicas. Madera y Bosques. 2004;10(2):3–16. Gascon C, da Fonseca GAB, Sechrest W, Billmark KA, Sanderson J. Biodiversity conservation in deforested and fragmented tropical landscapes: an overview. In: Schroth G, da Fonseca GAB, Harvey CA, Gascon C, Vasconcelos HL, Izac A-MN, editors. Agroforestry and biodiversity conservation in tropical landscapes. Washington, DC: Island Press; 2004. p. 15–33. Gentry AH. Patterns of neotropical plant species diversity. In: Hecht MK, Wallace B, Prance GT, editors. Evolutionary biology, vol. 15. Boston: Springer; 1982. p. 1–84. https://doi.org/10.1007/ 978-1-4615-6968-8_1. Gentry AH. Tree species richness of upper Amazonian forests. Proc Natl Acad Sci. 1988;85(1): 156–9. Gómez-Pompa A. Antecedentes de las investigaciones botánico-ecológicas en la región de Río Uxpanapa, Veracruz, México. Biótica 1979;4(3):127–33. Guevara SS, Laborde DJ, Sanchez Rios G. La fragmentación. In: Los Tuxtlas, El paisaje de la sierra. Veracruz, México: Instituto de Ecología, A.C., Unión Europea; 2006. p. 111–134. Halffter G, Llorente-Bousquets J, Morrone J. La perspectiva biogeográfica histórica. In: Sarukhán J, editor. Capital natural de México I. México, D.F.: Comisión Nacional para el Conocimiento y Uso de la Biodiversidad; 2008. p. 67–86. Harvey CA, Villanueva C, Villacís J, Chacón M, Muñoz D, López M, Ibrahim M, Gómez R, Taylor R, Martinez J, Navas A, Saenz J, Sánchez D, Medina A, Vilchez S, Hernández B,

926

J. C. Lo´pez-Acosta et al.

Perez A, Ruiz F, López F, Lang I, Sinclair FL. Contribution of live fences to the ecological integrity of agricultural landscapes in Central America. Agric Ecosyst Environ. 2005;111: 200–30. Harvey CA. La conservación de la biodiversidad en sistemas silvopastoriles. In: Ibrahim M, MoraDelgado J, Rosales M, editors. Potencialidades de los sistemas silvopastoriles para la generación de servicios ambientales. Turrialba: CATIE; 2003. p. 23–32. Hernández-Gómez I, Ellis-Edward A, Gallo-Gómez C. Deforestación y deterioro de las selvas tropicales en la región del Uxpanapa, Veracruz. México: Universidad Veracruzana; 2011. Hernández-Gómez IU. Evaluación de la transformación del paisaje para la determinación de áreas prioritarias de conservación en la región de Uxpanapa, Veracruz. México: Universidad Veracruzana; 2014. Holl KD. Effect of shrubs on tree seedling establishment in an abandoned tropical pasture. J Ecol. 2002;90(1):179–87. Ibrahim M, Villanueva C, Casasola F. Sistemas silvopastoriles como una herramienta para el mejoramiento de la productividad y rehabilitación ecológica de paisajes ganaderos en Centro América. Prod Anim. 2007;15(1):73–87. INEGI. Prontuario de información geográfica municipal de los Estados Unidos Mexicanos: Uxpanapa, Veracruz de Ignacio de la Llave. Aguascalientes: Instituto Nacional de Estadística y Geografía; 2010. Lamprecht H. Silvicultura en los trópicos: los ecosistemas forestales en los bosques tropicales y sus especies arbóreas; posibilidades y métodos para un aprovechamiento sostenido. Eschborn: FAO; 1990. Lascurain-Rangel M, Rodríguez-Rivas G, Gómez-Díaz JA, Alvarez-Palacios JL, Benitez-Badillo G, López-Binnquist C, Dávalos-Sotelo R, López-Acosta JC. Long-term enrichment with the camedor palm (Chamaedorea elegans Mart.) improved forest cover in an anthropogenic tropical landscape. For Ecol. Manage. 2019;450:117499. https://doi.org/10.1016/j.foreco.2019.117499. Levis C, Flores BM, Moreira PA, Luize BG, Alves RP, Franco-Moraes J, Lins J, Konings E, PeñaClaros M, Bongers F, Costa FRC, Clement CR. How people domesticated Amazonian forests. Front Ecol Evol. 2018;5:171. López JC, Dirzo R. Floristic diversity of sabal palmetto woodland: an endemic and endangered vegetation type from Mexico. Biodivers Conserv. 2007;16(3):807–25. López-Acosta JC, Lira-Noriega A, Cruz I, Dirzo R. La proliferación de luces nocturnas: Un indicador de actividad antrópica en México. Biodiversitas. 2011;94:1–5. López-Acosta JC, Zamora-Pedraza G, Mendoza-Ramírez E, MacSwiney MC. Recuperación de la diversidad florística en plantaciones de cítricos abandonadas en el sur de Veracruz, México: implicaciones para la conservación. Rev Mex Biodivers. 2018;89(4):1212–33. Monroy-Rivera C, Návar-Chaidez JDJ. Ecuaciones de aditividad para estimar componentes de biomasa de Hevea brasiliensis Muell. Arg., en Veracruz, México. Madera y Bosques. 2004;10(2):29–43. Montagnini F. Management for sustainability and restoration of degraded pastures in the neotropics. In: Myster RW, editor. Post-agricultural succession in the Neotropics. New York: Springer; 2008. p. 265–95. https://doi.org/10.1007/978-0-387-33642-8_13. Ningsih H, Rahayu S, Tata HL. Comparison of floristic composition and diversity in rubber plantations and their surroundings. In: Tata HL, editor. Recognising biodiversity in rubber plantations. Indonesia: World Agroforestry Centre (ICRAF); 2011. p. 265–96. Obeso-Granados V, Pérez-Méndez G. Desarrollar una propuesta para la mejora de la calidad en el tratamiento de postcosecha del hule hevea (Hevea brasiliensis) y sus productos derivados en el estado de Tabasco. Villahermosa: Instituto Tecnológico de Villahermosa; 2008. Ordóñez Candelaria VR, Martínez Castillo JL. Maquinado y susceptibilidad a la impregnación de la madera del hule (Hevea brasiliensis). Madera y Bosques. 1997;3(2):63–71. Ortiz-Hernández E. Paquete Tecnológico del Hule (Hevea brasiliensis Muell Arg.): Establecimiento y mantenimiento preoperativo. Tezonapa: SAGARPA, INIFAP; 2011. 22 p.

Floristic Diversity on Rubber Plantations and Their Importance. . .

927

Parroquín PJ. Densidad poblacional del Tepezcuintle Agouti paca en Uxpanapa, Veracruz. México: Instituto de Ecología, A.C; 2004. PRONATURA. Evaluación del estado de conservación de los ecosistemas forestales de la región denominada Uxpanapa. Xalapa, Veracruz, México: SEDARPA, Gobierno del Estado de Veracruz, PRONATURA A.C.; 2007. 237 p. Revel-Mouroz J. Aprovechamiento y colonización del trópico húmedo mexicano. México, D.F.: Fondo de la Cultura Económica; 1972. Rojo-Martínez G, Martínez Ruiz R, Jasso MJ. El cultivo del hule en México. Montecillo, México: Universidad Autónoma Indígena de México. Colegio de Posgraduados. 2011; 334 p Rojo-Martínez GE, Jasso-Mata J, Vargas-Hernández JJ, Palma-López DJ. Análisis de la problemática de caracter técnico-económico del proceso productivo del hule en México. Ra Ximhai. 2005;1(1):81–110. Ruiz-Guerra B, Velázquez-Rosas N, López-Acosta JC. Plant diversity in live fences and pastures, two examples from the Mexican humid tropics. Environ Manag. 2014;54(3):656–67. https://doi. org/10.1007/s00267-014-0307-x. Sánchez-Colón S, Flores-Martínez A, Cruz-Leyva IA, Velázquez A. Estado y transformación de los ecosistemas terrestres por causas humanas. In: Sarukhán J, editor. Capital natural de México, vol II: Estado de conservación y tendencias de cambio. México, D.F.: CONABIO; 2009. p. 75–129. Santiago del Valle MT. Dispersión de semillas por quirópteros en hulares de la zona de Uxpanapa, Veracruz, México. México: Universidad Veracruzana; 2014. SEMARNAT. Norma Oficial Mexicana NOM-059-SEMARNAT-2010, Protección ambientalEspecies nativas de México de flora y fauna silvestres-Categorías de riesgo y especificaciones para su inclusión, exclusión o cambio-Lista de especies en riesgo. México: SEMARNAT; 2010. Tobar L, Ibrahim M. ¿Las cercas vivas ayudan a la conservación de la diversidad de mariposas en paisajes agropecuarios? Rev Biol Trop. 2010;58(1):447–63. Velasco-Toro J, Vargas-Montero G. El Uxpanapa: Aproximación histórica a una experiencia de colonización agrícola con población Chinanteca relocalizada. In: Anuario VII. Instituto de Investigaciones Humanísticas; 1990. p. 229–53. Velasco-Toro J. Uxpanapa: reflexión historiográfica y proceso de cambio social. OPUS Soc Arte y Cult. 2011;1(1):5–15. Wendt T, van der Werff H. A new species of Ocotea (Lauraceae) from southeastern Mexico. Ann Missouri Bot Gard. 1987;74(2):413–5. Wendt T. Las selvas de Uxpanapa, Veracruz-Oaxaca, México: evidencia de refugios florísticos cenozoicos. An del Inst Biol UNAM Ser Botánica 1989;(58):29–54. Wiersum F. Domestication of trees or forests: development pathways for fruit tree production in South-East Asia. In: Akinnifesi FK, Leakey RRB, Ajayi OC, Sileshi G, Tchoundjew Z, Matakala P, Kwesiga FR, editors. Indigenous fruit trees in the tropics: domestication, utilization and commercialization. Wallingford, UK: CABI, World Agroforestry Centre (ICRAF); 2008. p. 70–83. Zamora PG. Caracterización de la vegetación secundaria a través de una cronosecuencia en plantaciones de cítricos abandonados en la región de Uxpanapa, Veracruz. México: Universidad Veracruzana; 2012.

Ethnobotany of the Highlands of Chiapas Felipe Ruan-Soto, Fausto Bolom-Ton, Ere´ndira J. Cano-Contreras, Lucia Domínguez-Torres, Fernando Guerrero-Martínez, and Ramo´n Mariaca-Me´ndez

Abstract

The state of Chiapas is one of the most biologically diverse and bioculturally rich areas of Mexico. The region known as the Highlands of Chiapas is a mountainous region that harbors vegetation types dominated by pine and oak species and human populations belonging to the Tsotsil, Tseltal, and Tojol-ab’al Indigenous groups. These people have named species, classifying them into groups reflecting their place in the universe according to their worldview and have generated knowledge about their biology, ecology, and the ways in which they may be used to satisfy both material and nonmaterial needs. People of the Highlands of Chiapas have produced knowledge that reflects the cultural significance of diverse species of plants. In general, there are scarce records about use, management, and perception of plants of the region from the Colonial period; however, anthropological and ethnobotanical researches were very active during the twentieth century to the present. It has been through ethnographies, dictionaries, vocabularies, and technical reports of reseach projects that ethnobotanical information has been documented. The forest is a fundamental source of resources for F. Ruan-Soto (*) Instituto de Ciencias Biológicas, Universidad de Ciencias y Artes de Chiapas, Tuxtla Gutiérrez, Chiapas, Mexico e-mail: [email protected] F. Bolom-Ton · F. Guerrero-Martínez Centro de Investigaciones Multidisciplinarias sobre Chiapas y la Frontera Sur, Universidad Nacional Autónoma de México, San Cristóbal de Las Casas, Mexico e-mail: [email protected]; [email protected] E. J. Cano-Contreras · L. Domínguez-Torres San Cristóbal de Las Casas, Mexico R. Mariaca-Méndez El Colegio de la Frontera Sur, San Cristóbal de Las Casas, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_17

929

930

F. Ruan-Soto et al.

the survival of Indigenous groups; rural people practice a multiple use of elements from forests, firmly based on empirical knowledge about them and the natural processes that mold them, and frequently using traditional management techniques. Furthermore, a great diversity of vascular and nonvascular, wild and domesticated, and native and exotic plants is employed to satisfy basic food, fuel, medicine, ornamental, and other needs. In this region, the use and knowledge about a great variety of botanical species used to treat diverse ailments is outstanding. A brief description of the role of plant species within the worldviews of some Mayan groups in this region is discussed.

Introduction The state of Chiapas is located in southeastern Mexico; it shares a border with Guatemala and is one of the most biologically diverse regions of the world. In Mexico, it is the second most diverse state of flora and fauna, harboring 8,248 species of vascular plants (Farrera et al. 2013). It includes ten hydrological basins in which flows over 30% of the freshwater of the Mexican territory. Because of its geographical position and its heterogeneous topography, Chiapas has different climate types, which generate diverse conditions, habitats, and specific ecosystems and plant associations. Along with this biological richness, Chiapas is a pluricultural state: In total, 11 Mayan languages and Zoque are spoken. There are also countless dialectal varieties, traditions, and ways of understanding the world (Martínez and Villasana 2018). The close relationship between Mayan and Zoque groups with their environment has caused, throughout history, the generation of strategies for the appropriate use and management of the available resources by these human groups, which has allowed their physical survival and cultural reproduction. The particular ways in which these groups relate to nature have additionally generated cultural expressions and cosmogonies that reflect the very biological diversity that continues to be a part of their day-to-day life (García 2018). The sum of biological, cultural, knowledge, and management richness is part of the biocultural heritage of Chiapas. The heterogeneity of its environment is reflected in the ten physiographic subprovinces the state is divided (Villalobos 2013). One of which is known as Los Altos de Chiapas (The Highlands of Chiapas). This region is a mountain area containing karstic valleys and spanning elevations from 800 to 2,800 m; the highest altitudes in the region are recorded for the Tsontehuits and Huitepec volcanoes (2,618 and 2,761 m, respectively) (Ochoa-Gaona and González-Espinosa 2000). Soils are thin and rocky, forming steep slopes. Climate is temperate and subhumid, with annual mean temperature of 18  C and mean annual rainfall of 1,700 mm, with the rainy season spanning from May to October (Enríquez et al. 2006; Nepomuceno and Ishiki 2010). There is no relevant superficial hydrological network (Villalobos 2013). The dominant vegetation is pine-oak forests and, less abundantly, pine forests, oak forests, cloud forests, and lowland deciduous forest in the lower regions (Quintana-Ascencio and González-Espinosa 1993). During the last few decades,

Ethnobotany of the Highlands of Chiapas

931

Fig. 1 Map of the Highlands of Chiapas, México

the natural vegetation areas have been deforested to become, cultivated fields, and spaces of plantations dominated solely by pine species (Cayuela et al. 2006). The Highlands of Chiapas is not only a physiographic area, it is considered a geographic, economic, and socio-cultural region integrated by 17 municipalities spanning 3,770 km2 (INAFED 2017) located in the central part of the state, between 16 300 and 17 000 North and 92 and 93 West (Robles-Porras et al. 2007) (Fig. 1). The region contains a population of 645,099, which is 57% Indigenous (INEGI 2015). In it, the main traditional activities are agriculture (especially milpa and small crops of commercially relevant plants), extensive sheep farming, and forest exploitation (wood, firewood, coal, and diverse wild nontimber forest products); more recently, secondary and tertiary activities such as mining, commerce, and tourism have gained relevance. In addition, in the last decades migration of many people to urban centers looking for employment in various activities has increased (Enríquez et al. 2006; Quintana-Ascencio and González-Espinosa 1993). The main Indigenous groups inhabiting this region and the neighboring areas are Tsotsil, Tseltal, and Tojol-ab’al. These groups are part of the western branch of the Mayan linguistic family (Kaufman 2017); Tsotsil and Tseltal are closely related languages, both within the Tseltalan group (Kaufman 2017). The three languages comprise about one million of speakers (INEGI 2015). Some of the most numerous towns where Tsotsil is spoken in the Highlands of Chiapas include Chamula, Chenalhó, Huixtán, Zinacantán, Pantelhó, Mitontic, and San Cristóbal de Las Casas. For Tseltal, the most populated towns in the Highlands are Amatenango del Valle, Aguacatenango, Tenejapa, Oxchuc, Cancuc, and Chanal. Archaeological information suggests the presence of the ancestors of these Mayan groups in the Highlands of Chiapas for at least 9,000 years (Mariaca et al. 2007). Throughout this long history, a broad and deep knowledge about plants of this

932

F. Ruan-Soto et al.

mountain region has been built up. These Indigenous groups have named particular species, ordered them into classification systems that reflect their place in the universe, and generated knowledge about their biology, their ecology, and how to manage them to satisfy material and nonmaterial needs- That is, they have produced knowledge reflecting the cultural significance of a high diversity of plant species. This chapter shows a general review of the ethnobotany of the region. It is an attempt to provide insight about how the Indigenous groups from the Highlands of Chiapas relate to plants, particularly regarding forest management, agrobiodiversity, and the role of plants in ethnomedicine, as well as the place of plants in the worldview of these Mayan peoples.

Background of Ethnobotanical Studies in the Highlands of Chiapas There are scarce records about the use, management, and perception of plants in the Highlands of Chiapas during the Colonial period. However, there is a document attributed to Fray Domingo de Ara, written during the second half of the sixteenth century just a few decades after the Spanish conquest entitled Bocabulario de lengua tzeldal según el orden de Copanabastla (roughly: “Vocabulary of the Tzeldal [Tseltal] language according to the Copanabastla order”). It includes over 70 Tseltal names of plants, as well as annotations about use of several of them for food, agriculture, construction, house items, and clothing, among the most relevant (Ruz 1985). In the area of agriculture, the terms relating to maize (Zea mays) are of special interest; these include its life cycle and matters relating to the sowing and harvesting of the milpa, while an outstanding topic is the production of cotton (Gossypium hirsutum) and its use in the relevant production of textiles in the region (Ruz 1985). For most of the twentieth century, despite the Highlands of Chiapas was one of the most explored areas of Mesoamerica from the perspectives of social sciences and the humanities, there was little research specifically dubbed “ethnobotany.” However, ethnographies, dictionaries, vocabularies, technical reports, and other documents include a great amount of ethnobotanical information that sheds light on how the Tsotsil and Tseltal related to, used, and managed the flora of the region. Two of the main projects that collected a great quantity of ethnobotanical data from the region were the Harvard Chiapas Project and the University of Chicago’s Man in Nature project, both of which were carried out in mid-twentieth century. The Harvard Chiapas Project was developed between 1957 and 1967 in the Zinacantan municipality with support from the Instituto Nacional Indigenista (National Indigenist Institute), and it became a field laboratory for the anthropology students of the Harvard University (Marzal 2016). Among the results obtained in this project, which provided ethnobiological information about the Zinacantan people and their plants, we can mention the work by Evon Z. Vogt (1966), in which aspects related to flower cultivation and medicinal plants were recorded.

Ethnobotany of the Highlands of Chiapas

933

The Man in Nature project carried out by the University of Chicago was conducted over two phases between 1956 and 1962. Its original objective was to investigate the relationship between Tseltal and Tsotsil communities and their environment from an interdisciplinary perspective. Consequently, information was gathered regarding habitat, vegetation, archaeology, linguistics, and ethno-history of the Highlands of Chiapas. In addition, valuable complete ethnographies from some of the Tseltal and Tsotsil villages, such as Amatenango del Valle, Aguacatenango, Chanal, Oxchuc, and Huixtan were produced (Barrera-Aguilar 2018). Some of the ethnobotanical data gathered in this Project included the following: (a) the studies by Lawrece Kaplan about the mountain and lowland milpas, in which he proposes differential productivity between altitudinal ecosystems; (b) the observations by John Hotchkiss about the three physiographic lands (grazing land, seasondependent, and plough land); and (c) the data gathered by Norman McQuown in a linguistic study about the distance between tseltal and tsotsil concerning specialized vocabulary about fauna, flora, and geography (Barrera-Aguilar 2018). From this research, it can be appreciated that linguistic documentation studies are regularly a source of ethnobotanical data and vice versa; this is due to the fact that plants have a great cultural significance in the day-to-day life of rural people. Regarding this fact, several studies conducted in the Highlands of Chiapas after the first half of the twentieth century must be considered important background for ethnobotany. For instance, the Tsotsil Dictionary, generated by the work by the Instituto Lingüístico de Verano (Summer Linguistics Institute), contains numerous terms used to name plants, specific vegetation associations, and ethnobotanical practices (Weathers and Weathers 1949). Furthermore, in the text “Sk’op Sotz’leb: El Tzotzil de San Lorenzo Zinacantan” (The Tsotsil of San Lorenzo Zinacantan) by John Haviland (1981), there is a vast information regarding the names of culturally significant plants that can be recovered from studies of linguistic documentation from the 1970s in the twentieth century. Lastly, the text “El Gran Diccionario Tzotzil de San Lorenzo Zinacantan” (The Great Tsotsil Dictionary from San Lorenzo Zinacantan) by Robert M. Laughlin (1975) is worthy of mention. In it, the Tsotsil names of vines, lianas, trees, and plants are recorded and supported by around 10,000 botanical specimens identified by Dennis E. Breedlove. Stemming from this work, the two researchers published a second book: “The Flowering of Man: A Tzotzil Botany of Zinacantan,” which delves into the botanical knowledge of zinacantecos (Zinacantan people) (Breedlove and Laughlin 1993). This work gathers fieldwork data from the 1960s in the twentieth century, and it collects 2,686 Tsotsil names for plants, 1,484 species that correspond to these names, as well as important information on their uses, management practices, and geographical, historical, cultural, and linguistic data related to flora. From the biological sciences, there are also relevant background studies to the ethnobotanical study of the Highlands of Chiapas. Faustino Miranda, in his work “La vegetación de Chiapas” (The vegetation of Chiapas) (Miranda 1952) expresses a constant admiration for the ancestral knowledge of the people from different Indigenous groups who were with him during his fieldwork, and he also stresses the importance of linguistics to study botanical aspects. In the Miranda’s work, the

934

F. Ruan-Soto et al.

Maya names for plants were recorded, as are the general lines of their classification systems, and information about their uses. An example is the Tseltal name of pine: taj, from which derives ichtaj (Pinus oocarpa), mocochtaj (Pinus pseudostrobus), and cantaj (Pinus tenuifolia). An important work that was driven by foreign researchers, like Berlin and Breedlove, along with some Mexican colleagues, was “La herbolaria médica Tzeltal-Tzotzil en los Altos de Chiapas” (the Tzeltal-Tzotzil herb medicine from the Highlands of Chiapas). It was published in 1990 as part of the Collaboration Programme “Programa de colaboración sobre medicina indígena tradicional y herbolaria” (PROCOMITH, A.C.). In this work, the ethnobotanical features of the 50 most frequent plants in Indigenous medicine are described. Lastly, one of the most relevant background studies for ethnobotany and ethnoecology, because of its consequences, were the works by Brent Berlin, in particular the work “Principles of Tzeltal plant classification” (Berlin et al. 1974). In it, the plant knowledge of Tseltal people from Tenejapa is studied. This research became a referent for ethnoscience because of the linguistic and botanical detail obtained in the collection of data: Each Tseltal name is backed up by an herbarium specimen. This study was decisive in the theoretical construction of the ways in which naming and classification of ethnobotanical systems work throughout the world based on general universal principles. While a full chapter could be dedicated to this topic, we can simply say that the ethnobotanical classification system employed by the Tseltal peoples of the Highlands of Chiapas is based on psychological principles of perceptual category-building on which the systematic principles of scientific botany lie. This contributed decisively to the theoretical formulation of the systematic principles that, Berlin proposes, rule the systems of traditional knowledge about living beings the world over (Berlin 2010).

Forest Management The mountain forests of the Highlands of Chiapas contain particular floristic and structural configurations. Phytogeographically, these forests are the result of the clash between the floristic compositions from the north and the south of the American continent (the Neartic and Neotropical regions, respectively) following the formation of the Central-America bridge about four million years ago. Thus, we can say that they are the encounter of two worlds decanted by physical and environmental factors such as climate, altitude, soil, and sun exposure, among others. Additionally, the vegetation from the Highlands has been known, appropriated, and used by human beings since they arrived into this land, which in turn has also influenced and modified its floristic and structural configuration. The most exhaustive research regarding floristic and ecological aspects of the mountain forests of Chiapas is possibly that carried out in the Highlands of Chiapas. The forests in this region exist in a landscape that looks like a mosaic of fragments of vegetation within a matrix of cultivated fields, grasslands, and settlements (González-Espinosa et al. 1997) (Fig. 2). The vegetation in general shows an

Ethnobotany of the Highlands of Chiapas

935

Fig. 2 Mosaic of fragments of vegetation, cultivated fields, grasslands, and settlements in Chiapas Highlands. (Photo by Felipe Ruan-Soto)

ample superposition of both its distribution and successional stage, which makes it difficult to limit in time and space. The relatively conserved communities are restricted to mountainous, cliffs or canyons, and the scarce protected areas that are managed by Indigenous communities, municipalities, or military authorities. Miranda (1952) and Breedlove (1973, 1981, 1986) carried out studies about the composition and structure of different forest associations in The Highlands, but, more recently, González-Espinosa et al. (1997) proposed the division of local forests in four general types: pine forest, oak forest, pine-oak forest, and cloud forest. The structural components of pine forests are species of conifers locally called ocote blanco and ocote rojo or, in Tsotsil, sakil toj and tsajal toj, respectively. These include species such as Pinus tecunumanii, P. montezumae, and P. pseudostrobus. This is the least diverse vegetation type in the region; it has the simplest composition and structure with practically nonexistent lower strata of trees and an abundance of grasses due to the intense anthropogenic activities carried out in past and present times. These are commonly located in areas that were cut down for agriculture or goat and sheep breeding, subjected to intense and periodic fires or to frequent tourist activities. It can be frequently found in the least humid places in the region, in sites like Rancho Nuevo and Chilil. Regarding oak forests (oaks are named tulan in Tsotsil), their structural base is formed by species like Quercus rugosa, Q. crassifolia, and Q. segoviensis, although the lower strata hold a considerable diversity and density of trees and shrubs and an abundance of epiphytes. It can be found in extensive areas in the northern slopes of the Huitepec mountain and in the Moxviquil private reserve, but there are a few smaller patches of it in the region as a product of the selective extraction of pines in pine-oak forests. Pine-oak forests are the most extended type of forest in the central highlands. They have a structural base of several pine and oak species (including those mentioned above and P. ayacahuite), although they also contain a diverse

936

F. Ruan-Soto et al.

underbrush and herbaceous stratum. Vascular epiphyte plants are common in the densest forests. Patches differing in size, with different degrees of preservation and connectivity, can be found throughout the region. A continuous forest that stands out because of its preservation state can be found in the area assigned to the 31st Military Zone. Cloud forests can be found near cold and humid summits (mostly above 2,400 m), in relatively small and isolated patches. The trees in these forests support many bromeliads, ferns, mosses, vines, and lianas. It has a great flora diversity, much higher than that of the other types of forest, and the canopy is frequently dominated by wide-leaf species, often Quercus laurina and Q. ocoteifolia, but the lower arboreal stratum and the herbaceous and shrub strata include species like Clethra macrophylla, Persea americana, Oreopanax xalapensis, and Magnolia spp., to mention a few. This is a type of forest that is restricted to the highest areas in the Huitepec, Tzontehuitz, and el Extranjero mountains, which have elevations of 2,600, 2,700 and 2,740 m, respectively. However, it can also be found in slightly lower areas, as is the case of the localities Bazom and Jokosik in the municipality of Huixtan, which are subject to a high exposure to humidity. The forests (known as monte or, in Tsotsil, te’etick-a’maltik) are essential sources of resources for the survival of the Indigenous groups of The Highlands. Highlander peoples use them in multiple ways with a solid base of empirical knowledge about the elements they contain and the natural processes that mold them, frequently using traditional means and techniques for their extraction (Bolom-Ton 2005) (Fig. 3). The yield is mainly destined to self-consumption, that is, to meet the needs for timber and house materials, as well as complementing people’s diets through collection (Alemán-Santillán 1985, 1989, 1997, De Jong and Ruíz 1997; González-Espinosa et al. 1997; Parra et al. 1993; Soto-Pinto et al. 1997). As it has been stated, these forests are far from being pristine, and their current aspect can be explained by the history and features of the productive activities of the neighboring human population and by the natural, social, and cultural conditions in which production takes place. In addition, the economic networks in which they exist (Alemán-Santillán 1985, 1989). The region, with the exception of San Cristobal de Las Casas which is the economic capital of the area, is eminently rural, with a population subjected to high indexes of marginalization and poverty, low levels of education, and, in some areas, dedicated mainly to primary activities (agricultural, livestock breeding, and forest sectors), especially the agriculture of maize and beans (INEGI 1994, 2001). In this framework, the activities related to the forests of the Highlands are carried out in an artisanal way, using manual tools, basically integrated in a greater or lesser degree to a system of self-subsistence of the highlander farmers. This subsistence pattern is characterized by an extensive agriculture of maize and beans, firewood extraction (frequently selective), wood extraction (often monospecific), coal production, livestock breeding activities within forests, collecting, and hunting, among others. Nonetheless, there are small private initiatives in sawmills and plywood factories, furniture or boxes, as well as small workshops and woodshops. In general, commercial forest activity is scarce and limited by the rugged topography of the region, as well as by the generalized basic infrastructure.

Ethnobotany of the Highlands of Chiapas

937

Fig. 3 Inhabitants of the highlands of Chiapas collecting plants in their environment. (Photo by Marisa Ordaz Velázquez)

Bolom-Ton (2005) distinguishes human activities like agriculture, that, while directly independent from a standing forest, may need a previous forest to obtain valuable nutrients for growing crops. Such is the case of milpa after the cut-slashburn, use, and fallow cycles. Similarly, extensive bovine, sheep, and goat stockbreeding depends on vegetation mass for grazing. Another group of activities can be grouped as extraction, which depends on the existence of the forest; examples of this is the obtention of wood, timber, and resources used as food, decoration, for ceremonial purposes, and medicine. Underlying the ways of relating to the forests, there are uses and practices that can be defined as traditional insofar as they have been carried out through time by the Indigenous population. We might say the forest is the natural environment and context for the highlander Mayan peoples since immemorial time. The forest is a space that circumscribes and delimits the territories of these peoples, while it also offers a natural context where their myths, expectations, and fears have a setting. The highlander peasants not only use forest, but they also distinguish it from other landscape units. Forests are classified by their density, by their successional state, or by the dominance of certain important species (Sánchez-Álvarez 2004). The parts of the tree can be distinguished, and the phenological stage of the plants, some life forms, and plant groups in the forest are differentiated (Bolom-Ton 2005) (Table 1).

938

F. Ruan-Soto et al.

Table 1 Examples of words in Tsotsil (Huixtan dialect from the Huixtan municipality, Chiapas) to name botanical categories and the parts of a tree. (Based on Bolom-Ton 2005) Botanical category

Tree part

Tsotsil ech’ te’ bik’tal te’ tsim chuch vomol tson te’ ak’ uch’al oh’ sni’ te’ te’lal yok te’ yi’bel spat te’ sk’ob te’ nichim sat yanal

Español bromelia árbol arbusto helecho hongo hierba musgo bejuco orquídea copa tallo contrafuerte raíz corteza rama flor fruto hoja

English Bromeliad Tree Shrub Fern Mushroom Herb Moss Liana Orchid Top Stem Buttress Root Bark Branch Flower Fruit Leaf

As mentioned before, several authors have compiled a vast amount of names for plant species, which reflect the enormous knowledge of plants by local people. The plant and forest elements are also known in terms of features such as shape, size, color, position, and consistency, or their use or appreciative meaning. Botanical classification, in this sense, is carried out based on daily life, with a high degree of specialization and specification. This local knowledge demonstrates centuries of interaction with the environment and a masterful familiarity with many natural processes, although it must be said that there are also many undeniable inclusions and adaptations to its body of knowledge since its encounter with the European world 500 years ago. Forests are a structural base for the Highland Maya, and their preservation implies the preservation of their culture. Because of this, modern techniques of exploitation and conservation must be in accordance with the social, productive, and cultural reality of the population, and they must be used in awareness of the multiple utility of the forest and its primordial importance for subsistence.

Useful Plants in the Highlands A great diversity of vascular and nonvascular plants, domesticated and wild, native and introduced, are used by the Indigenous groups of the Highlands of Chiapas. They come from the different elements that make up the heterogenous landscape: forests, shrubs, milpas, and orchards.

Ethnobotany of the Highlands of Chiapas

939

The Highland Maya make integral use of their environment, and the plant diversity that surrounds them provides them the elements they need to meet their needs of food, fuel, medicine, and ornament, to mention a few. The following is a brief account of the plants used, grouped by main categories of use. Fuel. The use of fuel from the forests in private homes has hardly changed over time, and firewood prevails as the basic fuel for cooking. In the Highlands, the most used and appreciated groups of plants for firewood are oaks (tulan, e.g., Quercus crassifolia, Q. crispipilis) and pines (toj, e.g., Pinus montezumae var. montezumae, P. pseudostrobus). Firewood from oaks is generally preferred because it takes longer to consume and it produces less smoke; however, pine wood is more frequently used because of its greater current availability and because pine trees are easier to cut, tear down, and transport (Bolom-Ton 2005). In general, a greater effort in the search of firewood is currently made due to the scarcity of tall trees with large diameters. However, there are other forest species that can be used for firewood alternatively. For example, in the Bazom area in the municipality of Huixtán, Martínez-Icó et al. (2015) report up to 94 species of woody plants that can be used as fuel, including Arbutus xalapensis, Clevera theaeoides, Prunus serotina, and Symplocos limoncillo (Martínez-Icó et al. 2015). Although larger trees are preferred, firewood is also obtained from shrubs and plants in the intermediate strata (Bolom-Ton 2005). Martínez-Ico et al. (2015) report some species from this stratum such as: Viburnum elatum, Holodiscus argenteus, and Zanthoxylum melanostictum. Furthermore, firewood is not exclusively obtained from mature forests, but also from secondary vegetation such as Baccharis vaccinoides and Verbesina perymenoides. The choice of trees for firewood depends on criteria such as species, height, width of the tree, ease to cut, and the means available for transport. The day-to-day management of firewood in domestic units is conducted by the housewives, who use it mainly for heating and cooking food, as well as heating and lighting the home. Although herbs are not used as fuel, the leaf litter is particularly dry needles from pines (xactoj or juncia). Corncobs are also used, as well as ocote (sappy wood fragments used for timber, from various species, mainly Pinus tecunumanii) to start fires. Charcoal is made from oaks, and in the Highlands of Chiapas, it is produced mainly in Zacualpa, El Pinar, Corazón de María, El Corralito, and in communities from the municipality of Teopisca to supply to the San Cristóbal de Las Casas market. The species used are the following: Quercus laurina, Q. rugosa, Q. crassifolia, Q. crispipilis, and, to a lesser degree, Q. candicans (Ramos-Martínez et al. 1999). This kind of fuel is manufactured to be sold in urban areas and other regions of the state, and it is rarely used by the manufacturing communities. Other Indigenous groups, like the Tseltal from Amatenango, also use wood for pottery (Calderón 2001). Wood and construction. Wood is a highly valued material because it can be used to obtain cash to buy other materials or goods (Bolom-Ton 2005). Currently, the Highland Maya, who own agricultural land and forests, are economically relieved because they are assured wood for use and sale in addition to diverse food and fuel. Unlike firewood, timber extraction is practically monospecific, with pine as the most widely used group of trees, including different species and varieties (white

940

F. Ruan-Soto et al.

pine: Pinus pseudostrobus, pinabeto Pinus ayacahuite, and the others mentioned above). Cypress wood (nukulpat: Cupressus lusitanica) is highly appreciated, but currently scarce in the region. To some Indigenous peoples, the activities for the extraction, processing, and transformation of wood are still rustic, mainly using pine to make storage furniture, bookcases, desks, and other furniture items. However, the wood extracted by Indigenous families is mostly for domestic use, mainly for the construction of houses or to make rustic furniture. Other forest varieties, such as Oreopanax xalapensis (yich’ak mut), Alnus acuminata (nok), Cornus excelsa (isbon), and the pine species, are used as rustic wood to make beams and scaffolding of houses, yoke yokes, ladders, and handles for working tools; meanwhile, oak (different species of Quercus) wood is used for fences (Martínez-Icó et al. 2015). In general, it could be said that wood is underutilized. Thus, large-scale extraction is carried out mainly by timber companies or by external individuals who use the wood for their own purposes. The use of pine wood to build the walls of houses of the Highland Maya is fairly recent, and its use has spread through the use of chainsaws. In the past, the houses of the Indigenous population were mainly made out of mud and bajareque walls, with grass roofs (mainly Muhlenbergia vaginata) tied with vines or, in some cases, with roofs made of tejamanil (thin, elongated sheets of pine wood). The wood used back then was not mainly pine, but rather species that were known to be resistant over time, such as oak (Quercus spp.) and ch’ix te’ (manzanita or tejocote: Crataegus pubescens). Among the forest trees, those with the shape of pitchforks are sought for pillars and straight-stemmed trees were sought for beams. These woody elements were used in a rustic way without any particular treatment other than bark removal (Fig. 4). On the other hand, the use of the tejamanil was a specialized and arduous work due to the care needed to obtain thin sheets of pine wood and wooden nails to make roofs. Food. Since the diet of the Highland Maya is based on maize and beans (Fig. 5), the milpa continues being a cultivation system that not only provides these basic grains, but also a great diversity of food elements such as gourds (Curcubita spp.), peppers (Capsicum annum), green tomatoes (Physalis philadelphica), and greens (chicory: Sonchus oleraceus; mustard: Brassica juncea; hierba mora: Solanum americanum; and turnip: Brassica rapa) (Greenberg 2015). Closer to homes, the huerto (vegetable patch) or the sitio provides both greens and a multitude of fruit trees, native and introduced, such as annona, wild avocado, peach, apple, perón, and prunes. This space also includes other plants used as condiment, such as cilantro (Coriandrum sativum), epazote (Dysphania ambrosioides), and peppermint (Mentha sativa) (Lara et al. 2019). On the other hand, Highlands Maya use the forest to supplement their diet. Within the forest spaces, two basic food obtention activities still occur: collection and hunting. For collection, mature forests provide foods that include fruits and berries, greens, edible flowers, leaves to wrap tamales with, and condiments, among others. Fruits include wild avocado (Persea americana, called on-te in tsotsil), wild annona (Annona sp., called k’evex), blackberries (Rubus ulmmifolius), manzanitas (Crataegus mexicana), and cherries (Prunus serotina); edible flowers include

Ethnobotany of the Highlands of Chiapas

941

Fig. 4 Rural Home in the Highland of Chiapas. (Photo by Felipe Ruan-Soto)

Fig. 5 Traditional management of maize seeds. (Photo by Felipe Ruan-Soto)

colorín (Erythrina chiapasana), tamal leaves include mum (Piper auritum), and condiments include tsis-uch (Litsea glaucescens) (Martínez-Icó et al. 2015; Ramírez-Marcial et al. 2010). Furthermore, during the rainy season, forests yield a great diversity of edible mushrooms (in Tsotsil yuy, tajchuch, moni, which correspond to Amanita hayalyuy, Neolentinus lepideus, and Agaricus spp., respectively) (Ruan-Soto and Ordaz-Velázquez 2015), as well as honey from both stingless and western honeybees (Melipona genus and Apis mellifera, respectively), beetle and bee larvae, and both terrestrial and aquatic snails. Additionally, animals that are hunted for food include gophers (mentes), opossums (uch), rabbits (t’ul), squirrels (chuch), field rats (carransa), deer (chij), and some birds. Ornamental and ceremonial. The forests in general, or the forest landscape, constitute spaces that Highlander Maya people appreciate and recognize as the

942

F. Ruan-Soto et al.

Fig. 6 Use of pine branches in ceremonies in Chamula, Chiapas. (Photo by Felipe Ruan-Soto)

Fig. 7 Use of different plants in Day of the Dead altar, Chamula, Chiapas. (Photo by Felipe Ruan-Soto)

context where they live. A great diversity of plants, whose foliage, flowers, and fruits are used, are also used as decoration for several life events and the dead of the Tsotsil (Figs. 6 and 7). Plants are used for ornamental purposes in political, scholarly, religious, and ritual ceremonies, as well as crosses in the roads and natural wells. Protestants also use plants to adorn their temples during special events. Bromeliads are the most sought for such festivities, and the most appreciated species are Tillandsia guatemalensis (ech’), T. eizii (kilon ech’, yok nene), and T. usneoides (hay or paste’) (Jiménez-López et al. 2016); other species that are also used are Pinus spp., mainly its needles ( juncia or xaktoj), Myrsine juergensenii (naranjillo or tilil), and Abies guatemalensis (romerillo) (Martínez-Icó et al. 2015; Ramírez-Marcial et al. 2010).

Ethnobotany of the Highlands of Chiapas

943

Medicinal. Because of the lack of medical centers, or due to cultural reasons, medicinal plants are commonly used by the highlander Maya and widespread in their territory. Many medicinal plants are not found in mature forests, but rather in disturbed areas, such as milpas, fallows, or homegardens. Such plants, commonly used by women and men, cure ailments ranging from diarrheas (e.g., Baccharis vaccinoides and Crataegus pubescens), intestine parasites (e.g., Dysphania ambrosioides), stomachaches (e.g., Litsea glaucescens, Tagetes nelsonii), cough (e.g., Salvia lavanduloides, Sambucus mexicana), and wounds (Solanum sp.), to anger and shame (e.g., Verbena litoralis), as has been documented in different research reports, which have allowed appreciating the vast knowledge that the highland Mayan peoples have about plants (e.g., Berlin et al. 2000; Caballero et al. 1998; Lara et al. 2019; Nepomuceno and Ishiki 2010; Soto-Pinto et al. 1988). Other uses and practices related to plants. As stated above, another important use of forests is as grazing lands. Many of their plants are used as forage for cattle. At this point, Martínez-Ico et al. (2015) identify up to 12 foraging species in the forests of Bazom. These include Equisetum myriochaetum, Oreopanax xalapensis, Piqueria pilosa, and Begonia oaxacana, among others. Lastly, forests provide trees used to build pens (Rhamnus caprifolia) and live fences (Salix spp.), reeds (Passiflora membranacea) used for elaborating mecates (rope), and herbs used as soap (Cyclanthera pedata) or as insecticide, among other varied uses (Martínez-Icó et al. 2015). However, it must be said that modern industrialized resources, many of which are more durable and affordable, have led to the lack of use of several of these plants.

Medicinal Ethnobotany in the Highlands of Chiapas From the earliest anthropological studies, in the Highlands of Chiapas started the recording of the local use and knowledge of a great diversity of botanical species for the treatment of several ailments. Knowledge related to the collection, processing, and use of local medicinal plants is found both in the focused knowledge of ethnomedical specialists and in the world of domestic attention of ailments. The information available about medicinal plants in the area has been recorded by research from different disciplines and from different methodological approaches, both quantitative and qualitative. These include areas such as social anthropology (Ayora-Diaz 2010; Holland 1963), ethnography (Vogt 1979), cultural anthropology (Pitarch 2000), ethnobiology and ethnobotany (Lara et al. 2019; Hernández-Alcázar et al. 2016; PROCOMITH 1990), linguistics applied to medical ethnobotany (Berlin and Berlin 1996; Berlin et al. 1974, 1990, 1999, 2000), and ethnomedicine (Page 2005). Within the Tsotsil and Tseltal ethnomedicine from the Highlands of Chiapas, people dedicated to cure and tend to ailments are known as jpoxtavanej, and this term groups diverse specialties. These include the Tsotsil j’iol (a person who is able to diagnose, predict, and treat ailments of diverse origins, mediate with deities, and

944

F. Ruan-Soto et al.

see through pulse), the jvetom or jtamol (midwife), j’ac’vomol (one who knows and cures through plants, herbalist), jtsac-bac (huesero or bone fixer), and jtoibits or coponej’vits (mountain prayer specialist) (Page 2005). In general terms, all that is deemed “medicine” is called pox, a word that includes first and foremost a traditional cane-distillated drink from the Highlands of Chiapas, but which also includes medicinal plants, prayers, industrialized products (supplements, vitamins), and patent drugs (Pitarch 2000). As discussed in other chapters of this book, the “medicinal” use category probably has the highest number of anthropocentric significance species in the Highlands of Chiapas. There are records of over one hundred of plant species and mixtures of plant species used as medicine (Berlin and Berlin 1993; Berlin et al. 1974; Hernández-Alcázar et al. 2016; Lara et al. 2019; Moscoso 1981; Nepomuceno and Ishiki 2010; Page 2005). These are used for the treatment of gastric, dermic, respiratory, cardiovascular, metabolic, chronic-degenerative, muscle-skeleton system, reproductive system, renal and urinary system, gynecologic, and birth- and puerperium-related ailments (Hernández-Alcázar et al. 2016; Lara et al. 2019; Page 2005; Sántiz-López et al. n.d.). However, as it is the case in many Indigenous ethnomedicines, the importance of carrying out ritual actions to ensure the efficacy of the plants used in the Tsotsil and Tseltal ethnomedicine has been widely documented (Holland 1963; Page 2005; Vogt 1979). However, it should be pointed out that these practices are among the most persecuted aspects by the increasingly widespread protestant religions and by the most conservative factions of the catholic church.1 Similarly, the botanical species used as medicine are placed within a symbolic referent that endows them with various features, which will ultimately determine their usefulness in treating specific ailments. The notion of the “hot” or “cold” nature of both illnesses and therapeutic elements is present in the Mayan ethnomedicine of the Highlands, as it is the case in numerous Amerindian ethnomedicines. Furthermore, the dual conception of cosmos, and the inherent duality of the elements within it, is a matrix of meaning in which diagnoses, procedures, and ethnomedical rituals are developed (Page 2005). Therefore, even though it has been frequently overlooked in ethnobotanical research, the comprehension and study of local worldviews is fundamental for the analysis and the documentation of the Tsotsil and Tseltal herbalist diversity of the Highlands of Chiapas. An important example of this is the detailed information provided by Evon Vogt (1979) in his study of rituals in Zinacantan, Chiapas. In this document, there is a description of an “affliction ritual,” destined to expel severe illnesses through diverse 1

In the Highlands of Chiapas, progressive currents of the catholic church, such as the Pastoral de la Tierra (Earth Pastoral) and the Teología India (Indigenous Theology), have flourished as a result of the work started by priests and nuns influenced by the Theology of Liberation and based on the premises set by the Second Vatican Council. In several churches from Indigenous communities in the region, the prayers and beliefs are respected, reaching even ecumenicism and a relationship with local worldviews, as is common in the Indigenous Theology. Based on these ideologies, there is a place for local ethnomedicines and their practice is respected.

Ethnobotany of the Highlands of Chiapas

945

offerings to specific deities, which includes the use of 13 ritual plants, whose taxonomy, botany, and symbolic significance, for the healing, is described in great detail.

Plants and Worldview The relevance of plants in the worldviews of the peoples from the Highlands of Chiapas is undeniable, and it includes profound and complex processes of symbolization, many of which are grounded on the Mesoamerican religious tradition (Espinosa 2015; Sánchez et al. 2018). In this sense, it is possible to approach the role of plant species within the worldviews of some of these Mayan peoples by considering the ethnographic fields of research related to cosmology or the geometry of the cosmos; the understanding and concept of human beings as persons; and the ritual processes, both in the field of health-illness and in matters of agriculture and specially corn (Medina 2000). In the following lines, we present elements of the human-plant link framed in these topics. Among the Highland Maya peoples, such as Tsotsil and Tseltal, there is a general concept of a world integrated by several spheres that are interconnected both horizontally and vertically. This cosmological order can be best appreciated in ritual processes, mythologies, prayers, and offerings for different purposes. In the horizontal plane, it is said that the world has four corners, directions, or regions and a center; each characterized by the presence of specific beings and qualities. Meanwhile, vertically, Indigenous conceptions indicate that there are three great spaces: the skies, the earth, and the subterranean or underworld (Figuerola 2010; Page 2005). In past decades, it was common to consider that each of the corners of the world contained ceibas (Ceiba pentandra, named inop in Tsotsil), which held the world in place and separated the earth from the sky. An enormous ceiba stood in the center of the universe, symbolizing a mountain or a tiered pyramid (Holland 1963). This conception is associated to the central significance of pines or ocotes (Pinus spp., called toj in Tsotsil) in their cosmology; these trees are arranged alongside wooden crosses or in their stead, in special spots of the sacred geography, symbolizing access doorways to the dwellings of the ancient gods, that is, connections between worlds or spaces (Vogt 1979). The altars that are configured in the spots where crosses and pines are found recreate the ascent to the “sacred mountain,” whose inside is deemed a sort of “pen” surrounded by pine trees that houses the companion animals of people, as well as the ch’ulel or soul of each of the inhabitants of the surrounding communities (Vogt 1979). This is linked to the fact that, in certain rituals, red geraniums tsajal nichim (Pelargonium x hortum) must be tied to the pine points that are set with the crosses; it is said that these geraniums are the “hearts” of people (Vogt 1979). As this information illustrates, there are profound associations between the botanical species and the spiritual human entities (frequently referred to as “souls” in literature), since, in general terms, the worldviews of the Mayan peoples coincide in allocating a similar spiritual entity to humans, animals, and plants, called in Tsotsil

946

F. Ruan-Soto et al.

and Tseltal ch’ulel. Its state determines the life of each person and may, under certain conditions, determine the life of a community (Guiteras 1986). This can be exemplified by data from the Tojolabal area, in which there is a concept of the soul of plants, called altsil in this language, which is found within specific mountains that are protected by the chawuk, lighting-humans who have a gift from birth that gives them the power of thunder and to whom divinities have assigned the care of the edible plants in each community (Guerrero and Álvarez 2018). The lightninghumans get their gift through a dream in which a seed of the plant under their care is given to them, while the altsil of this plant is housed within this person. If the altsil is stolen or the lightning-man dies, the plant that this person protected stops growing properly in the milpas of the community (Guerrero 2018). Interestingly, the soul or altsil of the plants is regularly found in each grain; it is the vegetable embryo, so this word may also be understood as the “vital principle” of plants. In a complementary fashion, plants, as well as humans, have a “maturation principle,” commonly called “heart” or k’ujol in Tojolabal. For woody species of plants, it is associated with the heartwood, the part of the plant that gives it firmness and structure. For humans, it is linked to the “spiritual heart” of a person, as its center in which emotions and cognition reside and which is acquired throughout the years and can be seen in the gray hair of the elderly (Guerrero 2018). Thus, analogies between human and plant anatomy are highly significant in the worldview of the Maya peoples of the Highlands of Chiapas. Among the Tsotsil, each plant was found to have a particular “inner soul,” which acquires a given singularity according to a mix of specific features, among which are activity or passivity, coldness or hotness, and whether it is red, white, black, yellow, or green-blue, the colors associated with the directions of the universe and its center (Vogt 1979). These attributes are fundamental for the use of ritual plants in a healing context, and they depend on the oneiric processes that each ritual specialist experiences. For example, in case someone suffers from xj’el or susto (literally, “scare”), an ailment caused by the partial or complete exiting or loss of the ch’ulel in a specific location because of a sudden event, the ritual specialist must employ 13 bouquets, each made up of four different species of ritual plants, for the healing process. In these bouquets, apart from Pinus spp., some relevant species include wild bay leaf tsis uch (Listea glaucescens), wild myrtle aja te’es (Gaultheria odorata); several species from the genus Peperomia, called wixobtakil in Tsotsil (a name literally translated as “older sisters,” but which refers to plants as helpers of the ancestral deities); bromeliads bats’j ech’ (Tillandsia guatemalensis); and the “spirit flower,” ch’ulelal nichim (Satureja mexicana). These, alongside other ritual plants, are used to round up the “pen” on which the patient lies, representing the forest in which the divinized forefathers, whose protection is requested, dwell (Vogt 1979). According to local conceptions, a total of 52 plant specimens symbolize the 13 parts of the soul or ch’ulel of the patient, which are directly associated with the 52 grains of maize that the healer uses in a mantric process to discern the origin and way to alleviate the ailment they suffer, as well as coinciding with the ancient Mesoamerican ritual calendar (Vogt 1979).

Ethnobotany of the Highlands of Chiapas

947

Another group of plants that is relevant in different fields of the worldview of peoples from the Highland of Chiapas is the squash, commonly called tecomates or pumpos and tsuj in several Mayan languages from the region. The main species that this refers to is Lagenaria siceraria, which is widely used as a vessel symbolizing sacred springs or water bodies, and which are fundamental in diverse healing or petition ritual, and they also appear in the mythology of several towns associated to deities or humans with powers, who pour these fruits filled with water onto the earth to make it rain in the world. These elements show the relevance of plant species in the conceptions people have on the world, as well as the diversity and depth of aspects in which this relationship is illustrated.

Final Thoughts As can be appreciated, the different studies carried out in the Highlands of Chiapas prove beyond doubt that the Maya peoples inhabiting this area possess a rich heritage of knowledge associated to the elements of their surrounding flora. They name and use them on a day-to-day basis within the realm of their needs and possibilities. This local knowledge is the result of centuries of interaction with their environment and the mastery of understanding of many natural processes; however, we may also state that it contains undeniable inclusions and adaptations. While a mythical relationship to natural elements remains, these have been modified after the colonization and, more recently, by educational, religious, and health campaigns. Regarding this, it may be said that local knowledge is founded on a balance between the management and use of resources and their conservation; however, this may only happen when the sociocultural conditions of a population are considered (including their high rates of poverty and marginalization) and their current economic, productive, political, and social relationships are acknowledged, which leads to a much wider framework. Acknowledgments We would like to thank the editors of this book for their invitation to participate. We would also like to thank Dikaryon Language Consultants for translating the original Spanish version of this chapter to English.

Referens Alemán-Santillán T. Chiapas: El bosque de pino-encino y los mayas de las tierras altas. México: IX Congreso Forestal Mundial; 1985. Alemán-Santillán T. Los sistemas de producción forestal y agrícola de roza. In: Parra R, coordinator. El subdesarrollo agrícola en Los Altos de Chiapas. Chapingo: Universidad Autónoma de Chapingo; 1989. Alemán-Santillán T. La explotación del bosque en las regiones indígenas: sus aportes y perspectivas en la generación de alternativas de uso sostenido de los recursos naturales. In: Parra MR, Díaz

948

F. Ruan-Soto et al.

Hernández BM, editors. . Los Altos de Chiapas: Agricultura y crisis rural, tomo I. San Cristóbal de Las Casas. El Colegio de la Frontera Sur; 1997. Ayora-Diaz SI. Modernidad alternativa: medicinas locales en los Altos de Chiapas. Nueva Antropol. 2010;23(72):11–31. Barrera-Aguilar O. El hombre en la naturaleza: los resultados del Proyecto Chicago en Chiapas, 1956–1959. Liminar. 2018;17(1):97–113. Berlin B. La clasificación etnobotánica maya de Los Altos de Chiapas. Un sistema comprensivo y natural. In: Perez-Farrera MA, Tejeda C, Rivera E, editors. Los bosques mesófilos de montaña en Chiapas, Situación actual, diversidad y conservación. Tuxtla Gutiérrez: UNICACH; 2010. Berlin EA, Berlin B. Enciclopedia etnomédica maya: bases científicas de la medicina tradicional en los Altos de Chiapas. Tuxtla Gutiérrez: Gobierno del Estado de Chiapas – Instituto Chiapaneco de Cultura; 1993. Berlin EA, Berlin B. Medical ethnobiology of the highland Maya of Chiapas, México. The gastrointestinal diseases. New Jersey: Princeton University Press; 1996. Berlin B, Breedlove D, Raven P. Principles of Tzeltal plant classification. New York: Academic; 1974. Berlin B, Berlin EA, Breedlove DE. La herbolaria médica tzeltal-tzotzil en Los Altos de Chiapas, vol. I. Tuxtla Gutiérrez: Gobierno de Estado de Chiapas – PROCOMITH; 1990. Berlin B, Berlin AE, Fernández-Ugalde JC, García-Barrios LE, Puett D, Nash R, GonzálezEspinoza M. The Maya ICBG: drug discovery, medical ethnobiology, and alternative forms of economic development in the Highland Maya region of Chiapas, Mexico. Pharm Biol. 1999;37(1):128–44. Berlin EA, Berlin B, Gómez-López S. Manual etnomédico de Oxchuc: guía básica y herbolaria/ Stalel sk’op ya’yejal bit’il ta pasel wamal poxil ta Oxchujk’: sk’opla bit’il ta meltsanel poxil sok wamaletik. San Cristóbal de Las Casas: El Colegio de la Frontera Sur; 2000. Bolom-Ton F. De bosques y saberes. Perspectivas de conservación de recursos forestales en el municipio de Huixtán, Chiapas. Tesis de maestría. San Cristóbal de Las Casas: Centro de Estudios Superiores de México y Centroamérica, Universidad de Ciencias y Artes de Chiapas; 2005. Breedlove DE. The phytogeography and vegetation of Chiapas (México). In: Graham A, editor. Vegetation and vegetational history of Northern Latin America. Amsterdam: Elsevier; 1973. Breedlove DE. Flora of Chiapas. Part I: introduction to the flora of Chiapas. San Francisco: California Academy of Sciences; 1981. Breedlove DE. Listado florístico de México. IV Flora de Chiapas. Ciudad de México: Instituto de Biología de la UNAM; 1986. Breedlove D, Laughlin R. The flowering of man: a Tzotzil botany of Zinacantan. Washington, DC: Smithsonian Institute Press; 1993. Caballero J, Casas M, Cortés L, Mapes C. Patrones en el conocimiento, uso y manejo de plantas en pueblos indígenas de México. Estud Atacameños. 1998;16:181–95. Calderón A. Uso y acceso a los recursos forestales en una comunidad indígena: La leña en Amatenango del Valle, Chiapas, México. Tesis de Maestría. San Cristóbal de Las Casas: El Colegio de la Frontera Sur; 2001. Cayuela L, Benayas JM, Echeverría C. Clearance and fragmentation of tropical montane forests in the highlands of Chiapas, Mexico (1975–2000). Forest Ecol Manag. 2006;226:208–18. De Jong B, Ruíz MJ. La investigación forestal y su contribución al manejo de bosques. In: Parra M, Díaz-Hernández BM, editors. Los Altos de Chiapas: Agricultura y crisis rural. México. San Cristóbal de Las Casas: El Colegio de la Frontera Sur; 1997. Enríquez P, Mariaca R, Retana O. Uso medicinal de la fauna silvestre en Los Altos de Chiapas, México. Interciencia. 2006;31(7):491–9. Espinosa G. La flora como esfera particular y, a la vez, generadora de la cosmovisión. Propuesta de cinco modelos para su estudio en un contexto mesoamericano. In: Albores B, editor. Flor-flora: su uso ritual en Mesoamérica. Estado de México: El Colegio Mexiquense, A. C., Gobierno del Estado de México; 2015.

Ethnobotany of the Highlands of Chiapas

949

Farrera O, Hernández F, Diaz M. Las Plantas con Flores. In: CONABIO, editor. La Biodiversidad en Chiapas: Estudio de Estado, vol. II. Ciudad de México: CONABIO/Gobierno del Estado de Chiapas; 2013. Figuerola H. Los dioses, los hombres y las palabras en la comunidad de San Juan Evangelista Cancuc en Chiapas. Mérida: Universidad Nacional Autónoma de México; 2010. García MA. Recursos Naturales y Desarrollo (¿sustentable?) y territorios indígenas de Chiapas. In: Del Val J, Sánchez C, Zolla C, Coordinators. Estado del Desarrollo Económico y Social de los Pueblos indígenas de Chiapas. Ciudad de México: Universidad Nacional Autónoma de México; 2018. González-Espinosa M, Ochoa-Gaona S, Ramírez-Marcial N, Quintana-Ascencio PF. Contexto vegetacional y florístico de la agricultura. In: Parra MR, Díaz-Hernández BM, editors. Los Altos de Chiapas: agricultura y crisis rural. San Cristóball de Las Casas: El Colegio de la Frontera Sur; 1997. Greenberg A. Estado del conocimiento y uso de las verduras silvestres y semidomésticadas en Los Altos de Chiapas. Tesis de maestría. San Cristóbal de Las Casas: El Colegio de la Frontera Sur; 2015. Guerrero F. Percepción, ambiente y persona entre los tojolabales de Las Margaritas y Altamirano, Chiapas. Tesis de Doctorado. Ciudad de México: Universidad Nacional Autónoma de México; 2018. Guerrero F, Álvarez RB. Etnobiología tojol-ab’al: síntesis y nuevos aportes. In: Mariaca R, Elizondo C, Ruan F, editors. Etnobiología y Patrimonio Biocultural de Chiapas, tomo I. San Cristóbal de Las Casas: El Colegio de la Frontera Sur; 2018. Guiteras C. Los peligros del alma. Visión del mundo de un tzotzil. México: Fondo de Cultura Económica; 1986. Haviland J. Sk’op Sotz’leb; El Tzotzil de San Lorenzo Zinacantán. Ciudad de México: UNAM; 1981. Hernández-Alcázar JA, Cruz-Cruz C, García-Martínez R, Gutiérrez-Bravo EJ, Urbina-Rojas FK, Rodríguez-Escobar JE. Plantas utilizadas por médicos tradicionales de la cabecera municipal de Pantelhó, Chiapas, México. Lacandonia. 2016;10(10):29–36. Holland RW. Medicina maya en los Altos de Chiapas; un estudio del cambio socio-cultural. Ciudad de México: Instituto Nacional Indigenista; 1963. INAFED. Instituto para el Federalismo y el Desarrollo Municipal. Secretaría de Gobernación. http://www.inafed.gob.mx/work/enciclopedia/EMM07chiapas/ Accessed 6 Oct 2017. INEGI. VII censo agrícola-ganadero. Chiapas resultados definitivos. Tomos I, II, III y IV. Aguascalientes: INEGI; 1994. INEGI. Principales resultados por localidad. Chiapas. Aguascalientes: INEGI; 2001. INEGI. Principales resultados de la Encuesta Intercensal. Aguascalientes: INEGI; 2015. Jiménez-López DA, Trejo-Cruz IA, Mó E. Uso ceremonial de bromelias epífitas en Las Margaritas, Chiapas, México. Herbario CICY. 2016;8:193–8. Kaufman T. Aspects of the lexicon of proto-mayan and its earliest descendants. In: Aissen J, England NC, Zavala R, editors. The Mayan languages. London: Routledge Language Family Series; 2017. Lara EA, Fernández E, Zepeda JM, Lara DJ, Aguilar A, Van Damme P. Etnomedicina en Los Altos de Chiapas, México. Bol Lat Car Plant Med Arom. 2019;18(1):42–57. Laughlin R. El Gran Diccionario Tzotzil de San Lorenzo Zinacantán. Ciudad de México: CIESAS; 1975. Mariaca R, Pérez J, León N, López A. La Milpa Tsotsil de los Altos de Chiapas y sus recursos genéticos. San Cristóbal de Las Casas: ECOSUR/UNICH; 2007. Martínez G, Villasana S. Dinámica Sociodemográfica. In: Del Val J, Sánchez C, Zolla C, Coordinators. Estado del Desarrollo Económico y Social de los Pueblos indígenas de Chiapas. Ciudad de México: Universidad Nacional Autónoma de México; 2018. Martínez-Icó M, Cetzal-Ix W, Noguera-Savelli E, Hernández-Juárez R. Flora vascular de la comunidad de Bazom, Los Altos de Chiapas, México. Bot Sci. 2015;93(1):53–72.

950

F. Ruan-Soto et al.

Marzal M. Historia de la Antropología: Antropología cultural, vol. II. Quito: Abya Yala; 2016. Medina A. En las cuatro esquinas, en el centro. Etnografía de la cosmovisión mesoamericana. Ciudad de México: Universidad Nacional Autónoma de México; 2000. Miranda F. La vegetación de Chiapas. Ciudad de México: UNAM; 1952. Moscoso P. La medicina tradicional de los Altos de Chiapas. San Cristóbal de Las Casas: Editorial Tradición; 1981. Nepomuceno A, Ishiki M. Las plantas empleadas para el tratamiento de las infecciones respiratorias en los altos de Chiapas (México). Etnobiología. 2010;8:11–30. Ochoa-Gaona S, González-Espinosa M. Land use and deforestation in the high land of Chiapas, Mexico. Appl Geogr. 2000;20:17–42. Page JT. El mandato de los dioses. Etnomedicina entre los tzotziles de Chamula y Chenalhó, Chiapas. San Cristóbal de Las Casas: Programa de Investigaciones Multidisciplinarias sobre Mesoamérica y el Sureste, Universidad Nacional Autónoma de México; 2005. Parra MR, Alemán-Santillán T, Hernández BD, García MC, García-Barrios LE, López-Meza A, Márquez A, Nahed J, Plascencia H, Pool L, Soto-Pinto L. La producción silvoagropecuaria en Los Altos de Chiapas: análisis de un sistema complejo. In: Colegio de Postgraduados, ORSTOM, CONACYT, editors. Sistemas de producción y desarrollo agrícola. Ciudad de México: Colegio de Postgraduados y ORSTOM; 1993. Pitarch P. En el Museo de la Medicina Maya. Mayab. 2000;1:79–91. PROCOMITH. La Herbolaria Médica Tzeltal-Tzotzil en los Altos de Chiapas: un ensayo preliminar sobre las cincuenta especies botánicas de uso más frecuente. Tuxtla Gutiérrez: Gobierno del Estado de Chiapas; 1990. Quintana-Ascencio P, González-Espinosa M. Afinidad fitogeográfica y papel sucesional de la flora leñosa de los bosques de pino-encino de Los Altos de Chiapas, México. A Bot Mex. 1993;21: 43–57. Ramírez-Marcial N, Camacho-Cruz A, Martínez-Icó M, Luna-Gómez A, Golicher D, GónzalezEspinosa M. Árboles y arbustos de los bosques de montaña en Chiapas. San Cristóbal de Las Casas: El Colegio de la Frontera Sur; 2010. Ramos-Martínez MM, Montoya G, Vásquez MA. Producción y comercialización de carbón vegetal en la microrregión Teopisca-San Cristóbal. Rev Geogr Agr. 1999;58:53–71. Robles-Porras L, Huerta G, Andrade Gallegos R, Ángeles H. Conocimiento tradicional sobre los macromycetes en dos comunidades de Oxchuc, Chiapas, México. Etnobiología. 2007;5:21–35. Ruan-Soto F, Ordaz-Velázquez M. Aproximaciones a la etnomicología maya. Revista Pueblos y fronteras digital. 2015;10(20):44–69. Ruz MH. Copanaguastla en un espejo. Un Pueblo Tzeltal en el Virreinato. San Cristóbal de Las Casas: Universidad Autónoma de Chiapas; 1985. Sánchez M, Sánchez MA, Vázquez TJ. Etnobiología de los tsotsiles de Chiapas. In: Mariaca R, Elizondo C, Ruan F, editors. Etnobiología y Patrimonio Biocultural de Chiapas, tomo I. San Cristóbal de Las Casas: El Colegio de la Frontera Sur; 2018. Sánchez-Álvarez M. Los sistemas tradicionales de producción de maíz y frijol en la vida de los tzotziles del municipio de Huixtán, Chiapas. Tesis de maestría. Oaxaca: Instituto Tecnológico de Oaxaca; 2004. Sántiz-López M, García-Pérez R, Gómez-Girón D. Las plantas medicinales utilizadas por los mayas de los Altos de Chiapas. San Cristóbal de Las Casas: Organización de Médicos Indígenas del Estado de Chiapas; n.d. Soto-Pinto L, López-Meza A, García MC. Etnobotánica y religión entre los chamulas en Los Altos de Chiapas, México. In: Uribe R, compilator. Medio Ambiente y Comunidades Indígenas del Sureste. Villahermosa: Comité Regional para la UNESCO; 1988. Soto-Pinto L, Jiménez G, de Jong BH. La agroforestería en Chiapas. El caso de la región Los Altos. In: Parra MR, Díaz Hernández BM, editors. Los Altos de Chiapas: Agricultura y crisis rural, tomo I. Los recursos naturales. San Cristóbal de Las Casas: El Colegio de la Frontera Sur; 1997.

Ethnobotany of the Highlands of Chiapas

951

Villalobos G. El Contexto Físico y su Importancia para la Preservación de la Biodiversidad. In: CONABIO, editor. La Biodiversidad en Chiapas: Estudio de Estado, vol. I. Ciudad de México: CONABIO/Gobierno del Estado de Chiapas; 2013. Vogt EZ. Los zinacantecos, un pueblo tzotzil de los altos de Chiapas. Ciudad de México: INI; 1966. Vogt EZ. Ofrendas para los dioses: análisis simbólico de rituales zinacantecos. Ciudad de México: Fondo de Cultura Económica; 1979. Weathers K, Weathers N. Diccionario tzotsil-español. Ciudad de México: Instituto Lingüistico de Verano; 1949.

Perspectives of the Ethnobotanical Research in Mexico Alejandro Casas, Jose´ Juan Blancas Vázquez, and Heike Vibrans

Abstract Ethnobotany is an integrative, multidimensional research field, whose main purpose is understanding interactions and relationships between humans and plants, and their cultural, ecological, and evolutionary consequences throughout history. This purpose is linked with the general interest of science for analyzing interactions between human societies and biodiversity, which is currently relevant not only from theoretical perspectives but also to design strategies to face environmental problems characterizing the global socio-ecological crisis. Ethnobotany, as ethnobiological sciences in general, has advanced in the construction of valuable theoretical and methodological frameworks, which have made visible the enormous value of the local experience of thousands of communities throughout the world. Their knowledge and techniques of managing plants and ecosystems may be the foundation of sustainable forms of biocultural interactions. In this chapter, some relevant theoretical and methodological advances and challenges for ethnobotanists working in Mexico are identified. We emphasize the importance of ethnobotany and ethnobiological sciences in establishing bridges of dialogue among different sectors of the societies that make decisions on biodiversity issues. Such role positions ethnobotany as a key transdisciplinary field for research and action to: (1) understanding A. Casas (*) Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] J. J. Blancas Vázquez Centro de Investigación en Biodiversidad y, Conservación (CIByC) - Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico e-mail: [email protected] H. Vibrans Posgrado en Botánica, Colegio de Postgraduados. Montecillo, Texcoco, Estado de México, Mexico e-mail: [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_57

953

954

A. Casas et al.

traditional botanical knowledge, (2) developing criteria to protect biodiversity and intellectual property rights, (3) understanding the origin, diversification, and diffusion of agrobiodiversity, (4) identifying the bases for the sustainable management of plants and ecosystems, and (5) constructing strategies for biocultural diversity conservation. These are central points in the agenda of ethnobotany, and it is intimately linked to anthropology, ecological economy, and ecological and evolutionary sciences. In addition, a main challenge is to link ethnobiological sciences with sustainability science. This is an emerging scientific approach based on new paradigms for conducting research, which are needed to understand and act in the context of the global environmental crisis. The chapters of this book reflect different perspectives and research approaches developed by ethnobotanists in Mexico, but also their views about the ways ahead. We emphasize the need of making explicit the insertion of ethnobotany with other research fields for constructing new theoretical and methodological perspectives, but, especially, to promote studies about cultures, ecosystems, and regions scarcely explored. We identify areas and cultures that are research priorities. Also, we emphasize the necessity to enhance the insertion of ethnobotany and ethnobiological sciences in educational programs and institutions that make decisions and public policies related to biocultural issues. This final chapter summarizes views emerging from the cases included in the book and provide some reflections we consider relevant to study the ethnobotany of the mountain regions of Mexico but that may also be helpful for ethnobotanists in other regions of the world.

Ethnobotany in Mexico Ethnobotany is an integrative, multidimensional research field, whose main purpose is understanding the different types of interactions and relationships between humans and plants and the ecosystems of which they are part. In particular, ethnobotany looks for answers about what kinds of interactions humans have with plants, when, how, and why these interactions have occurred, and their cultural, socioeconomic, ecological, and evolutionary consequences throughout history (HernándezXolocotzi 1971, 1979; Bye 1993; Pieroni and Quave 2014; Casas et al. 2016a; Vibrans and Casas 2022). Ethnobotany is part of more general scientific fields that analyze the interactions between human societies and biodiversity. It involves numerous research questions and hypotheses that integrate methods and concepts of the natural and social sciences to answer and test them. The construction of ethnobotany theoretical frameworks has been a main concern of scientists working in this field for decades (Schultes 1941; Hernández-Xolocotzi 1971, 1979, 1993; Berlin et al. 1974; Ford 1978; Caballero 1979; Alcorn 1984; Bye 1985, 1993; Martínez-Alfaro 1994; Schultes and von Reis 1995; Toledo 1995; Cotton 1997; Caballero et al. 1998; Alexiades 2003; Albuquerque and Hanazaki 2009; Casas et al. 2016a; Camou et al. 2016; Albuquerque et al. 2019; Vibrans and Casas 2022). Although biology and anthropology are the most related sciences and provide relevant theoretical and methodological approaches, ethnobotany has its own domains, processes, and techniques.

Perspectives of the Ethnobotanical Research in Mexico

955

Theoretical constructions of ethnobotany have considered, above all, that interactions and relations between humans and plants have particularities compared with those established among other species. The shared intentionality (O’Madagain and Tomasello 2021) and design of actions by humans, the transmission of their “cumulative culture,” their deliberate use of natural elements such as water and minerals, or phenomena like fire and wind to modify materials and systems, as well as their tools and other technologies have all mediated such interactions and their consequences (Vanberg 2006; Tattersall and Schwartz 2009; Stout 2008; Verbeek 2008; CelaConde and Ayala 2018). The “reason-based” forms of cultural transmission, and the ability to coordinate actions, according to O’Madagain and Tomasello (2021), are causes of the rapid innovation and cumulative cultural evolution of humans. And the particularities of interactions between humans and biodiversity are part of the also specific ways humans construct their ecological and cultural niches (Zeder 2012; Smith 2012; Boivin et al. 2016; Clement et al. 2021). Ethnobotany focuses its attention on analyzing the interactions and relations between humans and plants, and the ecosystems they are part of. The cultural and ecological contexts worldwide are highly variable and have changed throughout the history of humans on Earth (Boivin et al. 2016). Therefore, ethnobotany is immersed in diverse and changing constellations of interactions and impacts. In other words, the particularities of interactions between humans and plants (and ecosystems) are far from simple and have a plethora of expressions. These expressions are part of the biocultural diversity (Maffi and Woodley 2010) and understanding them is a main task of ethnobotany. Nearly half of the terrestrial ecosystems on Earth have been severely damaged by humans (MEA 2005; Barnosky et al. 2011, 2012), and these processes have been especially intensive during the last 300 years: the industrialization era. But they have been dramatic during the last 75 years (Barnosky et al. 2012), closely aligned with modern forms of capital accumulation and the hegemonic policies of economic growth (Pacheco et al. 2018). Sources of global cultural information like Ethnologue (Eberhard et al. 2022) have documented nearly 7000 languages currently existing worldwide but, as biodiversity, a significant proportion disappeared during different historical periods, especially since the European colonialism/imperialism epoch (Gilmartin 2009), and numerous others are in process of extinction in the era of globalization (Maffi and Woodley 2010; Eberhard et al. 2022). Current research tools allow the reconstruction of the diversity of ecosystems and biodiversity that have existed on the planet (Pereira et al. 2012; Navarro et al. 2017), and how landscapes have been transformed by humans at different times (Boivin et al. 2016; Balée 2018; Franco-Moraes et al. 2021). Linguistic approaches trace the history of language diversification, but also estimate their loss rates and monitor trends of both languages and culture loss (Cavalli-Sforza 1997; McMahon and McMahon 2005; Eberhard et al. 2022). The notion that biodiversity and culture are parts of complex biocultural systems has gained strength in this millennium (Maffi 2005; Maffi and Woodley 2010) and today, biodiversity and culture should be analyzed holistically. Luisa Maffi, one of the main pioneers and promoters of this

956

A. Casas et al.

trend, identified ethnobiology as a principal source of research and thinking on biocultural diversity, and the base for theoretical frameworks (Maffi 2005). The dramatic loss of biocultural diversity contrasts with its increase during millions and thousands of years of natural and cultural history, respectively. This information makes possible to dimension the magnitude of the catastrophic impact of the global socio-ecological crisis on diversity (Hamilton et al. 2015; O’Connor et al. 2020). Recognizing the characteristics of the interactions between humans and biodiversity, the diversity of interactions and interrelationships developed between humans and plants, and the processes causing their loss are all main challenges of ethnobotanical research. Such understanding is not only valuable from theoretical perspectives, but it is also crucial for designing strategies for biocultural conservation and restoration from the local to global scales. These general challenges inevitably lead the way ethnobotanical research should take in the near future, in Mexico and worldwide. The chapters of this book display a wide array of research approaches and knowledge on the diversity of interactions between cultures, plants, and ecosystems in Mexico. It complements the previous work Ethnobotany of Mexico (Lira et al. 2016), and surely further works will contribute to broadening its reach. The book includes case studies in communities and regions, general topics of ethnobiology related to conservation of biocultural diversity, as well as views on some plant genera representative of the biocultural diversity in Mexico. But it is only a sample of the research groups working in this country, the topics and methodological approaches we use, and a small portion of plant groups that form part of Mexican cultures. The chapters also identify some of the conceptual, methodological, and information gaps still needing to be filled to advance several research issues, and to cover more regions, ecosystems, and cultural groups. As part of the series Ethnobotany of the Mountain Regions, this book emphasizes studies of peoples living in the mountains, the predominant landscapes of the Mexican territory. However, some studies also include adjacent and connected lowlands and highland plateaus. Ethnobotanists working in Mexico are part of the growing international scientific community of this field, and its advances and limitations partly reflect worldwide trends. Therefore, these results may be relevant for colleagues working in other regions of the world. The theoretical and applied challenges of ethnobotany and ethnobiological sciences in the face of an overwhelming socio-ecological global crisis are great, but its approaches could play a role to mitigate its effects. Some reflections and concerns are shared in this final chapter of the book.

Contemporary Theoretical Challenges for Ethnobiological Sciences A number of ethnobotanists have discussed key questions and hypotheses for developing ethnobiological sciences. Here, we emphasize some lines of thought that we consider particularly relevant.

Perspectives of the Ethnobotanical Research in Mexico

957

Tracing theoretical frameworks provides insights on the origins and trajectories of disciplines, and their context. Victor Manuel Toledo in his work “New paradigms for a new ethnobotany: reflections on the case of Mexico” (Toledo 1995) examined the turn of ethnobotany (in Mexico, but with more general implications) based on the framework of the structure of scientific revolutions developed by Thomas S. Kuhn (1962). Toledo analyzed the stages of ethnobotany in Mexico and the need to connect its approaches with new paradigms emerging from sciences studying the nature of the environmental crisis, a proposal that continues to gain importance. According to Toledo, at the end of the twentieth century, ethnobotany in Mexico was transitioning from a “normal” stage of exploring new plant products for industry and other purposes, and analyzing the role of plants in the material and cognitive culture of humans. The transition to a new ethnobotany, according to Toledo, was fostered by ethnobotanists with critical views, for whom ethnobotany was not a neutral science but a research field with social implications that should lead to strong commitments with the indigenous and rural communities they worked with. Toledo envisioned ethnobotany as a research field with political implications, integrating social, economic, ecological, and political problems of rural communities. He examined the ways that the documented botanical and ecological knowledge might contribute to support local peoples’ goals and concerns. Stronger and explicit engagement with communities rather than documenting and extracting local knowledge should be a main feature of a “post-normal” ethnobotany, according to Toledo. The subsequent development of the field partially followed these premises, resulting in the emergence and development of agroecology and participatory research, among other approaches. Several authors have investigated the history of ethnobotany and ethnobiological sciences. Clément (1998) reconstructed the emergence of these fields from botany and anthropology, reminding us that ethnobotany developed in the context of the discoveries during the colonial expansion, a period marked by both disparagement of indigenous knowledge and the myth of the noble savage. However, this disparagement did not blind people to the advantages of prospecting plants, animals and other organisms, for benefiting the “more civilized world” and industry, as did other scientific disciplines. Clément (1998) described the first stages of ethnobiological sciences as part of the drama of colonialism and its mindset, discrediting all views, knowledge and ways of life different from the European ones, but seizing the opportunity to assimilate what was considered useful. This raiding of natural resources, including plants and animals, became a central issue in the Earth Summit of Río de Janeiro in 1992 and in subsequent meetings (Sánchez et al. 2019). The issue currently continues to be an important concern for ethnobiologists. This new perspective induced more ethical and integral views on the meaning of ecosystems for understanding local cultures and vice versa, as societies and biodiversity in interaction. Ethnobiologists became interested in working in favor of local communities and people, cooperating to resolve their problems, ensuring their rights, and supporting their views and projects for improving their lives. Also, ethnobotanists contributed to make visible the huge relevance of traditional cultures for building a sustainable future of the planet.

958

A. Casas et al.

Soon after Clément’s work, Miguel Alexiades (2003) highlighted the inherent conflict between Western views and systems of protection of intellectual property on the one hand, and traditional knowledge and useful plant products on the other. What was originally considered public goods became privately appropriated, and sometimes patented, clearly taking advantage of the diverse local knowledges (Alexiades 2003). Alexiades as other authors (see Brush 1993; Brush and Stabinsky 1996) suggested that the commoditization and politicization of genetic resources and local knowledge required fair systems of sharing knowledge and benefits. This subject continues to be a main concern in ethnobiological research. It is a complex issue for which there are no simple answers. As shown in several chapters of this book, people in traditional communities as well as in modern contexts, exchange products, knowledge and techniques within and between families, communities, and regions. This pattern allows visualizing that traditional knowledge is in reality highly dynamic and continually adapts to changing circumstances (Berkes et al. 2000). It includes processes of generation, conservation, and decrease of knowledge as documented by Hart and Salick (2017), or retention, erosion, adaptation, and hybridization, as proposed by Sharifan et al. (2022). The Western societies have frequently taken advantage of the ways of sharing and reciprocity norms in traditional societies. The Western property rules are commonly meaningless in the traditional communities’ context. What is the solution to this problem? There is no simple answer. If we leave traditional knowledge outside of the intellectual property regime, it will be taken advantage of and perhaps appropriated by various actors. On the other hand, implementing intellectual property for traditional knowledge requires identifying persons and/or collectivities to benefit from this asset or knowledge, which is a difficult task. It may cause conflicts and hinder the continuous improvement and adaptation which is a feature of much traditional knowledge or managed organisms. Bioprospection based on traditional knowledge and biopiracy by patenting products based on that bioprospection are both real. Compensation or retribution systems may be the fair way to legitimize the appropriation of both plant products and knowledge. But, compensate whom? Identifying individual innovators of current processes might be possible, although innovations in communities are rapidly diffused, tested, and subject to new innovations. The process of innovation is per se a topic of research (Blancas et al. 2010; Rangel-Landa et al. 2016). But identifying discoverers of plant properties or inventors of management techniques and the preparation of a plant used for long time periods is practically impossible. Traditional knowledge and experience are eminently collective, but collectivities are not restricted to a community or a region, their geography is rather complex. Contributing to developing fair systems of retribution continues to be a challenging topic for ethnobiologists, as well as for various social groups and organizations, governments, anthropologists, biologists, and ecologists. Alexiades points out that validating traditional knowledge has been one of the main concerns of ethnobotanists and continues to be so. Validation of local ecological knowledge was galvanized from the mid-twentieth century onwards, with intensifying pharmacological studies of traditional medicines, searching for new

Perspectives of the Ethnobotanical Research in Mexico

959

active principles in medicinal plants, with some success. However, success was uneven, as numerous ailments recognized by traditional practitioners cannot be understood with the principles of Western medicine (Plotkin 1993; Prance et al. 1994). In contrast, the search for new and improved edible plants in traditional contexts did have a good reputation as the crops sustaining the economy of the world originated there. Prospection was directed at interesting landraces of the main crops, crop wild relatives (Heywood et al. 2007), orphan crops, and important neglected and underutilized wild and weedy plants with promising value (Ulian et al. 2021). The called Green Revolution began to show its failures in the 1960s and 1970s in numerous local contexts. The standardized formulas of the technology spread from institutions through extensionists to rural people and local environments; however, they encountered a complex world. The outstanding work by Paulo Freire “Extension or Communication?” (Freire 1973) summarized the central reasons of the failure: (i) local agriculturalists were not passive receptors of technologies designed in laboratories and experimental fields; (ii) the local environmental and cultural contexts could not be incorporated into the standardized technologies; (iii) the local farmers knew better than technicians, promoters, and extensionists what could and should be done in the local contexts; and, among other issues, (iv) the “improved” hybrid varieties produced, patented, and commercialized as part of the programs not only generated dependence of their users but were not always productive or profitable; (v) the “improved” varieties favored displacement of local varieties, thus contributing to a loss of genetic variation historically shaped by local cultures, a process that in the 1970s was recognized as genetic erosion (Frankel and Benett 1970; Brush 2004). In Mexico, during the heyday of the Green Revolution, some agronomists, outstandingly Efraim Hernández-Xolocotzi (1959; HernándezXolocotzi and Ramos-Rodríguez 1977; Hernández-Xolocotzi et al. 1980), recognized that local knowledge and techniques were more effective than the modern technologies in traditional contexts, and these should be studied and understood. All these conclusions strongly reinforced the Paulo Freire vision. Hernández-Xolocotzi based his criticism on profound research of traditional agriculture. His vision propelled Mexican ethnobotany and established the bases of Mexican agroecology. Ethnobotanists exhibited evidence of the deficiencies of mainstream agricultural science, and also the extraordinary value of traditional knowledge for finding solutions. The same trajectory can be observed in environmental sciences. The contribution of local traditional knowledge to the solution of numerous and complex issues was increasingly recognized. The collaboration of different sectors, most importantly local managers, are crucial for attending long term problems (Grumbine 1994). The boost of the concept of traditional ecological knowledge (Berkes 1993, 1999), the governing of the commons (Ostrom 1990), agroecology (Altieri and Toledo 2011), and the emergence of sustainability science (Kates et al. 2001, 2011) became conceptual frameworks integrating local people’ knowledge in the design of sustainability (Berkes et al. 1994; Nelson and Shilling 2018). The

960

A. Casas et al.

transdisciplinary and participatory approaches of science as discussed below are based on this framework. Alexiades (2003) analyzed the challenges imposed by the global environmental crisis and the potential role of ethnobotany and ethnobiological sciences in interdisciplinary and intercultural programs, as well as participatory approaches of science. Local knowledge is much more than certain names and uses of organisms that industry can potentially use. It involves whole systems of knowledge, worldviews, and an ample diversity of forms of interaction (Berkes 1993, 1999; Berkes et al. 2000; Toledo and Barrera-Bassols 2008). In his reflection about main challenges to ethnobotany, Alexiades (2003) raised two important questions: (1) “How to construct, articulate and operationalize the relationship between different knowledge systems, actors, needs and views in the context of the inter-cultural and interdisciplinary (and we would add transdisciplinary) dialogue?”, and (2) “What opportunities and challenges lie beyond the rhetoric of participation and interdisciplinarity that permeates much of development, conservation and environmental scholarship?” These questions continue to be influential. In the last three decades, scholars, NGOs, and other sectors together with local people have developed numerous initiatives (Illsley-Granich et al. 2004, 2007; Bebbington 2007; Casas et al. 2017). One important result of that process is the recognition of the value of local knowledge and technical experience for developing better ways of interacting with biodiversity and building sustainable management systems (Nelson and Shilling 2018). The other result contributed avenues for participatory, interdisciplinary, and transdisciplinary processes beyond the rhetoric. There is no one single way, but initiatives in different regions of the world have created several areas of interactions: agroecology (Altieri 2002), non-timber forest products management (Illsley-Granich et al. 2004), agroforestry, ethnoagroforestry (Moreno-Calles et al. 2013, 2016), biocultural landscape conservation (Hong et al. 2014; Barrera-Bassols and Floriani 2018), food sovereignty, eco-technologies, governance of the commons (Ostrom 2007), sustainable food systems, resilient systems (Gunderson and Holling 2002; Walker and Salt 2012), among others. It is not our intention to review here all these approaches, issues and processes, which deserve their particular own analysis. But members of the academic community collaborating in social movements are studying and systematizing information at a global scale. For example, the International Forestry Resources and Institutions (IFRI 2013), founded by Elinor Ostrom, examines the ways people from different regions of the world reach agreements around forest management and their products with formal and informal institutions. Their aim is to provide policy makers criteria based on rigorous evidence-based experiences. Another important global initiative is the Resilience Alliance, which since 1999 conducts international research on the dynamics of socio-ecological systems, from multidisciplinary perspectives. The members of this organization document and systematize information, and conduct comparative research and synthesis of experiences (Resilience Alliance 2010). They have significantly contributed to the IPCC, the Millennium Ecosystem Assessment and the Future Earth projects, and have made extraordinary contributions to constructing frameworks on the sustainability of socio-ecological systems

Perspectives of the Ethnobotanical Research in Mexico

961

(Gunderson and Holling 2002; Walker and Salt 2012; Carpenter et al. 2012; Walker 2020). Other organizations are alliances of different sectors of the society. Probably the most common and successful are those of communities and social organizations, NGOs, and scholars, most commonly ethnobiologists, agroecologists, ecologists, and anthropologists (see for instance Illsley-Granich et al. 2007; Casas et al. 2017; Gavito et al. 2017). They often emerge as initiatives to address concrete environmental problems associated to mining, deforestation, over-exploitation of specific resources, dispossession of land and/or water, soil erosion and desertification, drug trafficking, and organized crime, among others. The local organizations commonly look for the collaboration of scholars and universities to obtain information to solve these problems. After initial contacts, the links and collaborations increase and search for appropriate ways of communication, ethical frameworks, and compromises. These developments have pushed the interdisciplinary, transdisciplinary, and participatory approaches beyond the rhetoric. However, the interactions are still often improvised, and a better understanding of the process is urgently needed. Some chapters of this book recount direct experiences on these issues; we will return to this topic further on. Another valuable reflection on challenges of ethnobotany was recently published by Albuquerque et al. (2019), based on a previous thinking by Albuquerque and Hanazaki (2009). The authors lay out some questions/issues they consider to be the most relevant challenging ethnobotanical research. One of them refers to how to adequately represent the views and knowledge of traditional people. Other subjects are about the appropriate methodological approaches to assess the continuous biocultural changes; the role of plants in socio-ecological systems and the mechanisms of intergenerational transmission of knowledge; the role of ethnobotany in supporting processes of biodiversity governance; the dialogue of ethnobotany with other sciences; the effect of human migration on distribution of plants and the effect on local flora; the properties of knowledge and plant composition in urban contexts and their influence on the surrounding areas; the effects of extraction of medicinal plants and other used plants, and how ethnobotany should interact with ecology to recommend sustainable ways of using these plants. All these are relevant issues and require significant effort to find solutions. Sustainability science is a general challenge for science and has been treated by several authors, outstandingly Kates et al. (2001, 2011), Swart et al. (2002), among others. The theoretical constructions associated with the concept of sustainability, such as resilience of socio-ecological systems (Holling 2001; Gunderson and Holling 2002), governance and institutions (Ostrom 1990, 2007, 2009), among others, have seen a renewed momentum during the first decade of this millennium. All these reflections emphasize that reducing the impact of the global crisis is still possible but requires multiple perspectives, based on the consideration that social and ecological systems are mutually inter-dependent and complex, with emerging properties, nonlinear responses, and high uncertainty (Young et al. 2006; Weible et al. 2010; Costanza 2014). Environmental problems require the interaction of natural and social sciences (Kates et al. 2001, 2011; García 2006, 2011; Casas

962

A. Casas et al.

et al. 2016b). Additionally, sustainability science has among its premises the need of science to interact with knowledge and technological experiences developed by local people and other sectors of society through transdisciplinary approaches (Kaufman et al. 2003; Brandt et al. 2013). Adaptive management is another relevant concept. It recognizes that the complexity of socio-ecological systems and their nonlinear behavior make it difficult if not impossible to predict the exact response of the system after interventions (Allen et al. 2011; Keith et al. 2011; Rist et al. 2013; Allen and Garmestani 2015). Therefore, the proposal emerging from this concept is to implement management practices based on the best and deepest knowledge and techniques available but considering them as provisional interventions. Then, the response of the system is evaluated, the effectiveness or failures of the practices analyzed, and a new stage of management initiated by adjusting the actions based on learning from the previous stage (Allen and Garmestani 2015). This concept recognizes that the contemporary scientific and technological systems are not infallible, and that local knowledge may have valuable solutions tested in practice. Ethnobotany is also concerned with the loss of languages, cultures and institutions (Gibson et al. 2000; Eberhard et al. 2022). Language, knowledge, worldviews, management practices, technologies, and institutions are all integrated in culture and their loss means the loss of human experience to solve environmental problems. Ecology, ethnobiology, and anthropology can build bridges to promote dialogues between different social sectors, and thus contributing to a better understanding of sustainability. Another main challenge of ethnobotany and ethnobiological sciences is how to consolidate their theoretical and methodological frameworks and integrate them with more general frameworks and emerging paradigms such as sustainable socialecological systems, biocultural diversity conservation, and restoration.

Methodological Approaches Studying interactions between humans, plants, and ecosystems may involve different types of questions, some of them answerable through quantitative methods, some others requiring qualitative approaches. One is not better than the other nor using one or the other makes ethnobotany a more authentic science. Methods are ways, systematically used by science, to answer questions and understanding phenomena. Quantitative approaches evaluate relations between variables, representativeness, and generalizability of phenomena, whereas qualitative methods help to analyze problems that are difficult to measure but necessary to understand the internal functioning of socioecological systems. Ethnobotany needs both approaches as discussed below. Documenting traditional knowledge. Ethnobotany requires methods for understanding traditional botanical and ecological knowledge, its connections with technologies and social organization and its ecological, economic, and evolutionary consequences. Berkes (1993) defined some general characteristics of traditional

Perspectives of the Ethnobotanical Research in Mexico

963

ecological knowledge: (i) it is mainly qualitative; (ii) it has an intuitive component; (iii) it is holistic; (iv) mind and matter are considered together; (v) it is moral (as opposed to supposedly value-free); (vi) it is spiritual (as opposed to mechanistic); (vii) it is based on empirical observations obtained by trial-and-error (as opposed to experimentation and systematic, deliberate accumulation of information); (viii) it is based on data generated by resource users themselves (not by specialized researchers); (ix) it is based on diachronic data, i.e., long time-series on information at one locality (as opposed to synchronic data, i.e., short time-series over a large area). These characteristics and others confer ethnobotany and other ethnobiological sciences the need for broad, flexible, adaptive, and contextual methodological approaches. Local systems of knowledge, worldviews, practices, and technologies are contextualized by the local social relations and particularities of ecosystems, all of which are dynamic. There is no single way to understand these systems. Albuquerque et al. (2019) suggested a multiple-evidence base (MEB) research, referring to a reflection by Tengö et al. (2014). Both qualitative and quantitative approaches are needed to understand these complex systems. Qualitative approaches are good windows for exploring the multiple relations and interconnected views within the systems whereas quantitative approaches help to test more specific hypotheses. Qualitative methods allow access to the meanings peoples give to processes, the quantitative approaches inform on the representativeness of ideas or practices in communities at different scales. Ethnobotanists have been criticized for their descriptive work which is common at the early stages of a research field, and in the last decades they have widely adopted and developed quantitative tools, partly to legitimize ethnobotany as a scientific field. However, it has now been recognized that the interactions between humans and plants are expressions of complex systems. Specific interactions may be better understood through quantitative approaches, but the study of complexity requires multiple approaches, including qualitative examination (Castillo et al. 2020). In social-ecological systems, documenting and analyzing the views (understood as the perceptions, feelings, and meanings to phenomena) of local people and the different sectors of the society are valuable sources of information for understanding problems and, therefore, are crucial bases for developing suitable actions. Thus, qualitative research are not only scientific but necessary to assess complex systems. In ethnobotanical research, quantitative research is mainly based on surveys, and qualitative studies are mainly based on in-depth interviews, participant observation, and other instruments. Surveys and in-depth interviews provide different kinds, but complementary information. One method is not better than the other; every method should be chosen according to the questions the researcher wants to answer. According to Drury et al. (2011), external validity evaluates if survey results can be generalized; it requires that the data be representative of the phenomena studied. Sample sizes can be large. Internal validity refers to representing “the diversity of individuals and groups being analysed and examining complex concepts in ambiguous and complex contexts.” In these cases, surveys are not commonly able to include locally important categories and other aspects defining the complexity, and,

964

A. Casas et al.

as the authors state, “Research investigating what people feel, think, plan and do commonly depends on asking respondents about their views or actions,” which are difficult to be caught through surveys. As Drury et al. (2011) state: Natural scientists accustomed to concentrating on external validity may perceive samples of respondents in qualitative studies as too small or inadequately selected to represent wider populations. But the fact that one cannot make statistical statements based on qualitative data does not render the findings invalid. Conversely, a large sample, with all informants receiving the same standardized research tool, does not automatically yield good data if that tool or sample is poorly designed or applied.

In other words, qualitative approaches may help to understand relationships, thinking, views, and other complex aspects in much more depth, and may facilitate the design of appropriate questionnaires for surveys. Therefore, qualitative and quantitative research techniques are commonly complementary, providing different types of information. Inventorying and systematizing the Mexican biocultural heritage. The documentation of the heritage of nomenclature, classification, use, and management of biodiversity has been the foundation of ethnobotanical research. This research avenue is helpful for many different types of studies. In the past this activity was undervalued as merely descriptive. However, having a complete inventory of the local biodiversity, the ecosystems where it occurs and the human experience of using and managing it constitutes valuable information for different types of analyses, in the same way that floristics is essential for botanical disciplines such as systematics and biogeography. Although this has been the dominant approach, the inventory is still incomplete for Mexico, as pointed out in chapter 2 by Caballero et al. (2022). Some estimations suggest Mexican people may use more than 11,500 vascular plant species, whereas databases contain information of less than 8000 species (Caballero et al. 2022). But much more is needed. First, not all ethnobotanical information recorded in the literature has been systematized. Also, uses and management forms of a species may differ from region to region, and the information for species is incomplete. It should be increased substantially by ethnobotanical studies in poorly explored regions. Nearly half of the main cultural groups of Mexico have been poorly or not studied, and some regions and vegetation types have been more studied than others. Based on the diagnosis of chapter 2 and the study by Camou et al. (2016) we identify in Table 1, those states, regions, cultural groups, and ecosystems of Mexico that require more research efforts. Most ethnobotanical information stored in databases is about plant uses, but inventorying uses is insufficient. We need to understand the nature of some uses and medicinal and nutritional aspects better. The effectiveness of traditional medicine, the nutraceutical properties of edible plants and the nutritional aspects of plants used as food and fodder, are examples of recurrent societal demand of information. Such information would substantially strengthen the role of traditional knowledge in both health and food systems. Management techniques, spatial information on distribution and abundance of the species, and biocultural aspects require more

Perspectives of the Ethnobotanical Research in Mexico

965

Table 1 Priority regions and cultural groups for ethnobotanical research in Mexico. The table enumerates regions and cultural groups with limited ethnobotanical studies of twenty states. The levels of relative priority (in italics in the last column) are the average of the information scarcity ranking proposed by Camou et al. (2016) and a ranking based on the number of records in the database BADEPLAM (Caballero et al. 2022). Numbers closer to 1 indivate higher level of priority for conducting studies. Regions and ethnic groups in bold are those considered exceptionally important in each state State Querétaro Colima Baja California

Region • Amealco • Tolimán • Comala • Ensenada • Mexicali • Tecate

Sinaloa

• Northern region • Fuerte River

Aguascalientes Guanajuato

• Western mountain region • Chichimeca, sierra de Guanajuato • Sierra Gorda • Mezquita • Buenaventura, • Jícoras • Mezquital • Guanaceví • Ocampo • San Bernardo • Melchor Músquiz • Valparaiso • Fresnillo

Durango

Coahuila Zacatecas

Campeche

• Throughout the state • Champotón • Edzná • Campeche • Eastern part of the state

Tlaxcala Jalisco

• Throughout the state • Mezquitic • Bolaños

Ethnic group • Otomí

Relative research priority (6, 1) ¼ 3.5

• Nahua • Ku’ahl • Pa ipai • Cochimí • Kiliwas • Cucapá • Kumiai • Yorome-Mayo • Tepehuan • Tarahumara • Nahua • Chichimeco Jonaz • Otomí • Tepehuanos del Sur • Nahua (Mexicaneros) • Huichol • Tarahumara

(10, 2) ¼ 6 (9, 4) ¼ 6.5

• Kikapú • Tepehuanos del Sur • Huichol • Maya • K’ich • Awakatec • Akatec • Chuj • Jakaltec • Kaqchikel • Mame • Ixil • Q’echi • Ch’ol • Nahua • Huichol

(5, 13) ¼ 9 (14, 5) ¼ 9.5

(2, 12) ¼ 7 (3, 11) ¼ 7 (1, 14) ¼ 7.5 (8, 7) ¼ 7.5

(12, 9) ¼ 10.5

(15, 6) ¼ 10.5 (17, 8) ¼ 12.5 (continued)

966

A. Casas et al.

Table 1 (continued) State Sonora

Region • Low Mayo River • Upper Mayo River • Lower Yaqui River • Sonoran Desert • Yécora • Colorado River • Bacerac

Chihuahua

• Guachochi • Guadalupe y Calvo • Madera • Uruachi

San Luis Potosí

• Eastern part of the state • South-eastern part of the state • Río Verde • Ciudad del Maíz • Arroyo Seco • Southern part of the state

Tamaulipas

Quintana Roo

• Maya Balam • Miguel Hidalgo • Kuchumatán • Throughout the state

Chiapas

• Amatenango • Las Margaritas • Mazapa • Motozintla • Ocosingo • San Cristóbal • Rayón • La Trinitaria

Nuevo León

• Southern part of the state

Ethnic group • Seri • Pima • Tohonó-Oódham • Cucapá • Ópata • Mayo • Guarijío • Kikapú • Yaqui • Tarahumara • Tepehuanos del Norte • Pima • Guarijó • Huastec • Nahua • Pame • Nahua • Huastec • Totonac • Akatec • Chuj • Ixil • Q’anjobal • Jakaltec • K’ich • Kaqchikel • Mame • Q’anjobal • Q’echi • Maya • Jakaltec • Kaqchikel • Mame • K’anjob’al • Tojolabal • Teko • Mochó • Tzeltal • Lacandon • Tzotzil • Zoque • Chuje • Huastec

Relative research priority (4, 23) ¼ 13.5

(7, 25) ¼ 16

(13, 21) ¼ 17

(19, 15) ¼ 17 (11, 27) ¼ 19

(18, 29) ¼ 23.5

(20, 28) ¼ 24

Perspectives of the Ethnobotanical Research in Mexico

967

emphasis. Ecological information on life history traits and interactions with other species should be incorporated into the ethnobotanical fact sheets as it is relevant for designing management plans. Ethnobotanists with a background in the biological sciences have the necessary skills to obtain them. Detailed information on the cultural, economic, and relational values of the species people interact with is extraordinarily important. Anthropologists, economists, biologists, and professionals trained in interdisciplinary programs have the capacities to document these aspects. Inventories of biocultural heritage also support the construction of theoretical and methodological frameworks as we discussed before. Therefore, inventories of biocultural information continue to be necessary. However, we stress that some aspects have been covered poorly. There are numerous efforts throughout the country constructing local or regional databases. The construction of a national database of ethnobotanical (and, in general, ethnobiological information) requires efforts to design and support ad hoc formats and the disposition to share the information. Establishing clear and fair rules of construction, operation and use, assigning the responsibilities of the curatorial work, the most appropriate institutional seat, and the coordination of work are priorities. After decades of these efforts, ethnobotanists and ethnobiologists must take steps toward more robust collaborative activities to systemize information at a national scale. Historical reconstruction approaches. Historical perspectives of culture and interactions with plants require collaborations with archaeological and ethnohistorical studies, as well as documental research in historical archives. Archaeological studies have advanced in documenting plant remains associated to ancient humans and thus reconstruct their historical interactions. Some classical studies referred to throughout this book are those conducted by MacNeish (1967, 1992), Flannery (1986) Smith (1997), Zeder (2017); Piperno and Pearsall (1993); Piperno et al. (2009); McClung-de Tapia et al. (2001); and Acosta-Ochoa (2008). Methods for identifying micro and macrofossils have improved and the combination of dating methods have increased precision. The recent and outstanding field of paleogenomics has increased accuracy in the identification of remains, as well as domestication imprints (see Lindqvist and Rajora 2019). By contributing evidence of the prehistoric interactions between humans and plants, the origin and diffusion of cultures, a number of views have changed recently. Maize, beans, peppers, squashes, and cacao have had much more complex interactions between North, Central, and South America than expected (see for instance Pease et al. 2016; Zarrillo et al. 2018; Kistler et al. 2020). Also, the earliest date of human occupation of the Americas has been pushed back repeatedly. Recently, Ardelean et al. (2020) dated human presence in the Chiquihuite Cave in Zacatecas as 24,000 years ago and probably earlier. Ethnohistorical and historical sources have been reviewed by several authors (see for instance Bye and Linares 2016; Camou et al. 2016). These extraordinary written and pictographic documents enrich the understanding of past and current patterns of interactions between humans and plants. This approach has a long tradition in Mexico and continues to provide interesting information and novel insights. History reconstruction of the last century is commonly possible through interviews with the oldest people in a community. Timelines are highly useful tools for

968

A. Casas et al.

analyzing aspects of ecosystem management and resilience. Life stories help to understand patterns of wider historical, social, environmental, and political contexts (Adriansen 2012). Interviews can be complemented with other sources of information, thus combining qualitative and quantitative sources. For instance, chronological analysis of a life history can be complemented with bibliographic information or statistical data. Several helpful software tools for qualitative analysis and specifically for timeline analysis are available. Ecological ethnobotany and sustainable management. The integration of ecological principles in ethnobotanical studies have resulted in new approaches and insights into the use patterns of forests or agricultural systems. Several authors have called this integration ecological ethnobotany (Caballero et al. 1998, 2022; DelgadoLemus et al. 2014; Camou et al. 2016). A number of studies from the last decades analyzed use patterns and population management, and their impact on genetic variation (Cruse-Sanders et al. 2013; Félix-Valdéz et al. 2016; Paz-Guerrero et al. 2019; Cabrera-Toledo et al. 2019; Alvarez-Ríos et al. 2020) and demographic rates (Martínez-Ballesté et al. 2005, 2006, 2008; Torres-García et al. 2015, 2020). Many of these studies tried to identify management forms that conserve genetic diversity and sustainable harvest rates based on local practices, through matrix analyses and integral projection models (see Torres-García et al. 2015, 2020). These approaches have been implemented with and by traditional communities for non-timber forest products (Illsley et al. 2004). Ecological methods for evaluating sustainability also include the biotic communities. They are particularly relevant for measuring species diversity (see PérezNegrón and Casas 2007) and interactions (pollination, facilitation, seed dispersal, herbivory; see for instance Torres-García et al. 2013; Rangel-Landa et al. 2015). These are affected by human practices and show pathways to ensure the permanence of the communities. Guides for the management of whole ecosystems are developed with multicriteria methods. These methods are based on the identification of critical points and attributes of the systems. Synchronic approaches explore the effect of interventions and design diachronic strategies of adaptive management. MESMIS is one of the most commonly used and highly versatile methods to analyze sustainability of forest, agroforestry, and agricultural systems (Masera et al. 2000). Evolutionary ethnobotany, domestication, and agrobiodiversity. Evolutionary ecology and relational anthropology are highly valuable for understanding interactions between humans and plants, management and domestication. These approaches are fundamental to understand the development of agrobiodiversity, including crops and non-crops. Evolutionary ecology documents the evolutionary consequences of human-plant interactions whereas relational anthropology and ethnographic studies describe and analyze how the interactions are. In Mexico, studies of plant management and domestication have increased during the last three decades. They show how the broad spectrum of management types of plant populations and communities influence the frequency of phenotypes and species in an area. This information is relevant to understand the human influence in molding landscapes and populations and their mutual influences (see Casas et al. 2016a).

Perspectives of the Ethnobotanical Research in Mexico

969

Ethnobotanical approaches identify useful species for different purposes, and their different qualities and values for people. These values (relational, cultural, and economic) drive human interactions with the species and communities. People create and identify phenotypic variation in attributes like size, color, texture, flavor, toxicity, and other features that enhance or limit use of plants as food, medicine, textiles, handcrafts, and others (Casas et al. 2007; Aguirre Dugua et al. 2012). They encourage favorable phenotypes through different forms of management and discourage others. These selective processes may act in several directions and with different intensity, thus influencing different degrees of domestication. But people also influence these processes through intervening on how the gene flow occurs. They move reproductive individuals from place to place, thus removing spatial barriers to pollination; also, they commonly move seeds or vegetative propagules from distant regions to others, creating reproductive bridges that influence plant evolution under domestication (Cruse-Sander et al. 2013). They also frequently create small, isolated populations which favor genetic drift. All these processes are also influenced by natural evolutionary forces (selection, genetic drift, breeding systems, and gene flow), which operate together with those guided by humans. By comparing wild and managed populations, we can characterize and evaluate the results of these processes that involve morphological, physiological, phytochemical, reproductive biology, and genetic variation in populations (Casas et al. 2016b). These approaches allowed identifying that some plants represent advanced stages of domestication, but plant populations exist on a continuum between wild and domesticated, depending on the type and intensity of management, but also on life cycle and other biological attributes of plants (Blancas et al. 2010, 2013; Rangel-Landa et al. 2016). New methods allow to measure morphological details of plants, or to monitor movements of pollinators and seed dispersers by telemetry. Techniques for the isolation, characterization, and identification of phytochemical compounds have increased their capacities; genomics and metabolomics are methods more rapid and efficient for screening chemical constituents of plant tissue. Chromatographic and spectroscopy methods (ultraviolet, infra-red, mass spectroscopy, and nuclear magnetic resonance) have modernized and recent methods like matrix-assisted laser desorption, electron impact, chemical ionization, atmospheric pressure ionization, among others, broaden the menu of options to characterize the compounds occurring in plants (Olufunke 2012). Advances in genomics are particularly relevant for strenghtening previously developed approaches like population genetics, phylogenetics, and phylogeography (Gepts 2014; Kantar et al. 2017; Chomicki et al. 2020). The level of resolution reached with single nucleotide polymorphism (SNPs) is extraordinary. Metagenomics helps to identify assemblages of organisms and their trends through time and strengthen the understanding of patterns related with human management. The capacity of genomics to identify coding regions and their relation to domestication traits is equally impressive (Zarrillo et al. 2018; Chomicki et al. 2020). Until a couple of decades ago, identifying coding regions required different approaches from quantitative genetics to field experiments to determine the heritability of characters. This approach helps to study evolutionary aspects of the most important crops of the

970

A. Casas et al.

world (about 150 species), but the number of species under domestication is much higher (more than 7000) and much work is still needed. These methods will contribute to their understanding faster than until now. Also, these methods are corroborating or correcting the hypotheses on the evolutionary influences of humans through the broad spectrum of interactions documented by ethnobotanical studies. Food systems and sovereignty. Ethnobotany contributes substantially to food sciences and food anthropology. Its holistic studies on traditional food and food security can support programs for food sovereignty. Ethnobotanical studies in Mexico have documented more than 2000 edible plant species (Mapes and Basurto 2016) and they are part of food systems. Most crops or managed plants are edible (Clement et al. 2021). Clearly, the main efforts of domestication have been directed to food plants in Mexico and elsewhere (Casas et al. 2022). Agrobiodiversity studies concentrate on edible plants and these studies are intimately connected with the study of diet patterns. Several chapters of this book treat problems related to food patterns and risks to food sovereignty. There are particular contexts in all cases but the trend to abandon traditional food and adopt industrialized food is clear and worrying. Multiple factors influence this complex situation. The urbanization of people migrating to cities of Mexico and the USA, which is commonly related to abandonment of agricultural practices and traditional ways of life. In addition, the food producing companies promote their products as a gateway to higher status. Also, the rural economy is becoming monetized, and people have much less time for agriculture and cooking. Another factor is racism and cultural discrimination, which rural people try to escape by adopting urban habits (Casas et al. 1994). Traditional food is tied to cultural identity and is a target of discrimination. Ethnobotanical studies commonly report that people omit to mention traditional food such as quelites or insects during interviews, considering them worthless or embarrassing, or refer to these elements as part of the diet of poor, indigenous, mountain people. Several social organizations, NGOs collaborating with ethnobotanists, anthropologists, and agroecologists, have been promoting traditional food with a certain amount of success. Some governmental research initiatives (like the Agrobiodiversity project, and others related to the use of biodiversity by CONABIO) have helped also, but there is still much to do. Local, regional and national collaborations linking research with promotion of traditional food are possible and necessary. Mexican ethnobotanists should include this priority line of research and action in their agendas. Transdisciplinary research and participation. As mentioned above, ethnobotanists and other scholars have been influenced by social movements defending their biodiversity and ecosystems, territory and ways of life, as well to those requiring innovations in management systems, internal organization and community institutions, public policies, development of educational programs, among other issues. All these experiences have represented different forms of establishing dialogues and interactions and participatory processes (Castillo et al. 2005, 2018). The modalities of all these ways of cooperation are numerous and require a careful systematization to be able to learn and use them to develop or reinforce current and future processes. These collaborations are based on: (1) the recognition of the importance of

Perspectives of the Ethnobotanical Research in Mexico

971

Fig. 1 General outline of inter and multidisciplinary collaboration of ethnobotany with other sciences when studying a shared problem

knowledge and experience of other participants; (2) respect and empathy with persons representing different sectors; (3) the disposition of each sector to acknowledge their own limitations; (4) the proclivity to imagine the actual and potential value of knowledge and experience of others; (5) the capacity to listen to the others and to transmit the own knowledge and skills; (6) the courage to put into practice tentative experiments to explore adaptive management processes; (7) the disposition to learn from the experience; (8) the capacity to communicate the failures and successes of the process to other members of the cooperating communities. Interaction networks based on these principles can deliver novel and helpful advances in rural systems. To consolidate ethnobotany as a research field, ethnobotanists should strengthen their ties with other academic disciplines (Fig. 1). Ethnobotany is connected with all sciences illustrated in the figure and others. It has its own domains but is supported and advanced by methodological and theoretical frameworks developed in those other disciplines. Interaction with other sectors is also needed (Fig. 2), and transdisciplinary approaches base on the recognition of the value of knowledge and experience of all these sectors. Each sector may in turn be composed by subsectors (S1, S2, S3, etc.) dialoguing within and between sectors. For instance, the rural communities may be composed by ejidatarios, comuneros, agriculturalists, ranchers, foresters, men, women, young, and elder people, among others. Scholars may be ethnobotanists, anthropologists, agronomists, ecologists, and others. NGOs may be regional, national, and international. Entrepreneurs may be merchants specialized in non-timber forests and/or agricultural products, those trading handcrafts or businesses purchasing rural products. Governmental agencies may involve those making decisions on conservation, forest management, land tenure, hunting and fishing, and biosphere reserve regulations.

972

A. Casas et al.

Fig. 2 Exchange of knowledge, experiences, needs, demands, visions, and techniques pertinent among different sectors related to biodiversity management

An example of these interaction networks is that built by the NGO Grupo de Estudios Ambientales (GEA for its acronym in Spanish) the Sanzekan tinemi (an organization of rural communities of the central region of the state of Guerrero) and scholars of different specialities (engineers, biologists, anthropologists, geologists, sociologists, environmental scientists, economists, among others). They have worked for several decades in programs of soil and water conservation, environmental restoration, sustainable management of mescal agaves and palms, and others. GEA has based its actions on the respect of communitarian institutions and has been propellant of initiatives to invite the scholars to collaborate (Illsley et al. 2004, 2007; Casas et al. 2017).

Final Comment Theoretical and methodological frameworks are research tools under continual revision in all sciences and ethnobotany is not the exception. The trends discussed in this chapter suggest reinforcing the efforts to strengthen the theory and generalizations on the broad spectrum of topics covered by ethnobotanical research. This is not only of scholarly importance but, as discussed above, highly relevant for guiding human activities toward a sustainable world. Acknowledgments The authors thank financial support from CONACYT, Mexico (project A1S-14306), the GEF Project ID 9380 770 CONABIO-GEF-FAO/RG023 “Manejo y domesticación de agrobiodiversidad en Mesoamérica: Bases para la soberanía alimentaria sustentable,” and PAPIT, UNAM (project IN206520 and IN224023). We also thank Alicia Castillo for her critical review and feedback.

Perspectives of the Ethnobotanical Research in Mexico

973

References Acosta-Ochoa G. La cueva de Santa Marta: y los cazadores-recolectores del Pleistoceno final Holoceno temprano en las regiones tropicales de México. PhD Thesis. Doctorado en Antropología. Mexico: Universidad Nacional Autónoma de México; 2008. Adriansen HK. Timeline interviews: a tool for conducting life history research. Qual Stud. 2012;3(1):40–55. Aguirre-Dugua X, González A, Eguiarte L, Casas A. Round and large: morphological and genetic consequences of artificial selection on the gourd tree Crescentia cujete by the Maya from the Yucatán peninsula, Mexico. Ann Bot. 2012;109:1307–16. Albuquerque UP, Hanazaki N. Five problems in current ethnobotanical research – and some suggestions for strength. Hum Ecol. 2009;37:653–61. Albuquerque UP, Nascimento ALB, Soldati GT, Feitosa IS, Campos JLA, Hurrell JA, Hanazaki N, Medeiros PM, Silva RRV, Ludwinsky RH, Júnior WSF, Reyes-García V. Ten important questions/issues for ethnobotanical research. Acta Bot Brasilica. 2019;33(2):376–85. Alcorn J. Huastec Mayan ethnobotany. Austin: University of Texas Press; 1984. Alexiades MN. Ethnobotany in the third millennium: expectations and unresolved issues. Delpinoa. 2003;45:15–28. Allen CR, Garmestani A. Adaptive management. In: Allen C, Garmestani A, editors. Adaptive management of social-ecological systems. Dordrecht: Springer; 2015. p. 1–10. Allen CR, Fontaine JJ, Pope KL, Garmestani AS. Adaptive management for a turbulent future. J Environ Manag. 2011;92(5):1339–45. Altieri M. “Part 1.” Agroecological innovations: increasing food production with participatory development. London: Earthscan Publications; 2002. Altieri M, Toledo VM. The agroecological revolution in Latin America: rescuing nature, ensuring food sovereignty and empowering peasants. J Peasant Stud. 2011;38(3):587–612. Álvarez-Ríos GD, Figueredo-Urbina CJ, Pacheco-Torres F, Casas A. Management and morphological and genetic variation of agaves for preparing pulque in Michoacán, México. J Ethnobiol Ethnomed. 2020;16(3):1–17. Ardelean CF, Becerra-Valdivia L, Pedersen MW, Schwenninger JL, Oviatt CG, Macías-Quintero JI, Arroyo-Cabrales J, Sikora M, Ocampo-Díaz YZE, Rubio-Cisneros II, Watling JG, Medeiros VB, Oliveira PE, Barba-Pingarón J, Ortiz-Butrón A, Blancas-Vázquez J, Rivera-González I, Solís-Rosales C, Rodríguez-Ceja M, Gandy DA, Navarro-Gutierrez Z, De La Rosa-Díaz JJ, Huerta-Arellano V, Marroquín-Fernández MB, Martínez-Riojas LM, López-Jiménez A, Higham T, Willerslev E. Evidence of human occupation in Mexico around the last glacial maximum. Nature. 2020;584:87–92. Balée W. Brief review of historical ecology. Les nouvelles de l'archéologie. 2018;152:7–10. Barnosky AD, Matzke N, Tomiya S, Wogan GOU, Swartz B, Quental TB, Marshall C, McGuire JL, Lindsey EL, Maguire KC, Mersey B, Ferrer EA. Has the Earth’s sixth mass extinction already arrived? Nature. 2011;471:51–7. Barnosky AD, Hadly EA, Bascompte J, Berlow EL, Brown JH, Fortelius M, Getz WM, Harte J, Hastings A, Marquet PA, Martinez ND, Mooers A, Roopnarine P, Vermeij G, Williams JW, Gillespie R, Kitzes J, Marshall C, Matzke N, Mindell DP, Revilla E, Smith AB. Approaching a state shift in Earth’s biosphere. Nature. 2012;486:52–8. Barrera-Bassols N, Floriani N, editors. Saberes locales, paisajes y territorios rurales en América Latina. Popayán: Universidad del Cauca; 2018. Bebbington A, editor. Investigación y cambio social: Desafíos para las ONG en Centroamérica y México. Guatemala: Editorial de Ciencias Sociales; 2007. Berkes F. Traditional ecological knowledge in perspective. In: Inglis JT, editor. Traditional ecological knowledge: concepts and cases. Ottawa: Canadian Museum of Nature/International Development Research Centre, International Program on Traditional Ecological Knowledge International Development Research Centre; 1993. p. 1–9.

974

A. Casas et al.

Berkes F. Sacred ecology: traditional ecological knowledge and resource management. London: Routledge; 1999. Berkes F, Folke C, Gadgil M. Traditional ecological knowledge, biodiversity, resilience and sustainability. In: Perrings CA, Mäler KG, Folke C, Holling CS, Jansson BO, editors. Biodiversity conservation. Ecology, economy & environment, vol. 4. Dordrecht: Springer; 1994. p. 269–87. Berkes F, Colding J, Folke C. Rediscovery of traditional knowledge as adaptive management. Ecol Appl. 2000;10:1251–62. Berlin B, Breedlove DE, Raven PH. Principles of Tzeltal plant classification. An introduction to the botanical ethnography of a Mayan-speaking, people of Highland, Chiapas. New York: Academic Press; 1974. Blancas J, Casas A, Rangel-Landa S, Moreno-Calles AI, Torres I, Pérez-Negrón E, Solís L, Delgado-Lemus A, Parra F, Arellanes Y, Caballero J, Cortés L, Lira R, Dávila P. Plant Management in the Tehuacán-Cuicatlán Valley, Mexico. Econ Bot. 2010;64:287–302. Blancas J, Casas A, Pérez-Salicrup D, Caballero J, Vega E. Ecological and socio- cultural factors influencing plant management in Nahuatl communities of the Tehuacan Valley, Mexico. J Ethnobiol Ethnomed. 2013;9(1):39. Boivin NL, Zeder MA, Fuller DQ, Crowther A, Larson G, Erlandson JM, Denham T, Petraglia MD. Ecological consequences of human niche construction: examining long-term anthropogenic shaping of global species distributions. Proc Natl Acad Sci. 2016;113(23):6388–96. Brandt P, Ernst A, Gralla F, Luederitz C, Lang DJ, Newig J, Reinert F, Abson DJ, Wehrden H. A review of transdisciplinary research in sustainability science. Ecol Econ. 2013;92:1–15. Brush SB. Indigenous knowledge of biological resources and intellectual property rights: the role of anthropology. Am Anthropol New Ser. 1993;95(3):653–71. Brush SB. Farmer’s bounty. Locating crop diversity in the contemporary world. New Haven: Yale University Press; 2004. Brush SB, Stabinsky D, editors. Valuing local knowledge: indigenous people and intellectual property rights. Washington, D.C.: Island Press; 1996. Bye RA. Botanical perspectives of ethnobotany of the greater southwest. Econ Bot. 1985;39: 375–86. Bye RA. The role of humans in the diversification of plants in Mexico. In: Ramamoorthy TP, Bye R, Lot A, Fa J, editors. Biological diversity of Mexico: origins and distribution. Oxford: Oxford University Press; 1993. p. 707–32. Bye R, Linares E. Ethnobotany and ethnohistorical sources of Mesoamerica. In: Lira R, Casas A, Blancas J, editors. Ethnobotany of Mexico. Interactions of people and plants in Mesoamerica. New York: Springer; 2016. p. 41–65. Caballero J. Perspectivas para el quehacer etnobotánico en México. In: Barrera A, editor. La etnobotánica: tres puntos de vista y una perspectiva. Xalapa: Instituto de Investigaciones sobre Recursos Bióticos; 1979. p. 27–30. Caballero J, Casas A, Cortés L, Mapes C. Patrones en el conocimiento, uso y manejo de plantas en pueblos indígenas de México. Estudios Atacameños. 1998;16:181–96. Caballero J, Cortés L, Mapes C, Blancas J, Rangel-Landa S, Torres-García I, Farfán-Heredia B, Martínez-Ballesté A, Casas A. Ethnobotanical diversity of Mexico. In: Casas A, Blancas J, editors. Ethnobotanical knowledge in Mexico: use, management and other interactions between people and plants. Springer; 2022. Cabrera-Toledo D, González-Astorga J, Vovides AP, Casas A, Vargas-Ponce O, Carrillo-Reyes P, Nolasco-Soto J, Vega E. Dynamics and local spatial genetic patterns of relictual populations of Dioon spp. (Zamiaceae): bases for their sustainable management. Popul Ecol. 2019;61(1): 62–73. Camou-Guerrero A, Casas A, Moreno-Calles AI, Aguilera-Lara J, Garrido-Rojas D, Rangel-LandaS, Torres I, Pérez-Negrón E, Solís L, Blancas J, Guillén S, Parra F, Rivera-Lozoya E. Ethnobotany in Mexico: history, development, and perspectives. In: Lira R, Casas A,

Perspectives of the Ethnobotanical Research in Mexico

975

Blancas J, editors. Ethnobotany of Mexico: interactions of people and plants in Mesoamerica. New York: Springer; 2016. p. 21–39. Carpenter SR, Arrow KJ, Barrett S, Biggs R, Brock WA, Crépin A-S, Engström G, Folke C, Hughes TP, Kautsky N, Li C-Z, McCarney G, Meng K, Mäler K-G, Polasky S, Scheffer M, Shogren J, Sterner T, Vincent JR, Walker B, Xepapadeas A, De Zeeuw A. General resilience to cope with extreme events. Sustainability. 2012;4(12):3248–59. Casas A, Viveros JL, Caballero J. Etnobotánica mixteca: sociedad, cultura y recursos naturales en la Montaña de Guerrero. Mexico: Instituto Nacional Indigenista CONACULTA; 1994. Casas A, Otero-Arnaiz A, Peréz-Negrón E, Valiente-Banuet A. In situ management and domestication of plants in Mesoamerica. Ann Bot. 2007;100:1101–15. Casas A, Blancas J, Lira R. Mexican ethnobotany. Interactions of people and plants in Mesoamerica. In: Lira R, Casas A, Blancas J, editors. Ethnobotany of Mexico. Interactions of peoples and plants in Mesoamerica, chapter 1. Utrecht: Springer; 2016a. p. 1–19. Casas A, Torres-Guevara J, Parra F. Domesticación en el continente americano. Volume 1. Manejo de biodiversidad y evolución dirigida por las culturas del Nuevo Mundo. Lima: Universidad Nacional Autónoma de México, Universidad Nacional Agraria La Molina; 2016b. Casas A, Torres I, Delgado-Lemus A, Rangel-Landa S, Félix-Valdés L, Cabrera D, Vargas O, Ilsley C, Moreno-Calles AI, Torres-Guevara J, Parra F, Farfán B, Castillo A, Ayala B, Blancas JJ, Vallejo M, Solís L, Ortíz T, Bullen A. Ciencia para la sustentabilidad: investigación, educación y procesos participativos. Rev Mex Biodivers. 2017;88(S1):113–28. Casas A, Farfán-Heredia B, Camou-Guerrero A, Torres I, Blancas J, Rangel-Landa S. Wild, weedy and domesticated plants for food security and sovereignty. In: Casas A, Blancas J, editors. Ethnobotany of the mountain regions of Mexico. Springer; 2022. Castillo A, Torres A, Bocco G, Velásquez A. The use of ecological science by rural producers: a case study in Mexico. Ecol Appl. 2005;15(2):745–56. Castillo A, Vega-Rivera JH, Pérez-Escobedo M, Romo-Díaz G, López-Carapia G, Ayala-Orozco B. Linking social-ecological knowledge with rural communities in Mexico: lessons and challenges towards sustainability. Ecosphere. 2018;9(10) https://doi.org/10.1002/ecs2.2470. Castillo A, Bullen-Aguiar AA, Peña-Mondragón JL, Gutiérrez-Serrano NG. The social component of social-ecological research: moving from the periphery to the center. Ecol Soc. 2020;25(1):6. Cavalli-Sforza LL. Genes, peoples, and languages. Proc Natl Acad Sci. 1997;94:7719–24. Cela-Conde CJ, Ayala F. Tools made us human. The role of technology in the biological and social evolution of the homo genus. Sociol Technosci. 2018;8(2):1–25. Chomicki G, Schaefer H, Renner SS. Origin and domestication of Cucurbitaceae crops: insights from phylogenies, genomics and archaeology. New Phytol. 2020;226(5):1240–55. Clément D. The historical foundations of ethnobiology. J Ethnobiol. 1998;18(2):161–87. Clement CR, Casas A, Parra-Rondinel FA, Levis C, Peroni N, Hanazaki N, Cortés-Zárraga L, Rangel-Landa S, Alves RP, Ferreira MJ, Cassino MF, Coelho SD, Cruz-Soriano A, PancorboOlivera M, Blancas J, Martínez-Ballesté A, Lemes G, Lotero-Velásquez E, Bertin VM, Mazzochini GG. Disentangling domestication from food production systems in the Neotropics. Quaternary. 2021;4(1):4. Costanza R. A theory of socio-ecological system change. J Bioecon. 2014;16:39–44. Cotton CM. Ethnobotany: principles and applications. Chichester: Wiley; 1997. Cruse-Sanders JM, Friar EA, Parker KC, Huang DI, Mashayekhi S, Prince LM, Otero-Arnaiz A, Casas A. Managing diversity: domestication and gene flow in Stenocereus stellatus Riccob. In Mexico. Ecol Evol. 2013;3(5):1340–55. Delgado-Lemus A, Casas A, Téllez-Valdés O. Distribution, abundance and traditional management of agave potatorum in the Tehuacán Valley, Mexico: perspectives for sustainable management of non-timber forest products. J Ethnobiol Ethnomed. 2014;10:63. Drury R, Homewood K, Randall S. Less is more: the potential of qualitative approaches in conservation research. Anim Conserv. 2011;14:18–24. Eberhard DM, Simons GF, Fennig CD, editors. Ethnologue: languages of the world. 25th ed. Dallas: SIL International. Online version: http://www.ethnologue.com; 2022.

976

A. Casas et al.

Félix-Valdéz L, Vargas O, Cabrera-Toledo D, Casas A, Cibrián-Jaramillo A, de la Cruz-Larios L. Effects of management for mescal production on the diversity and genetic structure of agave potatorum Zucc., in Central México. Genet Resour Crop Evol. 2016;63(7):1255–71. Flannery K, editor. Guilá Naquitz. New York: Academic Press; 1986. Ford RI. Ethnobotany: historical diversity and synthesis. In: Ford RI, editor. The nature and status of ethnobotany. Anthropological papers 67, Museum of Anthropology, University of Michigan. Ann Arbor: University of Michigan; 1978. p. 33–50. Franco-Moraes J, Clement CR, Oliveira JC, Oliveira AA. A framework for identifying and integrating sociocultural and environmental elements of indigenous peoples’ and local communities’ landscape transformations. Perspect Ecol Conserv. 2021;19:143–52. Frankel OH, Bennett E, editors. Genetic resources in plants–their exploration and conservation, IPB handbook no. 11. London: International Biological Programme; 1970. Freire P. ¿Extensión o comunicación? Mexico: La concientización en el medio rural, Siglo XXI Editores; 1973. García R. Sistemas complejos. Conceptos, método y fundamento de la investigación epistemológica. Barcelona: Gedisa; 2006. García R. Interdisciplinariedad y sistemas complejos. Rev Latinoam Metodol Cienc Soc. 2011;1(1): 66–101. Gavito ME, Van der Wal H, Aldasoro M, Ayala B, Bullen A, Cach M, Casas A, Fuentes A, González-Esquivel C, Jaramillo P, Martínez P, Masera O, Pascual F, Pérez-Salicrup D, Robles R, Ruiz I, Villanueva G. Ecología, tecnología e innovación para la sustentabilidad: Retos y perspectivas en México. Rev Mex Biodivers. 2017;88(S1):150–60. Gepts P. The contribution of genetic and genomic approaches to plant domestication studies. Curr Opin Plant Biol. 2014;18:51–9. Gibson CC, Ostrom E, Ahn TK. The concept of scale and the human dimensions of global change: a survey. Ecol Econ. 2000;32(2):217–39. Gilmartin M. Colonialism/imperialism. In: Gallaher C, Dahlman CT, Gilmartin M, Mountz A, Shirlow WT, editors. Key concepts in political geography. London: Sage; 2009. p. 115–23. Grumbine RE. What is ecosystem management? Conserv Biol. 1994;8(1):27–38. Gunderson LH, Holling CS. Panarchy. Understanding transformations in human and natural systems. Washington D. C: Island Press; 2002. Hamilton C, Bonneuil C, Gemenne F. The Anthropocene and the global environmental crisis. Rethinking modernity in a new epoch. London: Earthscan/Routledge; 2015. Hart R, Salick J. Dynamic ecological knowledge systems amid changing place and climate: Mt. Yulong rhododendrons. J Ethnobiol. 2017;37(1):21–36. Hernández-Xolocotzi E. La agricultura en la Península de Yucatán. In: Beltrán E, editor. Los recursos naturales del sureste y su aprovechamiento, vol. 3. Mexico: Instituto Mexicano de Recursos Naturales Renovables; 1959. p. 3–57. Hernández-Xolocotzi E. Apuntes sobre la exploración etnobotánica y su metodología. Chapingo: Colegio de Postgraduados; 1971. Hernández-Xolocotzi E. El concepto de etnobotánica. In: Barrera A, editor. La Etnobotánica: Tres puntos de vista y una perspectiva. Xalapa: INIREB A. C; 1979. p. 13–8. Hernández-Xolocotzi E. Aspects of plant domestication in Mexico: a personal view. In: Ramamoorthy TP, Bye R, Lot A, Fa J, editors. Biological diversity of Mexico: origins and distribution. Oxford: Oxford University Press; 1993. p. 733–53. Hernández-Xolocotzi E, Ramos-Rodríguez A. Metodología para el estudio de agroecosistemas con persistencia de tecnología agrícola tradicional. In: Hernández-Xolocotzi E, editor. Agroecosistemas de México. Chapingo: Colegio de Postgraduados, Escuela Nacional de Agricultura; 1977. p. 321–33. Hernández-Xolocotzi E, Inzunza-Mascareño F, Solano-Solano CB, Parra-Vázquez MR. Nuevos enfoques de la investigación en áreas agrícolas de ladera. In: Novoa AR, Posner JL, editors. Seminario internacional sobre producción agropecuaria y forestal en zonas de ladera de América

Perspectives of the Ethnobotanical Research in Mexico

977

tropical. Turrialba: Centro Agronómico Tropical de Investigación y Enseñanza, Rockefeller Foundation; 1980. p. 205–10. Heywood V, Casas A, Ford-Lloyd B, Kell S, Maxted N. Conservation and sustainable use of crop wild relatives. Agric Ecosyst Environ. 2007;121(3):245–55. Holling CS. Understanding the complexity of economic, ecological, and social systems. Ecosystems. 2001;4(5):390–405. Hong S-K, Bogaert J, Min Q. Biocultural landscapes: diversity, functions and values. New York: Springer; 2014. IFRI. International Forestry Resources and Institutions (IFRI) network: research methods. Available from: www.ifriresearch.net. 2013. Illsley-Granich C, Rivera G, Tlacotempa A, Morales P, García J, Gómez T, Martínez JM, Marcial J, Castro F, Calzada M, Mancilla S, García P, Casarrubias L, Hernández F, Flores J. Manual de manejo campesino de magueyes mezcaleros silvestres. Mexico: Grupo de Estudios Ambientales GEA A.C. Comisión Nacional para el Conocimiento y Uso de la Biodiversidad, CONABIO; 2004. Illsley-Granich C, Acosta J, Aguilar J, Aguilar M, Alatorre G, Díaz-León M, González A, Marielle C, Jayo A. El Grupo de Estudios Ambientales, AC: Entre la acción social y la consolidación institucional. In: Bebbington A, editor. Investigación y cambio social: Desafíos para las ONG en Centroamérica y México. Guatemala: Editorial de Ciencias Sociales; 2007. p. 175–208. Kantar MB, Nashoba AR, Anderson JE, Blackman BK, Rieseberg LH. The genetics and genomics of plant domestication. Bioscience. 2017;67(11):971–82. Kates RW. What kind of a science is sustainability science? Proc Natl Acad Sci. 2011;108(49): 19449–50. Kates RW, Clark WC, Corell R, Hall JM, Jaeger CC, Lowe I, McCarthy JJ, Schellnhuber HJ, Bolin B, Dickson NM, Faucheux S, Gallopin GC, Grübler A, Huntley B, Jäger J, Narpat SJ, Kasperson RE, Mabogunje A, Matson P, Mooney H, Moore-III B, O'riardan T, Svedin U. Sustainability science. Science. 2001;292(5517):641–2. Kaufman D, Moss DM, Osborn TA, editors. Beyond the boundaries. A transdisciplinary approach for learning and teaching. London: Praeger; 2003. Keith DA, Martin TG, McDonald-Madden E, Walters C. Uncertainty and adaptive management for biodiversity conservation. Biol Conserv. 2011;144(4):1175–8. Kistler L, Thakar HB, VanDerwarker AM, et al. Archaeological central American maize genomes suggest ancient gene flow from South America. PNAS. 2020;117:33124–9. Kuhn TS. The structure of scientific revolutions. Chicago: The University of Chicago Press; 1962. Lindqvist C, Rajora OP, editors. Paleogenomics. Genome-scale analysis of ancient DNA. Switzerland: Springer Nature; 2019. Lira R, Casas A, Blancas J. Ethnobotany of Mexico. Interactions of people and plants in Mesoamerica. New York: Springer; 2016. MacNeish RS. A summary of subsistence. In: Byers DS, editor. The prehistory of the Tehuacán Valley: environment and subsistence, vol. 1. Austin: University of Texas Press; 1967. p. 290–309. MacNeish RS. The origins of agriculture and settled life. Oklahoma University Press; 1992. Maffi L. Linguistic, cultural, and biological diversity. Annu Rev Anthropol. 2005;34:599–617. Maffi L, Woodley E. Biocultural diversity conservation. A global sourcebook. London: Earthscan; 2010. Mapes C, Basurto F. Biodiversity and edible plants of Mexico. In: Lira R, Casas A, Blancas J, editors. Ethnobotany of Mexico. Interactions of peoples and plants in Mesoamerica, chapter 1. Utrecht: Springer; 2016. p. 83–150. Martínez-Alfaro MA. Estado actual de las investigaciones etnobotánicas en México. Botanical Sciences. 1994;55:65–74.

978

A. Casas et al.

Martínez-Ballesté A, Martorell C, Martínez-Ramos M, Caballero J. Applying retrospective demographic models to assess sustainable use: the Maya management of xa’ an palms. Ecol Soc. 2005;10(2):1–35. Martínez-Ballesté A, Martorell C, Caballero J. Cultural or ecological sustainability? The effect of cultural change on sabal palm management among the lowland Maya of Mexico. Ecol Soc. 2006;11(2):27. Martínez-Ballesté A, Martorell C, Caballero J. The effect of Maya traditional harvesting on the leaf production, and demographic parameters of sabal palm in the Yucatán peninsula, Mexico. For Ecol Manag. 2008;256(6):1320–4. Masera O, Astier M, López-Ridaura S. Sustentabilidad y manejo de recursos naturales. El marco de evaluación MESMIS. México: Mundi-Prensa/Instituto de Ecología/UNAM; 2000. McClung de Tapia E, Martínez-Yrizar D, Acosta G, Zalaquet F, Robitaille EA. Nuevos fechamientos para las plantas domesticadas en el México prehispánico. Anales de Antropología. 2001;35:125–56. McMahon A, McMahon R. Language classification by numbers. Oxford: Oxford University Press; 2005. MEA Millennium Ecosystem Assessment. Ecosystems and human Well-being. Biodiversity synthesis. Washington, D.C.: World Resources Institute; 2005. Moreno Calles AI, Casas A, Vallejo M, Toledo VM, editors. Etnoagroforestería en México. Morelia: Universidad Nacional Autónoma de México; 2016. Moreno-Calles AI, Toledo VM, Casas A. Los sistemas agroforestales tradicionales de México: una aproximación biocultural. Botanical Sciences. 2013;91(4):375–98. Navarro LM, Fernández N, Guerra C, Guralnick R, Kissling WD, Londoño MC, Muller-Karger F, Turak E, Balvanera P, Costello MJ, Delavaud A, El Serafy GY, Ferrier S, Geijzendorffer I, Geller GN, Jetz W, Kim ES, Kim H, Martin CS, McGeoch MA, Mwampamba TH, Nel JL, Nicholson E, Pettorelli N, Schaepman ME, Skidmore A, Pinto IS, Vergara S, Vihervaara P, Xu H, Yahara T, Gill M, Pereira HM. Monitoring biodiversity change through effective global coordination. Curr Opin Environ Sustain. 2017;29:158–69. Nelson MK, Shilling D, editors. Traditional ecological knowledge. Learning from indigenous practices for environmental sustainability. Cambridge, MA: Cambridge University Press; 2018. O’Madagain C, Tomasello M. Shared intentionality, reason-giving and the evolution of human culture. Philos Trans R Soc B. 2021;377:20200320. O’Connor B, Bojinski S, Röösli C, Schaepman ME. Monitoring global changes in biodiversity and climate essential as ecological crisis intensifies. Eco Inform. 2020;55:101033. Olufunke MD. Developments in phytochemistry. In: Vallisuta O, editor. Drug discovery. Research in pharmacognosy, chapter 2. London: IntechOpen Limited; 2012. p. 11–22. Ostrom E. Governing the commons. The evolution of institutions for collective action. Cambridge, MA: Cambridge University Press; 1990. Ostrom E. A diagnostic approach for going beyond panaceas. Proc Natl Acad Sci. 2007;104(39): 15181–7. Ostrom E. A general framework for analyzing sustainability of social-ecological systems. Science. 2009;325(5939):419–22. Pacheco LF, Altrichter M, Beck H, Buchori D, Owusu EH. Economic growth as a major cause of environmental crisis: comment to ripple et al. Bioscience. 2018;68(4):238. Paz-Guerrero F, Casas A, Alvarado-Sizzo H. Habitat disturbance and population genetics of Stenocereus quevedonis (Cactaceae) in Michoacán, México: bases for in situ conservation of silvicultural managed genetic resources. Genet Resour Crop Evol. 2019;66(3):633–43. Pease JB, Haak DC, Hahn MW, Moyle LC. Phylogenomics reveals three sources of adaptive variation during a rapid radiation. PLoS Biol. 2016;14:e1002379. Pereira HM, Navarro LM, Martins IS. Global biodiversity change: the bad, the good, and the unknown. Annu Rev Environ Resour. 2012;37:25–50.

Perspectives of the Ethnobotanical Research in Mexico

979

Pérez-Negrón E, Casas A. Use, extraction rates and spatial availability of plant resources in the Tehuacán-Cuicatlán Valley, Mexico: the case of Santiago Quiotepec, Oaxaca. J Arid Environ. 2007;70(2):356–79. Pieroni A, Quave CL. Ethnobotany in the Balkans: quo Vadis? In: Pieroni A, Quave CL, editors. Ethnobotany and biocultural diversities in the Balkans: perspectives on sustainable rural development and reconciliation. New York: Springer; 2014. p. 1–9. Piperno D, Pearsall D. Phytoliths in the reproductive structures of maize and teosinte: implications for the study of maize evolution. J Archaeol Sci. 1993;20:337–62. Piperno DR, Ranere AJ, Holst I, Iriarte J, Dickau R. Starch grain and phytolith evidence for early ninth millennium BP maize from the central Balsas River valley, Mexico. Proc Natl Acad Sci. 2009;106:5019–24. Plotkin MJ. Tales of a Shaman’s apprentice. An ethnobotanist searches for new medicines in the Amazon rainforest. New York: Penguin Books; 1993. Prance GT, Chadwick DJ, Marsh J, editors. Ethnobotany and the search for new drugs. New York: Wiley; 1994. Rangel-Landa S, Dávila P, Casas A. Facilitation of agave potatorum: an ecological approach for assisted population recovery. For Ecol Manag. 2015;347:57–74. Rangel-Landa S, Casas A, Rivera-Lozoya E, Torres-García I, Vallejo M. Ixcatec ethnoecology: biocultural principles of plant management in Oaxaca, Mexico. J Ethnobiol Ethnomed. 2016;12:30. Resilience Alliance. Assessing resilience in social-ecological systems: workbook for practitioners. Version 2.0. Online: http://www.resalliance.org/3871.php 2010. Rist L, Felton A, Samuelsson L, Sandström C, Rosvall O. A new paradigm for adaptive management. Ecol Soc. 2013;18(4):63. Sánchez J, Domínguez R, León M, Samaniego J, Sunkel O. Recursos naturales, medio ambiente y sostenibilidad. 70 años de pensamiento de la CEPAL. Santiago: Comisión Económica para América Latina y el Caribe (CEPAL); 2019. Schultes RE. La etnobotánica: su alcance y sus objetos. Caldasia. 1941;3:7–12. Schultes RE, von Reis S, editors. Ethnobotany: evolution of a discipline. Portland: Dioscorides Press; 1995. Sharifian A, Fernández-Llamazares Á, Wario HT, Molnár Z, Cabeza M. Dynamics of pastoral traditional ecological knowledge: a global state-of-the-art review. Ecol Soc. 2022;27(1):14. Smith BD. The initial domestication of Cucurbita pepo in the Americas 10,000 years ago. Science. 1997;276:932–4. Smith BD. A cultural niche construction theory of initial domestication. Biol Theory. 2012;6(3): 260–71. Stout D. Technology and human brain evolution. Gen Anthropol. 2008;15(2):1–5. Swart R, Raskin P, Robinson J. Critical challenges for sustainability science. Science. 2002;297(5589):1994–6. Tattersall I, Schwartz JH. Evolution of the genus homo. Annu Rev Earth Planet Sci. 2009;37:67–92. Tengö M, Brondizio ES, Elmqvist T, Malmer P, Spierenburg M. Connecting diverse knowledge systems for enhanced ecosystem governance: the multiple evidence base approach. Ambio. 2014; https://doi.org/10.1007/s13280-014-0501-3. Toledo VM. New paradigms for a new ethnobotany: reflections on the case of Mexico. In: Schultes RE, Reis SV, editors. Ethnobotany: evolution of a discipline. Portland: Discorides; 1995. Toledo VM, Barrera-Bassols N. La memoria biocultural. Barcelona: España; 2008. Torres-García I, Casas A, Delgado-Lemus A, Rangel-Landa S. Aprovechamiento, demografía y establecimiento de Agave potatorum en el Valle de Tehuacán, México: Aportes etnobiológicos y ecológicos para su manejo sustentable. Zonas Áridas. 2013;15(1):92–109. Torres-García I, Casas A, Vega E, Martínez-Ramos M, Delgado-Lemus A. Population dynamics and sustainable management of mescal agaves in Central Mexico: agave potatorum in the Tehuacán-Cuicatlán Valley. Econ Bot. 2015;69(1):26–41.

980

A. Casas et al.

Torres-García I, León A, Vega E, Moreno-Calles AI, Casas A. Integral projection models and sustainable forest management of agave inaequidens in western Mexico. Front Plant Sci. 2020;11:1224. Ulian T, Pritchard HW, Casas A, Mattana E, Liu U, Way M, Dávila P, Lira R. Neglected and underutilized species of Mexico: status and priorities for their conservation and sustainable use. In: Padulosi S, King IO, Hunter D, Swaminathan MS, editors. Orphan crops for sustainable food and nutrition security: promoting neglected and underutilized species. Issues in agricultural biodiversity, chapter 9. London: Taylor and Francis Routledge; 2021. p. 118–27. Vanberg VJ. 10 human intentionality and design in cultural evolution. In: Schubert C, von Wangenheim G, editors. Evolution and design of institutions. London: Routledge; 2006. p. 197–211. Verbeek PP. Cyborg intentionality: rethinking the phenomenology of human–technology relations. Phenomenol Cogn Sci. 2008;7:387–95. Vibrans H, Casas A. Roads travelled and roads ahead: the consolidation of Mexican ethnobotany in the new millennium. An essay. Bot Sci. 2022;100 Walker B. Resilience: what it is and is not. Ecol Soc. 2020;25(2):11. Walker BH, Salt D. Resilience practice. Washington, D.C., USA: Island; 2012. Weible CM, Pattison A, Sabatier PA. Harnessing expert-based information for learning and the sustainable management of complex socio-ecological systems. Environ Sci Policy. 2010;13(6): 522–34. Young OR, Berkhout F, Gallopin GC, Janssen MA, Ostrom E, van der Leeuw S. The globalization of socio-ecological systems: an agenda for scientific research. Glob Environ Chang. 2006;16(3): 304–16. Zarrillo S, Gaikwad N, Lanaud C, Powis T, Viot C, Lesur I, Fouet O, Argout X, Guichoux E, Salin F, Solorzano RL, Bouchez O, Vignes H, Severts P, Hurtado J, Yepez A, Grivetti L, Blake M, Valdez F. The use and domestication of Theobroma cacao during the mid-Holocene in the upper Amazon. Nat Ecol Evol. 2018;2:1879–88. Zeder MA. The broad spectrum revolution at 40: resource diversity, intensification, and an alternative to optimal foraging explanations. J Anthropol Archaeol. 2012;31(3):241–64. Zeder MA. Domestication as a model system for the extended evolutionary synthesis. Interface Focus. 2017;7:20160133.

Part II Plant Profiles

Agastache spp. LAMIACEAE. Important Species of Hyssop in Mexico Guadalupe Carrillo-Galva´n and Robert Bye

Synonyms Agastache aurantiaca (A. Gray) Lint & Epling: Cedronella aurantiaca A. Gray; Brittonastrum aurantiacum (A. Gray) Briquet Agastache mexicana (Kunth) Lint & Epling subsp. mexicana: Dracocephalum mexicanum Kunth; Cedronella mexicana (Kunth) Bentham; Brittonastrum mexicanum (Kunth) Briquet; Gardoquia betonicoides (Lindley); Brittonastrum betonicoides (Lindley) Briquet; Dekinia coccinea Martens & Galeotti Agastache mexicana (Kunth) Lint & Epling subsp. xolocotziana Bye, Linares & Ramamoorthy Agastache micrantha (A. Gray) Wooton & Standley: Cedronella micrantha A. Gray; Brittonastrum micranthum (A. Gray) Briquet Agastache pallida (Lindley) Cory: Cedronella pallida Lindley; Brittonastrum pallidum (Lindley) Briquet Agastache palmeri (B. L. Robinson) Lint & Epling var. breviflora (Regel) R. W. Sanders: Cedronella mexicana (Kunth) Bentham var. breviflora Regel Agastache palmeri (B. L. Robinson) Lint & Epling var. palmeri: Brittonastrum palmeri B. L. Robinson Agastache pringlei (Briquet) Lint & Epling var. pringlei: Brittonastrum pringlei Briquet

G. Carrillo-Galván (*) · R. Bye Jardín Botánico, Instituto de Biología, Universidad Nacional Autónoma de México, Ciudad de México, Mexico e-mail: [email protected]; [email protected] © Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_24

983

984

G. Carrillo-Galva´n and R. Bye

Local Names Agastache aurantiaca: In the northern Sierra Madre Occidental (SMOc), the Tarahumara refer to this plant as “úmichi” (Bye 1985). Agastache mexicana subsp. mexicana: Since prehispanic times, it has been known in Nahuatl, the language of the Aztecs, as “tzompili-huitz-patli” [catarrh, medicine], “tzompilihuiz-xihuitl” [catarrh, herb], and “Tlalahuehuetl” (De la Cruz and Badiano 1964; Martínez 1979; Cruz-Hernández 2014). Today throughout Mexico, it is known as “toronjil” (Mexican hyssop), “toronjil morado” (purple Mexican hyssop), and “toronjil rojo” (red Mexican hyssop) (Argueta et al. 1994; Farmacopea Herbolaria de los Estados Unidos Mexicanos 2014; Martínez 1979, 1990; SantillánRamírez et al. 2008); “toronjil” is based upon “toronja,” which is grapefruit in Spanish and refers to the citrus fragrances of the herb. Mestizos of Nahuatl descent from central Mexico located in the foothills of the Popocatépetl and Iztaccíhuatl volcanoes (Mexico City, State of Mexico, and Morelos) traditionally recognize wild and cultivated plants such as “toronjil morado de monte” (field purple Mexican hyssop) and “toronjil morado de casa” (house purple Mexican hyssop), respectively (Carrillo-Galván et al. 2020). The Otomí people from Puebla refer to it as “tama,” “torojí,” and “toronjil.” The Tepehua people from Hidalgo, Puebla and Veracruz know it as “pinkil.” In Michoacán it receives the names “noritén” and “toronjillo” (Martínez 1979). Agastache mexicana subsp. xolocotziana: In central Mexico it is known as “toronjil blanco” (white Mexican hyssop) and “toronjil blanco de casa” (white house Mexican hyssop) (Carrillo-Galván et al. 2020; Martínez 1990; Santillán et al. 2008). Agastache micrantha: The Tarahumara of northern Mexico refer to this species in Spanish as “té de menta” and “poleo” while in Rarámuri as “húpachi,” “júpichi,” or “júpisi” (Bye 1985, 1986; Cardenal 1993; Olivas-Sánchez 1999; Wyndham 2010). Agastache pallida: As with the previous species, the Tarahumara apply the names “poleo” and “húpachi” (Bye 2007). Agastache palmeri var. breviflora: The Huastecos of the lowlands in southern Sierra Madre Oriental (SMOr) know it as “toronjil rosa” (pink Mexican hyssop) or “toronjil cimarrón” (wild or maroon Mexican hyssop) (Carrillo-Galván et al. 2017). Agastache palmeri var. palmeri: “Betónica” is the name applied to this popular herb in the markets of San Luis Potosi and in the Sierra de Álvarez to the east (Bye 1979; Colín 2006). Agastache pringlei: The Tarahumara people apply a local variant of the Spanish term for anise, “anisco,” as well as the indigenous name “húpachi” (Bye 2007).

Botany and Ecology The genus Agastache (Lamiaceae – Nepeteae) consists of two sections: Agastache with eight species in boreal North America and eastern Asia, and Brittonastrum with 14 species in of the mountains of Mexico and adjacent southwestern United

Agastache spp. LAMIACEAE. Important Species of Hyssop in Mexico

985

States. They are perennial herbs with simple quadrangular stems that extend from horizontal rhizomes, often with woody caudex. The simple, opposite leaves have short petioles and simple blades, 2–9 cm long, that vary in shape from ovate to lanceolate and with marginal indentation from crenate to serrate. The inflorescence consists of cymose clusters of 5–30 flowers arranged in verticils that may be partitioned or condensed at the apex of each stem. The cylindrical calyx tube terminates with five short teeth at the mouth which can be transverse or oblique. The oblique mouth of the funnelform corolla tube exhibits lobes less than one-fifth the tube length and consist of 2 upper lobes that are fused to form one lip, a lower medium lobe and 2 lateral lobes; the length and color (from white, golden-orange, magenta to red) of the corolla varies according to the species. The four exserted stamens arranged in pairs of unequal length may be included under or exceed the upper corolla lip. The exserted style extends from a 4-lobed ovary surround by nectarifeorus disc and terminates with two branches. The fruit consists of 4 mericarps, 1–2 mm long, trigonous-ovoid in shape, and enclosed by the calyx tube. The members of Agastache section Brittonastrum are distributed as follows: one throughout the Trans Mexican Volcanic Belt or Eje Volcánico Transversal (EVT) of central Mexico, two along the Sierra Madre Oriental (SMOr), six in the SMOc, four in the southern Rocky Mountains of the USA, and two in the northern SMOc and contiguous southern RM with 64% being Mexican endemic species (Martínez-Gordillo et al. 2013, 2017; Sanders 1987). The morphological description and distribution patterns listed below are based upon Sanders (1987) and Bye et al. (1987) as well as herbarium specimens deposited in the National Herbarium of Mexico (MEXU). Agastache aurantiaca: This strongly aromatic herb grows from 0.5 to 0.9 m high from a woody base. The leaves are 2–4 cm long, triangular-ovate in shape, with a cordate to truncate base, acute apex, and crenate-serrate margin. The interrupted inflorescence consists of 1–10 flowered paired cymes each subtended by lanceolate to lance-elliptic bracts and separated by cymal internodes 1.2–3 cm long. The orangish-gray colored calyx is 1.5–4 mm long with a tube of 3–4.5 mm and teeth of 1–1.5 mm in length. The golden-orange colored corolla has a tube 25–28 mm long and median lobe 4.5–5 mm long with shallowly few-toothed margin. Both pairs of stamens and style are exserted beyond upper corolla-lip (Fig. 1). Phenology: Flowering and fruiting occur from August to November. Distribution: Grows pineoak woodlands on volcanic outcrops of the mountains of SMOc in southwestern Chihuahua and western Durango, at elevations 2000–2500 m. Agastache mexicana subsp. mexicana: This attractive, strongly aromatic herb grows from 0.5 to 1.5 m high. The leaves are 2–9 cm long, triangular-lanceolate in shape, with an obtuse to subcordate base, acute to attenuate apex, and crenateserrate margin. The interrupted inflorescence consists of 5–20 flowered paired cymes each subtended by linear-lanceolate to linear-elliptic bracts and with cymal internodes 2–4 cm long. The green to magenta colored calyx is 12–16 mm long with a tube of 7–10 mm and teeth of 2.5–3.5 in length. The deep pink to deep magenta or bright red colored corolla has a tube 21–27 mm long and median lobe 4–5 mm long with shallowly crenate margin. Both pairs of stamens and style are

986

G. Carrillo-Galva´n and R. Bye

Fig. 1 (a) A. aurantiaca growing in the Sierra Tarahumara of the SMOc and b) detail of its inflorescence. (Photos: R. Bye)

exserted beyond upper corolla-lip. Phenology: Spontaneous plants flower from June to November and, under cultivation, flowering extends to February. Distribution: It grows spontaneously in temperate pine and pine-oak forests in the mountains of the EVT, from central Michoacán to central Veracruz; 2800–3200 m asl. Cultivated populations are found in the temperate zones of Mexico City, State of Mexico, Hidalgo, Michoacán, Morelos, Puebla, Tlaxcala, and Baja California. It was apparently introduced into European gardens shortly after its discovery by Humboldt and Bonpland (Sanders 1987). Recently its domestication process has been documented for plants cultivated and employed medicinally in the foothills of the Popocatépetl volcano of EVT. The domestication syndrome indicates organoleptic differentiation according to a phytochemical dissimilarity, gigantism in vegetative structures (foliar and rhizome) and floral (corolla, tube, stamens, and style), as well as an intensification of floral pigmentation (Fig. 2) (Carrillo-Galván et al. 2020). Agastache mexicana subsp. xolocotziana: These herbs are similar to plants of the typical species but differ by having shorter, weaker stems (usually to 1.5 m tall) derived from thicker, spreading rhizomes. The shorter leaves with ovate-lanceolate blades, 2–2.5 cm long, have an obtuse to cuneate base, acute apex, and crenateserrate margin except near tip which is entire. The inflorescences of interrupted verticils of 5–15 flowers have cymal internodes 3–6 cm in length and flowers with a green calyx 1–1.3 cm long with a tube 7–8 mm long, triangular acute teeth to 3 mm long; the white corolla measures about 24 mm in length consisting of a tube to 18 mm long and a medium lobe bearing scattered trichomes on the upper surface. Distribution: This subspecies only occurs in a cultivated state apparently having been domesticated in the central eastern region of the EVT; it has been suggested that

Agastache spp. LAMIACEAE. Important Species of Hyssop in Mexico

987

Fig. 2 A. m. subsp. mexicana. (a) Spontaneous plant, (b) domesticated plants growing around the Popocatépetl volcano, and (c) approximation of its inflorescence. The morphological dissimilarity between spontaneous and domesticated plants is appreciated. (Photos: G. Carrillo-Galván)

it is possibly a hybrid between A. m. subsp. mexicana and A. palmerí var. breviflora, given the proximity of allopatric populations and the ability of humans to break geographical barriers (Bye et al. 1987). The domestication syndrome includes organoleptic differentiation along with phytochemical dissimilarity, gigantism in flowers, seeds and rhizome, floral albinism, and partially sterile seeds (Fig. 3) (Carrillo-Galván et al. 2020). Agastache micrantha: This strongly aromatic herb grows from 0.3 to 1.2 m high from a woody base. The leaves are 2–5 cm long, lanceolate-triangular in shape, with an obtuse to truncate base, attenuate apex, and serrate margin. The compact, continuous inflorescence consists of 5–15 flowered paired cymes each subtended by linear bracts and separated by cymal internodes less than 0.3 cm long. The green colored calyx is 2.5 to 3.5 mm long with a tube of 2–3 mm and teeth of 0.5–1 mm in length. The white colored corolla (sometimes with purple spots) has a tube 2.5–3.5 mm long and median lobe 0.5–0.9 mm long with entire or bluntly few-toothed margin. The upper pair of stamens and style are exserted beyond upper corolla-lip (Fig. 4). Phenology: Flowering and fruiting occur August to October. Distribution: It grows in the lower mountains and foothills of the SMOc from northern Arizona and New Mexico to central Chihuahua, northwestern Coahuila, and Durango, at elevations 1600–2300 m. Agastache pallida: This herb grows from 0.3 to 1.3 m high. The leaves are 2–6 cm long, broadly ovate to ovate-triangular in shape, with a cordate to truncate base, obtuse to attenuate apex, and crenate-serrate margin. The interrupted inflorescence consists of 5–20 flowered paired cymes each subtended by triangularlanceolate to elliptic-linear bracts and separated by cymal internodes 1–2 cm long. The green to pink colored calyx is 8–12 mm long with a tube of 6–10 mm and teeth of 1.5–3.5 mm in length. The rose-pink colored corolla has a tube 18–28 mm long and median lobe 4–7 mm long with several sharply toothed margin. Both pairs of stamens and style are slightly exserted beyond upper corolla-lip (Fig. 5). Phenology: Flowering and fruiting occur August to October. Distribution: It is commonly found on volcanic soils in pine-oak forests of the SMOc from southern Arizona and Chihuahua to southern Durango, at elevations 1800–2500 m.

988

G. Carrillo-Galva´n and R. Bye

Fig. 3 (a) Domesticated plants of A. m. subsp. xolocotziana in the foothills of the Popocatépetl volcano and (b) its inflorescence, albinism is observed in the flower. (Photos: G. Carrillo-Galván)

Fig. 4 (a) A. micrantha and (b) detail of its inflorescence. (Photos: R. Bye)

Agastache palmeri var. breviflora: This variety is similar to the typical variety below but differs in having shorty cymal internodes (the inflorescence appears more compact) and the calyx tube is shorter, 4–6.5 mm in legth (Fig. 6). Phenology: Flowering and fruiting extend from June to November. Distribution: Wild plants grow in mesic pine or pine-deciduous forest and on igneous or volcanic rock in the

Agastache spp. LAMIACEAE. Important Species of Hyssop in Mexico

989

Fig. 5 (a) A. pallida growing in the mountains of the SMOc, where the Tarahumara are located and make use of it and (b) inflorescence of the plant. (Photos: R. Bye)

Fig. 6 Plants of A. palmeri var. breviflora growing in a pine-oak forest of the lower Huasteca belonging to the SMOr, and (b) detail of the inflorescence. (Photos: G. Carrillo-Galván)

SMOr stretching from Querétaro south to northern Oaxaca, at elevations 1600–2400 m. Agastache palmeri var. palmeri: This aromatic herb grows from 0.3 to 1.2 m high. The leaves are 2–6 cm long, ovate to ovate-triangular in shape, with an obtuse to cordate base, acute apex, and crenate-serrate margin that is entire at the apex. The interrupted inflorescence consists of 20–30 flowered paired cymes each subtended

990

G. Carrillo-Galva´n and R. Bye

by linear-lanceolate to linear-elliptic bracteoles and separated by cymal internodes less than 1 cm long. The green to magenta colored calyx is 6–13 mm long with a tube of 6.5–9 mm and teeth of 1.5 mm in length. The magenta to reddish purple colored corolla has a tube 8–16 mm long and median lobe 2.5–3.5 mm long with regular dentate-serrate margin. The upper pair of stamens and style are slightly exserted beyond upper corolla-lip (Fig. 7). Phenology: Flowering and fruiting occur from August to November. Distribution: Grows in the oak woodlands and limestone outcrops of the mountains in southern San Luis Potosí, at elevations 1900–2700 m. Agastache pringlei: This herb grows from 0.2 to 0.8 m high. The leaves are 0.5–4 cm long, broadly deltate-ovate to triangular-lanceolate in shape, with an obtuse-truncate base, acute apex, and crenate-serrate margin. The interrupted inflorescence consists of 15–40 flowered paired cymes each subtended by lanceolateelliptic bracts and separated by cymal internodes 1–5 cm long. The green to pink colored calyx is 5–8 mm long with a tube of 3.5–5 mm and teeth of 1.5–3 mm in length. The pink to rose-purple colored corolla has a tube 8.5–14 mm long and median lobe 1.5–2.3 mm long with several toothed margin. The upper pair of stamens and style are slightly exserted beyond upper corolla-lip (Fig. 8). Phenology:

Fig. 7 (a) A. palmeri var. palmeri growing in the Sierra de Álvarez, San Luis Potosí, and (b) approximation of its inflorescence. (Photos: R. Bye)

Agastache spp. LAMIACEAE. Important Species of Hyssop in Mexico

991

Fig. 8 (a) A. pringlei var. pringlei growing on the slopes of the SMOc and (b) approximation of its inflorescence. (Photos: R. Bye)

Flowering and fruiting occurs July to October. Distribution: It grows on igneous rocks in pine-oak woodlands of the SMOc in western Chihuahua, and adjacent New Mexico, at elevations 1900–2400 m.

Phytochemistry Phytochemical studies have focused primarily on the following three Agastache taxa revealing that the main constituents are terpenoid and flavonoid compounds with known pharmacological activity (Fuentes-Granados et al. 1998; Zielinska and Matkowski 2014). Other compounds have also been reported from other species of Agastache section Brittonastrum and may exhibit bioactivity that explains their employment in Traditional Mexican Medicine. Agastache mexicana subsp. mexicana: Of the eleven compounds isolated from its essential oil, the major ones are estragole, limonene, and linalool (Estrada-Reyes et al. 2004). Estragole, geraniol, linalool, menthone, and pulegone have been found in cultivated plants (Carrillo-Galván et al. 2020). The main constituent reported in its extract is the flavonoid glycoside, acacetin-7-O-β-glucoside; it has a greater abundance of polyphenols and total flavonoids than the A. mexicana subsp. xolocotziana, possibly related to the purple pigmentation of its flowers (Estrada-Reyes et al. 2014). Agastache mexicana subsp. xolocotziana: Of the 28 compounds recorded in its essential oil, pulegone, menthone, and isopulegone represent the majority (EstradaReyes et al. 2004). Estragole, geraniol, and linalool have also been found (CarrilloGalván et al. 2020). In its aqueous extract, as in A. mexicana subsp. mexicana, the main constituent reported is the flavonoid glycoside, acacetin-7-O-β-glucoside (Estrada-Reyes et al. 2014). Agastache palmeri var. breviflora: Geraniol, menthone, and pulegone have been documented in its essential oil (Carrillo-Galván et al. 2020).

992

G. Carrillo-Galva´n and R. Bye

Local Medicinal Uses The “toronjil medicinal complex” is one of the more distinctive ethnobotanical complexes in Mexico (Bye and Linares 2015). Such a complex is defined as an assemblage of taxonomically distinct plants that share certain features including the local or trade name. In addition, the plants may hold in common medicinal uses, morphological features of the utilized part, organoleptic assets and phytochemical composition or pharmacological properties. Each complex consists of a dominant or signature species that is commercialized beyond its natural phytogeographic range while the subordinate species tend to be used within their respective natural phytogeographic ranges or as a substitute for the signature taxon. In addition to the species of Agastache listed here, the toronjil ethnobotanical complex includes the European introduced Dracocephalum moldavica L. (“toronjil azul” and “toronjil chino”) and Melissa officinalis L. (“toronjil” and “melisa”). The signature species is A. mexicana which cultivated and available in markets throughout Mexico. All the species are employed to relieve gastrointestinal ailments, to attend respiratory illnesses and to promote calmness. These herbs share such morphological features square stems, opposite leaves with simple blades and short petioles, and condensed inflorescences of white to purple-colored flowers. The aromatic and citrus-flavored herbs have been employed as condiments and infusions and contain varying concentrations of essential oils well known for their carminative and antibacterial properties. Agastache aurantiaca: In the Sierra Tarahumara, infusion of the leaves is drunk when one has bronchial inflammations, fevers, and various gastrointestinal ailments such as indigestion and colic (Bye 1985). Agatache mexicana subsp. mexicana: In the central states of Mexico such as Mexico City, State of Mexico, Tlaxcala, Morelos, Puebla, Veracruz, and Hidalgo, the infusion of the aerial parts (stem, leaves, and inflorescence) of “toronjil morado” along with “toronjil blanco” and “toronjil azul” is drunk to treatment of gastrointestinal ailments, menstrual pains, and nervousness (Farmacopea Herbolaria de los Estados Unidos Mexicanos 2014; Linares et al. 1999; Martínez 1990). It is also used to combat intense anger or disgust, as well as to treat Diseases of Cultural Affiliation (DCA) such as “scaredness” or “fright” (Argueta et al. 1994; Bye et al. 1987; Carrillo-Galván et al. 2020). “Tlalahuehuetl” was employed as an infusion as an antidysentery and to treat bruises as well as applied as a poultice for healing burns (De la Cruz and Badiano 1964). Agastache mexicana subsp. xolocotziana: It is used in infusion together with “toronjil morado” by residents of Mexico City, State of Mexico, Puebla, Tlaxcala, and Morelos, for the conditions described above. Martínez (1990) reports it as antispasmodic. Agastache micrantha: The Tarahumara living the northern SMOc consume an infusion of the aerial plant parts when treating various illnesses related to the circulatory system, gastrointestinal problems, bronchitis, colic, and to treat insomnia (Bye 1985; Bye 2007; Cardenal 1993; Olivas-Sánchez 1999). Agastache pallida: Inhabitants of the Sierra Tarahumara drink an infusion of the aerial plant parts to alleviate colds, fevers, and coughs as well as to treat various heart

Agastache spp. LAMIACEAE. Important Species of Hyssop in Mexico

993

ailments, gastrointestinal problems (such as infections, diarrhea, gastric ulcers, indigestion, and colic) (Bye 1985). The fresh leaves are placed in the nostrils to facilitate nasal decongestion (Bennet and Zingg 1935; Ford 1975). Agastache palmeri var. breviflora: The residents of southeastern SMO use the infusion aerial parts as a tea and in night baths to treat the “scaredness” and “fright” (DCA) and to calm the nerves (Carrillo-Galván et al. 2017, 2020). Agastache palmeri var. palmeri: In San Luis Potosí, it is used to relieve colds through baths and in infusion it is taken two or three times a day to regulate the menstrual cycle (Bye 1979; Colín 2006). In the nineteenth century, it was infused in alcohol and syrups to produce “mustela,” a popular remedy for indigestion and hiccoughs (Bye 1979). Agastache pringlei: The Tarahumara of Chihuahua employ the whole plant in infusion that is drunk to relieve gastrointestinal pain and fever as well as to induce sleep (Bye 1979, 2007).

Pharmacological Studies Pharmacological studies have been performed using on the two subspecies of Agastache mexicana. Agastache mexicana subsp. mexicana: Laboratory studies have demonstrated antinociceptive and anti-inflammatory effects (González-Ramírez et al. 2012) as well as such properties as spasmogenic (Ventura-Martínez et al. 2017), antifungal (Juárez et al. 2015), and antioxidant (Ibarra-Alvarado et al. 2010). Other research shown its relaxant effect (Navarrete et al. 2017), vasorelaxant effect (HernándezAbreu et al. 2011), and anxiolytic action (Estrada-Reyes et al. 2014). Agastache mexicana subsp. xolocotziana: Similarly, these studies also documented its spasmolytic activity (Ventura-Martínez et al. 2017) as well as sedative properties and anxiolytic action (Estrada-Reyes et al. 2014).

Horticulture and Other Uses In central Mexico, two species are found in traditional homegardens due to their ornamental and medicinal values. In towns along the EVT, A. mexicana subsp. mexicana is popular (Carrillo-Galván et al. 2020; Sanders 1987). In the Huasteca region of southern SMOr, A. palmeri var. breviflora is a valued attractive herb in homegardens while in Chihuahua, the aromatic A. micrantha is a prized among other home-grown aromatic plants used for teas and medicine. Because of the showiness of the inflorescence, many species of Agastache are used in the international horticultural settings; Mexican species available through international commerce include A. mexicana, A. micrantha, and A. pallida (Bailey and Bailey 1976; Huxley and Griffiths 1999). These ornamental species are becoming popular in native landscaping because of they attract insect and bird pollinators (Cheatham and Johnston 1996). Because their natural propensity to hybridize

994

G. Carrillo-Galva´n and R. Bye

(Sanders 1987), more spectacular horticultural selections are becoming available in the markets.

References Argueta A, Cano L, Rodarte M. Atlas de las Plantas de la Medicina Tradicional Mexicana. México, DF: Instituto Nacional Indigenista; 1994. Bailey LH, Bailey EZ. Hortus third – a concise dictionary of plants cultivated in the United States and Canada New York. New York: Macmillan Publishing Co; 1976. Bennet WC, Zingg RM. The Tarahumara, an Indian tribe of Northern Mexico. Chicago: University of Chicago Press; 1935. Bye R. An 1878 ethnobotanical collection from San Luis Potosi: Dr. Edward Palmer’s first major Mexican collection. Econ Bot. 1979;33:135–62. Bye R. Medicinal plants of the Tarahumara Indians of Chihuahua, México. In: Tyson RA, Elerick DV, editors. Two Mummies from Chihuahua: a multidisciplinary study. San Diego: San Diego Museum Paper. No. 19; 1985. p. 77–104. Bye R. Medicinal plants of the Sierra Madre: comparative study of Tarahumara and Mexican market plants. Econ Bot. 1986;40:103–24. Bye R, Linares E, Ramamoorthy TP, García O, Collera F, Palomino G, Corona V. Agastache mexicana subsp. xolocotziana (Lamiaceae), new taxon from the Mexican medicinal plants. Phytologia. 1987;62:157–63. Bye R. Un estudio de la flora y sus usos en la Sierra Tarahumara, Chihuahua, México – 1885. México, DF: Reporte Comisión Nacional para el Conocimiento y Uso de la Biodiversidad X006. 2007. Bye R, Linares E. Perspectives on ethnopharmacology in Mexico. In: Heinrich M, Jäger A, editors. Ethnopharmacology. New York: Wiley; 2015. p. 393–404. Carrillo-Galván G, Bye R, Eguiarte L. Domesticación de plantas medicinales aromáticas. In: Casas A, Torres-Guevara J, Parra-Rondinel F, editors. Domesticación en el Continente Americano. CDMX: UNAM-UNALM: México; 2017. p. 431–49. Carrillo-Galván G, Bye R, Eguiarte L, Cristians S, López-Pérez P, Vergara-Silva F, Luna-Cavazos M. Domestication of aromatic medicinal plants in Mexico: Agastache (Lamiaceae) – an ethnobotanical, morpho-physiological, and phytochemical analysis. J Ethnobiol Ethnomed. 2020;16:1–16. Cardenal F. Remedio y practica curativas en la sierra Tarahumara. Chihuahua: Editorial Camino; 1993. Cheatham S, Johnston MC. The useful wild plants of Texas, the southeastern and southwestern United States, the southern plains, and northern Mexico. Austin: Useful Wild Plants, Inc; 1996. Colín M. Contribución al Estudio Etnobotánico de la Flora Medicinal en la Comunidad de San Francisco, Municipio de Zaragoza, Sierra de Álvarez, San Luis Porosí, México. Tesina. México, Distrito Federal: Universidad Autónoma Metropolitana-Xochimilco; 2006. Cruz-Hernández, A. S. Manual para el Manejo Sustentable de Plantas Medicinales. Sierras Verdes, Instituto Nacional de Desarrollo Social. 2014. https://doi.org/CS-09-F-DI-308-14. De la Cruz M, Badiano J. Libellus de medicinalibus indorum herbis. Manuscrito azteca de 1552. Según traducción latina de Juan Badiano. Instituto Mexicano del Seguro Social. Segunda ed. Fondo de Cultura Economica: México; 1964. Estrada-Reyes R, Aguirre Hernández E, García-Argáez A, Soto Hernández M, Linares E, Bye R, Heinze G, Martínez-Vázquez M. Comparative chemical composition of Agastache mexicana subsp. mexicana and A. mexicana subsp. xolocotziana. Biochem Syst Ecol. 2004;32:685–94. Estrada-Reyes R, López-Rubalcava C, Ferreyra-Cruz OA, Dorantes-Barrón A, Heinze G, Moreno Aguilar J, Martínez-Vázquez M. Central nervous system effects and chemical composition of

Agastache spp. LAMIACEAE. Important Species of Hyssop in Mexico

995

two subspecies of Agastache mexicana; an ethnomedicine of Mexico. J Ethnopharmacol. 2014;153:98–110. Farmacopea Herbolaria de los Estados Unidos Mexicanos. 2014. (FHEUM). Ford KC. La Yerbas de la Gente: a study of Hispano-Aamerican medicinal plants. Museum of Anthropology, Anthropological Papers no 60. Ann Arbor: Unvieristy of Michigan; 1975. Fuentes-Granados RG, Widrlechner M, Wilson L. An overview of Agastache research. Int J Geogr Inf Syst. 1998;6:1–8. González-Ramírez A, González-Trujano M, Pellicer F, López-Muñoz F. Anti-nociceptive and antiinflammatory activities of the Agastache mexicana extracts by using several experimental models in rodents. J Ethnopharmacol. 2012;142:700–5. Hernández-Abreu O, Durán-Gómez L, Best-Brown R, Villalobos-Molina R, Rivera-Leyva J, Estrada-Soto S. Validated liquid chromatographic method and analysis of content of tilianin on several extracts obtained from Agastache mexicana and its correlation with vasorelaxant effect. J Ethnopharmacol. 2011;138:487–91. Huxley A, Griffiths M. The new Royal Horticultural Society Dictionary of gardening. London: Macmillan Reference Ltd; 1999. Ibarra-Alvarado C, Rojas A, Mendoza S, Bah M, Gutiérrez DM, Hernández- Sandoval L, Martínez M. Vasoactive and antioxidant activities of plants used in Mexican traditional medicine for the treatment of cardiovascular diseases. Pharm Biol. 2010;48:732–9. Juárez Z, Hernández L, Bach H, Sánchez-Arreola E, Bach H. Antifungal activity of essential oils extracted from Agastache mexicana ssp xolocotziana and Porophyllum linaria against postharvest pathogens. Ind Crops Prod. 2015;74:178–82. https://doi.org/10.1016/j.indcrop.2015. 04.058. Linares E, Bye R, Flores B. Plantas Medicinales de México/Medicinal Plants of Mexico. México, Distrito Federal: Sistema de Información Geográfica, SA de CV, Universidad Nacional Autónoma de México; 1999. Martínez M. Catálogo de nombres vulgares y científicos de plantas mexicanas. México, Distrito Federal: Fondo de Cultura Económica; 1979. Martínez M. Las Plantas Medicinales de México. México, Distrito Federal: Ediciones Botas; 1990. Martínez-Gordillo M, Bedolla-García B, Cornejotenorio G, Fragoso-Martínez I, García-Peña MDR, González-Gallegos JG, Lara-Cabrera SI, Zamudio S. Lamiaceae de México. Bot Sci. 2017;95: 780–806. Martínez-Gordillo M, Fragoso-Martínez I, García-Peña MDR, Montiel O. Géneros de Lamiaceae de México, diversidad y endemismo. Rev Mex Biodivers. 2013;84:30–86. Navarrete A, Ávila-Rosas N, Majín-León M, Balderas-López JL, Alfaro-Romero A, TavaresCarvalho JC. Mechanism of action of relaxant effect of Agastache mexicana ssp. mexicana essential oil in guinea-pig trachea smooth muscle. Pharm Biol. 2017;55:96–100. Olivas-Sánchez P. Plantas Medicinales del Estado de Chihuahua. Cd. Juárez, Chihuahua: Universidad Autónoma de Ciudad Juárez; 1999. Sanders RW. Taxonomy of Agastache section Brittonastrum (Lamiaceae-Nepeteae). Syst Bot Monogr. 1987;15:1–92. Santillán-Ramírez M, López-Villafranco ME, Aguilar-Rodríguez S, Aguilar-Contreras A. Estudio etnobotánico, arquitectura foliar y anatomía vegetativa de Agastache mexicana ssp. mexicana y A. mexicana ssp. xolocotziana. Rev Mex Biodivers. 2008;79:513–24. Ventura-Martínez R, Rodríguez R, González-Trujano ME, Ángeles-López GE, Déciga-Campos M, Gómez C. Spasmogenic and spasmolytic activities of Agastache mexicana ssp. mexicana and A. mexicana ssp. xolocotziana methanolic extracts on the guinea pig ileum. J Ethnopharmacol. 2017;196:58–65. Wyndham FS. Environments of learning: Rarámuri children’s plant knowledge and experience of schooling, family, and landscapes in the Sierra Tarahumara, Mexico. J Hum Ecol Interdiscip. 2010;38:87–99. Zielinska S, Matkowski A. Phytochemistry and bioactivity of aromatic and medicinal plants from the genus Agastache (Lamiaceae). Phytochem Rev. 2014;13:391–416.

Agave americana L. Agave angustifolia Haw. Agave atrovirens Karw. ex Salm-Dyck. Agave asperrima Jacobi. Agave bovicornuta Gentry. Agave cupreata Trel. & A. Berger. Agave hookeri Jacobi. Agave inaequidens K. Koch. Agave karwinskii Zucc. Agave kerchovei Lem. Agave lechuguilla Torr. Agave mapisaga Trel. Agave marmorata Roezl Agave maximiliana Baker Agave montana Villareal Agave potatorum Zucc. Agave rhodacantha Trel. Agave salmiana Otto ex Salm-Dyck Agave scaposa Gentry Agave tequilana F.A.C. Weber Agave victoriae-reginae A. Berger ASPARAGACEAE

© Springer Nature Switzerland AG 2023 A. Casas, J. J. Blancas Vázquez (eds.), Ethnobotany of the Mountain Regions of Mexico, Ethnobotany of Mountain Regions, https://doi.org/10.1007/978-3-030-99357-3_26

997

998

I. Torres-García et al.

Ignacio Torres-García, Ame´rica Minerva Delgado-Lemus, Alejandro Casas, Gonzalo D. A´lvarez-Ríos, Selene Rangel-Landa, Raymundo Martínez-Jime´nez, Carmen Julia Figueredo-Urbina, Ofelia Vargas-Ponce, Guadalupe Casarrubias-Herna´ndez, Oassis Huerta-Galva´n, Da´nae Cabrera-Toledo, and Nancy Va´zquez-Pe´rez Synonyms Agave americana L.: Agave altissima Zumagl.; Agave americana var. marginata Trel.; Agave americana var. mediopicta (Hovey) Trel.; Agave americana var. picta (Salm-Dyck) A.Terracc.; Agave americana f. picta (Salm-Dyck) Voss.; Agave americana var. striata (Hovey) Trel.; Agave cordillerensis Lodé & Pino (Gentry 1982; González-Elizondo et al. 2009; Tropicos 2020; POWO 2020). Agave angustifolia Haw.: Agave aboriginum Trel.; Agave pacifica Trel.; A.Terracc.; Agave vivipara var. nivea (Trel.) P.I.Forst (Gentry 1982, García-Mendoza & E. J. Lott, 1994; Tropicos 2020; POWO 2020). Agave asperrima Jacobi: Agave caeciliana A.Berger.; Agave scabra SalmDyck.; Agave wislizeni Engelm (Gentry 1982; González-Elizondo et al. 2009; Tropicos 2020; POWO 2020). Agave atrovirens Karw. ex Salm-Dyck: Agave macroculmis Tod.; Agave mirabilis Trel.; Agave ottonis Jacobi (Gentry 1982; Tropicos 2020; POWO 2020). Agave bovicornuta Gentry (Gentry 1982; González-Elizondo et al. 2009; Villaseñor 2016; POWO 2020). Agave cupreata Trel. & A. Berger (González-Elizondo et al. 2009; Villaseñor 2016; POWO 2020).

I. Torres-García (*) · A. M. Delgado-Lemus · A. Casas · S. Rangel-Landa MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico e-mail: [email protected]; [email protected]; [email protected] G. D. Álvarez-Ríos MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico Laboratorio Manejo y Evolución de Recursos Genéticos, Instituto de Investigaciones en Ecosistemas y Sustentabilidad, Universidad Nacional Autónoma de México, Morelia, Michoacán, Mexico e-mail: [email protected] R. Martínez-Jiménez Investigador independiente, Villa Sola de Vega, Oaxaca, Mexico

Agave americana L.. . .

999

Agave hookeri Jacobi: (Gentry 1982; Tropicos 2020; POWO 2020). Agave inaequidens K. Koch: Agave bourgaei Trel.; Agave crenata Jacobi.; Agave fenzliana Jacobi (Gentry 1982; POWO 2020). Agave karwinskii Zucc.: Agave bakeri H.Ross; Agave corderoyi Verschaff. ex Baker; Agave laxa Salm-Dyck (Gentry 1982; Tropicos 2020; POWO 2020). Agave kerchovei Lem.: Agave convallis Trel.; Agave dissimulans Trel.; Agave expatriata Rose (Gentry 1982, CONABIO 2009; Tropicos 2020; POWO 2020). Agave lechuguilla Torr.: Agave lecheguilla Torr.; Agave heteracantha Jacobi; Agave lophantha var. poselgeri (Salm-Dyck) (Gentry 1982; González-Elizondo et al. 2009; Tropicos 2020; POWO 2020). Agave mapisaga Trel.: Agave mapisaga var. lisa Gentry.; Agave salmiana var. angustifolia A.Berger (Gentry 1982; Tropicos 2020; POWO 2020). Agave marmorata Roezl: Agave todaroi Baker (Gentry 1982; Tropicos 2020; POWO 2020). Agave maximiliana Baker: Agave conjuncta A. Berger.; Agave gustaviana IDEMaage & IDEMchmidt.; Agave katharinae A. Berger (POWO 2020). Agave montana Villareal (POWO 2020).

C. J. Figueredo-Urbina Cátedra CONACYT. Centro de Investigaciones Biológicas, Instituto de Ciencias Básicas e Ingeniería, Universidad Autónoma del Estado de Hidalgo, Mineral de la Reforma, Hidalgo, Mexico O. Vargas-Ponce Laboratorio Nacional de Identificación y Caracterización Vegetal (LaniVeg-CONACYT), Departamento de Botánica y Zoología, Centro Universitario de Ciencias Biológicas y Agropecuarias, Universidad de Guadalajara, Zapopan, Jalisco, Mexico e-mail: [email protected] G. Casarrubias-Hernández Facultad de Ciencias Químico-Biológicas, Universidad Autónoma de Guerrero, Chilpancingo, Guerrero, Mexico O. Huerta-Galván MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico Maestría en Biosistemática y Manejo de Recursos Forestales y Agrícolas, Centro Universitario de Ciencias Biológicas y Agropecuarias, Universidad de Guadalajara, Zapopan, Jalisco, Mexico D. Cabrera-Toledo MILPA (Manejo Integral y Local de Productos Agroforestales) Asociación Civil, Morelia, Michoacán, Mexico Laboratorio Nacional de Identificación y Caracterización Vegetal (LaniVeg-CONACYT), Departamento de Botánica y Zoología, Centro Universitario de Ciencias Biológicas y Agropecuarias, Universidad de Guadalajara, Zapopan, Jalisco, Mexico N. Vázquez-Pérez Centro de Investigación en Biodiversidad y Conservación, Universidad Autónoma del Estado de Morelos, Cuernavaca, Morelos, Mexico

1000

I. Torres-García et al.

Agave potatorum Zucc.: Agave amoena Lem. ex Jacobi.; Agave scolymus Karw. ex Salm-Dyck; Agave verschaffeltii Lem. ex Jacobi (Gentry 1982; Tropicos 2020; POWO 2020). Agave rhodacantha Trel. (POWO 2020). Agave salmiana Otto ex Salm-Dyck: Agave chinensis F.P.Sm.; Agave coarcta Jacobi; Agave cochlearis Jacobi (Gentry 1982; POWO 2020). Agave scaposa Gentry (POWO 2020). Agave tequilana F.A.C. Weber: Agave angustifolia subsp. tequilana (F.A.C. Weber) Valenz.-Zap. & Nabhan; Agave pes mulae Trel.; Agave pseudotequilana Trel. (POWO 2020). Agave victoriae-reginae A. Berger: Agave ferdinandi-regis A. Berger; Agave nickelsii Rol.-Goss.; Agave victoriae-reginae var. laxior Berger (Gentry 1982; CONABIO 2009; González-Elizondo et al. 2009; Tropicos 2020; POWO 2020).

Local Names Agave americana: Cenizo, Chichimeco de castilla, Maguey coyote, Maguey de castilla, Maguey de coyote, Maguey manso, Maguey meco, Maguey meco rayado, Maguey pinto, Maguey serrano, Maguey verde, Mezcal (Spanish), Mecometl, Metl, Mexoxochitl, Teometl (Nahua), Hok’uada, Mbängät’ax’uada, T’ax’uada, Taxi huada, Taxihuada, Uadä, Uadá, Uanthe (Otomí), Cachrots’ánta (Popoloca), I’gok pinto, Sorril i’gok (Tepehuano). Aguascalientes: Pita (Spanish). Baja California Sur: Maguey blanco. Chiapas: Muk’ta met, Muk’ta j’alnom met, Yax-met (Tzotzil). Chihuahua: Galime (Tarahumara). Coahuila: Maguey cenizo (Spanish). Durango: Maguey blanco, Maguey cenizo, Maguey chino (Spanish), I’gok (Tepehuano). Hidalgo: Vanthe, Nbandat’ax’uada (Otomí). Jalisco: Maguey azul, Maguey blanco, Maguey de pulque (Spanish). Michoacán: Maguey blanco, Maguey carrizaleño, Maguey cenizo (Spanish). Morelos: Maguey azul, Maguey blanco, Maguey pulquero (Spanish). Nuevo León: Maguey, Agave achimeco, Agave amarillo, Agave meco, Agave teometl, Agave amarillo, Pita, Maguey aguamielero (Spanish). Oaxaca: Arroqueño, Maguey pulquero, Maguey sierrudo, Maguey Ceniza, Maguey Coyote, Ruqueño, Maguey Sierrudo, Maguey sierra negra, Xolo, Maguey de rayo, Maguey xolo blanco (Spanish), Yavi cuam (Mixteco), Doba nupy, Dòb-dzìn, Dòb-mpiè, Dobncjp, Dua-bsug, Dua-bzog, Dua-pchez, Dua-ya-do, Dua-yesh (Zapoteco). Puebla: Maguey chichimeco (Spanish). Querétaro: Maguey fino (Spanish). San Luis Potosí: Maguey chino (Spanish). Tamaulipas: Maguey cenizo (Spanish) (Álvarez-Ríos et al. 2020a; Caballero et al. 1984–2020; Hunn 2008, Colunga-GarcíaMarín et al. 2007; González-Elizondo et al. 2009; García-Mendoza and Franco-Martínez 2018; González-Elizondo et al. 2009; Martínez-Jiménez et al. 2019; Ríos-Colín 2017). Agave angustifolia: Maguey, Maguey de la barranca, Maguey de los bosques, Mezcal bravo, Árbol xixique, Apanguero, Prieto apanguero, Amole, Alista, Mezcal casero, Maguey cincoañero, Mezcal criollo, Chacaleño, Maguey delgado, Garabato, Ixtle manso, Maguey de ixtle, Mezcal de la ladera, Maguey de lechuguilla, Maguey

Agave americana L.. . .

1001

liso, Maguey manso, Maguey mescaler, Agave mescalero, Mezcalillo, Tepemete, Maguey pintillo, Peruano gigante, Saguayo, Sigüin, Sihuin, Xigüin, Maguey teteleño, Verde cenizo, Zapul, Zapupe (Spanish). Metl (Nahua). Weey (Huasteco) Chihuahua: Gusime, Ku’uri (Tarahumara). Guerrero: Zacatoro (Nahua), Jalisco: Tepemete, Gubuk, Guvúkai (Tepehuano), Brocha, Azul telcruz, Lineño, Presa grande, Cenizo, Ciriaco, Cimarron, Cuaquesoca, Cabresto, Negro cimarron, Media penca, Negro telcruz, Pinole, Pencudo, Chancuellar, Ixtero verde, Enano, Delgadillo, Mezcal amarillo, Cimarrón espinoso, Chautesoca, Soca, Mezcal grande, Palmero, Alineño, Motosierra, Mezcal blanco, Sierrilla (Spanish). Michoacán: Espadancillo, Espadilla, Cuijillo (Spanish). Morelos: Maguey Espadín (Spanish). Oaxaca: Espadín, Espadancillo, Maguey de gobierno, Maguey de lumbre Mexicanito, Maguey de campo, Mexicano azul espadilla, Espadilla blanca, Espadilla colorada, Maguey de ixtle, Maguey quixe (Spanish), Zu va eh-zha (Ixcateco), Dòbmèzcâl, Dob yee, Doba yej (Zapoteco). Yavi chucu, Yavi tachucu, Yahui xiuco, Yavi incoyo (Mixteco) Puebla: Agave mescalero, Mahuey mescalillo. Quintana Roo: Tapemete, Babki, Xtuk-ki, Xixki (Maya). Sonora: Hamoc, Coptoj (Seri), Lechuguilla, Magí (Guarijío), Maguey de mesca, Maguey bacanora, Bacanora (Spanish) Cu0 u, Juya cu’u (Mayo). San Luis Potosí: Sapupe, Zapupe, Weey (Huasteco). Sinaloa: Maguey cimarrón, Mezcal (Spanish-Mayo). Tamaulipas: Caña de jabalí (Spanish). Veracruz: Agave zapupe, Ixtle, Maguey de ixtle, Quiote, Zapupe de Tepetzintla, Zapul, Zapupe de Tantoyuca, Zapupe silvestre, Zapupe verde (Spanish) Yucatán: Chelem amarillo, Chelem blanco, Chelem verde, Che’elem-ki o, Ch’ekem ki o, Ch’ekemki o, Chelem-ki, Chucum ci, Chucum-ci, Chucumqui, Chukum ki, Chukumki, Chukunki,Kitam ki, Kitamki, Quitan ci, Quitamqui, Citam, Citamci, Pita ki, Pitaci, Xtu ki (Maya) (Caballero et al. 1984–2020, Colunga-GarcíaMarín et al. 2007, García-Mendoza and Franco-Martínez 2018; González-Elizondo et al. 2009). Agave asperrima: Maguey bruto, Maguey cenizo, Maguey de cerro, Lechuguilla, Maguey mescal, Maguey serrano, Maguey rasposo (Spanish). Durango: Lamparillo (Spanish) (Colunga-GarcíaMarín et al. 2007; González-Elizondo et al. 2009). Agave atrovirens: Chichimeco, Comiteco, Maguey divino, Cimarrón fino, Cimarrón inferior, Cimarrón morado, Maguey jabalín, Maguey de venado, Maguey de penca larga, Lechuguilla, Maguey manco fino, Maguey manco verdadero, Maguey manso fino, Maguey manso, Maguey de pulque, Agave pulquero, Tuah (Spanish). Centemetl, Mazametl, Mecuametl, Mecoametl, Metl, Metlchichitl, Teometl, Tepalcametl, Tepeme, Tepemetl, Tepemexcatl, Tepemexcall, Tlacametl (Nahua). Tobalá (Zapoteco).Tsaats (Mixe). Tziyé (Otomí). Puebla: Mezote, Maguey manso, Maguey pulquero, Maguey blanco, Maguey fino, Maguey bronco, Maguey cenizo (Spanish). Matehonti, Metehontic (Nahua). Oaxaca: Maguey Blanco, Maguey Manso, Maguey pulquero, Maguey cenizo, Maguey de cumbre, Maguey de Montaña (Spanish). Dòb-bææl, Dojb, Dob-buEl,” Dob-cuEl (Zapoteco). Estado de México: Maguey pulquero. Hidalgo: Maguey manso (Spanish) (Caballero et al. 1984–2020; Hunn 2008, Colunga-GarcíaMarín et al. 2007; García-Mendoza and Franco-Martínez 2018).

1002

I. Torres-García et al.

Agave bovicornuta: Chihuahua: Lechuguilla, Maguey (Spanish), Sapuli, Watusa, Imé (Tarahumara). Durango: masparilla (Spanish), Sapulh (Tepehuano). Sinaloa: Lechuguilla, Lechuguilla verde (Spanish), Sabali, Sapuli, Noriba (Mayo). Sonora: Sapari, Temuchí, Mezcal (Guarijío) (Caballero et al. 1984–2020; ColungaGarcíaMarín et al. 2007; González-Elizondo et al. 2009). Agave cupreata: Michoacán: Maguey chino (Spanish). Guerrero: Maguey papalote, Maguey ancho, Maguey cimarrón, Maguey bravo, Maguey de mescal, Tuchi (Spanish), Papalometl, Papalome (Nahua). Yaabendisii, Yaave ndishi (Mixteco) (Caballero et al. 1984–2020; Colunga-GarcíaMarín et al. 2007). Agave hookeri: Maguey (Spanish). Ixquitécatl (Nahua). Michoacán: Maguey Manso, Maguey pulquero (Spanish), Vaca Akamba (Purhépecha) (ColungaGarcíaMarín et al. 2007; Colunga-GarcíaMarín et al. 2017). Agave inaequidens: Agave mescalero, Saguayo (Spanish). Maguey hocimetl, Hocimetl (Nahua). Durango: Maguey cimarrón, Maguey mayero, Maguey de la barranca, Maguey negro, Mezcalillo (Spanish). Michoacán: Maguey alto, Maguey bruto, Mezcal bruto, Maguey negro, Mezcal (Spanish) Akamba, Toro Akamba (Purhépecha). Jalisco: Maguey bravo, Mezcal bravo (Spanish). Guerrero: Maguey mescalero (Spanish), Yaave ndishi (Mixteco). Morelos: Maguey (Spanish) (Caballero et al. 1984–2020; Colunga-GarcíaMarín et al. 2007; González-Elizondo et al. 2009; Torres et al. 2015). Agave karwinskii: Oaxaca: Madrecuishe, Mezcal, Bicuishe, Bicuixe, Candelillo, Cuishe, Kuish, Marteño, Barril, Maguey barril verde, Maguey barril amarillo, Maguey barril chino, Maguey barril gordo, Cirial, Espina negra, Sierrudo, Maguey San Martín, San Martinero, Largo, Tripón, Tobasiche, Tabaziche, Tobaxiche, Dòb-cuǐx, Chamisa, Doba-cirial, Dob-cirial, Do-cirial, Maguey cirial, Siriaar, Doba-sh’che (Spanish-Zapoteco). Al-mal-bi-cuish (Chontal). Puebla: Ixtle (Nahua), Cachitún, Cachutum, Cacaya de cachutum (Popoloca) (Caballero et al. 1984–2020; Hunn 2008; García-Mendoza and Franco-Martínez 2018; VázquezPérez et al. 2020, Figueredo-Urbina et al. 2017; Colunga-GarcíaMarín et al. 2007). Agave kerchovei: Puebla: Cacaya, Maguey de conejo, Maguey escobeta, Maguey de pasmo, Rabo de león, Huitzotzi, Cacaya de tunecho, Tunecho de cacaya, Tunecho de ixcle, Ixtle, Ixtle amarillo, Lechuguilla dura, Cacallas (Spanish-Nahua). Cachrojá (Popoloca). Oaxaca: Maguey tieso, Ixtle, Lechuguilla dura, Maguey jabalí, Rabo de león (Spanish) (Caballero et al. 1984–2020; Colunga-GarcíaMarín et al. 2007; García-Mendoza and Franco-Martínez 2018; Palma-Cruz 2000). Agave lechuguilla: Lechuguilla amol, Lechuguilla corriente, Lechuguilla de cerro, Lechuguilla verde (Spanish). Jarcia (Tepehuano). Metometl (Nahua). Coahuila: Lechuguilla, Maguey, Pita, Tzeth, Tzuta. Hidalgo: Lechuguilla, Amole (Spanish), Ts’u’ta, Hogäts’u’ta, Tzita (Otomí). Nuevo León: Lechuguilla, Amole, Amole dulce, Tzuta, Ixtle (Spanish). Puebla: Agave lechuguilla (Spanish). San Luis Potosí: Lechuguilla (Spanish) (Caballero et al. 1984–2020; Colunga-GarcíaMarín et al. 2007; González-Elizondo et al. 2009). Agave mapisaga: Agave pulquero, Cimarrón, Maguey de pitzahua, Maguey flaco, Maguey verde penca larga, Mepichahua, Mepichahuac, Mepizahua, Mepizahuac, Xayametl (Spanish-Náhuatl). Ciudad de México: Maguey

Agave americana L.. . .

1003

aguamielero, Maguey liso, Maguey mano larga, Maguey manos largas, Maguey manso, Maguey pulquero (Spanish). Estado de México: Maguey carrizo, Maguey penca larga, Maguey verde (Spanish), Xilometl (Nahua). Hidalgo: Maguey mexicano, Maguey penca larga (Spanish), Ma’ye (Otomí). Jalisco: Maguey de pulque (Spanish). Michoacán: Maguey listoncillo, Maguey tarímbaro (Spanish). Oaxaca: Maguey penca larga, Maguey manos largas, Maguey mano delgada, Maguey mexicano, Maguey poblano, Pulquero (Spanish). San Luis Potosí: Maguey blanco, Maguey de quiotillo, Maguey gris, Maguey jilotillo, Maguey liso, Maguey manso, Maguey mexicano, Maguey mielero, Maguey penca larga (Spanish), May’e (Otomí). Oaxaca: Maguey penca larga, Maguey manos largas, Maguey mano delgada, Maguey mexicano (Spanish). Puebla: Maguey sabililla (Spanish), Xilometl (Nahua). Tamaulipas: Maguey manso (Spanish), Maguey xa’mini (Otomí). Tlaxcala: Maguey ayoteco, Maguey carrizo, Maguey morado, Maguey palmilla, Maguey sabililla (Spanish), Xilometl (Nahua) (Álvarez-Ríos et al. 2020b; Caballero et al. 1984–2020; Colunga-GarcíaMarín et al. 2007; García-Mendoza and Franco-Martínez 2018; Mora-López et al. 2011; Reyes-Agüero et al. 2019; Trejo et al. 2020). Agave marmorata: Oaxaca: Becuela, Du cual, Dòb-pcuêl (Zapoteco). Tepeztate, Tepextate (Nahua). Maguey caballo, Maguey curandero (Spanish). Puebla: Pichomel, Picho, Pichu, Pitzometl, Huiscole (Spanish-Nahua) (Caballero et al. 1984–2020; Colunga-GarcíaMarín et al. 2007; Flores-Maya et al. 2015). Agave maximiliana: Jalisco: Lechuguilla, Raicilla, Tecolote, Maguey de la sierra, Maguey cenizo, Maguey manso fino, Maguey manso, Masparillo (Spanish). A’hl mai, Ahl may (Tepehuano). Maxo (Otomí). Teometl (Nahua) (Caballero et al. 1984–2020; Colunga-GarcíaMarín et al. 2007; Vázquez-García et al. 2007). Agave potatorum: Arruqueño, Magueycillo, Maguey de mescal, Maguey del monte, Maguey de cochi, Maguey de pasmo (Spanish). Puebla: Papalometl, Papalomé, Cacaya, Cantarorocacaya (Nahua). Magueycillo (Spanish). Oaxaca: Maguey papalome (Spanish-Nahua). Tobalá, Toabalá, Tobalán, Toeb bá lá, Dóbbè, Dòb-pùrêl, Dòb-tòbàlâ, Doba-lá, Dob-lá, Do-lá, Dob bé, Batobpaz (Zapoteco). Ya vií, Ya vi ticuchi, Yauiticushi (Mixteco). Zu nllishe (Ixcateco). Jä näk tsääjts (Mixe) (Caballero et al. 1984–2020; Hunn 2008; García-Mendoza 2010; ColungaGarcíaMarín et al. 2007). Agave rhodacantha: Criollo del cerro, Mezcal cimarrón, Cuchara, Cucharo, Mezcal cucharo, Maguey del monte, Quixe, Sierrilla, Sierrilla amarillo, Sierrilla criollo, Sierrilla del cerro, Sierrilla verde, Verde negro, Verde rápido, Cimarrón verde rápido, Verde, Cimarrón, Criollo (Spanish). Jalisco: Ixtero amarillo, Mexicano, Maguey mexicano, Mezcal ixtero, Mezcal amarillo (Spahish-Nahua). Nayarit: Mezcal cimarrón (Spanish). Oaxaca: Cuishe, Dobada (Zapoteco). Barril (Spanish) (Caballero et al. 1984–2020; Colunga-GarcíaMarín et al. 2007; González-Elizondo et al. 2009). Agave salmiana: Maguey bueno, Maguey cornudo, Maguey corriente toluqueño, Maguey de púa, Maguey grande, Maguey manos largas, Maguey puya larga, Maguey santo domingo, Maguey toluqueño, Uña de gato (Spanish), Metl, Teometl, Tlacametl (Nahua). Aguascalientes: Maguey blanco, Maguey pulquero (Spanish).

1004

I. Torres-García et al.

Ciudad de México: Maguey chalqueño, Maguey manso, Maguey pulquero (Spanish). Coahuila: Maguey blanco, Maguey cenizo, Maguey verde, Maguey verde cuerudo (Spanish). Durango: Maguey cimarrón, Maguey manso, Maguey verde, Maguey verde silvestre (Spanish), Mbänuada (Otomí), I’gok jiguiarum (Tepehuano). Estado de México: Maguey amapola, Maguey ayoteco, Maguey manso, Maguey negro, Maguey verde, Maguey pico de gorrión (Spanish), Mäx’o´ uada, Mudha (Otomí). Guanajuato: Maguey cimarrón, Maguey fino, Maguey pulquero (Spanish). Hidalgo: Maguey amarillo, Maguey blanco, Maguey chino, Maguey chino cenizo, Maguey corriente, Maguey corriente cenizo, Maguey corriente colorado, Maguey cuerno, Maguey espina china, Maguey manso, Maguey manso de zoqui, Maguey penca ancha, Maguey penca larga, Maguey poblano, Maguey púa larga, Maguey pulquero, Maguey verde, Maguey verde cimarrón (Spanish), Xilemetl, Xilomén, Xilometl (Nahua), Bo’ta, Dämni, Doo´ mini, Gää’mini, Gääx’mini, Hbank’uada, Hok ´uada, Ithui chico, Ithui grande, K0 ank’uada, Mäx’o´uada, Määxo, Mayeé, Mbän k’uada, Mbänuada, M’ondat’ax’uada, Mu’ta, Muthá, Nta’ä’mini, Ntsamni, Tash huadá, Taxi huada, Tsam’niuada Xa’mini, Xhi’ñu, Xinye (Otomí). Amapola, Ayoteco, Chalqueño, Chino, Chino cenizo, Dämni, Ithui grande, Hok’uada, K0 ank’uada, Mäx’o´uada, Muda, Mutha, Púa larga, Xaminí, Gääx’mini,Ithui chico, Verde, Pico de gorrión, Xinye (Hñähñú and Spanish) (Reyes-Agüero et al. 2019). Jalisco: Maguey cenizo, Maguey de pulque, Maguey macho, Maguey verde (Spanish), Mezcal (Nahua). Michoacán: Maguey negro, Maguey verde (Spanish). Nuevo León: Maguey de pulque, Maguey pulquero (Spanish). Morelos: Maguey pulquero, Maguey verde (Spanish). Oaxaca: Maguey calzonudo, Maguey cimarrón, Maguey de pulque, Maguey manso, Maguey mano ancha (Spanish), Exmo´ tstsaajts (Mixe), Yavi cui, Yavi incoyo (Mixteco), Dòb-gú-lò, Dòb-lò (Zapoteco). Puebla: Maguey blanco, Maguey bronco, Maguey del monte, Maguey pinto, Maguey de puya, Maguey pulquero, Maguey rayado, Maguey verde (Spanish), Mexcalli mateuonti (Nahua), Chachro, Káçü, Káçüyüá (Popoloca). Querétaro: Maguey cimarrón, Maguey manso, Maguey verde (Spanish). San Luis Potosí: Maguey blanco, Maguey bronco, Maguey chino, Maguey chino cenizo, Maguey chino venadito, Maguey chino verde, Maguey cimarrón, Maguey cuernavaca, Maguey cuerno, Maguey manso venado, Maguey pulquero, Maguey verde (Spanish). Tamaulipas: Maguey manso (Spanish). Tlaxcala: Maguey amarillo, Maguey ayoteco, Maguey blanco, Maguey chalqueño, Maguey chino, Maguey colorado, Maguey manso, Maguey matecón, Maguey prieto, Maguey púa larga, Maguey verde, Manso listado, Prieto, Prieto silvestre, Prieto listado (Spanish), Xilometl, Wexomelt, Tepezorra (Nahua) (Trejo et al. 2020). Veracruz: Amarillo (Spanish). Zacatecas: Maguey aguamielero, Maguey pulquero (Spanish) (Álvarez-Ríos et al. 2020b; Caballero et al. 1984–2020; Hunn 2008; Colunga-GarcíaMarín et al. 2007; González-Elizondo et al. 2009; Mora-López et al. 2011; Reyes-Agüero et al. 2019; Trejo et al. 2020). Agave scaposa: Puebla: Maguey de potrero (Spanish). Oaxaca: Maguey de potrero, Maguey cimarrón, Maguey de caballo, Maguey de pitzorra, Tequiol, Maguey verde (Spanish) (Caballero et al. 1984–2020; García-Mendoza and Franco-Martínez 2018; Torres 2004).

Agave americana L.. . .

1005

Agave tequilana: Jalisco: Maguey azul, Agave azul Azul criollo, Azul listado, Azul, Tequila, Azulillo, Maguey Coyote blanco, Chino azul, Mezcal chino, Cimarrón cenizo, Mezcal, Maguey tequilero, Mezcal azul tequilero, Tequila azul, Bermejo, Espadín, Mano, Mano achanque, Mano de achanque, Mano de mula, Mano larga, Moraleño, Moraneño, Pata de mula, Pie de mula, Perempís, Siriaco, Raicilla, Saguayo, Següin, Sehuin, Siguín, Sigüín, Sihuin, Siquin, Xigüin, Zigguin, Zingüin, Telcruz, Maguey verde, Zapalote, Zapupe, Zopilochino, Zopilote (Spanish). Michoacán: Maguey azulillo, Mezcal azulillo, Maguey tequilero. Estado de México: Maguey tequilero (Caballero et al. 1984–2020; Colunga-GarcíaMarín et al. 2007). Agave victoriae-reginae: Nuevo León: Lechuguilla, Maguey noha, Maguey de roca, Noha, Noa, Agave de la reyna Victoria, Espadín, Pintilla, Maguey mescalero (Spanish) (Caballero et al. 1984–2020, González-Elizondo et al. 2011).

Botany and Ecology For thousands of years, plants of the genus Agave L. have represented crucial resources for practically all human cultures that have developed in the Mexican territory. From prehistoric times until the present, these plants have been indispensable in many rural communities, for covering almost all the basic needs (ColungaGarcíaMarín et al. 2017; Torres-García et al. 2019a). The most ancient evidence of its use are remains of chewed fibers of the roasted plant tissue found preserved together with other archeobotanical remains, including human coprolites with agave tissue remains, in caves of the Tehuacan Valley in south central México; these findings are dated from about 10,000 years ago (MacNeish 1967; Flannery 1986). Later remains of objects weaved with agave fiber were found by archaeologists. Like those remains, almost every part of an agave has a use, from their roots, the stem, leaves and cuticles, terminal spine, their inflorescences, flower buds, seeds and sap, their dry leaves and stems used as fuel, and even their ashes. Primary and secondary needs are satisfied with agaves. From the basic needs, like food, dress, firewood, shelter construction, medicine, tools, to secondary needs directed to satisfy spiritual and cultural aspects. Agaves are among the plants with the largest number of uses reported in literature (Torres-García et al. 2019a). Throughout the entire territory of Mexico, each culture has appointed terms in their native language to refer to this important group of plants, their parts and products. Some cultures even have associated agaves to deities that are related to abundance and fertility, making clear the importance and symbiosis that these plants have had in the development of their cultures and identity. Just as diverse are their uses, diverse are the species that inhabit the also diverse habitats of Mexico. In this country, around 159 species have been reported, and new species, with ancient uses, are being discovered by the specialists every year (García-Mendoza et al. 2019a). In this work we review information on just a sample of species that are distributed in mountainous or highland habitats in various areas of the country that are remarkable in cultural and economic aspects. Although we only delve into this sample, in the Appendix

1006

I. Torres-García et al.

1 we provide an updated list of all the useful agave species (114) reported in the literature, including the use categories referred to in this review. Agave americana: According to Thiede (2019), this species possesses none stems or short trunks 10 cm, with 14–18 flowers. Flowers are erect, 60–80 mm; tepals yellow, tube cylindrical, 13–18  14–16 mm, lobes erect, lanceolate-linear, involute, subequal, 18–26  4–6 mm, outer 2 mm longer, inner broadening at the base, with narrow, prominent keel, both cucullate; stamens long-exerted; filament erect, slender, tapering, 50–65 mm, yellow, inserted just above mid-tube; anthers centric, 24–30 mm, yellow. The ovary slender, 30–40 mm, greenish, neck constricted, 4–6 mm; style longer than the stamens after anthesis, broadly clavate toward the apex; stigma 3-lobate. Fruits are oblong, 40–50  17–20 mm, beaked, shortly stipitate; seeds crescent shaped, 6–7  5 mm, margin with wavy wing. This species is distributed from South USA to North and Central Mexico, from Chihuahua to Querétaro where it can be found growing wild. Four intraspecific taxa are accepted A. asperrima ssp. asperrima

Agave americana L.. . .

1009

(Gentry) B. Ullrich, A. asperrima ssp. maderensis (Gentry) B. Ullrich, A. asperrima ssp. potosiensis (Gentry) B. Ullrich, and A. asperrima ssp. zarcensis (Gentry) B. Ullrich. The IUCN considers its conservation status as least concern (LC) (Hernández-Sandoval et al. 2019) (Fig. 3). Agave atrovirens: According to Gentry (1982) and Thiede (2019), this species is acaulescent, the rosette openly spreading, large to very large, 1.5–2.5  3–4 m, solitary or caespitose. Leaves 40–70, are erect to recurved, broadly lanceolate, thick at the base (to 25 cm), usually narrowed below the middle, convex below, openly concave above, mostly 150–200(220)  20–40 cm, dark to blackish-green to light glaucous or glaucous-variegated, margins straight or slightly wavy; marginal teeth regular, straight, antrorse, 4–8(12)  10–13(17) mm (mid-leaf), dark brown, grayish-brown or reddish, with low broad bases, 10–20(40) mm apart, sometimes bases fused into a continuous band. The terminal spine straight, conical, strong, broad at the base, widely openly grooved above, (30–) 60–70 mm, decurrent for 10–20 cm, keel rounded below and markedly intruding into the leaf tip for 1.5–2.5 cm, dark brown. The inflorescence 6–12 m, paniculate, peduncle with long triangular dark brown, bracts 40–50 cm long, fertile part narrow, oblong, part-inflorescence congested in the upper of the inflorescence, globose, (18–) 35–60, 20–30 cm long, with >600 flowers, axis reddish. Flowers are thickly fleshy,

Fig. 3 Inflorescence, flowers (photo: Rafael Humberto Cardenas Ollivier), and rosette of Agave asperrima (Photo: Manuel Nevarez)

1010

I. Torres-García et al.

(70–) 80–90(100) mm; pedicels 10–20 mm; tepals red to purple in bud, at anthesis yellowish to yellowish-green within, tip reddish, tube deeply furrowed, thick-walled, 10–15  10–20 mm, bulging at the insertion of the stamens, lobes triangular or oblong, erect, conduplicate-revolute, thickly fleshy, unequal, 30–40  4–6(9) mm, incurved at the tips. The outer tepals slightly larger; interior tepals with very thick keel and thin margins, red with paler margins; filaments large, flattened, 70–80(90) mm, purplish or red-spotted, inserted in the middle or at the mouth of the tube, yellowish or with purple dots; anthers excentric, straight to sinuous, (25–) 30–43 mm, yellow(ish) to bronze. The ovary is cylindrical, tapering at the base, 3- to 6-angled, 30–50  (4–) 6–8 mm, neck thick, furrowed, 4–7 mm; Style 80–110 mm, much longer than the stamens after anthesis; stigma clavate, 4–5 mm diameter. Fruits are ovoid, ellipsoid or clavate, 40–50(65)  20–30 mm, 5–12 mm stipitate, tapering from the apex; seeds crescent-shaped, 7–9  4–6 mm, shiny, with inconspicuous wing. This species is distributed in high montane habitats of the states of Guerrero, Oaxaca, and Veracruz. Two intraspecific taxa are recognized, being A. atrovirens var. atrovirens Gentry the wild and managed form and A. atrovirens var. mirabilis (Trelease) Gentry the cultivated form. The IUCN considers its conservation status as least concern (LC) (García-Mendoza et al. 2019d) (Fig. 4). Agave bovicornuta: According to Gentry (1982) and Thiede (2019), this species is acaulescent, rosette 0.6–1  1.5–2 m but often smaller, solitary. Leaves lanceolate to spatulate, much narrowed toward the base, widest at or above the middle, smooth, 45–80  10–17 cm, yellowish-green to light or deep green, younger leaves frequently shining glaucous, with conspicuous imprints from the central bud, margins crenate. Marginal teeth are dimorphic, larger teeth mostly 8–12 mm, flexuous and slender above a broad base, mostly 20–40 mm apart, on prominent prominences, smaller teeth mostly 2–5 mm, 1 to several between the larger teeth, all chestnut brown or dark brown to grayish-brown with age. The terminal spine is strong, 30–50 mm; The inflorescence 2.5–7 m, paniculate, peduncle short, with triangular acuminate reflexed bracts 20–30 cm long, fertile part narrow, part-inflorescence short, compact, 20–30 in the upper of the inflorescence; pedicels 5–10 mm. Flowers are 55–65 mm; tepals greenish-yellow, tube 6–9  12–14 mm, lobes linearlanceolate, ascending spreading, conduplicate, involute, sinuses broadly overlapping, 15–21  4–7 mm, inner tepals broader and 2-costate within, yellow; filament 30–45 mm, yellow, inserted (3–) 4–5 mm above the tube base; Anthers 18–24 mm, yellow. The ovary 26–29 mm, pale green, neck constricted, 4–6 mm; style hardly longer than the tepals; stigma capitate, 3-lobate. Fruits are oblong, 40–50  15–20 mm, stipitate, valves thin; seed finely punctate, 15  7 mm, curved side with a flange or wing, hilar notch shallow. This species is commonly found growing wild in Northern Sierra Madre Occidental. The IUCN considers its conservation status as vulnerable (VU) (Gonzáles-Elizondo et al. 2019) (Fig. 5). Agave cupreata: According to Gentry (1982), Thiede (2019) and AvendañoArrazate et al. (2015) this species is caulescent, the rosette openly spreading, 0.8–1.6 m diameter, solitary. Leaves are broadly lanceolate or ovate, thickly fleshy, strongly narrowed at the base, flat to slightly concave above, 40–80  18–20 cm, bright shiny green, with conspicuous patterns from the central bud, margins deeply

Agave americana L.. . .

1011

Fig. 4 Mature individual with floral stalk, rosette, and flower buds of Agave atrovirens (photo: Pablo Carrillo-Reyes) Fig. 5 Rosette of Agave bovicornuta (photo: Matt Brooks)

1012

I. Torres-García et al.

crenate-mamillate. The marginal teeth are straight to curved, strongly flattened, dimorphic, larger teeth 10–15 mm on distinct prominences, 30–60 mm apart, with smaller intermittent teeth of varying sizes, copper-colored to gray. The terminal spine is slender, sinuous, openly grooved above, 30–50 mm, light brown to grayish, with a sharp border decurrent to the upper teeth. The inflorescence 4–7 m, paniculate, fertile part rather broad, part-inflorescence lax diffuse “umbels,” 14–25 in the upper of the inflorescence, with dark small bracts. Flowers are 55–60 mm; tepals brown-red in bud, at anthesis orange-yellow, tube broadly funnel-shaped, 6–7  14–15 mm, lobes erect, linear-lanceolate, acute, subequal, outer 20–21 mm, apex rust-colored, broader and thicker than the inner, inner with narrow keel and thin involute margins; filament thickened toward the base, 35–40 mm, inserted at mid-tube; anthers excentric, curved, 23–24 mm, yellow. The ovary is fusiform, knobby, grooved, thickwalled, 30–35 mm, olive-green, neck constricted, doubly 3-grooved; style hardly longer than the tepals; stigma capitate. Fruits are oblateelliptic 36–57  19–21 mm, flattened-pinched apex; and seed semi-triangular to arc shaped. This species is endemic to the Balsas River basin in the states of Michoacán, Guerrero, and Oaxaca. The IUCN considers its conservation status as endangered (EN), derived mainly of its ongoing use in the elaboration of mescal (Torres-García et al. 2020) (Fig. 6). Agave hookeri: According to Gentry (1982) and Thiede (2019) this species present short thick stems; the rosette 2 m, solitary. Leaves lanceolate, arching in age, thickly fleshy, gradually narrowed toward base and tip, generally concave above, 120–175  20–25 cm, glaucous to green or yellow-green, margins repand to crenate. The marginal teeth straight or curved, 8–12 mm (middle of the lamina), dark brown to grayish-brown, 20–50 mm apart, with few smaller intermittent teeth, much reduced and closely spaced toward the base, broadly based on fleshy prominences. The terminal spine subulate, shortly openly grooved above, roundly keeled below, and with linguiform protrusion into the leaf, 35–60 mm, edges decurrent as toothless smooth horny border for 15–20 cm. Inflorescence 7–8 m, paniculate, peduncle with narrow, triangular, reflexed bracts, part-inflorescence compact, 20–40 in the upper half of the inflorescence; floral bracts caducous; pedicels long. Flower slender, 63–80 mm; tepalous yellow, red to pink in bud, tube grooved, 5–8  13–14 mm, lobes red to pink, linear, erect, elongate, conduplicate, unequal, outer 28–32 mm, rounded on the back, cucullate, inner 2–3 mm shorter, with prominent keel; filament slender, flattened, 50–60 mm, inserted above mid-tube; anther excentric, 26–34 mm, yellow. The ovary cylindrical, angled, smooth, 34–41 mm, neck long, grooved, constricted; style almost as long as the filaments; stigma capitate. Fruits are oblong, rounded at the tip, thick-walled, 50–55_25 mm, stipitate; seed broadly crescent-shaped, rugose, 8–9  6–7 mm, lucid black, hilar notch broad, marginal wing broad but little raised. This species is presumed to be a domesticated form of Agave inaequidens. It can be found only cultivated or eventually scaped from culture, in the mountain regions in the states of Jalisco and Michoacán. Its cultural importance is declining because of the introduction of other pulque producer species and the decreased pulque activity (Figueredo-Urbina et al. 2017) (Fig. 7).

Agave americana L.. . .

1013

Fig. 6 Floral stalk, rosette, and flowers of Agave cupreata. (Photos by the authors)

Agave inaequidens: According to Gentry (1982) and Thiede (2019), this species present short stems; the rosette openly spreading, 1.5–3.5 m in diameter, solitary. Leaves variable, broadly or narrowly lanceolate or oblanceolate, ascending to outcurving, thickly fleshy, concave above, toward the rounded base, mostly 75–150  11–21 cm, light green to yellow-green, rarely faintly glaucous, margins undulate to repand and crenate. Marginal teeth dimorphic, straight or variously curved, the flattened bases longer than the height of the teeth, commonly 8–10 mm long, chestnut-brown to dark brown, 25–40 mm apart, with few smaller intermittent teeth, larger teeth on broad prominences. The terminal spine stout, broadly deeply channeled above, 25–55 mm, dark brown, protruding into the leaf tissue below, sharply decurrent to the uppermost teeth. Inflorescence 5–8 (12) m, paniculate, peduncle short, fertile part narrow in outline, part-inflorescence compact, 30–50 in the upper of the inflorescence. Flowers 60–90 mm; tepals yellow, buds reddish-purple, tube deeply grooved, 5–12(15) mm, thick-walled, bulging at the filament insertions, lobes linear, erect, narrow, conduplicate, involute, strongly cucullate and papillate within at the tip, unequal, 25–30(34) mm, reddish; filament stout, ovate in cross section, 50–60 mm, inserted above mid-tube; anther centric to excentric, 26–34 mm. The ovary is cylindrical, trigonous, 30–40 mm, neck short, furrowed; style slightly longer than the tepals; stigma capitate. Fruits are oblong,

1014

I. Torres-García et al.

Fig. 7 Detail of rosette, rosette, theeth, and flowers of Agave hookeri. (Photos by the authors)

40–45  20 mm, rounded to apiculate at the apex, without beak, stipitate, brown; seed semicircular, 6–7.5  4.5–5.5 mm, shiny black, finely punctate, the marginal wing half-curved or straight, hilar notch broad. The most abundant Agave species in the Trans-Mexican Volcanic Belt, the mountains from Puebla to Jalisco. Two infraspecific taxa are recognized A. inaequidens ssp. inaequidens and A. inaequidens ssp. barracencis (Gentry) from Durango and Sinaloa. The IUCN considers its conservation status as least concern (LC) (Torres-García et al. 2019b) (Fig. 8). Agave karwinskii: According to García-Mendoza (2011) and Thiede (2019) this species is arborescent, stems 0.3–0.7 m (to 3 m in cultivation), forming clonal colonies with spreading rhizomes. The rosette dense, 0.6–1  1–1.5 m, extending down the stem from the stem tip with leaves reflexing along the stems with age. Leaves 80–100, linear to linear-lanceolate, rigid, fibrous, ascending to radiately spreading or erect, narrowed and thickened toward the base, acuminate, involute at the tip toward the base of the terminal spine, convex below, guttered or concave

Agave americana L.. . .

1015

Fig. 8 Flowers, rosette, and mature individual with massive floral stalk of Agave inaequides. (Photos by the authors)

above, (33–) 40–65(80)  2–7(8) cm, dark-green to yellowish-green, margins straight. Marginal teeth delicate, nearly straight or antrorse to cuspidate and flexuous, pyramidal with broad base, 2–4(6) mm (at mid-leaf), blackish, dark brown or reddish, (10–) 20–40(70) mm apart. The terminal spine is variable, subulate or conical with thickened base, narrowly to broadly grooved above, 15–30(45) mm, blackish or dark brown to grayish and corroding at the base, decurrent or not. Inflorescence 3–4(6) m, paniculate, peduncle green, with chartaceous triangular bracts 4–11 cm long, fertile part openly diffuse, oval to oblong, part-inflorescence lax, 10–15 in the upper, 15–30 cm; pedicels 2 mm. Flower 40–50(57) mm; tepals greenish to pale yellow, tube bulging between the grooves, rather thin-walled, 10–11  8–11 mm, lobes linear-spatulate to oblong, erect, involute, incurving before anther dehiscence, quickly wilting, unequal, 11–19  2–3 mm, with ferrugineous tinge; filament 35–40 mm, inserted irregularly at mid-tube or slightly above, greenish, dotted reddish; anther regular, centric, 15–22 mm, yellow(ish). The ovary angularly cylindrical, 20–30  7–9 mm, neck 1–2 mm, slightly 6-grooved; style 45 mm, dotted reddish. Fruits oblong or obpyriform to subglobose, 35–50  25–30 mm; seed 7–10  6–10 mm, winged for