Vibrational Communication in Animals 9780674273825

In creatures as different as crickets and scorpions, mole rats and elephants, there exists an overlooked channel of comm

186 106 6MB

English Pages 272 Year 2008

Report DMCA / Copyright

DOWNLOAD PDF FILE

Recommend Papers

Vibrational Communication in Animals
 9780674273825

  • 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

Vibrational Communication in Animals

Vibrational Communication in Animals ◆ ◆ ◆

PEGGY S. M. HILL

HARVARD UNIVERSITY PRESS

Cambridge, Massachusetts London, England 2008

Copyright © 2008 by the President and Fellows of Harvard College All rights reserved Printed in the United States of America Library of Congress Cataloging-in-Publication Data Hill, Peggy S. M., 1948– Vibrational communication in animals / Peggy S. M. Hill. p. cm. Includes bibliographical references and index. ISBN-13: 978-0-674-02798-5 (alk. paper) 1. Animal communication. 2. Bioacoustics. 3. Vibration. I. Title. QL776.H55 2008 591.59—dc22 2007038954

Contents

Preface

vii

ONE

Vibration as a Channel for Information Transfer

TWO

Communication and the Medium

THREE FOUR FIVE SIX

Receiving Signals Sending Signals

39 86

Predator-Prey Interactions Mating

119

135

SEVEN

Group Information Transfer

EIGHT

Why Vibration?

References

203

211

Species Index

249

Subject Index

259

183

12

1

Preface

grew into this book was planted during a symposium on vibration as a communication channel held at the annual meeting of the Society for Integrative and Comparative Biology (SICB) in Chicago on January 4, 2001. This symposium was sponsored and funded by the National Science Foundation, SICB’s Divisions of Animal Behavior and Neurobiology, and the Animal Behaviour Society. Papers were given and later published in one of the last issues of American Zoologist (the predecessor of Integrative and Comparative Biology) by myself and my University of Tulsa collaborator, John R. Shadley, Jan A. Randall, Caitlin O’Connell-Rodwell and Lynette Hart, Peter M. Narins, Edwin R. Lewis, Reginald B. (Rex) Cocroft, Randy E. Hunt, and Philip H. Brownell. The discussions after the final paper and over dinner stimulated us all, I think, to speak and publish more often and more widely to educate the scientific community and general public about this fascinating world about which so little has been written. Anyone from the symposium could have taken up the challenge of writing this book, so I hope my efforts will meet with their approval. Vibration communication is ubiquitous indeed, and this book is essentially a broad overview of the literature by someone who is hardly an expert on any aspect. Even with the best of intentions, I am sure there are glaring omissions, and so I apologize in advance for my lapses and failures. I hope each of you still likes this book. I think I do. I would first like to thank my editor, Ann Downer-Hazell, for all her enthusiasm and encouragement of this project. It took Ann’s vision

THE SEED THAT

viii

Preface

of a book, and about five years of her patience, before I agreed to tackle the project. The book would literally not have been written if not for Ann’s advocacy. Special thanks also go to Ted Lewis, Rex Cocroft, and one anonymous reader whose comments and suggestions helped to improve the manuscript during revision. I am grateful to Vanessa Hayes, editorial assistant, and Elizabeth Gilbert, senior editor, at HUP, as well as Susan Rescigno, production editor at Westchester Book Services, and Pat Cattani, copy editor, for all their encouragement and technical help on the production side of the project. I would not have been able to think about vibration with any clarity, much less to understand any of the physics or to sort out the field experiments, if it had not been for the patience, tutelage, and kind mentoring of my professional mechanical engineer, friend, colleague, and field companion, John R. Shadley. I thank the University of Tulsa and its College of Engineering and Natural Sciences and Faculty of Biological Sciences for supporting me during a semester sabbatical leave while I wrote. I further thank the outstanding scientists working in the field of vibration communication, whose work I hope to have accurately summarized in these pages. Special thanks go to Peter Narins, Rex Cocroft, Caitlin O’Connell-Rodwell, Lynette Hart, Philip Brownell, Glenn Morris, and Andrew Mason. I also thank Karen Warkentin, Becky Talyn, David Clayton, Kevin Johnson, and Clint Kelly for their sometimes long-distance comments and encouragement. I further thank my two stalwart PhD students, Dan Howard and Charlotte Sanderson, who spent many hours listening and encouraging, and for Dan’s presubmission comments. I also thank my extended professional family who have listened and prodded: Janice Burnett, Sallie Trecek, William Rosche, Harrington Wells, and Randy Wymore. Lastly, the project could not have been finished without the careful attention and professional care of Ricardo Ortega, who edited and prepared the figures for publication, thus removing a major burden from my shoulders. I wish to dedicate this book to my family, who have given me the space and time to complete the project, as well as a great deal of love and laughter along the way: my husband, Robert; my children, Angela and Yevgeni Kachuyevski and Bobi and Joe Deere; my grandchildren,

Preface

ix

Jordan P. H. and R. J. Yonah Hill Deere, Alexander Y. and Nikolai Y. Kachuyevski, Kristina M. and Timothy J. Whatley; and my mother, Cara Staten Morrison. They have taught me to love my life without taking myself too seriously, and Bobi and Jordan have been a tremendous help for years with my field research.

ONE ◆ ◆ ◆

Vibration as a Channel for Information Transfer

most compelling outcomes of the process of science is that we tend to find what we seek. This observation is not meant to suggest that we search for support for a pet hypothesis while intentionally ignoring other alternatives, but that we tend to look for things when we already have a pretty good idea that they are there. The “normal science” of Kuhn (1996) is spent in working on small questions within the larger framework of a paradigm. We have a good general idea of how things work, but we question the smaller details, and we follow leads, looking for inconsistencies—the anomalies that lead us to other questions. Answers bring a new doorway with a heretofore-unknown world on the other side. We say, “The more I learn, the less I know.” However, rarely does an answer shift a paradigm. Only with major insights, or the use of new tools, or with a question asked of exactly the right person who is prepared to think of it in a novel way do we break away from the nested boxes that define most of our work in normal science. Such is the story of animal communication through substrate vibration: the idea that individuals send and receive signals using the substrate as the medium. Our study of vibrational communication is in its infancy, but the glimpse we have of the mechanisms employed leads us to believe that this is a very ancient system, that it is perhaps the primary channel of communication in some animal groups, and that it is ubiquitous, at least in vertebrates and arthropods (Hill 2001a,b). For example, the very loud signaling via airborne sounds in cicadas is actually a derived modality that evolved from a substrate vibration precursor

ONE OF THE

1

2

Vibrational Communication in Animals

still employed by its “little cicada” relatives in the Hemiptera (Claridge, Morgan, and Moulds 1999). Cocroft and Rodríguez (2005) conservatively suggest that 150,000 described species of insects communicate exclusively through substrate vibration, and another 45,000 species use vibration signals together with other forms of mechanical signaling. Those who work outside the field of animal communication may justifiably be skeptical that humans could have missed stumbling on this phenomenon before, even if by accident, were it really as widespread and primary as we now argue. The answer, of course, is that we have encountered it, but we were not looking for it. We described and documented the function of all sorts of mechanical receptors that could detect substrate vibrations in a large number of arthropod and vertebrate taxa, and we observed animals drumming or thumping appendages, heads, and abdomens against various substrates while producing no sounds audible to humans. We attributed the anomalies that led us to suspect something else was going on to anything more plausible than communication through vibration signals—or the anomalies persisted as anomalies, to be examined again later when it was more practical to do so. For example, Jones and Dambach (1973) reported that the cricket Gryllus campestris was able to respond to loud airborne signals even after the tympanal organs, antennae, and cerci had been removed. They did not speculate that the airborne sounds might have been picked up by the substrate and channeled to the cricket nervous system as vibration, but two decades later Kalmring et al. (1997) cite the Jones and Dambach paper as reporting on a vibratory communication system—not just vibration detection, but vibratory communication. Likewise, an early description of maternal care in the short-tailed cricket, Anurogryllus muticus, mentioned that the very aggressive female only retreated in response to “violent substrate vibration” (21) and that in her aggressive displays she “shook her body in the manner characteristic of aggression in many kinds of crickets” (West and Alexander 1963, 21). This behavior might strongly suggest the use of substrate vibration as a communication channel, but at that time, this interpretation was not really as intuitive as it might appear today. The community bound by a paradigm does not purposefully make square pegs fit round holes, but through the carrot-and-stick approach of peer review of publications and grant proposals, practitioners determine

Vibration as a Channel for Information Transfer

3

which questions are worth pursuing. Sometimes the weight of accumulating data is required to shift interpretations of events. Some investigators in the first half of the twentieth century were studying vibrational communication, especially in invertebrates, but the scientific community was not fully engaged. For example, Ossiannilsson (1949) published an extensive treatise on leafhoppers that suggested vibrations produced by the tymbal apparatus were conducted through the substrate. He summarized the literature to date, including a 1907 paper that reported “sounds” of the otherwise “silent” leafhoppers that had been transferred from a leaf to the air. Ossiannilsson himself used musical notation to describe substrate vibration of some leafhoppers that he was able to hear by inducing individuals to signal from leaves in a test tube that he held to his ear. He also analyzed recordings of signaling species and published the waveforms in his monograph, including one of two alternating males. He described substrate vibration signals used in a chorus, suggested that male-female duets were important mating signals, and determined that pitch varied with temperature (1949). However, up through the 1970s, conventional wisdom held that substrate-borne vibration could not transfer any biologically meaningful information among organisms, especially the very small. The medium of soil or water or plant stems was too inelastic, and propagation speeds were too great to provide anything more than a nonspecific alerting mechanism to inform the animal of a disturbance nearby (Schwartzkopff 1974). Then technology in the form of faster, cheaper computers and seismic sensors, such as the geophones that had been developed for oil exploration (and detecting advancing enemy troops in the jungles of Southeast Asia), became accessible to more scientists. Affordable transducers made detection, simulation, and manipulation of signals possible. Faster, smaller, cheaper computers provided the opportunity for rigorous analyses of the data gathered (Hill 2001b). Sometimes the right person was in a position to think of a new way to test an anomaly (see Miles et al. 2001). At the same time, perhaps, the anomalies encountered in asking other questions became too massive to be set aside for another time. My own interest in vibration began with a frustrating experience trying to manipulate male prairie mole crickets (Gryllotalpa major) in the field using a playback experiment (see Hill and Shadley 2001). These rare burrowing endemics of the south-central United States construct

4

Vibrational Communication in Animals

burrows from which they project airborne sexual-advertisement calling songs to attract flying females for mating. Males sing from a fixed location inside an underground “acoustic horn” for about 30 min just at sunset on spring evenings when it is not too hot or too cold, not raining and not too windy. I wanted to test a scramble competition model that required me to evoke a response from a male when I simulated a competitor. Anecdotal reports were that prairie mole cricket males did not respond to conspecific song when the playback methods used to study other orthopterans and frogs were employed, but to test my hypothesis, I had to try. Males ignored the taped sounds I played for them, but as I moved about the area as darkness settled in night after night, males stopped singing whenever I was within 1–3 m of their burrow openings. It was entirely within my worldview at the time that these males could detect my footfalls and consider my presence a threat, but it took some investigation before I could ask whether they could also use this same sort of information to interact with conspecific neighbors. Could a person in the twenty-first century a.d. hope to write an account of animal communication through any other sensory modality? Would it be possible to describe visual or hearing or olfactory pathways that included a short history of use of the channel, background of the nature of the signal, production of signals as well as reception of them, and the contexts within which the signals were used by species known to use them across Kingdom Animalia? Yet even with the groundswell of interest in vibration since the dawn of the twenty-first century, we still speculate about the breadth and depth of the importance of substrate vibration as a communication pathway. We may know something about a behavior, but nothing yet about the signal or the neurophysiology of how the behavior is elicited and controlled. We may understand the technical aspects of a receiver, but not the ecological context that might have promoted its selection, or the selection for production of the signal, or even the behavioral interactions that were influenced by its selection. The ecologist with questions is not likely to be trained as a neuroethologist, a histologist, or a neurophysiologist. The functional morphologist may not have the tools to pursue questions beyond a point, and luring colleagues with these tools away from their own chosen research lines to help pursue a potential illusion is difficult, at best. An examination of the literature reveals early descriptions of anomalies and guarded suggestions of a link to substrate vibration. As recently

Vibration as a Channel for Information Transfer

5

as the 1980s, experiments in anatomy, physiology, ecology, and behavior still bolstered support for only tentative interpretations that animals really are using substrate vibration detection in as many contexts as they would use any other major sensory pathway. The pioneering work of Philip Brownell and Roger Farley (Brownell 1977; Brownell and Farley 1979a,b,c) showed us that scorpions not only can detect prey but can use information contained in substrate-borne vibrations created by the prey themselves to locate and capture them. It was Brownell and Farley’s carefully controlled and executed experiments, published in the best mainstream journals, that encouraged others to examine the possibility that baffling behaviors observed in their study species might be linked to sensory input via a substrate vibration channel. Hadley and Williams had already reported in 1968 that scorpions appeared to use tactile senses rather than vision to recognize prey and would attack forceps drawn along the substrate nearby (Hadley and

BT

H

SS B T

PS 1 mm

LC

B MC

Figure 1.1. Dorsal-anterior view of right fourth leg of the sand scorpion, Paruroctonus mesaensis, showing tarsal hairs (H) and the basitarsal slit sensillum (SS). B = bristle hairs, BT = basitarsus, LC = lateral claw, MC = medial claw, PS = pedal spur, T = tarsus.

Adapted and reprinted with kind permission of Springer Science and Business Media and The Journal of Comparative Physiology from Brownell and Farley (1979a, 24).

6

Vibrational Communication in Animals

Williams 1968). Yet Brownell knew that the sandy substrate on which he observed scorpions was supposed to absorb, or damp, low-frequency sound rather than transmit it, because sand is granular and would not support propagation of a wave from one particle to the next in the fashion of a more homogenous substrate. Further, waves traveling at higher frequencies would have a wavelength similar to the size of the sand grains, and these waves would be scattered as they traveled through the medium. What he found, however, was that sand was a good conductor of both compressional (P) waves and Rayleigh waves, that these waves traveled much slower in sand than predicted by current theory, and that in the range of several decimeters, animals with the right sort of receivers should be able to use information contained in these waves to make decisions that promote fitness. Specifically, he concluded that Rayleigh waves were detected by the basitarsal compound slit sensilla (BCSS) in scorpion legs (Figure 1.1) and used by scorpions to determine source direction, whereas P-waves were detected by mechanosensitive tarsal hairs and used to determine source distance (Brownell 1984). We now have uncovered evidence of use of the substrate vibration channel, especially in the Arthropoda, in contexts from predation or pathogen defense, to foraging coordination, to mate location and choice, to family dynamics. Of special note is the observation that if we find an unusual morphological feature that empirical evidence supports as a sender or receiver in substrate vibration communication, we then can find behavioral contexts in which this communication can be shown to occur, and vice versa. In examining related taxa, we often will find the same behavior or morphology, and sometimes an evolutionary trail can be followed by studying a feature in a phylogenetic framework. Again, if we find that a species gathers substrate vibration cues in one context, we often can find that it also uses vibration signals in conspecific communication. When our worldview expands to suspect that substrate vibration might be a key to explaining a behavior or morphology, we very often find that it is. What the Fossils Tell Us The fossil record and comparative morphology studies tell us a great deal about the possibilities for sensory input in ancient taxa, even if the central nervous system (CNS) does not fossilize. We can fall back on

Vibration as a Channel for Information Transfer

7

comparative evidence from the skeletal material in which the nervous system has been housed, as we do with modern forms, and modern taxa may not be that much different from their extinct ancestors. Morphologists can see that ancient amphibians, the first vertebrates on land, would have been able to detect substrate vibrations as their massive heads lay on the ground. Their lower jaws, coupled with the soil, could conduct vibrations through the quadrate bone to the inner ear through a bony pathway (Hildebrand and Goslow 2001). Thus, the earliest land vertebrates could “hear” substrate vibrations before they had the acoustic apparatus required to hear airborne sounds as terrestrial forms do today. Furthermore, the vibration pathway is still in use by amphibians, and the sacculus and lagena of the inner ear of modern frogs are the sites of their exquisite sensitivity to substrate-borne vibrations (Ashcroft and Hallpike 1934; Caston, Precht, and Blanks 1977; Lewis 1984). The story of the evolution of vertebrate animals and their head skeleton is a fascinating one (see review in Northcutt and Gans 1983). The process of cephalization that accompanied the earliest bilateral symmetry in invertebrates not only defined the head end of an individual, but it focused sensory structures, and then feeding structures and a brain, in this region. Details of the head so define vertebrates as a group, and in turn the classes into which they are sorted, that a wellreasoned and logical argument can be made to call them the Craniata. All vertebrate animals have a brain enclosed by a skull, but not all have jaws, and the earliest jaws in the fishes were integrated into the rest of the head skeleton in a different way than the jaws later seen in terrestrial forms. Elements of the visceral skeleton, which supported gills in the fish, were freed up by terrestrial life where the gill arches had no role. Some of these elements were incorporated into new hearing structures, the auditory ossicles of the middle ear, and others were incorporated into the vocal apparatus of the larynx and hyoid apparatus, which anchors the tongue and supports the floor of the mouth (Figure 1.2). At the same time, the number of elements in the lower jaw, or mandible, was dramatically decreased in a manner that gained stability but lost flexibility. The jaw joint itself was “reinvented” and was moved forward, while the muscles that closed the jaw shifted to a position where they attached outside the skull, rather than inside, as they had in the earlier fishes. The skull developed windows, or fenestrae, to accommodate these external muscles, and the options for food choices were

8

Vibrational Communication in Animals

rather suddenly, for geologic time, increased. We might also argue that the options for communication also were multiplied by these changes in the head skeleton. In primitive bony fish (Class Osteichthyes) skulls, the visceral skeleton includes an element, the hyomandibula, in association with the quadrate bone in the jaw joint. In the early amphibians, this hyomandibula was reduced and took on a new role as an early ear ossicle, the stapes, or columella (Hildebrand and Goslow 2001). Modern amphibians and

Figure 1.2. Phylogeny of the vertebrate jaw articulation and ear ossicles.

Note angular/ectotympanic, articular/malleus, quadrate/incus, and hyomandibula/stapes relationships between lobe-finned fish and mammal. Reprinted with kind permission of John Wiley & Sons, Inc., from Hildebrand and Goslow (2001, 128).

Vibration as a Channel for Information Transfer

9

reptiles that lie flat on the soil in direct contact with the Earth over much of their body surface have not developed a tympanum (ear drum) or middle ear cavity that promotes the most efficient conduction of airborne sound. They, like their ancient, extinct ancestors, are more suited to receiving information from vibration in the substrate than from sounds transmitted through the atmosphere. Aberrant Behavior and Vibration Sensitivity Humans have actually taken note of various animal responses to substrate vibration for at least 3,000 years (Snarr 2005). Tributsch (1982) rather extensively discussed documented reports of anomalous animal behavior that occurred minutes to days prior to an earthquake, beginning with a detailed description of the 373 b.c. destruction of Helice in ancient Greece for which the primary source has been lost. Many people have heard that animals “act funny” before, after, and during earthquakes (see Anderson 1973), and I, myself, have been interviewed by the BBC on how animals might be able to predict earthquakes. However, after the 2004 tsunami in Southeast Asia, most of us were surprised to hear anecdotal reports that animals such as elephants and tigers did not die from the wave as humans did (i.e., Kenneally 2004, 2005), and even the indigenous people did not die as the more “civilized” ones did (Parker 2004; McAdoo et al. 2006; Nowak 2006). Before the tsunami arrived in Thailand, elephants carrying tourists reportedly broke free from handlers to move away from the area. Elephants, their riders, and the handlers who chased them survived, while other humans who stayed did not (Kenneally 2005). Of the survivors of the tsunami in Thailand who responded to a formal survey within three months of the event, only 21% remembered any unusual animal behavior prior to the arrival of the first wave, but only 24% recalled noticing unusual shaking of the ground (Gregg et al. 2006). Even though interest in this phenomenon is high, empirical evidence is sparse, especially in what would be considered natural, nondomestic, or noncaptive settings. Two radio-collared Asian elephants (Elephas maximus) being monitored with their herds at Yala National Park in Sri Lanka provided an unplanned opportunity to gather data on 26 December 2004. No evidence in their movement behavior suggested either was using anything other than immediate environmental cues as they avoided the wave

10

Vibrational Communication in Animals

(Wikramanayake, Fernando, and Leimgruber 2006). However, the wave did not reach either area where the two herds were actually located, unlike the situation of the elephants in Thailand in the anecdotal reports, whose position was inundated and human lives were lost. Mantled howlers (Alouatta palliata) in northeast Honduras that were being studied in Cuero y Salado Wildlife Refuge provided an opportunity to gather opportunistic data when an earthquake occurred in El Salvador in February 2001 with its epicenter at a distance of 341 km from the field site. Because behavioral data were already being collected on these habituated, but free-ranging, individuals at 10 min intervals, trained observers were on hand to document precisely the sequence of population responses to the surprising occurrence of the earthquake. Individuals responded much as they had been noted to do to a threat from the ground and rapidly moved higher into the trees, switching from a resting to an alert state. This behavior was in response to Modified Mercalli Intensity Scale Level IV motion at the site, or an incident where automobiles would rock visibly and dishes, windows, and doors would be disturbed but not broken. Although individuals did not display any anomalous preseismic activity, seismic body waves were detected by the howlers almost instantly, as “the P-waves arrived at the field site approximately 60 s post-origin, with the S-waves arriving 87 s post-origin” (Snarr 2005, 281). Kirschvink (2000) made an interesting argument that even with a short life span that would preclude an animal’s being able to “remember” a seismic event and respond to it, individuals might “inherit” an ability to respond appropriately to a seismic event. “A population-genetic model indicates that such a seismic-escape response system can be maintained against random mutations as a result of episodic selection that operates with time scales comparable to that of strong seismic events” (312). The oldest known fossils of mammals that survived the Permian extinction are most commonly found as pairs in a collapsed burrow; therefore, selection pressure for a response to seismic events was present at least 250 million years ago (Kirschvink 2000). We speculate that those burrowing mating pairs that did not respond were killed where their remains were found, whereas those that exited the burrow complex survived to be parents. If individuals detect P-waves, as reported for the mantled howlers by Snarr (2005), they might avoid death by responding rapidly before

Vibration as a Channel for Information Transfer

11

the S-waves arrived. The greater the distance from the epicenter of the earthquake, the more time individuals would have to respond because P-waves travel through the crust at about 2–4 km/s faster than S-waves (Kirschvink 2000). The response time provided by the difference in the arrival of these two wave types would have been sufficient to allow the escape from the 2004 tsunami, as well. We thus know a bit about a lot with regard to the substrate as a communication channel. More data must be gathered before we can convert anecdotal reports into questions with testable hypotheses in many areas. Evidence from the fossil record and behavior associated with earthquakes has mostly been incidentally gathered, rather than being in response to the question driving the research. Very few systems have even been superficially studied. A handful of research groups are linking communication studies with morphology and neurophysiology and asking questions about selection for a particular type of signal in a given habitat or social environment. However, an entire planet full of potential research subjects awaits our interest. One purpose of this book is to document what we do know, in the broadest possible sense, from others’ work to identify promising areas for future research. Another is to simply provide a glimpse into a fascinating world of animal interactions that we otherwise might happily continue not to see.

TWO ◆ ◆ ◆

Communication and the Medium

a broad range of potential signals merely as incidental events, or byproducts, or epiphenomena, of routine activities. These events can only be labeled signals after empirical testing to confirm that they transfer information among or between individuals (see Bradbury and Vehrencamp 1998). Animals generate airborne, or waterborne, or substrate-borne vibrations; display colors or patterns; and emit chemicals as they live their lives. However, even if an event is stereotypical and species specific (Doherty and Gerhardt 1984), it cannot be classified as a signal unless it functions to transfer information from the sender to a receiver and then modifies the behavior of the receiver in a predictable way that has adaptive value for one or both (Markl 1983). This concept of a signal, though by all means not the only one available in the literature, will be used throughout this book. The definition clearly applies to conspecific communication used in mating, parental care, and coordinated group living, but it would also include information transferred incidentally between prey and predators. The signal need not be directed toward a specific individual; in other words, it may be broadcast in hopes of its encountering a receiver. Further, the time delay between sending and receiving may be rather short, or the signal could be persistent, as in scent marking of a mammal’s territory. A signal could even be a thread spun and deposited by a spider (Krafft 1982). The essential requirements, then, of any communication system include that an individual acts to initiate an event that encodes information (intentionally or incidentally), that the informa-

ANIMALS GENERATE

12

Communication and the Medium

13

tion is carried through some medium in a format that will deliver it to a receiving individual, and that the receiver is able to both detect and decode the message, ultimately acting on the information received. All of these steps must be present and detectable by a researcher for the event to be formally classified as a signal. However, signals may also be multimodal, and scientists must remain open to the idea that cryptic signals, or certain aspects of signals, may be repeatedly overlooked. Determining the range of information that can be transmitted via vibration in solids of all kinds, translating the signals, and estimating the contribution of such signaling to overall fitness will likely be a focus of study for some time, as will determining which anatomical features are specialized for sending and receiving the signals, and whether the same behaviors and features are present or absent in related taxa. A further challenge is to communicate to the reader a concise view of a complex, hidden process, where, for example, signal intensity “can be expressed either in terms of acceleration (m/s2), velocity (m/s), or displacement (m)” (Dambach 1989, 179). Still, the search for these answers and our attempts to incorporate them into our communication paradigm are rather young. Ehrenberg’s mole-rat, or the Palestine mole-rat, Spalax ehrenbergi (but work also reported under the generic name Nannospalax, or Nanospalax: see Mason and Narins 2001), has been observed to bang its flat head against the top of its subterranean tunnel in a way that could be explained as part of tunnel building activity. However, when individuals were shown to respond in a highly repeatable way to playback stimuli of thumping by conspecifics (Heth et al. 1987) and to respond to simulations in artificial tunnels as well as in the field (Rado et al. 1987), this head banging could be referred to as signaling. Thus, the first confirmation of vibrational communication in terrestrial mammals was as recent as 1987. Our information on invertebrate communication via vibration signals is currently restricted essentially to the literature of Arthropods, but behavioral observations of other taxa at least suggest that looking for such signals might be worthwhile (i.e., Beesley, Glasby, and Ross 2000 on the extreme vibration sensitivity of feather-duster worms in Annelida, and Newbury 1972 on use of vibration in prey detection by the Chaetognatha). Vibrational communication has only been a research focus of primarily the last 30 years, even though we have been

14

Vibrational Communication in Animals

able to build on century-old observations of novel, or anomalous, behaviors. One example is our study of the fiddler crabs in the genus Uca, which inhabit intertidal beaches around the world. Males have long been of interest to those studying communication because they have one larger claw, or chela, which they are seen to wave and also to rap on the ground. Male Uca pugilator have been observed in both laboratory and field settings as they drummed their large chelae against the substrate. Teasing apart the waveforms showed that drumming created three types of waves. Study of this and related species is providing a great deal of information on how substrate vibrations contribute to animal communication. If conspecific animals, in particular the females targeted by male courtship, are able to perceive waves with different properties and propagation velocities as different vibration events, they should be able to use the information encoded in the drumming signal to locate the male who is producing them (Aicher and Tautz 1990). However, what does this drumming say to other males, or to any members of the population in a context other than mating? Again, work continues to reveal a fascinating world of information transfer. Distinguishing strictly between sound and vibration may not be a useful enterprise. We can somewhat distinguish them based on sensory perception, if we hold that sound is detected by pressure, or pressure difference, receivers, and vibration by particle displacement in a medium (Kalmring 1985). Vibrations that propagate through both the air and the substrate are produced simultaneously by the same action (Gogala 1985a). Especially in the Orthoptera (Markl 1983), it appears that airborne sound and substrate-borne vibration signals are used in combination in both mating and aggressive interactions. The tympanal organ used for “hearing” and the subgenual organ that detects substrate vibration are located adjacent to each other in the cricket prothoracic legs. The sensory cells in the subgenual organ are “structurally indistinguishable from those in the ear, and these sensory cells respond to substrate vibration as well as to low-frequency sound” (Ball, Oldfield, and Rudolph 1989, 391). Not only is teasing apart the signal functions a daunting task, but it is entirely possible that the two function in combination, as well. Therefore, the more we learn about the channels animals use to gather information about the environment,

Communication and the Medium

15

the less valuable a distinction between sound and vibration becomes. Ossiannilsson (1949) commented on this distinction as follows: This discussion has been made more complicated by many workers having felt obliged to separate a perception of air-born vibrations by a specific auditory sense from a perception of the vibrations by the tactile sense. This presentation of the problem will, in my opinion, very easily turn into a battle of words. Even in insects with a well developed tympanal organ apparently specially constructed for the interception of vibrations of the air, we shall of course never be able to gain a real conception of how the animal subjectively apprehends these . . . I believe that the vibrations produced by the tymbal organ of one specimen are conducted to other individuals mainly by the solid substratum—as a rule some part of a plant—and only in a less degree by air. If it could be established that the animals do in this way apprehend calls of each other as such—is this to be termed hearing or not? A matter of taste! (136)

The Problem of Ambiguous Terminology Within the context of human interactions, we all have a fair understanding of what is meant by words such as sound, hearing, vibration, song, acoustic, and audible. However, one of the greatest challenges of trying to organize and review what is known about communication via substrate signaling has been the lack of a consistent, cohesive vocabulary that is used by researchers studying disparate taxa. A common experience of any emerging academic discipline is to borrow terminology from mostly compatible, or closely related, fields, and then to develop jargon when the established vocabulary fails. In the case of vibrational communication studies, terms have been imported from, for example, physics and mechanical engineering, but their use has evolved and definitions have been sculpted in isolation to the point that communication among groups of researchers is impeded. In many cases first descriptions were published before substrate signaling was suspected, and then the same terms were subsequently used, even after confirmation that the signals were carried via the substrate. Merely repeating the terms that authors have used to describe their systems creates a jumble of roadblocks for the intended audience of a review. Trying to translate the authors’ descriptions by substituting standard terms carries with it

16

Vibrational Communication in Animals

all the dangers inherent in any translation and may unintentionally alter or misrepresent the original observer’s perspective. However, after first relying on individual authors’ use of terms for this book, the decision was made to convert them, and accept the criticism, by using terminology based on the considerations discussed here. A large and comprehensive literature exists on senders and receivers used in animal communication via airborne or waterborne compressional waves (i.e., the Springer Handbook of Auditory Research volumes), but that literature and other conventional literature on hearing do not include substrate-borne vibration communication. The topics covered in this book are mostly distinguished from that literature by the medium (substrates of one kind or another) and the direction of particle motion detected by receivers (perpendicular to the direction of propagation), where detection of the particle motion usually involves some sort of inertial motion sensor, or sensing mass (see Lewis 1984). However, attempting to strictly separate airborne sound from substrate-borne vibration signaling leaves a gray, or fuzzy, area of behaviors that we have yet to successfully categorize, especially when the same words are used to describe very different phenomena. The first distinction must be made between sound and vibration. Use of these terms is not as clear-cut as we might think (e.g., Webster 1992; Lewis and Fay 2004), even though they are generally understood to be distinct when used in a popular connotation. Sound is typically used to describe airborne or waterborne compressional acoustic waves that can be “heard” after detection by some specialized sensory organ (such as an ear), but authors may also use sounds to describe substrateborne vibrations detected by other sensory organs. The use of sound has thus been minimized in this book, even when it is the term most used in the source material to describe signals. Using the strictest definition, vibration can be airborne, waterborne, or substrate-borne in that it represents particle motion in a fluid or elastic body. Particle motion, rather than the medium, is the issue in defining vibration. Further, vibrations in a fluid always accompany vibrations in a solid with which it is in contact. We tend to use vibration to refer to substrate-borne waves, and this convention is followed in this book, even though this is an ambiguous usage that engineers will find appalling. Vibration signals that are propagated via the substrate, atmosphere, or surface waves of water, as well as airborne and waterborne signals propa-

Communication and the Medium

17

gated as compressional waves, could all be considered acoustic, and so acoustic is not used in this book except in direct quotations. Likewise, for those who study animals whose signals are carried in soil or sand, seismic and substrate are synonymous. However, vibrations carried in plants are not seismic, by strict definition, because they are not carried through the earth. For all those authors who wrote of seismic signals, the alternatives vibration signal or substrate-borne vibration signal have been used. The terms song and noise are also not restricted to medium. Signals can be songs, or they can be noisy, but they are distinguished by whether they contain a fundamental frequency with harmonic overtones (song) or whether energy is broadly represented in frequencies that are not whole number multiples of the fundamental. Songs can be propagated via the substrate, air, or water and are distinguished in this book with terms such as airborne song or vibrational song. Somewhat easier to elucidate, but not significantly so, is the distinction among the types of waves produced and detected by animals in their communication. A Primer on Wave Theory Markl (1983) distinguishes among vibration events that induce contact vibrations (when one individual directly touches another with varying pressure: tapping, drumming, antennating, etc.), nearfield medium motions (generated in the fluid medium surrounding the sender and perceived by movement detectors such as hairs or cerci), or boundary vibrations (substrate vibrations at the interface between solid/air, water/air, solid/water, solid/solid, etc.). Most of our focus is on boundary vibrations; however, examples of both contact and nearfield motion are included in subsequent pages, especially when they are difficult to tease apart from other events, or where an evolving story would be truncated and incomplete without their inclusion. Markl (1983) subdivides boundary waves into subcategories: pure longitudinal, quasi-longitudinal, transverse, torsional, and bending waves. However, we are cautioned to remember that “different wave types are characterized by their different kinds of motions in relation to the direction of wave energy propagation, by their different speeds of propagation and—depending on the conducting material and its geometry—by the different attenuations with which they are propagated” (Markl 1983, 338) (Figure 2.1). Likewise, one disturbance

18

Vibrational Communication in Animals

Figure 2.1. Types of waves in elastic solids: (1) compressional, or P-waves,

(2) shear, or S-waves, (3) Rayleigh waves, and (4) Love waves. The arrow shows the direction of propagation. Reprinted with kind permission of Springer Science and Business Media and The Journal of Comparative Physiology from Aicher and Tautz (1990, 346).

Communication and the Medium

19

event may excite different types of vibrations in the same substrate, and when passing through some sort of junction (e.g., where a twig meets a larger stem) a given wave may even be converted into an entirely different type (Markl 1983). Pure longitudinal waves, also known as compressional waves (Brownell 1977), are restricted to “the interior of a medium that extends over many wavelengths in all directions” (Markl 1983, 338), and particle oscillation is in the direction that the wave is propagating. A variety of surface waves are produced when pure longitudinal waves arrive at boundaries, and the ideal vibration pattern is changed (Gogala 1985a). Quasi-longitudinal waves, also known as symmetric waves, are produced by pure longitudinal waves at the boundary with rod- or plate-like structures and have a slightly lower propagation velocity than the pure longitudinal waves (Gogala 1985a). Quasi-longitudinal waves are often propagated through structures such as plant stems or leaves, or spider web strands, where one or two dimensions are small in relation to the wavelength. The displacement perpendicular to the surface, where an animal might be found, would only be 1/100 to 1/1000 times the displacement in the longitudinal wave’s propagation direction, but very sensitive receptors could perceive that level of displacement (Markl 1983). Quasi-longitudinal waves are thought to be a likely candidate when looking for a mechanism for vibrational signaling (Gogala 1985a). Transverse waves only exist in solids (Gogala 1985a) and “can only be propagated in solid materials, which are able to sustain and transmit elastic shearing forces. Particle motion in them is perpendicular to the direction of wave propagation and in the plane of the surface of a body boundary” (Markl 1983, 339). The propagation velocity of plane transverse waves is about 60% that of quasi-longitudinal waves (Markl 1983). Rayleigh waves are produced by a combination of longitudinal and transverse waves at boundaries of solids and have a shear component that is perpendicular to the plane of the body surface. Their propagation velocity is even lower: 4–13% less than transverse waves (Gogala 1985a). Love waves are transverse surface waves (Markl 1983). Torsional waves, which appear similar to transverse waves and are likely to be seen in parts of plants (Gogala 1985a), have not been explored as a source of biologically relevant information. They are “the type of transverse vibrations transmitted in long structures whose diameters are small in comparison to their length. If the structure has rotational symmetry

20

Vibrational Communication in Animals

(as in a circular rod) there is no component of motion perpendicular to the surface and wave propagation along the rod’s axis is the same as that for a plane transverse wave” (Markl 1983, 339). Bending waves appear similar to transverse waves and are sometimes referred to as transverse waves, but in the strictest interpretation they are quite different. One difference is that their propagation velocity is frequency dependent, and bending waves are propagated with dispersion (Markl 1983). This means that the different frequency components in dispersion waves travel at different speeds, are not in phase with each other, and arrive at a given distance from the source at different times. Animals might be able to use these phase differences in locating the signal source (Gogala 1985a). A second difference between bending waves and plane transverse waves is that “the phase and group velocity are not the same, the latter being twice the phase velocity” (Gogala 1985a, 119). The group velocity “can be regarded as the propagation velocity of the carrier wave envelope of a wave bundle or wave packet” (Markl 1983, 339). Bending waves occur in a structure with a small diameter in relation to either the structure’s length or the wavelength: the diameter must be less than 1/6 the wavelength (Markl 1983). “Particle motion in bending waves is perpendicular to the structural surface at rest because the whole structure is rhythmically bent” (Markl 1983, 339). Boundary waves on the surface of water present a special case. A disturbance event will generate waves in which particle movement (in ellipses or circles) is dependent on water depth in relation to the wavelength. Likewise, propagation velocity at wavelengths greater than or equal to 10 cm (or frequency of less than or equal to 4 Hz) is dependent on gravity (gravity waves), whereas at higher frequencies (greater than or equal to 46 Hz or wavelength less than or equal to 0.6 cm) it is dependent on surface tension (capillary waves). In between, propagation velocity is dependent on both (capillary-gravity waves). Gravity waves in shallow water, like bending waves, are propagated with dispersion, while the dispersive waves in deeper water are more complicated. Attenuation of surface waves is also frequency dependent (Markl 1983). Since this book is intended as a study of vibrational communication in animals, the focus has been on substrate vibrations, whether those are via the surface of water, soils of various kinds, plant tissues, or spider webs. These vibration signals all involve at least some particle motion that is perpendicular to the direction of propagation. The

Communication and the Medium

21

distinction between sound and vibration underwater is not a clear one, and the discussion of communication through water has been limited to very specific cases, as has that of infrasound and tactile (contact) communication signaling. The Issue of the Medium Animals create substrate vibrations, including those used in communication, through a variety of behaviors. Substrates can be soils, plants, water, honeycombs, and spider webs, but even within these broad groupings, great variation often exists in the characteristics of any given substrate (i.e., sand vs. clay soils, wet vs. dry sand, soft vs. rigid stems, leaves with parallel veins vs. leaves with netted veins, or even leaves vs. stems of the same plant). How much damping or filtering of the emitted signal occurs during propagation? How much of the signal is reflected, and do vibrations composed of energy across broad frequency bands propagate with dispersion, or are they phase locked? We have made progress in 30 years of study of vibration signaling, but much is still unknown. Further, the same behavior generates different kinds of waveforms, depending on the substrate, and different behaviors create waves of varying types even within the same substrate. For example, drumming is a percussive event, and we would expect that a drumming individual would produce longitudinal, or compression, waves (Morris 1980). Drumming on sand, or tree bark, or dry leaves are similar events from a behavioral perspective, but the transmission of the signals may be quite different, or they may not. Do drumming on a plant stem and stridulating while standing on a plant leaf create the same type of wave? Should and might are still part of the discussion until more data are gathered. Can we account for the tendency of animals to tremulate on plants but drum on sand because of the types of waves created by the behavior? Perhaps. Soils Soil is not a static substrate. Propagation properties of soil change from day to day with changes in temperature and moisture content (Hill and Shadley 2001). Further, generalizations and extrapolations

22

Vibrational Communication in Animals

from one habitat to another are difficult to make (and to support empirically) because of variations in such things as particle size, degree of heterogeneity, and overall complexity of a soil. Some animals, such as the jumping spider Habronattus dossenus, routinely traverse an environmentally heterogenous substrate on a daily basis. H. dossenus can be located on sand, rock, and leaf litter in the Sonoran desert of North America and will court on all three substrates. The three substrates have very different filtering properties, and most successful courtship was on leaf litter, which was also the most favorable signaling substrate of the three (Elias, Mason, and Hoy 2004). In observing animals and designing simulations to test their responses, characteristics of the medium must be considered and incorporated into the design because “seismic waveforms (observed with geophones) produced by impulsive mechanical stimuli (including those produced by the thumper) are largely determined by the physical properties of the ground and largely independent of the stimulus source” (Lewis et al. 2001, 1195). In studying terrestrial or burrowing vertebrates or invertebrates, in rain forest or desert or prairie, the issue of the medium is nontrivial. As long as the temporal pattern and intensity of the simulated signal are a good approximation of a naturally produced one, the receiving animal will respond, whether the signal is produced by a living source or a mechanical one (Brownell and Farley 1979c; Hill and Shadley 2001; Lewis et al. 2001). In grasslands soils, attenuation of vibration signals is strongly dependent on frequency. For example, in a test of signal propagation at a frequency of 142.5 Hz, “the vibration level fell off at a rate of 44 dB per decade increase in distance” (Hill and Shadley 1997, 462). In addition, the soil is not a noise-free environment. Soil organisms produce vibrations as they move, forage, and perhaps communicate. Anyone who has ever listened to the Earth using a geophone in a desert or grassland when the wind is blowing knows the sounds of the tugging of grasses against the soil, which are much like the creaking and groaning of a wooden ship, or an old house, in the wind. There is also incidental noise in most inhabited regions due to vehicle traffic, and engineers are aware, even if biologists are not, of a rather ubiquitous 60 Hz vibration signal (50 Hz in Europe) from transducers in both substrate and atmosphere that could be interpreted by the unwary as evidence of vibration produced by animals. “It is not uncommon for recordings

Communication and the Medium

23

from electrical transducers to contain noise at 60 Hz and multiples of 60 Hz. In outdoor measurement, these are often due to emissions from electrical power transmission lines” (Hill and Shadley 2001, 1204). Thus, collaborative research across disciplinary lines is a wise choice for all those hoping to disentangle the part of a putative signal that is from the animal and the part that is from its environment. Even though Brownell (1977) knew that sand was supposed to absorb low- frequency vibrations instead of propagate them, he identified two separate waves propagating through desert sand when a piezoelectric crystal generated displacement pulses on the sand surface (Figure 2.2). A fast wave varied in velocity with sand compaction from 95 to 120 m/sec and was determined to be a compressional body wave, or P-wave. The slower wave velocity was 40 m/sec in loose sand and 50 m/sec in lightly compacted sand. This slower wave was determined to be a Rayleigh wave, and both wave types traveled at about 1/10th their documented velocities in other natural substrates. Both waves were detected when the receiver axis was placed radially to the source, but amplitude of the Rayleigh wave increased, and that of the P-wave decreased, when the receiver was oriented perpendicular to the surface. Likewise, the Rayleigh wave could be damped by placing an absorbant object between source and receiver, whereas the P-wave was not affected (Brownell 1977). Rayleigh waves and Love waves, which are boundary waves, are the principal components of the vibrations produced by drumming fiddler crabs, and the surface P-waves, or primary waves, carried much lower energy (Aicher and Tautz 1990).

Vibrational Wave Transmission in Sand Source

R

sand

V (m/s) Freq (Hz) (cm) G(r) 10(db/cm) Surface waves (>70%) 45–50 300–400 ~10–15 cm ~1/r 0.20–0.40 Rayleigh (R)

----~1/r Love (L)

P Body waves ( 60 kHz) airborne sexual advertisement calls (Morris et al. 1994), and an undescribed species in the genus Arachnoscelis from Columbia actually sings a pure-tone song at 129 kHz (Montealegre-Z, Morris, and Mason 2006). Bushcrickets sing these very high-frequency calls with significant intervals of silence that result in a reduced duty cycle for the song. The signal bursts are short, but at these very high frequencies they would also attenuate rapidly with distance and could thwart eavesdropping predators. The signals would be of less importance in long-range advertisement to females, but they are thought to have evolved in areas where katydids are heavily preyed upon by bats, and so the high-frequency signals would also be less conspicuous, though not undetectable, to nearby bats. Many of the same katydid species have developed elaborate vibration signaling that employs a mechanism separate from stridulation, and in some cases the vibration event is the only calling song produced by males. Tremulation of the male body often produces the substrate signal, but in contrast to temperate zone tremulation, neotropical species actually employ the behavior in “rhetorical” sexual advertisement in the absence of females, rather than for close-range pairing. Trade-offs are always present, however, because in gaining privacy from bat eavesdropping through the evolution of substrate-borne vibration signals, the door would have been opened for eavesdropping on vibration signals by wandering spiders in the genus Cupiennius (Morris et al. 1994). It should be noted that the concept of privacy is not unique to communication via the vibration channel and may simply be a human construct used to explain phenomena about which we are short on information. The neotropical katydid Copiphora rhinoceros may also have evolved tremulation-induced substrate signaling in response to bat predation.

208

Vibrational Communication in Animals

However, an additional selective mechanism may have been avoidance of cuckoldry. Males of this species produce very large spermatophores, which are passed to females during copulation, and they invest both time and body mass in spermatophore assembly. One male was observed to spend five hours in courtship and copulation, whether or not this is standard for the species. Males within a refractory period after mating can still copulate and pass sperm packets, but the tremulation signals they produce with and without the burden of a spermatophore are substantively different. Signals produced by males should thus provide an honest signal to females of ability to invest direct nutritional benefits because tremulation signals would be changed for any given male with the weight loss that follows passing a spermatophore. Neighboring males that have already passed spermatophores could take advantage of the airborne signal of a male with a spermatophore to offer, but tremulation signals would only be detectable by a female on the same branch of vegetation. Thus, cuckoldry could be avoided by the male who actually is investing nutrients for the benefit of future offspring (Morris 1980). Likewise, in a species such as the black-horned tree cricket, Oecanthus nigricornis, where the male feeds his mate secretions from a metanotal gland, attracting too many females could be a problem. Immediately following mating and before he is capable of inseminating others, females that were able to find him based on his airborne or pheromonal signaling could harvest his nutritional contribution without providing him the benefit of parenthood. Use of vibrational signals in courtship at the last stage of pair formation thus affords some privacy (Bell 1980a). Epiphenomena Associated with Other Behaviors Randall (2001) suggested that footdrumming in mammals evolved independently in different lineages, possibly as a ritualization of other behaviors. The conflict between running away from a threat and moving toward it, in inspection and for risk assessment, could have led to drumming as a displacement activity. This hypothesis has stronger support than a suggestion that footdrumming evolved from a digging behavior because some kangaroo rats drum just before they run to chase an intruder. Evolution of a signal, however, is perhaps linked in

Why Vibration?

209

an unexplored way to body mass in kangaroo rats: large-bodied territory defenders drum, whereas medium-sized species have a rudimentary signal, and the smallest species do not drum (Randall 2001). The days of arguing in a brew pub whether or not communication through the substrate vibration channel even exists are happily gone. Scientists are working across vertebrate and arthropod taxa to document and test hypotheses for behaviors and to describe sending and receiving mechanisms. When a baffling behavior is encountered, it is no longer rare for someone to suggest that vibration signaling may be involved. However, a great deal of alpha-level work is still required before a unified theory of vibrational communication can be formulated. So much has been accomplished in the last five years that we are justifiably encouraged to hope that energetic young scientists will accept the challenge to use their interdisciplinary training and creativity to make a name for themselves through revealing the mysteries surrounding this communication channel. Einstein was chided near the end of his life for “wasting” so much time searching for answers that would reveal the unified field theory. We are told he replied that he could afford the time because he had already made his reputation, whereas young scientists were expected to work on problems that held promise that a solution actually did exist. Now the standard model in physics includes an explanation for everything but the gravitational field. Perhaps the study of vibration communication will develop in the same way. Those who have worked so hard to study vibration as side projects to other, sponsored research are seeing younger scientists begin to attract funding for projects where vibrational communication plays a key role. The torch is being passed. Do not let it fall.

References

Abbott, J. C., and K. W. Stewart. 1993. “Male Search Behavior of the Stonefly, Pteronarcella badia (Hagen) (Plecoptera: Pteronarcyidae), in Relation to Drumming.” Journal of Insect Behavior 6: 467–481. Aicher, B., H. Markl, W. M. Masters, and H. L. Kirschenlohr. 1983. “Vibration Transmission Through the Walking Legs of the Fiddler Crab, Uca pugilator (Brachyura, Ocypodidae) as Measured by Laser Doppler Vibrometry.” Journal of Comparative Physiology A 150: 483–491. Aicher, B., and J. Tautz. 1990. “Vibrational Communication in the Fiddler Crab, Uca pugilator. I. Signal Transmission Through the Substratum.” Journal of Comparative Physiology A 166: 345–353. Anderson, C. J. 1973. “Animals, Earthquakes, and Eruptions.” Field Museum Natural History Bulletin 44: 9–11. Andrade, M. C. B., and A. C. Mason. 2000. “Male Condition, Female Choice, and Extreme Variation in Repeated Mating in a Scaly Cricket, Ornebius aperta (Orthoptera: Gryllidae: Mogoplistinae).” Journal of Insect Behavior 13: 483–497. Arnason, B. T., C. E. O’Connell, and L. A. Hart. 1998. “Long Range Seismic Characteristics of Asian Elephant (Elephus maximus) Vocalizations and Locomotion.” Journal of the Acoustical Society of America 104: 1810. Arnqvist, G. 1989. “Multiple Mating in a Water Strider: Mutual Benefits or Intersexual Conflict?” Animal Behaviour 38: 749–756. ———. 1997. “The Evolution of Water Strider Mating Systems: Causes and Consequences of Sexual Conflicts.” In The Evolution of Mating Systems in Insects and Arachnids, eds. J. C. Choe and B. J. Crespi, pp. 146–163. Cambridge: Cambridge University Press.

212

References

Ashcroft, D. W., and C. S. Hallpike. 1934. “Action Potentials in the Saccular Nerve of the Frog.” Journal of Physiology (London) 81: 23P. Autrum, H., and W. Schneider. 1948. “Vergleichende Untersuchungen über den Erschütterungssinn der Insekten.” Zeitschrift für vergleichende Physiologie 31: 77–88. Bach-y-Rita, P. 2004. “Tactile Sensory Substitution Studies.” Annals of the New York Academy of Sciences 1013: 83–91. Bach-y-Rita, P., C. C. Collins, F. A. Saunders, B. White, and L. Scadden. 1969. “Vision Substitution by Tactile Image Projection.” Nature 221: 963–964. Bacher, S., J. Casas, and S. Dorn. 1996. “Parasitoid Vibrations as Potential Releasing Stimulus of Evasive Behaviour in a Leafminer.” Physiological Entomology 21: 33–43. Bacher, S., J. Casas, F. Wäckers, and S. Dorn. 1997. “Substrate Vibrations Elicit Defensive Behaviour in Leafminer Pupae.” Journal of Insect Physiology 10: 945–952. Bailey, W. J. 1985. “Acoustic Cues for Female Choice in Bushcrickets (Tettigoniidae).” In Acoustic and Vibrational Communication in Insects, eds. K. Kalmring and N. Elsner, pp. 101–110. Berlin: Paul Parey. ———. 2003. “Insect Duets: Underlying Mechanisms and Their Evolution.” Physiological Entomology 28: 157–174. Ball, E. E., B. P. Oldfield, and K. M. Rudolph. 1989. “Auditory Organ Structure, Development, and Function.” In Cricket Behavior and Neurobiology, eds. F. Huber, J. E. Moore, and W. Loher, pp. 391–422. Ithaca, NY: Cornell University Press. Barnett, K. E., R. B. Cocroft, and L. J. Fleishman. 1999. “Possible Communication by Substrate Vibration in a Chameleon.” Copeia 1999: 225–228. Baroni-Urbani, C., M. W. Buser, and E. Schilliger. 1988. “Substrate Vibration During Recruitment in Ant Social Organization.” Insectes Sociaux 35: 241–250. Barth, F. G. 1982. “Spiders and Vibratory Signals: Sensory Reception and Behavioral Significance.” In Spider Communication, eds. P. N. Witt and J. S. Rovner, pp. 67–122. Princeton, NJ: Princeton University Press. ———. 1997. “Vibratory Communication in Spiders: Adaptation and Compromise at Many Levels.” In Orientation and Communication in Arthropods, ed. M. Lehrer, pp. 247–272. Basel: Birkhäuser Verlag. ———. 2002. “Spider Senses—Technical Perfection and Biology.” Zoology 105: 271–285. Barth, F. G., H. Bleckmann, J. Bohnenberger, and E.-A. Seyfarth. 1988. “Spiders of the Genus Cupiennius Simon 1891 (Araneae, Ctenidae):

References

213

II. On the Vibratory Environment of a Wandering Spider.” Oecologia 77: 194–201. Barth, F. G., and Geethabali. 1982. “Spider Vibration Receptors: Threshold Curves of Individual Slits in the Metatarsal Lyriform Organ.” Journal of Comparative Physiology A 148: 175–185. Barth, F. G., and A. Schmitt. 1991. “Species Recognition and Species Isolation in Wandering Spiders (Cupiennius spp.; Ctenidae).” Behavioral Ecology and Sociobiology 29: 333–339. Baurecht, D., and F. G. Barth. 1992. “Vibratory Communication in Spiders I. Representation of Male Courtship Signals by Female Vibration Receptor.” Journal of Comparative Physiology A 171: 231–243. ———. 1993. “Vibratory Communication in Spiders II. Representation of Parameters Contained in Synthetic Male Courtship Signals by Female Vibration Receptor.” Journal of Comparative Physiology A 173: 309–319. Beesley, P. L., C. J. Glasby, and G. J. B. Ross. 2000. Polychaetes and Allies: The Southern Synthesis. Collingwood, Victoria, AU: CSIRO Publishing. Bell, P. D. 1980a. “Multimodal Communication by the Black-horned Tree Cricket Oecanthus nigricornis (Walker) (Orthoptera: Gryllidae).” Canadian Journal of Zoology 58: 1861–1868. ———. 1980b. “Transmission of Vibrations Along Plant Stems: Implications for Insect Communication.” Journal of the New York Entomological Society 88: 210–216. Belwood, J. J., and G. K. Morris. 1987. “Bat Predation and Its Influence on Calling Behavior in Neotropical Katydids.” Science 238: 64–67. Bender, H. 2006. “Structure and Function of the Eastern Grey Kangaroo (Macropus giganteus) Foot Thump.” Journal of Zoology (London) 268: 415–422. Bennett, N. C., and J. U. M. Jarvis. 1988. “The Reproductive Biology of the Cape Mole-rat, Georychus capensis (Rodentia, Bathyergidae).” Journal of Zoology (London) 214: 95–106. Birch, M. C., and J. J. Keenlyside. 1991. “Tapping Behavior Is a Rhythmic Communication in the Death-watch Beetle, Xestobium rufovillosum (Coleoptera: Anobiidae).” Journal of Insect Behavior 4: 257–263. Bleckmann, H., and F. G. Barth. 1984. “Sensory Ecology of a Semi-aquatic Spider (Dolomedes triton). II. The Release of Predatory Behavior by Water Surface Waves.” Behavioral Ecology and Sociobiology 14: 303–312. Bliss, J. 1962. “Kinesthetic-tactile communications.” IEEE Transactions on Information Theory 8: 92–99. Blumstein, D. T., J. C. Daniel, A. S. Griffin, and C. S. Evans. 2000. “Insular Tammar Wallabies (Macropus eugenii) Respond to Visual but Not Acoustic Cues from Predators.” Behavioral Ecology 11: 528–535.

214

References

Boake, C. R. B., and T. Poulsen. 1997. “Correlates Versus Predictors of Courtship Success: Courtship Song in Drosophila silvestris and D. heteroneura.” Animal Behaviour 54: 699–704. Bouley, D. M., C. N. Alarcón, T. Hildebrandt, and C. E. O’Connell-Rodwell. 2007. The distribution, density and three-dimensional histomorphology of Pacinian corpuscles in the foot of the Asian elephant (Elephas maximus) and their potential role in seismic communication. Journal of Anatomy 211: 428–435. Bradbury, J. W., and S. L. Vehrencamp. 1998. Principles of Animal Communication. Sunderland, MA: Sinauer Associates. Breidbach, O. 1986. “Studies on the Stridulation of Hylotrupes bajulus (L.) (Cerambycidae, Coleoptera): Communication Through Support Vibration—Morphology and Mechanics of the Signal.” Behavioural Processes 12: 169–186. Brennan, B. J. 2005. “Vibratory Communication in the Social Paper Wasp Polistes dominulus (Hymenoptera: Vespidae).” PhD diss., Cornell University. Brill, R. L., M. L. Sevenich, T. J. Sullivan, J. D. Sustman, and R. E. Witt. 1988. “Behavioral Evidence for Hearing Through the Lower Jaw by an Echolocating Dolphin (Tursiops truncatus).” Marine Mammal Science 4: 223–230. Broad, G. R., and D. L. J. Quicke. 2000. “The Adaptive Significance of Host Location by Vibrational Sounding in Parasitoid Wasps.” Proceedings of the Royal Society of London B 267: 2403–2409. Brooks, S. J. 1987. “Stridulatory Structures in Three Green Lacewings (Neuroptera: Chrysopidae).” International Journal of Insect Morphology and Embryology 16: 237–244. Brownell, P., and R. D. Farley. 1979a. “Detection of Vibrations in Sand by Tarsal Sense Organs of the Nocturnal Scorpion, Paruroctonus mesaensis.” Journal of Comparative Physiology A 131: 23–30. ———. 1979b. “Orientation to Vibrations in Sand by the Nocturnal Scorpion Paruroctonus mesaensis: Mechanism of Target Localization.” Journal of Comparative Physiology A 131: 31–38. ———. 1979c. “Prey-localizing Behaviour of the Nocturnal Desert Scorpion, Paruroctonus mesaensis: Orientation to Substrate Vibrations.” Animal Behaviour 27: 185–193. Brownell, P. H. 1977. “Compressional and Surface Waves in Sand Used by Desert Scorpions to Locate Prey.” Science 197: 479–482. ———. 1984. “Prey Detection by the Sand Scorpion.” Scientific American 251: 86–97.

References

215

Brownell, P. H., and J. L. van Hemmen. 2001. “Vibration Sensitivity and a Computational Theory for Prey-localizing Behavior in Sand Scorpions.” American Zoologist 41: 1229–1240. Buchmann, S. L. 1980. “Preliminary Anthecological Observations on Xiphidium caeruleum Aubl. (Monocotyledoneae: Haemodoraceae) in Panama.” Journal of the Kansas Entomological Society 53: 685–699. Burda, H., V. Bruns, and E. Nevo. 1989. “Middle Ear and Cochlear Receptors in the Subterranean Mole-rat, Spalax ehrenbergi.” Hearing Research 39: 225–230. Burger, J. 1998. “Antipredator Behaviour of Hatchling Snakes: Effects of Incubation Temperature and Simulated Predators.” Animal Behaviour 56: 547–553. Butlin, R. K. 1993. “The Variability of Mating Signals and Preferences in the Brown Planthopper, Nilaparvata lugens (Homoptera: Delphacidae).” Journal of Insect Behavior 6: 125–140. Calne, D. B., and C. A. Pallis. 1966. “Vibratory Sense: A Critical Review.” Brain 89: 723–746. Cardoso, A. J., and W. R. Heyer. 1995. “Advertisement, Aggressive, and Possible Seismic Signals of the Frog Leptodactylus syphax (Amphibia, Leptodactylidae).” Alytes 13: 67–76. Caro, T. M., L. Lombardo, A. W. Goldizen, and M. Kelly. 1995. “Tail-flagging and Other Antipredator Signals in White-tailed Deer: New Data and Synthesis.” Behavioral Ecology 6: 442–450. Casas, J. 1989. “Foraging Behaviour of a Leafminer Parasitoid in the Field.” Ecological Entomology 14: 257–265. Casas, J., S. Bacher, J. Tautz, R. Meyhöfer, and D. Pierre. 1998. “Leaf Vibrations and Air Movements in a Leafminer-parasitoid System.” Biological Control 11: 147–153. Castellanos, I., and P. Barbosa. 2006. “Evaluation of Predation Risk by a Caterpillar Using Substrate-borne Vibrations.” Animal Behaviour 72: 461–469. Caston, J., W. Precht, and R. H. I. Blanks. 1977. “Responses of Lagena Afferents to Natural Stimuli.” Journal of Comparative Physiology 118: 273–289. Catania, K. C. 1999. “A Nose that Looks Like a Hand and Acts Like an Eye: The Unusual Mechanosensory System of the Star-nosed Mole.” Journal of Comparative Physiology A 185: 367–372. Christensen-Dalsgaard, J., and A. Elepfandt. 1995. “Biophysics of Underwater Hearing in the Clawed Frog, Xenopus laevis.” Journal of Comparative Physiology A 176: 317–324.

216

References

Christensen-Dalsgaard, J., and M. B. Jørgensen. 1988. “The Response Characteristics of Vibration-sensitive Saccular Fibers in the Grassfrog, Rana temporaria.” Journal of Comparative Physiology A 162: 633–638. ———. 1996. “Sound and Vibration Sensitivity of VIIIth Nerve Fibers in the Grassfrog, Rana temporaria.” Journal of Comparative Physiology A 179: 437–445. Christensen-Dalsgaard, J., and P. M. Narins. 1993. “Sound and Vibration Sensitivity of VIIIth Nerve Fibers in the Frogs Leptodactylus albilabris and Rana pipiens pipiens.” Journal of Comparative Physiology A 172: 653–662. Claridge, M. F. 1985. “Acoustic Signals in the Homoptera: Behavior, Taxonomy, and Evolution.” Annual Review of Entomology 30: 297–317. ———. 1990. “Acoustic Recognition Signals: Barriers to Hybridization in Homoptera Auchenorrhyncha.” Canadian Journal of Zoology 68: 1741–1746. Claridge, M. F., and P. E. Howse. 1968. “Songs of Some British Oncopsis Species (Hemiptera: Cicadellidae).” Proceedings of the Royal Entomological Society of London A 43: 57–61. Claridge, M. F., J. C. Morgan, and M. S. Moulds. 1999. “Substrate-transmitted Acoustic Signals of the Primitive Cicada, Tettigarcta crinita Distant (Hemiptera Cicadoidea, Tettigarctidae).” Journal of Natural History 33: 1831–1834. Clayton, D. 2001. “Acoustic Calling in Four Species of Ghost Crabs: Ocypode jousseaumei, O. platytarsus, O. rotundata and O. saratan (Brachyura: Ocypodidae).” Bioacoustics 12: 37–55. ———. 2005. “Substrate (Acoustic/vibrational) Communication and Ecology of the Ghost Crab Ocypode jousseaumei (Brachyura: Ocypodidae).” Marine and Freshwater Behaviour and Physiology 38: 53–70. Cocroft, R. B. 1996. “Insect Vibrational Defence Signals.” Nature 382: 679–680. ———. 1999a. “Offspring-parent Communication in a Subsocial Treehopper (Hemiptera: Membracidae: Umbonia crassicornis).” Behaviour 136: 1–21. ———. 1999b. “Parent-offspring Communication in Response to Predators in a Subsocial Treehopper (Hemiptera: Membracidae: Umbonia crassicornis).” Ethology 105: 553–568. ———. 2001. “Vibrational Communication and the Ecology of Group-living, Herbivorous Insects.” American Zoologist 41: 1215–1221. ———. 2002. “Antipredator Defense as a Limited Resource: Unequal Predation Risk in Broods of an Insect with Maternal Care.” Behavioral Ecology 13: 125–133.

References

217

———. 2003. “The Social Environment of an Aggregating, Ant-attended Treehopper (Hemiptera: Membracidae: Vanduzea arquata).” Journal of Insect Behavior 16: 79–95. ———. 2005. “Vibrational Communication Facilitates Cooperative Foraging in a Phloem-feeding Insect.” Proceedings of the Royal Society of London B 272: 1023–1029. Cocroft, R. B., and R. L. Rodríguez. 2005. “The Behavioral Ecology of Insect Vibrational Communication.” BioScience 55: 323–334. Cocroft, R. B., H. J. Shugart, K. T. Konrad, and K. Tibbs. 2006. “Variation in Plant Substrates and Its Consequences for Insect Vibrational Communication.” Ethology 112: 779–789. Cocroft, R. B., T. D. Tieu, R. R. Hoy, and R. N. Miles. 2000. “Directionality in the Mechanical Response to Substrate Vibration in a Treehopper (Hemiptera: Membracidae: Umbonia crassicornis).” Journal of Comparative Physiology A 186: 695–705. Cˇokl, A. 1983. “Functional Properties of Vibroreceptors in the Legs of Nezara viridula (L.) (Heteroptera, Pentatomidae).” Journal of Comparative Physiology A 150: 261–269. ˇ okl, A., H. L. McBrien, and J. G. Millar. 2001. “Comparison of SubstrateC borne Vibrational Signals of Two Stink Bug Species, Acrosternum hilare and Nezara viridula (Heteroptera: Pentatomidae).” Annals of the Entomological Society of America 94: 471–479. ˇ Cokl, A., and M. Virant-Doberlet. 2003. “Communication with Substrateborne Signals in Small Plant-dwelling Insects.” Annual Review of Entomology 48: 29–50. Cˇ okl, A., M. Virant-Doberlet, and A. McDowell. 1999. “Vibrational Directionality in the Southern Green Stink Bug, Nezara viridula (L.), Is Mediated by Female Song.” Animal Behaviour 58: 1277–1283. Cˇokl, A., M. Virant-Doberlet, and N. Stritih. 2000a. “Temporal and Spectral Properties of the Songs of the Southern Green Stink Bug Nezara viridula (L.) from Slovenia.” European Journal of Physiology Suppl. 439: R168–R170. ———. 2000b. “The Structure and Function of Songs Emitted by Southern Green Stink Bugs from Brazil, Florida, Italy and Slovenia.” Physiological Entomology 25: 196–205. Cˇokl, A., J. Presˇern, M. Virant-Doberlet, G. J. Bagwell, and J. G. Millar. 2004. “Vibratory Signals of the Harlequin Bug and Their Transmission Through Plants.” Physiological Entomology 29: 372–380. ˇ ˇ unic, and M. Virant-Doberlet. 2005. “Tuning of Cokl, A., M. Zorovic´, A. Z Host Plants with Vibratory Songs of Nezara viridula L (Heteroptera: Pentatomidae).” Journal of Experimental Biology 208: 1481–1488.

218

References

Connétable, S., A. Robert, F. Bouffault, and C. Bordereau. 1999. “Vibratory Alarm Signals in Two Sympatric Higher Termite Species: Pseudacanthotermes spiniger and P. militaris (Termitidae, Macrotermitinae).” Journal of Insect Behavior 12: 329–342. Cortopassi, K. A., and E. R. Lewis. 1998. “A Comparison of the Linear Tuning Properties of Two Classes of Axons in the Bullfrog Lagena.” Brain, Behavior and Evolution 51: 331–348. Costa, J. T., T. D. Fitzgerald, A. Pescador-Rubio, J. Mays, and D. H. Janzen. 2004. “Social Behavior of Larvae of the Neotropical Processionary Weevil Phelypera distigma (Boheman) (Coleoptera: Curculionidae: Hyperinae).” Ethology 110: 515–530. Crabb, W. D. 1948. “The Ecology and Management of the Prairie Spotted Skunk in Iowa.” Ecological Monographs 18: 201–232. Credner, S., H. Burda, and F. Ludescher. 1997. “Acoustic Communication Underground: Vocalization Characteristics in Subterranean Social Molerats (Cryptomys sp., Bathyergidae).” Journal of Comparative Physiology A 180: 245–255. Cummings, D. L. D., G. J. Gamboa, and B. J. Harding. 1999. “Lateral Vibrations by Social Wasps Signal Larvae to Withhold Salivary Secretions (Polistes fuscatus, Hymenoptera: Vespidae).” Journal of Insect Behavior 12: 465–473. Daly, M., and S. Daly. 1975. “Behavior of Psammomys obesus (Rodentia: Gerbillinae) in the Algerian Sahara.” Zeitschrift für Tierpsychologie 37: 298–321. Dambach, M. 1989. “Vibrational Responses.” In Cricket Behavior and Neurobiology, eds. F. Huber, J. E. Moore, and W. Loher, pp. 178–197. Ithaca, NY: Cornell University Press. DeLuca, P. A., and G. K. Morris. 1998. “Courtship Communication in Meadow Katydids: Female Preference for Large Male Vibrations.” Behaviour 135: 777–794. Devetak, D. 1998. “Detection of Substrate Vibration in Neuropteroidea: A Review.” Acta Zoologica Fennica 209: 87–94. Devetak, D., and T. Amon. 1997. “Substrate Vibration Sensitivity of the Leg Scolopidial Organs in the Green Lacewing, Chrysoperla carnea.” Journal of Insect Physiology 43: 433–437. Devetak, D., and M. A. Pabst. 1994. “Structure of the Subgenual Organ in the Green Lacewing, Chrysoperla carnea.” Tissue and Cell 26: 249–257. Devetak, D., M. A. Pabst, and S. L. Delakorda. 2004. “Leg Chordotonal Organs and Campaniform Sensilla in Chrysoperla Steinmann 1964 (Neuroptera): Structure and Function.” Denisia 13: 163–171.

References

219

DeVries, P. J. 1990. “Enhancement of Symbioses Between Butterfly Caterpillars and Ants by Vibrational Communication.” Science 248: 1104–1106. DeVrijer, P. W. F. 1986. “Species Distinctiveness and Variability of Acoustic Calling Signals in the Planthopper Genus Javesella (Homoptera: Delphacidae).” Netherlands Journal of Zoology 36: 162–175. DeWinter. A. J. 1992. “The Genetic Basis and Evolution of Acoustic Mate Recognition Signals in a Ribautodelphax Planthopper (Homoptera, Delphacidae). I. The Female Call.” Journal of Evolutionary Biology 5: 249–265. Díaz-Fleischer, F. 2005. “Predatory Behaviour and Prey-Capture DecisionMaking by the Web-weaving Spider Micrathena sagittata.” Canadian Journal of Zoology 83: 268–273. Dierkes, S., and F. G. Barth. 1995. “Mechanism of Signal Production in the Vibratory Communication of the Wandering Spider Cupiennius getazi (Arachnida, Araneae).” Journal of Comparative Physiology A 176: 31–44. Djemai, I., J. Casas, and C. Magal. 2001. “Matching Host Reactions to Parasitoid Wasp Vibrations.” Proceedings of the Royal Society of London B 268: 2403–2408. Doherty, J. A., and H. C. Gerhardt. 1984. “Evolutionary and Neurobiological Implications of Selective Phonotaxis in the Spring Peeper (Hyla crucifer).” Animal Behaviour 32: 875–881. Donahoe, K., L. A. Lewis, and S. S. Schneider. 2003. “The Role of the Vibration Signal in the House-hunting Process of Honey Bee (Apis mellifera) Swarms.” Behavioral Ecology and Sociobiology 54: 593–600. Dorward, P. K., and A. K. McIntyre. 1971. “Responses of Vibration-sensitive Receptors in the Interosseous Region of the Duck’s Hind Limb.” Journal of Physiology (London) 219: 77–87. Ehret, G., J. Tautz, and B. Schmitz. 1990. “Hearing Through the Lungs: Lung-Eardrum Transmission of Sound in the Frog Eleutherodactylus coqui.” Naturwissenschaften 77: 192–194. Eijkman, E. G. J. 1989. “Black Box Analysis of the Skin Senses as a Multiple Communication Channel.” Biological Cybernetics 61: 163–170. Eisenberg, J. F., L. R. Collins, and C. Wemmer. 1975. “Communication in the Tasmanian Devil (Sarcophilus harrisii) and a Survey of Auditory Communication in the Marsupialia.” Zeitschrift für Tierpsychologie 37: 379–399. Elias, D. O., E. A. Hebets, and R. R. Hoy. 2006. “Female Preference for Complex/novel Signals in a Spider.” Behavioral Ecology 17: 765–771. Elias, D. O., N. Lee, E. A. Hebets, and A. C. Mason. 2006. “Seismic Signal Production in a Wolf Spider: Parallel Versus Serial Multi-component Signals.” Journal of Experimental Biology 209: 1074–1084.

220

References

Elias, D. O., A. C. Mason, and R. R. Hoy. 2004. “The Effect of Substrate on the Efficacy of Seismic Courtship Signal Transmission in the Jumping Spider Habronattus dossenus (Araneae: Salticidae).” Journal of Experimental Biology 207: 4105–4110. Elias, D. O., A. C. Mason, W. P. Maddison, and R. R. Hoy. 2003. “Seismic Signals in a Courting Male Jumping Spider (Araneae: Salticidae).” Journal of Experimental Biology 206: 4029–4039. Emerson, A. E., and R. C. Simpson. 1929. “Apparatus for the Detection of Substratum Communication Among Termites.” Science 69: 648–649. Engel, M. S., and F. Dingemans-Bakels. 1980. “Nectar and Pollen Resources for Stingless Bees (Meliponinae, Hymenoptera) in Surinam (South America).” Apidologie 11: 341–350. Esch, H. 1961. “Über die Schallerzeugung beim Werbetanz der Honigbiene.” Zeitschrift für vergleichende Physiologie 45: 1–11. Ewing, A. W. 1989. Arthropod Bioacoustics: Neurobiology and Behaviour. Ithaca, NY: Cornell University Press. Fairbairn, D. J. 1993. “Costs of Loading Associated with Mate-carrying in the Waterstrider, Aquarius remigis.” Behavioral Ecology 4: 224–231. Fernandez-Montraveta, C., and A. Schmitt. 1994. “Substrate-borne Vibrations Produced by Male Lycosa tarentula fasciiventris (Araneae, Lycosidae) During Courtship and Agonistic Interactions.” Ethology 97: 81–93. Field, L. H. 1993. “Observations on Stridulatory, Agonistic, and Mating Behaviour of Hemideina ricta (Stenopelmatidae: Orthoptera), the Rare Banks Peninsula Weta.” New Zealand Entomologist 16: 68–74. Field, L. H., and W. J. Bailey. 1997. “Sound Production in Primitive Orthoptera from Western Australia: Sounds Used in Defence and Social Communication in Ametrus sp. and Hadrogryllacris sp. (Gryllacrididae: Orthoptera).” Journal of Natural History 31: 1127–1141. Field, L. H., and F. C. Rind. 1992. “Stridulatory Behaviour in a New Zealand Weta, Hemideina crassidens.” Zoology (London) 228: 371–394. Field, S. A., and M. A. Keller. 1993. “Courtship and Intersexual Signaling in the Parasitic Wasp Cotesia rubecula (Hymenoptera: Braconidae).” Journal of Insect Behavior 6: 737–750. Finck, A. 1981. “The Lyriform Organ of the Orb-weaving Spider Araneous sericatus: Vibration Sensitivity is Altered by Bending the Leg.” Journal of the Acoustical Society of America 70: 231–233. Fine, M. L. 1983. “Frequency Response of the Swimbladder of the Oyster Toadfish.” Comparative Biochemistry and Physiology A 74: 659–663. Fischer, S., J. Samietz, F. L. Wäckers, and S. Dorn. 2001. “Interaction of Vibrational and Visual Cues in Parasitoid Host Location.” Journal of Comparative Physiology A 187: 785–791.

References

221

Fletcher, L. E., J. E. Yack, T. D. Fitzgerald, and R. R. Hoy. 2006. “Vibrational Communication in the Cherry Leaf Roller Caterpillar Caloptilia serotinella (Gracillarioidea: Gracillariidae).” Journal of Insect Behavior 19: 1–18. Forrest, T. G. 1987. “Sinistrality in the Southern and Tawny Mole Crickets (Gryllotalpidae: Scapteriscus).” Florida Entomologist 70: 284–286. Forster Cooper, C. 1928. “On the Ear Region of Certain of the Chrysochloridae.” Philosophical Transactions of the Royal Society of London B 216: 265–283. Foxe, J. J., G. R. Wylie, A. Martinez, C. E. Schroeder, D. C. Javitt, D. Guilfoyle, W. Ritter, and M. M. Murray. 2002. “Auditory-somatosensory Multisensory Processing in Auditory Association Cortex: An fMRI Study.” Journal of Neurophysiology 88: 540–543. Frings, H., and F. Little. 1957. “Reactions of Honey Bees in the Hive to Simple Sounds.” Science 125: 122. Frohlich, C., and R. E. Buskirk. 1982. “Transmission and Attenuation of Vibration in Orb Spider Webs.” Journal of Theoretical Biology 95: 13–36. Fuchs, S. 1976. “The Response to Vibrations of the Substrate and Reactions to the Specific Drumming in Colonies of Carpenter Ants (Camponotus, Formicidae, Hymenoptera).” Behavioral Ecology and Sociobiology 1: 155–184. Gans, C., and E. G. Wever. 1972. “The Ear and Hearing in Amphisbaenia (Reptilia).” Journal of Experimental Zoology 179: 17–34. Gogala, M. 1985a. “Vibrational Communication in Insects (Biophysical and Behavioural Aspects).” In Acoustic and Vibrational Communication in Insects, eds. K. Kalmring and N. Elsner, pp. 117–126. Berlin: Paul Parey. ———. 1985b. “Vibrational Songs of Land Bugs and Their Production.” In Acoustic and Vibrational Communication in Insects, eds. K. Kalmring and N. Elsner, pp. 143–150. Berlin: Paul Parey. Gogala, M., A. Cˇokl, K. Drasˇlar, and A. Blazˇevic. 1974. “Substrate-borne Sound Communication in Cydnidae (Heteroptera).” Journal of Comparative Physiology 94: 25–31. Göpfert, M. C., H. Briegel, and D. Robert. 1999. “Mosquito Hearing: Soundinduced Antennal Vibrations in Male and Female Aedes Aegypti.” Journal of Experimental Biology 202: 2727–2738. Goulson, D., M. C. Birch, and T. D. Wyatt. 1994. “Mate Location in the Deathwatch Beetle, Xestobium rufovillosum De Geer (Anobiidae): Orientation to Substrate Vibrations.” Animal Behaviour 47: 899–907. Greenfield, M. D. 2002. Signalers and Receivers: Mechanisms and Evolution of Arthropod Communication. New York: Oxford University Press. Gregg, C. E., B. F. Houghton, D. Paton, R. Lachman, J. Lachman, D. M. Johnston, and S. Wongbusarakum. 2006. “Natural Warning Signs of Tsunamis: Human Sensory Experience and Response to the 2004 Great

222

References

Sumatra Earthquake and Tsunami in Thailand.” Earthquake Spectra 22: S671–S691. Gregory, J. E. 1973. “An Electrophysiological Investigation of the Receptor Apparatus of the Duck’s Bill.” Journal of Physiology (London) 229: 151–164. Gregory, J. E., A. K. McIntyre, and U. Proske. 1986. “Vibration-evoked Responses from Lamellated Corpuscles in the Legs of Kangaroos.” Experimental Brain Research 62: 648–653. Gu˝nther, R. H., C. E. O’Connell-Rodwell, and S. L. Klemperer. 2004. “Seismic Waves from Elephant Vocalizations: A Possible Communication Mode?” Geophysical Research Letters 31: 1–4. Gwynne, D. T. 2004. “Reproductive Behavior of Ground Weta (Orthoptera: Anostostomatidae): Drumming Behavior, Nuptial Feeding, Postcopulatory Guarding and Maternal Care.” Journal of the Kansas Entomological Society 77: 414–428. Hadley, N. F., and S. C. Williams. 1968. “Surface Activities of Some North American Scorpions in Relation to Feeding.” Ecology 49: 726–734. Hanrahan, S. A., and W. H. Kirchner. 1994. “Acoustic Orientation and Communication in Desert Tenebrionid Beetles in Sand Dunes.” Ethology 97: 26–32. Hartline, P. H. 1971. “Physiological Basis for Detection of Sound and Vibration in Snakes.” Journal of Experimental Biology 54: 349–371. Haskell, P. T. 1955. “Vibrations of the Substrate and Stridulation in a Grasshopper.” Nature 175: 639–640. Hawkins, A. D., and A. A. Myrberg, Jr. 1983. “Hearing and Sound Communication Under Water.” In Bioacoustics: A Comparative Approach, ed. B. Lewis, pp. 347–405. London: Academic Press. Heady, S. E., and R. F. Denno. 1991. “Reproductive Isolation in Prokelisia Planthoppers (Homoptera: Delphacidae): Acoustic Differentiation and Hybridization Failure.” Journal of Insect Behaviour 4: 367–390. Henaut, A., and J. Guerdoux. 1982. “Location of a Lure by the Drumming Insect Pimpla instigator (Hymenoptera, Ichneumonidae).” Experientia 38: 346–347. Henry, C. S. 1979. “Acoustical Communication During Courtship and Mating in the Green Lacewing Chrysopa carnea (Neuroptera: Chrysopidae).” Annals of the Entomological Society of America 72: 68–79. ———. 1980. “The Importance of Low-frequency, Substrate-borne Sounds in Lacewing Communication (Neuroptera: Chrysopidae).” Annals of the Entomological Society of America 73: 617–621. ———. 1985. “The Proliferation of Cryptic Species in Chrysoperla Green Lacewings Through Song Divergence.” Florida Entomologist 68: 18–38.

References

223

———. 1994. “Singing and Cryptic Speciation in Insects.” Trends in Ecology and Evolution 9: 388–392. Henry, C. S., S. J. Brooks, P. Duelli, and J. B. Johnson. 2003. “A Lacewing with the Wanderlust: The European Song Species ‘Maltese,’ Chrysoperla agilis, sp. n., of the carnea group of Chrysoperla (Neuroptera: Chrysopidae).” Systematic Entomology 28: 131–147. Henry, C. S., and M. L. M. Wells. 2004. “Adaptation or Random Change? The Evolutionary Response of Songs to Substrate Properties in Lacewings (Neuroptera: Chrysopidae: Chrysoperla).” Animal Behaviour 68: 879–895. Henschel, J. R. 2002. “Long-distance Wandering and Mating by the Dancing White Lady Spider (Leucorchestris arenicola) (Araneae, Sparassidae) Across Namib Dunes.” Journal of Arachnology 30: 321–330. Hergenröder, R., and F. G. Barth. 1983a. “The Release of Attack and Escape Behavior by Vibratory Stimuli in a Wandering Spider (Cupiennius salei Keys).” Journal of Comparative Physiology A 152: 347–358. ———. 1983b. “Vibratory Signals and Spider Behavior: How do the Sensory Inputs from the Eight Legs Interact in Orientation?” Journal of Comparative Physiology A 152: 361–371. Heth, G., E. Frankenberg, H. Pratt, and E. Nevo. 1991. “Seismic Communication in the Blind Subterranean Mole-rat: Patterns of Head Thumping and of Their Detection in the Spalax ehrenbergi Superspecies in Israel.” Journal of Zoology (London) 224: 633–638. Heth, G., E. Frankenberg, A. Raz, and E. Nevo. 1987. “Vibrational Communication in Subterranean Mole Rats (Spalax ehrenbergi).” Behavioral Ecology and Sociobiology 21: 31–33. Hetherington, T. E. 1985. “Role of the Opercularis Muscle in Seismic Sensitivity in the Bullfrog Rana catesbeiana.” Journal of Experimental Zoology 235: 27–34. ———. 1988. “Biomechanics of Vibration Reception in the Bullfrog, Rana catesbeiana.” Journal of Comparative Physiology A 163: 43–52. ———. 1989. “Use of Vibratory Cues for Detection of Insect Prey by the Sandswimming Lizard Scincus scincus.” Animal Behaviour 37: 290–297. Hildebrand, M., and G. E. Goslow, Jr. 2001. Analysis of Vertebrate Structure, 5th ed. New York: John Wiley. Hill, P. S. M. 1998. “Environmental and Social Influences on Calling Effort in the Prairie Mole Cricket (Gryllotalpa major).” Behavioral Ecology 9: 101–108. ———. 1999. “Lekking in Gryllotalpa major, the Prairie Mole Cricket (Insecta: Gryllotalpidae).” Ethology 105: 531–545. ———. 2001a. “Vibration as a Communication Channel: A Review.” American Zoologist 41: 1135–1142.

224

References

———. 2001b. “Vibration as a Communication Channel: A Synopsis.” American Zoologist 41: 1133–1134. Hill, P. S. M., and J. R. Shadley. 1997. “Substrate Vibration as a Component of a Calling Song.” Naturwissenschaften 84: 460–463. ———. 2001. “Talking Back: Sending Soil Vibration Signals to Lekking Prairie Mole Cricket Males.” American Zoologist 41: 1200–1214. Hirschberger, P. 2001. “Stridulation in Aphodius Dung Beetles: Behavioral Context and Intraspecific Variability of Song Patterns in Aphodius ater (Scarabaeidae).” Journal of Insect Behavior 14: 69–88. Hoch, H., J. Deckert, and A. Wessel. 2006. “Vibrational Signalling in a Gondwanan Relict Insect (Hemiptera: Coleorrhyncha: Peloridiidae).” Biology Letters 2: 222–224. Hoch, H., and A. Wessel. 2006. “Communication by Substrate-bone Vibrations in Cave Planthoppers.” In Insect Sounds and Communication: Physiology, Behaviour, Ecology and Evolution, eds. S. Drosopoulos and M. F. Claridge, pp. 187–197. Boca Raton, FL: Taylor & Francis. Hölldobler, B. 1999. “Multimodal Signals in Ant Communication.” Journal of Comparative Physiology A 184: 129–141. Hölldobler, B., U. Braun, W. Gronenberg, W. H. Kirchner, and C. Peeters. 1994. “Trail Communication in the Ant Megaponera foetens (Fabr.) (Formicidae, Ponerinae).” Journal of Insect Physiology 40: 585–593. Howse, P. E. 1962. “The Perception of Vibration by the Subgenual Organ in Zootermopsis angusticollis Emerson and Periplaneta americana L.” Experientia 18: 457–458. ———. 1964a. “An Investigation into the Mode of Action of the Subgenual Organ in the Termite, Zootermopsis angusticollis Emerson, and in the Cockroach, Periplaneta americana L.” Journal of Insect Physiology 10: 409–424. ———. 1964b. “The Significance of Sound Produced by the Termite Zootermopsis angusticollis Hagen.” Animal Behaviour 12: 284–300. ———. 1965. “On the Significance of Certain Oscillatory Movements of Termites.” Insectes Sociaux 12: 335–346. Howse, P. E., and M. F. Claridge. 1970. “The Fine Structure of Johnston’s Organ of the Leaf-hopper, Oncopsis flavicollis.” Journal of Insect Physiology 16: 1665–1675. Hoy, R. R., A. Hoikkala, and K. Kaneshiro. 1988. “Hawaiian Courtship Songs: Evolutionary Innovation in Communication Signals of Drosophila.” Science 240: 217–219. Hrncir, M., V. M. Schmidt, D. L. P. Shorkopf, S. Jarau, R. Zucchi, and F. G. Barth. 2006. “Vibrating the Food Receivers: A Direct Way of Signal Transmission in Stingless Bees (Melipona seminigra).” Journal of Comparative Physiology A 192: 879–887.

References

225

Hunt, C. C. 1961. “On the Nature of Vibration Receptors in the Hind Limb of the Cat.” Journal of Physiology (London) 155: 175–186. Hunt, R. E. 1993. “Role of Vibrational Signals in Mating Behavior of Spissistilus festinus (Homoptera: Membracidae).” Annals of the Entomological Society of America 86: 356–361. ———. 1994. “Vibrational Signals Associated with Mating Behavior in the Treehopper, Enchenopa binotata Say (Hemiptera: Homoptera: Membracidae).” Journal of the New York Entomological Society 102: 266–270. Hunt, R. E., J. P. Fox, and K. F. Haynes. 1992. “Behavioral Response of Graminella nigrifrons (Homoptera: Cicadellidae) to Experimentally Manipulated Vibrational Signals.” Journal of Insect Behavior 5: 1–13. Hunt, R. E., and T. L. Morton. 2001. “Regulation of Chorusing in the Vibrational Communication System of the Leafhopper Graminella nigrifrons.” American Zoologist 41: 1222–1228. Hunt, R. E., and L. R. Nault. 1991. “Roles of Interplant Movement, Acoustic Communication, and Phototaxis in Mate-location Behavior of the Leafhopper Graminella nigrifrons.” Behavioral Ecology and Sociobiology 28: 315–320. Hutchings, M., and B. Lewis. 1983. “Insect Sound and Vibration Receptors.” In Bioacoustics: a Comparative Approach, ed. B. Lewis, pp. 181–205. London: Academic Press. Ichikawa, T. 1976. “Mutual Communication by Substrate Vibrations in the Mating Behavior of Planthoppers (Homoptera: Delphacidae).” Applied Entomology and Zoology 11: 8–21. ———. 1982. “Density-related Changes in Male-male Competitive Behavior in the Rice Brown Planthopper, Nilaparvata lugens (Stål) (Homoptera: Delphacidae).” Applied Entomology and Zoology 17: 439–452. Ichikawa, T., and S. Ishii. 1974. “Mating Signal of the Brown Planthopper, Nilaparvata lugens Stål (Homoptera: Delphacidae): Vibration of the Substrate.” Applied Entomology and Zoology 9: 196–198. Ichikawa, T., M. Sakuma, and S. Ishii. 1975. “Substrate Vibrations: Mating Signal of Three Species of Planthoppers Which Attack the Rice Plant (Homoptera: Delphacidae).” Applied Entomology and Zoology 10: 162–171. Ishay, J., and E. M. Landau. 1972. “Vespa Larvae Send Out Rhythmic Hunger Signals.” Nature 237: 286–287. Ishay, J., A. Motro, S. Gitter, and M. B. Brown. 1974. “Rhythms in Acoustical Communication by the Oriental Hornet, Vespa orientalis.” Animal Behaviour 22: 741–744. Ishay, J., and A. Schwartz. 1973. “Acoustical Communication Between the Members of the Oriental Hornet (Vespa orientalis) Colony.” Journal of the Acoustical Society of America 53: 640–649.

226

References

Jaslow, A. P., T. E. Hetherington, and R. E. Lombard. 1988. “Structure and Function of the Amphibian Middle Ear.” In The Evolution of the Amphibian Auditory System, eds. B. Fritzsch, M. J. Ryan, W. Wilczynski, T. E. Hetherington, and W. Walkowiak, pp. 69–91. New York: John Wiley & Sons. ˇ okl. 1996. “Mechanoreceptors in Insects: Johnston’s Jeram, S., and A. C Organ in Nezara viridula (L.) (Pentatomidae, Heteroptera).” European Journal of Physiology Suppl. 431: R281–R282. Jeram, S., and M. A. Pabst. 1996. “Johnston’s Organ and Central Organ in Nezara viridula (L.) (Heteroptera, Pentatomidae).” Tissue & Cell 28: 227–235. Johansson, R. S., and Å. B. Vallbo. 1983. “Tactile Sensory Coding in the Glabrous Skin of the Human Hand.” Trends in Neurosciences 6: 27–32. Johnson, J. B., D. Saenz, C. K. Adams, and R. N. Conner. 2003. “The Influence of Predator Threat on the Timing of a Life-history Switch Point: Predator-induced Hatching in the Southern Leopard Frog (Rana sphenocephala).” Canadian Journal of Zoology 81: 1608–1613. Johnson, V., and W. P. Morrison. 1979. “Mating Behavior of Three Species of Coniopterygidae (Neuroptera).” Psyche 86: 395–398. Jones, M. D. R., and M. Dambach. 1973. “Response to Sound in Crickets Without Tympanal Organs (Gryllus campestris L.).” Journal of Comparative Physiology 87: 89–98. Jørgensen, M. B., and J. Christensen-Dalsgaard. 1991. “Peripheral Origins and Functional Characteristics of Vibration-Sensitive VIIIth Nerve Fibers in the Frog Rana temporaria.” Journal of Comparative Physiology A 169: 341–347. Kalmring, K. 1985. “Vibrational Communication in Insects (Reception and Integration of Vibratory Information).” In Acoustic and Vibrational Communication in Insects, eds. K. Kalmring and N. Elsner, pp. 127–134. Berlin: Paul Parey. Kalmring, K., E. Hoffmann, M. Jatho, T. Sickmann, and M. Grossbach. 1996. “The Auditory-vibratory Sensory System of the Bushcricket Polysarcus denticauda (Phaneropterinae, Tettigoniidae). II. Physiology of Receptor Cells. Journal of Experimental Zoology 276: 315–329. Kalmring, K., M. Jatho, W. Rössler, and T. Sickmann. 1997. “Acoustovibratory Communication in Bushcrickets (Orthoptera: Tettigoniidae).” Entomologia Generalis 21: 265–291. Kalmring, K., and R. Kühne. 1983. “The Processing of Acoustic and Vibrational Information in Insects.” In Bioacoustics: A Comparative Approach, ed. B. Lewis, pp. 261–282. London: Academic Press.

References

227

Kämper, G., and M. Dambach. 1979. “Communication by Infrasound in a Non-stridulating Cricket.” Naturwissenschaften 66: 530. Kanmiya, K. 1990. “Acoustic Properties and Geographic Variation in the Vibratory Courtship Signals of the European Chloropid Fly, Lipara lucens Meigen (Diptera, Chloropidae).” Journal of Ethology 8: 105–119. ———. 2006a. “Communication by Vibratory Signals in Diptera.” In Insect Sounds and Communication: Physiology, Behaviour, Ecology and Evolution, eds. S. Drosopoulos and M. F. Claridge, pp. 381–396. Boca Raton, FL: Taylor & Francis. ———. 2006b. “Mating Behaviour and Vibratory Signals in Whiteflies (Hemiptera: Aleyroididae).” In Insect Sounds and Communication: Physiology, Behaviour, Ecology and Evolution, eds. S. Drosopoulos and M. F. Claridge, pp. 365–379. Boca Raton, FL: Taylor & Francis. Kasper, J., and P. Hirschberger. 2005. “Stridulation in Aphodius Dung Beetles: Songs and Morphology of Stridulatory Organs in North American Aphodius Species (Scarabaeidae).” Journal of Natural History 39: 91–99. Kelly, C. D. 2006. “Resource Quality or Harem Size: What Influences Male Tenure at Refuge Sites in Tree Weta (Orthoptera: Anostostomatidae)?” Behavioral Ecology and Sociobiology 60: 175–183. Kenneally, C. 2004. “Surviving the Tsunami: What Sri Lanka’s Animals Knew That Humans Didn’t.” Slate, December 30, 2004. www.slate.com (accessed March 7, 2007). ———. 2005. “Do They Know Something We Don’t? Animals’ Senses May Have Helped Them Survive the Tsunami.” The Boston Globe, January 11, C1, C4. Keuper, A., and R. Kühne. 1983. “The Acoustic Behaviour of the Bushcricket Tettigonia cantans. II. Transmission of Airborne-sound and Vibration Signals in the Biotope.” Behavioural Processes 8: 125–145. Keuper, A., C. Otto, W. Latimer, and A. Schatral. 1985. “Airborne Sound and Vibration Signals of Bushcrickets and Locusts: Their Importance for the Behaviour in the Biotope.” In Acoustic and Vibrational Communication in Insects, eds. K. Kalmring and N. Elsner, pp. 135–142. Berlin: Paul Parey. Kilpinen, O., and J. Storm. 1997. “Biophysics of the Subgenual Organ of the Honeybee, Apis mellifera.” Journal of Comparative Physiology A 181: 309–318. Kimchi, T., M. Reshef, and J. Terkel. 2005. “Evidence for the Use of Reflected Self-generated Seismic Waves for Spatial Orientation in a Blind Subterranean Mammal.” Journal of Experimental Biology 208: 647–659.

228

References

King, M. J., and S. L. Buchmann. 1995. “Bumble Bee-initiated Vibration Release Mechanism of Rhododendron Pollen.” American Journal of Botany 82: 1407–1411. Kirchner, W. H. 1993. “Vibrational Signals in the Tremble Dance of the Honeybee, Apis mellifera.” Behavioral Ecology and Sociobiology 33: 169–172. ———. 1997. “Acoustical Communication in Social Insects.” In Orientation and Communication in Arthropods, ed. M. Lehrer, pp. 273–300. Basel: Birkhäuser Verlag. Kirchner, W. H., I. Broecker, and J. Tautz. 1994. “Vibrational Alarm Communication in the Damp-wood Termite Zootermopsis nevadensis.” Physiological Entomology 19: 187–190. Kirchner, W. H., and C. Dreller. 1993. “Acoustical Signals in the Dance Language of the Giant Honeybee, Apis dorsata.” Behavioral Ecology and Sociobiology 33: 67–72. Kirchner, W. H., C. Dreller, A. Grasser, and D. Baidya. 1996. “The Silent Dances of the Himalayan Honeybee, Apis laboriosa.” Apidologie 27: 331–339. Kirchner, W. H., and M. Lindauer. 1994. “The Causes of the Tremble Dance of the Honeybee, Apis mellifera.” Behavioral Ecology and Sociobiology 35: 303–308. Kirschvink, J. L. 2000. “Earthquake Prediction by Animals: Evolution and Sensory Perception.” Bulletin of the Seismological Society of America 90: 312–323. Klaaßen, F. 1973. “Stridulation und Kommunikation Durch Substratschall bei Gecarcinus lateralis (Crustacea Decapoda).” Journal of Comparative Physiology 83: 73–79. Klärner, D., and F. G. Barth. 1982. “Vibratory Signals and Prey Capture in Orb-weaving Spiders (Zygiella x-notata, Nephila clavipes; Araneidae).” Journal of Comparative Physiology A 148: 445–455. Klauer, G., H. Burda, and E. Nevo. 1997. “Adaptive Differentiations of the Skin of the Head in a Subterranean Rodent, Spalax ehrenbergi.” Journal of Morphology 233: 53–66. Kolmes, S. A. 1985. “Surface Vibrational Cues in the Precopulatory Behavior of Whirligig Beetles.” Journal of the New York Entomological Society 93: 1137–1140. Kotiaho, J., R. V. Alatalo, J. Mappes, and S. Parri. 1996. “Sexual Selection in a Wolf Spider: Male Drumming Activity, Body Size, and Viability.” Evolution 50: 1977–1981. Koyama, H., E. R. Lewis, E. L. Leverenz, and R. A. Baird. 1982. “Acute Seismic Sensitivity in the Bullfrog Ear.” Brain Research 250: 168–172. Krafft, B. 1982. “The Significance and Complexity of Communication in Spiders.” In Spider Communication, eds. P. N. Witt and J. S. Rovner, pp. 15–66. Princeton, NJ: Princeton University Press.

References

229

Kristensen, L., and K. E. Zachariassen. 1980. “Behavioural Studies on the Sensitivity to Sound in the Desert Tenebrionid Beetle Phrynocolus somalicus Wilke.” Comparative Biochemistry and Physiology A 65: 223–226. Kuhn, T. S. 1996. The Structure of Scientific Revolutions, 3rd ed. Chicago: Chicago University Press. Landolfa, M. A., and F. G. Barth. 1996. “Vibrations in the Orb Web of the Spider Nephila clavipes: Cues for Discrimination and Orientation.” Journal of Comparative Physiology A 179: 493–508. Lang, H. H. 1980. “Surface Wave Discrimination Between Prey and Nonprey by the Back Swimmer Notonecta glauca L. (Hemiptera, Heteroptera).” Behavioral Ecology and Sociobiology 6: 233–246. Latimer, W., and A. Schatral. 1983. “The Acoustic Behaviour of the Bushcricket Tettigonia cantans. I. Behavioural Responses to Sound and Vibration.” Behavioural Processes 8: 113–124. Lawrence, P. O. 1981. “Host Vibration—A Cue to Host Location by the Parasite, Biosteres longicaudatus.” Oecologia 48: 249–251. Lay, D. M. 1972. “The Anatomy, Physiology, Functional Significance and Evolution of Specialized Hearing Organs of Gerbilline Rodents.” Journal of Morphology 138: 41–120. Levänen, S., and D. Hamdorf. 2001. “Feeling Vibrations: Enhanced Tactile Sensitivity in Congenitally Deaf Humans.” Neuroscience Letters 301: 75–77. Levänen, S., V. Jousmäki, and R. Hari. 1998. “Vibration-induced Auditorycortex Activation in a Congenitally Deaf Adult.” Current Biology 8: 869–872. Lewis, E. R. 1984. “Inertial Motion Sensors.” In Comparative Physiology of Sensory Systems, eds. L. Bolis, R. D. Keynes, and S. H. P. Maddrell, pp. 587–610. Cambridge: Cambridge University Press. ———. 1992. “Convergence of Design in Vertebrate Acoustic Sensors.” In The Evolutionary Biology of Hearing, eds. D. B. Webster, R. R. Fay, and A. N. Popper, pp. 163–184. New York: Springer-Verlag. Lewis, E. R., R. A. Baird, E. L. Leverenz, and H. Koyama. 1982. “Inner Ear: Dye Injection Reveals Peripheral Origins of Specific Sensitivities.” Science 215: 1641–1643. Lewis, E. R., and R. R. Fay. 2004. “Environmental Variables and the Fundamental Nature of Hearing.” In Evolution of the Vertebrate Auditory System, eds. G. A. Manley, A. N. Popper, and R. R. Fay, pp. 27–54. New York: Springer-Verlag. Lewis, E. R., and R. E. Lombard. 1988. “The Amphibian Inner Ear.” In The Evolution of the Amphibian Auditory System, eds. B. Fritzsch, M. J. Ryan, W. Wilczynski, T. E. Hetherington, and W. Walkowiak, pp. 93–123. New York: John Wiley & Sons.

230

References

Lewis, E. R., and P. M. Narins. 1985. “Do Frogs Communicate with Seismic Signals?” Science 227: 187–189. Lewis, E. R., P. M. Narins, K. A. Cortopassi, W. M. Yamada, E. H. Poinar, S. W. Moore, and X.-L. Yu. 2001. “Do Male White-lipped Frogs Use Seismic Signals for Intraspecific Communication?” American Zoologist 41: 1185–1199. Lewis, E. R., P. M. Narins, J. U. M. Jarvis, G. Bronner, and M. J. Mason. 2006. “Preliminary Evidence for the Use of Microseismic Cues for Navigation by the Namib Golden Mole.” Journal of the Acoustical Society of America 119: 1260–1268. Lewis, L. A., and S. S. Schneider. 2000. “The Modulation of Worker Behavior by the Vibration Signal During House Hunting in Swarms of the Honeybee, Apis mellifera.” Behavioral Ecology and Sociobiology 48: 154–164. Lewis, L. A., S. S. Schneider, and G. Degrandi-Hoffman. 2002. “Factors Influencing the Selection of Recipients by Workers Performing Vibration Signals in Colonies of the Honeybee, Apis mellifera.” Animal Behaviour 63: 361–367. Li, D. 2002. “Hatching Responses of Subsocial Spitting Spiders to Predation Risk.” Proceedings of the Royal Society of London B 269: 2155–2161. Lighton, J. R. B. 1987. “Cost of Tokking: The Energetics of Substrate Communication in the Tok-tok Beetle, Psammodes striatus.” Journal of Comparative Physiology B 157: 11–20. Loewenstein, W. R., and M. Mendelson. 1965. “Components of Receptor Adaptation in a Pacinian Corpuscle.” Journal of Physiology (London) 177: 377–397. Loher, W., and M. Dambach. 1989. “Reproductive Behavior.” In Cricket Behavior and Neurobiology, eds. F. Huber, J. E. Moore, and W. Loher, pp. 43–82. Ithaca, NY: Cornell University Press. Magal, C., M. Schöller, J. Tautz, and J. Casas. 2000. “The Role of Leaf Structure in Vibration Propagation.” Journal of the Acoustical Society of America 108: 2412–2418. Maketon, M., and K. W. Stewart. 1984a. “Drumming Behavior in Four North American Perlodidae (Plecoptera) Species.” Annals of the Entomological Society of America 77: 621–626. ———. 1984b. “Further Studies of the Drumming Behavior of North American Perlidae (Plecoptera).” Annals of the Entomological Society of America 77: 770–778. ———. 1988. “Patterns and Evolution of Drumming Behavior in the Stonefly Families Perlidae and Peltoperlidae.” Aquatic Insects 10: 77–98. Maketon, M., K. W. Stewart, B. C. Kondratieff, and R. F. Kirchner. 1988. “New Descriptions of Drumming and Evolution of the Behavior in North

References

231

American Perlodidae (Plecoptera).” Journal of the Kansas Entomological Society 61: 161–168. Maklakov, A. A., T. Bilde, and Y. Lubin. 2003. “Vibratory Courtship in a Webbuilding Spider: Signalling Quality or Stimulating the Female?” Animal Behaviour 66: 623–630. Manrique, G., and P. E. Schilman. 2000. “Two Different Vibratory Signals in Rhodnius prolixus (Hemiptera: Reduviidae).” Acta Tropica 77: 271–278. Mappes, J., R. Alatalo, J. Kotiaho, and S. Parri. 1996. “Viability Costs of Condition-dependent Sexual Male Display in a Drumming Wolf Spider.” Proceedings of the Royal Society of London B 263: 785–789. Markl, H. 1967. “Die Verständigung Durch Stridulationssignale bei Blattschneiderameisen. I. Die Biologische Bedeutung der Stridulation.” Zeitschrift für vergleichende Physiologie 57: 299–330. ———. 1968. “Die Verständigung Durch Stridulationssignale bei Blattschneiderameisen. II. Erzeugung und Eigenschaften der Signale.” Zeitschrift für vergleichende Physiologie 60: 103–150. ———. 1969. “Verständigung Durch Vibrationssignale bei Arthropoden.” Naturwissenschaften 56: 499–505. ———. 1970. “Die Verständigung Durch Stridulationssignale bei Blattschneiderameisen. III. Die Empfindlichkeit für Substratvibrationen.” Zeitschrift für vergleichende Physiologie 69: 6–37. ———. 1983. “Vibrational Communication.” In Neuroethology and Behavioral Physiology, eds. F. Huber and H. Markl, pp. 332–353. Berlin: SpringerVerlag. Markl, H., and B. Hölldobler. 1978. “Recruitment and Food-retrieving Behavior in Novomessor (Formicidae, Hymenoptera). II. Vibration Signals.” Behavioral Ecology and Sociobiology 4: 183–216. Markl, H., B. Hölldobler, and T. Hölldobler. 1977. “Mating Behavior and Sound Production in Harvester Ants (Pogonomyrmex, Formicidae).” Insectes Sociaux 24: 191–212. Maschwitz, U., and P. Schönegge. 1983. “Forage Communication, Nest Moving Recruitment, and Prey Specialization in the Oriental Ponerine Leptogenys chinensis.” Oecologia 57: 175–182. Mason, M. J. 2001. “Middle Ear Structures in Fossorial Mammals: A Comparison with Non-fossorial species.” Journal of Zoology 255: 467–486. ———. 2003. “Bone Conduction and Seismic Sensitivity in Golden Moles (Chrysochloridae).” Journal of Zoology (London) 260: 405–413. Mason, M. J., and P. M. Narins. 2001. “Seismic Signal Use by Fossorial Mammals.” American Zoologist 41: 1171–1184. ———. 2002. “Seismic Sensitivity in the Desert Golden Mole (Eremitalpa granti): A Review.” Journal of Comparative Psychology 116: 158–163.

232

References

Masters, W. M. 1980. “Insect Disturbance Stridulation: Characterization of Airborne and Vibrational Components of the Sound.” Journal of Comparative Physiology 135: 259–268. ———. 1984. “Vibrations in the Orbwebs of Nuctenea sclopetaria (Araneidae). I. Transmission Through the Web.” Behavioral Ecology and Sociobiology 15: 207–215. Masters, W. M., and H. Markl. 1981. “Vibration Signal Transmission in Spider Orb Webs.” Science 213: 363–365. Masters, W. M., J. Tautz, N. H. Fletcher, and H. Markl. 1983. “Body Vibration and Sound Production in an Insect (Atta sexdens) Without Specialized Radiating Structures.” Journal of Comparative Physiology A 150: 239–249. McAdoo, B. G., L. Dengler, M. Eeri, G. Prasetya, and V. Titov. 2006. “Smong: How an Oral History Saved Thousands on Indonesia’s Simeulue Island During the December 2004 and March 2005 Tsunamis.” Earthquake Spectra 22: S661–S669. McBrien, H. L., A. Çokl, and J. G. Millar. 2002. “Comparison of Substrateborne Vibrational Signals of Two Congeneric Stink Bug Species, Thyanta pallidovirens and T. custator accerra (Heteroptera: Pentatomidae).” Journal of Insect Behavior 15: 715–738. McBrien, H. L., and J. G. Millar. 2003. “Substrate-borne Vibrational Signals of the Consperse Stink Bug (Hemiptera: Pentatomidae).” Canadian Entomologist 135: 555–567. McCormick, C. A. 1988. “Evolution of Auditory Pathways in the Amphibia.” In The Evolution of the Amphibian Auditory System, eds. B. Fritzsch, M. J. Ryan, W. Wilczynski, T. E. Hetherington, and W. Walkowiak, pp. 587–612. New York: John Wiley & Sons. McIntyre, A. K. 1962. “Cortical Projection of Impulses in the Interosseous Nerve of the Cat’s Hind Limb.” Journal of Physiology (London) 163: 46–60. ———. 1980. “Biological Seismography.” Trends in Neuroscience 3: 202–205. McNett, G. D., R. N. Miles, D. Homentcovschi, and R. B. Cocroft. 2006. “A Method for Two-dimensional Characterization of Animal Vibrational Signals Transmitted Along Plant Stems.” Journal of Comparative Physiology A 192: 1245–1251. McVean, A., and L. H. Field. 1996. “Communication by Substratum Vibration in the New Zealand Tree Weta, Hemideina femorata (Stenopelmatidae: Orthoptera).” Journal of Zoology (London) 239: 101–122. Menzel, J. G., and J. Tautz. 1994. “Functional Morphology of the Subgenual Organ of the Carpenter Ant.” Tissue and Cell 26: 735–746. Meyhöfer, R., and J. Casas. 1999. “Vibratory Stimuli in Host Location by Parasitic Wasps.” Journal of Insect Physiology 45: 967–971.

References

233

Meyhöfer, R., J. Casas, and S. Dorn. 1994. “Host Location by a Parasitoid Using Leafminer Vibrations: Characterizing the Vibrational Signals Produced by the Leafmining Host.” Physiological Entomology 19: 349–359. ———. 1997. “Vibration-mediated Interactions in a Host-parasitoid System.” Proceedings of the Royal Society of London B 264: 261–266. Michelsen, A., F. Fink, M. Gogala, and D. Traue. 1982. “Plants as Transmission Channels for Insect Vibrational Songs.” Behavioral Ecology and Sociobiology 11: 269–281. Michelsen, A., W. H. Kirchner, B. B. Andersen, and M. Lindauer. 1986. “The Tooting and Quacking Vibration Signals of Honeybee Queens: A Quantitative Analysis.” Journal of Comparative Physiology A 158: 605–611. Michelsen, A., W. H. Kirchner, and M. Lindauer. 1986. “Sound and Vibrational Signals in the Dance Language of the Honeybee, Apis mellifera.” Behavioral Ecology and Sociobiology 18: 207–212. Miklas, N., A. Cˇokl, M. Renou, and M. Virant-Doberlet. 2003. “Variability of Vibratory Signals and Mate Choice Selectivity in the Southern Green Stink Bug.” Behavioural Processes 61: 131–142. Miklas, N., T. Lasnier, and M. Renou. 2003. “Male Bugs Modulate Pheromone Emission in Response to Vibratory Signals from Conspecifics.” Journal of Chemical Ecology 29: 561–574. ˇ okl, M. Virant-Doberlet, and M. Renou. 2001. “The Miklas, N., N. Stritih, A. C Influence of Substrate on Male Responsiveness to the Female Calling Song in Nezara viridula.” Journal of Insect Behavior 14: 313–332. Miles, R. N., R. B. Cocroft, C. Gibbons, and D. Batt. 2001. “A Bending Wave Simulator for Investigating Directional Vibration Sensing in Insects.” Journal of the Acoustical Society of America 110: 579–587. Milum, V. G. 1955. “Honey Bee Communication.” American Bee Journal 95: 97–104. Miranda, X. 2006. “Substrate-borne Signal Repertoire and Courtship Jamming by Adults of Ennya chrysura (Hemiptera: Membracidae).” Annals of the Entomological Society of America 99: 374–386. Mitomi, M., T. Ichikawa, and H. Okamoto. 1984. “Morphology of the Vibration-producing Organ in Adult Rice Brown Planthopper, Nilaparvata lugens (Stål) (Homoptera: Delphacidae).” Applied Entomology and Zoology 19: 407–417. Montealegre-Z, F., G. K. Morris, and A. C. Mason. 2006. “Generation of Extreme Ultrasonics in Rainforest Katydids.” Journal of Experimental Biology 209: 4923–4937. Morris, G. K. 1980. “Calling Display and Mating Behaviour of Copiphora rhinoceros Pictet (Orthoptera: Tettigoniidae).” Animal Behaviour 28: 42–51.

234

References

Morris, G. K., and M. Beier. 1982. “Song Structure and Description of Some Costa Rican Katydids (Orthoptera: Tettigoniidae).” Transactions of the American Entomological Society 108: 287–314. Morris, G. K., D. E. Klimas, and D. A. Nickle. 1988. “Acoustic Signals and Systematics of False-leaf Katydids from Ecuador (Orthoptera, Tettigoniidae, Pseudophyllinae).” Transactions of the American Entomological Society 114: 215–263. Morris, G. K., A. C. Mason, P. Wall, and J. J. Belwood. 1994. “High Ultrasonic and Tremulation Signals in Neotropical Katydids (Orthoptera: Tettigoniidae).” Journal of Zoology (London) 233: 129–163. Narhardiyati, M., and W. J. Bailey. 2005. “Biology and Natural Enemies of the Leafhopper Balclutha incisa (Matsumura) (Hemiptera: Cicadellidae: Deltocephalinae) in South-western Australia.” Australian Journal of Entomology 44: 104–109. Narins, P. M., G. Ehret, and J. Tautz. 1988. “Accessory Pathway for Sound Transfer in a Neotropical Frog.” Proceedings of the National Academy of Science of the United States of America 85: 1508–1512. Narins, P. M., and E. R. Lewis. 1984. “The Vertebrate Ear as an Exquisite Seismic Sensor.” Journal of the Acoustical Society of America 76: 1384–1387. Narins, P. M., E. R. Lewis, J. U. M. Jarvis, and J. O’Riain. 1997. “The Use of Seismic Signals by Fossorial Southern African Mammals: A Neuroethological Gold Mine.” Brain Research Bulletin 44: 641–646. Narins, P. M., O. J. Reichman, J. U. M. Jarvis, and E. R. Lewis. 1992. “Seismic Signal Transmission Between Burrows of the Cape Mole-rat, Georychus capensis.” Journal of Comparative Physiology A 170: 13–21. Nevo, E. 1990. “Evolution of Nonvisual Communication and Photoperiodic Perception in Speciation and Adaptation of Blind Subterranean Mole Rats.” Behaviour 114: 249–276. Nevo, E., G. Heth, and H. Pratt. 1991. “Seismic Communication in a Blind Subterranean Mammal: A Major Somatosensory Mechanism in Adaptive Evolution Underground.” Proceedings of the National Academy of Science of the United States of America 88: 1256–1260. Newbury, T. K. 1972. “Vibration Perception by Chaetognaths.” Nature 236: 459–460. Nieh, J. C. 1993. “The Stop Signal of Honey Bees: Reconsidering Its Message.” Behavioral Ecology and Sociobiology 33: 51–56. ———. 1998. “The Honey Bee Shaking Signal: Function and Design of a Modulatory Communication Signal.” Behavioral Ecology and Sociobiology 42: 23–36.

References

235

Nieh, J. C., and J. Tautz. 2000. “Behaviour-locked Signal Analysis Reveals Weak 200–300 Hz Comb Vibrations During the Honeybee Waggle Dance.” Journal of Experimental Biology 203: 1573–1579. Northcutt, R. G., and C. Gans. 1983. “The Genesis of Neural Crest and Epidermal Placodes: A Reinterpretation of Vertebrate Origins.” Quarterly Review of Biology 58: 1–28. Nowak, R. 2006. “How a Lullaby Can Warn of an Approaching Tsunami.” New Scientist 191: 14. Nummelin, M. 1987. “Ripple Signals of the Waterstrider Limnoporus rufoscutellatus (Heteroptera, Gerridae).” Annales Entomologici Fennici 53: 17–22. O’Connell, C. E., B. T. Arnason, and L. A. Hart. 1997. “Seismic Transmission of Elephant Vocalizations and Movement.” Journal of the Acoustical Society of America 102: 3124. O’Connell-Rodwell, C. E., B. T. Arnason, and L. A. Hart. 2000. “Seismic Properties of Asian Elephant (Elephas maximus) Vocalizations and Locomotion.” Journal of the Acoustical Society of America 108: 3066–3072. O’Connell-Rodwell, C. E., L. A. Hart, and B. T. Arnason. 2001. “Exploring the Potential Use of Seismic Waves as a Communication Channel by Elephants and Other Large Mammals.” American Zoologist 41: 1157–1170. O’Connell-Rodwell, C. E., J. D. Wood, T. C. Rodwell, S. Puria, S. R. Partan, R. Keefe, D. Shriver, B. T. Arnason, and L. A. Hart. 2006. “Wild Elephant (Loxodonta africana) Breeding Herds Respond to Artificially Transmitted Seismic Stimuli.” Behavioral Ecology and Sociobiology 59: 842–850. O’Connell-Rodwell, C. E., J. D. Wood, C. Kinzley, T. C. Rodwell, J. H. Poole, and S. Puria. 2007. “Wild African Elephants (Loxodonta africana) Discriminate Between Familiar and Unfamiliar Conspecific Seismic Alarm Calls.” Journal of the Acoustical Society of America 122: 823–830. Ohya, E., and H. Kinuura. 2001. “Close Range Sound Communications of the Oak Platypodid Beetle Platypus quercivorus (Murayama) (Coleoptera: Platypodidae).” Applied Entomology and Zoology 36: 317–321. Ossiannilsson, F. 1949. “Insect Drummers. A Study on the Morphology and Function of the Sound-producing Organ of Swedish Homoptera Auchenorrhyncha with Notes on Their Sound Production.” Opuscula Entomologica Suppl. 10: 1–146. Ota, D., and A. Cˇokl. 1991. “Mate Location in the Southern Green Stink Bug, Nezara viridula (Heteroptera: Pentatomidae), Mediated Through Substrate-borne Signals on Ivy.” Journal of Insect Behavior 4: 441–447. Otten, H., F. Wäckers, M. Battini, and S. Dorn. 2001. “Efficiency of Vibrational Sounding in the Parasitoid Pimpla turionellae is Affected by Female Size.” Animal Behaviour 61: 671–677.

236

References

Parker, S. 2004. “Indigenous Tribes on India’s Andaman Islands Survive Tsunami Disaster.” VOA News, December 31, 2004. www.voanews.com (accessed March 7, 2007). Parmentier, E., P. Vandewalle, and J. P. Lagardère. 2003. “Sound-producing Mechanisms and Recordings in Carapini Species (Teleostei, Pisces).” Journal of Comparative Physiology A 189: 283–292. Parri, S., R. V. Alatalo, J. Kotiaho, and J. Mappes. 1997. “Female Choice for Male Drumming in the Wolf Spider Hygrolycosa rubrofasciata.” Animal Behaviour 53: 305–312. Parri, S., R. V. Alatalo, J. S. Kotiaho, J. Mappes, and A. Rivero. 2002. “Sexual Selection in the Wolf Spider Hygrolycosa rubrofasciata: Female Preference for Drum Duration and Pulse Rate.” Behavioral Ecology 13: 615–621. ˇ okl. 2001. “Songs of Holcostethus strictus (Fabricius): Pavlovcˇicˇ, P., and A. C A Different Repertoire Among Landbugs (Heteroptera: Pentatomidae).” Behavioural Processes 53: 65–73. Pearman, J. V. 1928. “On Sound Production in the Psocoptera and on a Presumed Stridulatory Organ.” Entomological Monograph Magazine 64 (3rd ser., v. 14): 179–186. Pearson, G. A., and D. M. Allen. 1996. “Vibrational Communication in Eusattus convexus LeConte (Coleoptera: Tenebrionidae).” Coleopterists Bulletin 50: 391–394. Penna, M., and A. Veloso. 1987. “Vocalizations by Andean Frogs of the Genus Telmatobius (Leptodactylidae).” Herpetologica 43: 208–216. Percy, D. M., G. S. Taylor, and M. Kennedy. 2006. “Psyllid Communication: Acoustic Diversity, Mate Recognition and Phylogenetic Signal.” Invertebrate Systematics 20: 431–445. Pfannenstiel, R. S., R. E. Hunt, and K. V. Yeargan. 1995. “Orientation of a Hemipteran Predator to Vibrations Produced by Feeding Caterpillars.” Journal of Insect Behavior 8: 1–9. Popper, A. N., M. Salmon, and K. W. Horch. 2001. “Acoustic Detection and Communication by Decapod Crustaceans.” Journal of Comparative Physiology A 187: 83–39. Pratte, M., and R. L. Jeanne. 1984. “Antennal Drumming Behavior in Polistes Wasps (Hymenoptera: Vespidae).” Zeitschrift für Tierpsychologie 66: 177–188. Proctor, H. C. 1991. “Courtship in the Water Mite Neumania papillator: Males Capitalize on Female Adaptations for Predation.” Animal Behaviour 42: 589–598. Proske, U. 1969a. “Nerve Endings in Skin of the Australian Black Snake.” Anatomical Record 164: 259–266. ———. 1969b. “Vibration-sensitive Mechanoreceptors in Snake Skin.” Experimental Neurology 23: 187–194.

References

237

Quirici, V., and F. G. Costa. 2005. “Seismic Communication During Courtship in Two Burrowing Tarantula Spiders: An Experimental Study on Eupalaestrus weijenberghi and Acanthoscurria suina.” Journal of Arachnology 33: 159–166. Rado, R., M. Himelfarb, B. Arensburg, J. Terkel, and Z. Wollberg. 1989. “Are Seismic Communication Signals Transmitted by Bone Conduction in the Blind Mole Rat?” Hearing Research 41: 23–30. Rado, R., J. Terkel, and Z. Wollberg. 1998. “Seismic Communication Signals in the Blind Mole-rat (Spalax ehrenbergi): Electrophysiological and Behavioral Evidence for Their Processing by the Auditory System.” Journal of Comparative Physiology A 183: 503–511. Rado, R. N., N. Levi, H. Hauser, J. Witcher, N. Adler, N. Intrator, A. Wollberg, and J. Terkell. 1987. “Seismic Signalling as a Means of Communication in a Subterranean Mammal.” Animal Behaviour 35: 1249–1251. Randall, J. A. 1989. “Individual Footdrumming Signatures in Banner-tailed Kangaroo Rats Dipodomys spectabilis.” Animal Behaviour 38: 620–630. ———. 1993. “Behavioural Adaptations of Desert Rodents (Heteromyidae).” Animal Behaviour 45: 263–287. ———. 1994a. “Convergences and Divergences in Communication and Social Organization of Desert Rodents.” Australian Journal of Zoology 42: 405–433. ———. 1994b. “Discrimination of Footdrumming Signatures by Kangaroo Rats, Dipodomys spectabilis.” Animal Behaviour 47: 45–54. ———. 1997. “Species-specific Footdrumming in Kangaroo Rats, Dipodomys ingens, D. deserti, D. spectabilis.” Animal Behaviour 54: 1167–1175. ———. 2001. “Evolution and Function of Drumming as Communication in Mammals.” American Zoologist 41: 1143–1156. Randall, J. A., and D. K. Boltas King. 2001. “Assessment and Defence of Solitary Kangaroo Rats under Risk of Predation by Snakes.” Animal Behaviour 61: 579–587. Randall, J. A., and E. R. Lewis. 1997. “Seismic Communication Between the Burrows of Kangaroo Rats, Dipodomys spectabilis.” Journal of Comparative Physiology 181: 525–531. Randall, J. A., and M. D. Matocq. 1997. “Why Do Kangaroo Rats (Dipodomys spectabilis) Footdrum at Snakes?” Behavioral Ecology 8: 404–413. Randall, J. A., K. A. Rogovin, and D. M. Shier. 2000. “Antipredator Behavior of a Social Desert Rodent: Footdrumming and Alarm Calling in the Great Gerbil, Rhombomys opimus.” Behavioral Ecology and Sociobiology 48: 110–118. Reuter, T., S. Nummela, and S. Hemilä. 1998. “Elephant Hearing.” Journal of the Acoustical Society of America 104: 1122–1123. Richardson, P. W. 1982. “Foot-slapping by Coots.” British Birds 75: 126–127.

238

References

Rivero, A., R. V. Alatalo, J. S. Kotiaho, J. Mappes, and S. Parri. 2000. “Acoustic Signalling in a Wolf Spider: Can Signal Characteristics Predict Male Quality?” Animal Behaviour 60: 187–194. Roces, F., and B. Hölldobler. 1995. “Vibrational Communication Between Hitchhikers and Foragers in Leaf-cutting Ants (Atta cephalotes).” Behavioral Ecology and Sociobiology 37: 297–302. ———. 1996. “Use of Stridulation in Foraging Leaf-cutting Ants: Mechanical Support During Cutting or Short-range Recruitment Signal?” Behavioral Ecology and Sociobiology 39: 293–299. Roces, F., and G. Manrique. 1996. “Different Stridulatory Vibrations During Sexual Behaviour and Disturbance in the Blood-sucking Bug Triatoma infestans (Hemiptera: Reduviidae). Journal of Insect Physiology 42: 231–238. Roces, F., J. Tautz, and B. Hölldobler. 1993. “Stridulation in Leaf-cutting Ants: Short-range Recruitment Through Plant-borne Vibrations.” Naturwissenschaften 80: 521–524. Rodríguez, R. L., K. Ramaswamy, and R. B. Cocroft. 2006. “Evidence that Female Preferences Have Shaped Male Signal Evolution in a Clade of Specialized Plant-feeding Insects.” Proceedings of the Royal Society of London B 273: 2585–2593. Rodríguez, R. L., L. E. Sullivan, and R. B. Cocroft. 2004. “Vibrational Communication and Reproductive Isolation in the Enchenopa binotata Species Complex of Treehoppers (Hemiptera: Membracidae).” Evolution 58: 571–578. Röhrig, A., W. H. Kirchner, and R. H. Leuthold. 1999. “Vibrational Alarm Communication in the African Fungus-growing Termite Genus Macrotermes (Isoptera, Termitidae).” Insectes Sociaux 46: 71–77. Rohrseitz, K., and O Kilpinen. 1997. “Vibration Transmission Characteristics of the Legs of Freely Standing Honeybees.” Zoology 100: 80–84. Rosengaus, R. B., C. Jordan, M. L. Lefebvre, and J. F. A. Traniello. 1999. “Pathogen Alarm Behavior in a Termite: A New Form of Communication in Social Insects.” Naturwissenschaften 86: 544–548. Ross, R. J., and J. J. B. Smith. 1978. “Detection of Substrate Vibrations by Salamanders: Inner Ear Sense Organ Activity.” Canadian Journal of Zoology 56: 1156–1162. Rovner, J. S. 1975. “Sound Production by Nearctic Wolf Spiders: A Substratum-coupled Stridulatory Mechanism.” Science 190: 1309–1310. ———. 1980. “Vibration in Heteropoda venatoria (Sparassidae): A Third Method of Sound Production in Spiders.” Journal of Arachnology 8: 193–200. Rovner, J. S., and F. G. Barth. 1981. “Vibratory Communication Through Living Plants by a Tropical Wandering Spider.” Science 214: 464–466.

References

239

Rowe, L. 1994. “The Costs of Mating and Mate Choice in Water Striders.” Animal Behaviour 48: 1049–1056. Rupprecht, R. 1968. “Das Trommeln der Plecopteren.” Zeitschrift für vergleichende Physiologie 59: 38–71. ———. 1974. “Vibrationssignale bei der Paarung von Panorpa (Mecoptera/Insecta).” Experientia 30: 340–341. ———. 1975. “Die Kommunikation von Sialis (Megaloptera) durch Vibrationsignale.” Journal of Insect Physiology 21: 305–320. ———. 1982. “Drumming Signals of Danish Plecoptera.” Aquatic Insects 4: 93–103. Russell, E. M. 1984. “Social Behaviour and Social Organization of Marsupials.” Mammal Review 14: 101–154. Ryan, M. A., A. Cˇokl, and G. H. Walter. 1996. “Differences in Vibratory Sound Communication Between a Slovenian and an Australian Population of Nezara viridula (L.) (Heteroptera: Pentatomidae).” Behavioural Processes 36: 183–193. Ryan, M. A., and G. H. Walter. 1992. “Sound Communication in Nezara viridula (L.) (Heteroptera: Pentatomidae): Further Evidence that Signal Transmission is Substrate-borne.” Experientia 48: 1112–1115. Salmon, M., and K. W. Horch. 1972. “Acoustic Signalling and Detection by Semiterrestrial Crabs of the Family Ocypodidae.” In Behavior of Marine Animals, vol. 1, eds. H. E. Winn and B. L. Olla, pp. 60–96. New York: Plenum Press. Salmon, M., K. Horch, and G. W. Hyatt. 1977. “Barth’s Myochordotonal Organ as a Receptor for Auditory and Vibrational Stimuli in Fiddler Crabs (Uca pugilator and U. minax).” Marine Behavior and Physiology 4: 187–194. Sandeman, D. C., J. Tautz, and M. Lindauer. 1996. “Transmission of Vibration Across Honeycombs and its Detection by Bee Leg Receptors.” Journal of Experimental Biology 199: 2585–2594. Sato, M. 1961. “Response of Pacinian Corpuscles to Sinusoidal Vibration.” Journal of Physiology 159: 391–409. Satou, M., A. Shiraishi, T. Matsushima, and N. Okumoto. 1991. “Vibrational Communication During Spawning Behavior in the Himé Salmon (Landlocked Red Salmon, Oncorhynchus nerka).” Journal of Comparative Physiology A 168: 417–428. Satou, M., H.-A. Takeuchi, J. Nishii, M. Tanabe, S. Kitamura, N. Okumoto, and M. Iwata. 1994. “Behavioral and Electrophysiological Evidences that the Lateral Line Is Involved in the Inter-sexual Vibrational Communication of the Himé Salmon (Landlocked Red Salmon, Oncorhynchus nerka).” Journal of Comparative Physiology A 174: 539–549.

240

References

Sattman, D. A., and R. B. Cocroft. 2003. “Phenotypic Plasticity and Repeatability in the Mating Signals of Enchenopa Treehoppers, with Implications for Reduced Gene Flow Among Host-shifted Populations.” Ethology 109: 981–994. Savoyard, J. L., G. J. Gamboa, D. L. D. Cummings, and R. L. Foster. 1998. “The Communicative Meaning of Body Oscillations in the Social Wasp, Polistes fuscatus (Hymenoptera, Vespidae).” Insectes Sociaux 45: 215–230. Saxena, K. N., and H. Kumar. 1980. “Interruption of Acoustic Communication and Mating in a Leafhopper and a Planthopper by Aerial Sound Vibrations Picked up by Plants.” Experientia 36: 933–936. ———. 1984. “Acoustic Communication in the Sexual Behaviour of the Leafhopper, Amrasca devastans.” Physiological Entomology 9: 77–86. Schatral, A., W. Latimer, and K. Kalmring. 1985. “The Role of the Song for Spatial Dispersion and Agonistic Contacts in Male Bushcrickets.” In Acoustic and Vibrational Communication in Insects, eds. K. Kalmring and N. Elsner, pp. 111–116. Berlin: Paul Parey. Schilman, P. E., C. R. Lazzari, and G. Manrique. 2001. “Comparison of Disturbance Stridulations in Five Species of Triatominae Bugs.” Acta Tropica 79: 171–178. Schmid, A. 1997. “A Visually Induced Switch in Mode of Locomotion of a Spider.” Zeitschrift für Naturforschung C 52: 124–128. Schmitt, A., T. Friedel, and F. G. Barth. 1993. “Importance of Pause Between Spider Courtship Vibrations and General Problems Using Synthetic Stimuli in Behavioural Studies.” Journal of Comparative Physiology A 172: 707–714. Schmitt, A., M. Schuster, and F. G. Barth. 1992. “Male Competition in a Wandering Spider (Cupiennius getazi, Ctenidae).” Ethology 90: 293–306. ———. 1994. “Vibratory Communication in a Wandering Spider, Cupiennius getazi: Female and Male Preferences for Features of the Conspecific Male’s Releaser.” Animal Behaviour 48: 1155–1171. Schneider, S. S. 1986. “The Vibration Dance Activity of Successful Foragers of the Honeybee, Apis mellifera (Hymenoptera: Apidae).” Journal of the Kansas Entomological Society 59: 699–705. ———. 1987. “The Modulation of Worker Activity by the Vibration Dance of the Honeybee, Apis mellifera.” Ethology 74: 211–218. ———. 1989. “Dance Behaviour of Successful Foragers of the African Honeybee, Apis mellifera scutellata (Hymenoptera: Apidae).” Journal of Apicultural Research 28: 150–154. ———. 1991. “Modulation of Queen Activity by the Vibration Dance in Swarming Colonies of the African Honey Bee, Apis mellifera scutellata (Hymenoptera: Apidae).” Journal of the Kansas Entomological Society 64: 269–278.

References

241

Schneider, S. S., and L. A. Lewis. 2004. “The Vibration Signal, Modulatory Communication and the Organization of Labor in Honey Bees, Apis mellifera.” Apidologie 35: 117–131. Schneider, S. S., L. A. Lewis, and Z. Y. Huang. 2004. “The Vibrational Signal and Juvenile Hormone Titers in Worker Honeybees, Apis mellifera.” Ethology 110: 977–985. Schneider, S. S., S. Painter-Kurt, and G. DeGrandi-Hoffman. 2001. “The Role of the Vibration Signal During Queen Competition in Colonies of the Honeybee, Apis mellifera.” Animal Behaviour 61: 1173–1180. Schneider, S. S., J. A. Stamps, and N. E. Gary. 1986a. “The Vibration Dance of the Honey Bee. I. Communication Regulating Foraging on Two Time Scales.” Animal Behaviour 34: 377–385. ———. 1986b. “The Vibration Dance of the Honey Bee. II. The Effects of Foraging Success on Daily Patterns of Vibration Activity.” Animal Behaviour 34: 386–391. Schroeder, C. E., R. W. Lindsley, C. Specht, A. Marcovici, J. F. Smiley, and D. C. Javitt. 2001. “Somatosensory Input to Auditory Association Cortex in the Macaque Monkey.” Journal of Neurophysiology 85: 1322–1327. Schüch, W., and F. G. Barth. 1985. “Temporal Patterns in the Vibratory Courtship Signals of the Wandering Spider Cupiennius salei Keys.” Behavioral Ecology and Sociobiology 16: 263–271. ———. 1990. “Vibratory Communication in a Spider: Female Responses to Synthetic Male Vibrations.” Journal of Comparative Physiology A 166: 817–826. Schwartzkopff, J. 1974. “Mechanoreception.” In The Physiology of Insecta, vol. 2, ed. M. Rockstein, pp. 273–352. New York: Academic Press. Seeley, R. R., T. D. Stephens, and P. Tate. 2007. Essentials of Anatomy & Physiology, 6th ed. Boston: McGraw-Hill. Seeley, T. D. 1992. “The Tremble Dance of the Honey Bee: Message and Meanings.” Behavioral Ecology and Sociobiology 31: 375–383. Seeley, T. D., S. Kühnholz, A. Weidenmüller. 1996. “The Honey Bee’s Tremble Dance Stimulates Additional Bees to Function as Nectar Receivers.” Behavioral Ecology and Sociobiology 39: 419–427. Seeley, T. D., A. Weidenmüller, and S. Kühnholz. 1998. “The Shaking Signal of the Honey Bee Informs Workers to Prepare for Greater Activity.” Ethology 104: 10–26. Seyfarth, E.-A., and F. G. Barth. 1972. “Compound Slit Sense Organs on the Spider Leg: Mechanoreceptors Involved in Kinesthetic Orientation.” Journal of Comparative Physiology 78: 176–191. Shaw, K. C. 1976. “Sounds and Associated Behavior of Agallia constricta and Agalliopsis novella (Homoptera: Auchenorrhyncha: Cicadellidae).” Journal of the Kansas Entomological Society 49: 1–50.

242

References

Shaw, K. C., and O. V. Carlson. 1979. “Morphology of the Tymbal Organ of the Potato Leafhopper Empoasca fabae Harris (Homoptera: Cicadellidae).” Journal of the Kansas Entomological Society 52: 701–711. Shaw, K. C., A. Vargo, and O. V. Carlson. 1974. “Sounds and Associated Behavior of Some Species of Empoasca (Homoptera: Cicadellidae).” Journal of the Kansas Entomological Society 47: 284–307. Shaw, S. R. 1994a. “Detection of Airborne Sound by a Cockroach ‘Vibration Detector’: A Possible Missing Link in Insect Auditory Evolution.” Journal of Experimental Biology 193: 13–47. ———. 1994b. “Re-evaluation of the Absolute Threshold and Response Mode of the Most Sensitive Known ‘Vibration’ Detector, the Cockroach’s Subgenual Organ: A Cochlea-like Displacement Threshold and a Direct Response to Sound.” Journal of Neurobiology 25: 1167–1185. Shier, D. M., and S. I. Yoerg. 1999. “What Footdrumming Signals in Kangaroo Rats (Dipodomys heermanni).” Journal of Comparative Psychology 113: 66–73. Shimizu, I., and F. G. Barth. 1996. “The Effect of Temperature on the Temporal Structure of the Vibratory Courtship Signals of a Spider (Cupiennius salei Keys.).” Journal of Comparative Physiology A 179: 363–370. Simmons, L. W., and W. J. Bailey. 1993. “Agonistic Communication Between Males of a Zaprochiline Katydid (Orthoptera: Tettigoniidae).” Behavioral Ecology 4: 364–368. Singer, F., S. E. Riechert, H. Xu, A. W. Morris, E. Becker, J. A. Hale, and M. A. Noureddine. 2000. “Analysis of Courtship Success in the Funnel-web Spider Agelenopsis aperta.” Behaviour 137: 93–117. Sismondo, E. 1980. “Physical Characteristics of Drumming of Meconema thalassinum.” Journal of Insect Physiology 26: 209–212. Skoglund, C. R. 1960. “Properties of Pacinian Corpuscles of Ulnar and Tibial Location in Cat and Fowl.” Acta Physiologica Scandinavica 50: 385–386. Slobodchikoff, C. N., and H. G. Spangler. 1979. “Two Types of Sound Production in Eupsophulus castaneus (Coleoptera: Tenebrionidae).” Coleopterists Bulletin 33: 239–243. Smith, J. J. B. 1968. “Hearing in Terrestrial Urodeles: A Vibration-sensitive Mechanism in the Ear.” Journal of Experimental Biology 48: 191–205. Smith, J. W., Jr., and G. P. Georghiou. 1972. “Morphology of the Tymbal Organ of the Beet Leafhopper, Circulifer tenellus.” Annals of the Entomological Society of America 65: 221–226. Snarr, K. A. 2005. “Seismic Activity Response as Observed in Mantled Howlers (Alouatta palliata), Cuero y Salado Wildlife Refuge, Honduras.” Primates 46: 281–285.

References

243

Sohmer, H., S. Freeman, M. Geal-Dor, C. Adelman, and I. Savion. 2000. “Bone Conduction Experiments in Humans—A Fluid Pathway from Bone to Ear.” Hearing Research 146: 81–88. Spangler, H. G., and D. G. Manley. 1978. “Sounds Associated with the Mating Behavior of a Mutillid Wasp.” Annals of the Entomological Society of America 71: 389–392. Standing Bear, L. 1933. Land of the Spotted Eagle. Boston: Houghton Mifflin. Stewart, I. C. F., and P. J. Setchell. 1974. “Seismic Recordings of Kangaroo Activity.” Search 5: 107–108. Stewart, K. W. 1997. “Vibrational Communication in Insects: Epitome in the Language of Stoneflies?” American Entomologist 43: 81–91. Stewart, K. W., J. C. Abbott, R. F. Kirchner, and S. R. Moulton II. 1995. “New Descriptions of North American Euholognathan Stonefly Drumming (Plecoptera) and First Nemouridae Ancestral Call Discovered in Soyedina carolinensis (Plecoptera: Nemouridae).” Annals of the Entomological Society of America 88: 234–239. Stewart, K. W., R. L. Bottorff, A. W. Knight, and J. B Moring. 1991. “Drumming of Four North American Euholognathan Stonefly Species, and a New Complex Signal Pattern in Nemoura spiniloba Jewett (Plecoptera: Nemouridae).” Annals of the Entomological Society of America 84: 201–206. Stewart, K. W., and M. Maketon. 1991. “Structures Used by Nearctic Stoneflies (Plecoptera) for Drumming, and Their Relationship to Behavioral Pattern Diversity.” Aquatic Insects 13: 33–53. Stewart, K. W., S. W. Szczytko, and M. Maketon. 1988. “Drumming as a Behavioral Line of Evidence for Delineating Species in the Genera Isoperla, Pteronarcys, and Taeniopteryx (Plecoptera).” Annals of the Entomological Society of America 81: 689–699. Stewart, K. W., S. W. Szczytko, and B. P. Stark. 1982. “Drumming Behavior of Four Species of North American Pteronarcyidae (Plecoptera): Dialects in Colorado and Alaska Pteronarcella badia.” Annals of the Entomological Society of America 75: 530–533. Stewart, K. W., S. W. Szczytko, B. P. Stark, and D. D. Zeigler. 1982. “Drumming Behavior of Six North American Perlidae (Plecoptera) Species.” Annals of the Entomological Society of America 75: 549–554. Stewart, K. W., and D. D. Zeigler. 1984a. “Drumming Behavior of Twelve North American Stonefly (Plecoptera) Species: First Descriptions in Peltoperlidae, Taeniopterygidae and Chloroperlidae.” Aquatic Insects 6: 49–61.

244

References

———. 1984b. “The Use of Larval Morphology and Drumming in Plecoptera Systematics, and Further Studies of Drumming Behavior.” Annals of Limnology 20: 105–114. Stiedl, O., and K. Kalmring. 1989. “The Importance of Song and Vibratory Signals in the Behaviour of the Bushcricket Ephippiger ephippiger Fiebig (Orthoptera, Tettigoniidae): Taxis by Females.” Oecologia 80: 142–144. Stölting, H., T. E. Moore, and R. Lakes-Harlan. 2002. “Substrate Vibrations During Acoustic Signalling in the Cicada Okanagana rimosa.” Journal of Insect Science 2: 1–7. Stratton, G. E. 1997. “Investigation of Species Divergence and Reproductive Isolation of Schizocosa stridulans (Araneae: Lycosidae) from Illinois.” Bulletin of the British Arachnological Society 10: 313–321. ———. 2005. “Evolution of Ornamentation and Courtship Behavior in Schizocosa: Insights from a Phylogeny Based on Morphology (Araneae, Lycosidae).” Journal of Arachnology 33: 347–376. Stratton, G. E., and G. W. Uetz. 1983. “Communication via Substratumcoupled Stridulation and Reproductive Isolation in Wolf Spiders (Araneae: Lycosidae).” Animal Behaviour 31: 164–172. Stritih, N., M. Virant-Doberlet, and A. Cˇokl. 2000. “Green Stink Bug Nezara viridula Detects Differences in Amplitude Between Courtship Song Vibrations at Stem and Petiolus.” European Journal of Physiology Suppl. 439: R190–R192. Strübing, H. 2006. “Vibratory Communication and Mating Behaviour in the European Lantern Fly, Dictyophara europea (Dictyopharidae, Hemiptera).” In Insect Sounds and Communication: Physiology, Behaviour, Ecology and Evolution, eds. S. Drosopoulos and M. F. Claridge, pp. 351–355. Boca Raton, FL: Taylor & Francis. Szczytko, S. W., and K. W. Stewart. 1979. “Drumming Behavior of Four Nearctic Isoperla (Plecoptera) Species.” Annals of the Entomological Society of America 72: 781–786. Szlep, R., and T. Jacobi. 1967. “The Mechanism of Recruitment to Mass Foraging in Colonies of Monomorium venustum Smith, M. subopacum ssp. Phoenicium Em., Tapinoma israelis For. and T. simothi v. Phoenicium Em.” Insectes Sociaux 14: 25–40. Talyn, B. C., and H. B. Dowse. 2004. “The Role of Courtship Song in Sexual Selection and Species Recognition by Female Drosophila melanogaster.” Animal Behaviour 68: 1165–1180. Tarsitano, M., R. R. Jackson, and W. H. Kirchner. 2000. “Signals and Signal Choices Made by the Araneophagic Jumping Spider Portia fimbriata While Hunting the Orb-weaving Web Spiders Zygiella x-notata and Zosis geniculatus.” Ethology 106: 595–615.

References

245

Tautz, J. 1977. “Reception of Medium Vibration by Thoracal Hairs of Caterpillars of Barathra brassicae L. (Lepidoptera, Noctuidae). I. Mechanical Properties of the Receptor Hairs.” Journal of Comparative Physiology 118: 13–31. Tautz, J., F. Roces, and B. Hölldobler. 1995. “Use of a Sound-based Vibratome by Leaf-cutting Ants.” Science 267: 84–87. Thei;sz, J. 1982. “Generation and Radiation of Sound by Stridulating Water Insects as Exemplified by the Corixids.” Behavioral Ecology and Sociobiology 10: 225–235. ———. 1983. “An Acoustic Duet is Necessary for Successful Mating in Corixa dentipes.” Naturwissenschaften 70: 467–468. Tindo, M., E. Francescato, and A. Dejean. 1997. “Abdominal Vibrations in a Primitively Eusocial Wasp Belonogaster juncea juncea (Vespidae: Polistinae).” Sociobiology 29: 255–261. Tishechkin, D. Y. 2006. “Vibratory Communication in Psylloidea (Hemiptera).” In Insect Sounds and Communication: Physiology, Behaviour, Ecology and Evolution, eds. S. Drosopoulos and M. F. Claridge, pp. 357–363. Boca Raton, FL: Taylor & Francis. Torr, P., S. Heritage, and M. J. Wilson. 2004. “Vibrations as a Novel Signal for Host Location by Parasitic Nematodes.” International Journal for Parasitology 34: 997–999. Towne, W. F. 1985. “Acoustic and Visual Cues in the Dances of Four Honey Bee Species.” Behavioral Ecology and Sociobiology 16: 185–187. Travassos, M. A., and N. E. Pierce. 2000. “Acoustics, Context and Function of Vibrational Signalling in a Lycaenid Butterfly-ant Mutualism.” Animal Behaviour 60: 13–26. Tributsch, H. 1982. When the Snakes Awake: Animals and Earthquake Prediction. Cambridge, MA: MIT Press. Tschinkel, W. R., and J. T. Doyen. 1976. “Sound Production by Substratal Tapping in Beetles of the Genus Eusattus (Tentyriidae: Coniontini).” Coleopterists Bulletin 30: 331–335. Uematsu, K., and K. Yamamori. 1982. “Body Vibration as a Timing Cue for Spawning in Chum Salmon.” Comparative Biochemistry and Physiology A 72: 591–594. Uetz, G. W., and G. E. Stratton. 1982. “Acoustic Communication and Reproductive Isolation in Spiders.” In Spider Communication, eds. P. N. Witt and J. S. Rovner, pp. 123–159. Princeton, NJ: Princeton University Press. Virant-Doberlet, M., and A. Cˇokl. 2004. “Vibrational Communication in Insects.” Neotropical Entomology 33: 121–134. Virant-Doberlet, M., A. Cˇokl, and N. Stritih. 2000. “Vibratory Songs of Hybrids from Brazilian and Slovenian Populations of the Green Stink

246

References

Bug Nezara viridula.” European Journal of Physiology Suppl. 439: R196–R198. Visscher, P. K., J. Shepardson, L. McCart, and S. Camazine. 1999. “Vibration Signal Modulates the Behavior of House-hunting Honey Bees (Apis mellifera).” Ethology 105: 759–769. Vollrath, F. 1979. “Vibrations: Their Signal Function for a Spider Kleptoparasite.” Science 205: 1149–1151. Von Hagen, H.-O. 2000. “Vibration Signals in Australian Fiddler Crabs—A First Inventory.” Beagle 16: 97–106. Von Mayer, A., G. O’Brien, and E. E. Sarmiento. 1995. “Functional and Systematic Implications of the Ear in Golden Moles (Chrysochloridae).” Journal of Zoology (London) 236: 417–430. Walker, T. J., and D. E. Figg. 1990. “Song and Acoustic Burrow of the Prairie Mole Cricket, Gryllotalpa major (Orthoptera: Gryllidae).” Journal of the Kansas Entomological Society 63: 237–242. Warkentin, K. M. 1995. “Adaptive Plasticity in Hatching Age: A Response to Predation Risk Trade-offs.” Proceedings of the National Academy of Sciences of the United States of America 92: 3507–3510. ———. 2000. “Wasp Predation and Wasp-induced Hatching of Red-eyed Treefrog Eggs.” Animal Behaviour 60: 503–510. ———. 2005. “How Do Embryos Assess Risk? Vibrational Cues in Predatorinduced Hatching of Red-eyed Treefrogs.” Animal Behaviour 70: 59–71. Warkentin, K. M., M. S. Caldwell, and J. G. McDaniel. 2006. “Temporal Pattern Cues in Vibrational Risk Assessment by Embryos of the Red-eyed Treefrog, Agalychnis callidryas.” Journal of Experimental Biology 209: 1376–1384. Webster, D. B. 1992. “Epilogue to the Conference on the Evolutionary Biology of Hearing.” In The Evolutionary Biology of Hearing, eds. D. B. Webster, R. R. Fay, and A. N. Popper, pp. 787–793. New York: Springer-Verlag. Weidemann, S., and A. Keuper. 1987. “Influence of Vibratory Signals on the Phonotaxis of the Gryllid Gryllus bimaculatus DeGeer (Ensifera: Gryllidae).” Oecologia 74: 316–318. Weissman, D. B. 2001. “Communication and Reproductive Behaviour in North American Jerusalem Crickets (Stenopelmatus) (Orthoptera: Stenopelmatidae).” In The Biology of Wetas, King Crickets and Their Allies, ed. L. H. Field, pp. 351–375. Wallingford, UK: CAB International. Wells, M. M., and C. S. Henry. 1992a. “Behavioural Responses of Green Lacewings (Neuroptera: Chrysopidae: Chrysoperla) to Synthetic Mating Songs.” Animal Behaviour 44: 641–652. ———. 1992b. “The Role of Courtship Songs in Reproductive Isolation Among Populations of Green Lacewings of the Genus Chrysoperla (Neuroptera: Chrysopidae).” Evolution 46: 31–42.

References

247

Wenner, A. M. 1962a. “Communication with Queen Honey Bees by Substrate Sound.” Science 138: 446–448. ———. 1962b. “Sound Production During the Waggle Dance of the Honey Bee.” Animal Behaviour 10: 79–95. West, M. J., and R. D. Alexander. 1963. “Sub-social Behavior in a Burrowing Cricket Anurogryllus muticus (DeGeer) Orthoptera: Gryllidae.” Ohio Journal of Science 63: 19–24. West-Eberhard, M. J. 1978. “Temporary Queens in Metapolybia Wasps: Nonreproductive Helpers Without Altruism?” Science 200: 441–443. Wever, E. G. 1968. “The Ear of the Chameleon: Chamaeleo senegalensis and Chamaeleo quilensis.” Journal of Experimental Zoology 168: 423–436. ———. 1969. “The Ear of the Chameleon: Chamaeleo höhnelii and Chamaeleo jacksoni.” Journal of Experimental Zoology 171: 305–312. Whang, A., and J. Janssen. 1994. “Sound Production Through the Substrate During Reproduction in the Mottled Sculpin, Cottus bairdi (Cottidae).” Environmental Biology of Fishes 40: 141–148. Wiese, K. 1974. “The Mechanoreceptive System of Prey Localization in Notonecta. II. The Principle of Prey Localization.” Journal of Comparative Physiology 92: 317–325. Wikramanayake, E., P. Fernando, and P. Leimgruber. 2006. “Behavioral Response of Satellite-collared Elephants to the Tsunami in Southern Sri Lanka.” Biotropica 38: 775–777. Wilcox, R. S. 1972. “Communication by Surface Waves.” Journal of Comparative Physiology 80: 255–266. ———. 1979. “Sex Discrimination in Gerris remigis: Role of a Surface Wave Signal.” Science 206: 1325–1327. Wilcox, R. S., and J. Di Stefano. 1991. “Vibratory Signals Enhance MateGuarding in a Water Strider (Hemiptera: Gerridae).” Journal of Insect Behavior 4: 43–50. Wilcox, R. S., R. R. Jackson, and K. Gentile. 1996. “Spiderweb Smokescreens: Spider Trickster Uses Background Noise to Mask Stalking Movements.” Animal Behaviour 51: 313–326. Wilczynski, W., C. Resler, and R. R. Capranica. 1987. “Tympanic and Extratympanic Sound Transmission in the Leopard Frog.” Journal of Comparative Physiology A 161: 659–669. Willi, U. B., G. N. Bronner, and P. M. Narins. 2006. “Ossicular Differentiation of Airborne and Seismic Stimuli in the Cape Golden Mole (Chrysochloris asiatica).” Journal of Comparative Physiology A 192: 267–277. Yack, J. E., M. L. Smith, and P. J. Weatherhead. 2001. “Caterpillar Talk: Acoustically Mediated Territoriality in Larval Lepidoptera.” Proceedings of

248

References

the National Academy of Sciences of the United States of America 98: 11371–11375. Yodlowski, M. L., M. L. Kreither, and W. T. Keeton. 1977. “Detection of Atmospheric Infrasound by Homing Pigeons.” Nature 265: 725–726. Young, B. A. 2003. “Snake Bioacoustics: Toward a Richer Understanding of the Behavioral Ecology of Snakes.” Quarterly Review of Biology 78: 303–325. Young, B. A., and M. Morain. 2002. “The Use of Ground-borne Vibrations for Prey Localization in the Saharan Sand Vipers (Cerastes).” Journal of Experimental Biology 205: 661–665. Yu, X., E. R. Lewis, and D. Feld. 1991. “Seismic and Auditory Tuning Curves from Bullfrog Saccular and Amphibian Papillar Axons.” Journal of Comparative Physiology A 169: 241–248. Zachariassen, K. E. 1977. “Communication by Sound Between Desert Tenebrionids.” Norwegian Journal of Entomology 24: 35–36. Zeigler, D. D., and K. W. Stewart. 1977. “Drumming Behavior of Eleven Nearctic Stonefly (Plecoptera) Species.” Annals of the Entomological Society of America 70: 495–505. ———. 1985a. “Age Effects on Drumming Behavior of Pteronarcella badia (Plecoptera) Males.” Entomological News 96: 157–160. ———. 1985b. “Drumming Behavior of Five Stonefly (Plecoptera) Species from Central and Western North America.” Annals of the Entomological Society of America 78: 717–722. ———. 1986. “Female Response Thresholds of Two Stonefly (Plecoptera) Species to Computer-simulated and Modified Male Drumming Calls.” Animal Behaviour 34: 929–931. Zuk, M., and G. R. Kolluru. 1998. “Exploitation of Sexual Signals by Predators and Parasitoids.” Quarterly Review of Biology 73: 415–438.

Species Index

MAMMALIA Order Monotremata Ornithorhynchus anatinus Gray, 63 Tachyglosss aculeatus (Shaw), 63 Order Dasyuromorpha Sarcophilus harrisii Boitard, 91–92 Order Diprotodontia Macropus eugenii Desmarest, 63–64, 91–92 Macropus giganteus Shaw, 91–92 Order Afrosoricida Chrysochloris asiatica (Linnaeus), 46 Eremitalpa granti Broom, 44–46, 120–121, 205 Order Proboscidea Elephus maximus Linnaeus, 9–10, 43–44, 61, 89–91, 178 Loxodonta africana Blumenbach, 89–91, 178 Order Primates Alouatta palliata Gray, 10 Homo sapiens sapiens Linnaeus, 56, 60 Order Rodentia Cryptomys Gray sp., 205 Dipodomys deserti Stephens, 46, 88, 128 Dipodomys heermanni LeConte, 88–89 Dipodomys ingens (Merriam), 46, 88, 128 Dipodomys spectabilis Merriam, 46, 88, 127–128, 183 Georychus capensis (Pallas), 44, 87–88 Heterocephalus glaber Rüppell, 47 Nannospalax ehrenbergi, 13, 43, 87. See also Spalax ehrenbergi Nehring Notomys Lesson sp., 47 Pseudomys Gray sp., 47 Rhombomys opimus (Lichtenstein), 129

249

250

Species Index

Spalax ehrenbergi Nehring, 13, 43–44, 47, 61–62, 87, 136, 205. See also Nannospalax ehrenbergi Order Soricomorpha Condylura cristata Linnaeus, 62–63 Order Carnivora Spilogale interrupta (Linnaeus), 89 Vulpes macrotus Merriam, 128 Order Artiodactyla Odocoileus virginianus Zimmermann, 89 Order Cetacea Tursiops truncatus Montagu, 42 AVES Columba livia Gmelin, 41 Fulica atra Linnaeus, 92 REPTILIA Order Squamata Cerastes Laurenti sp., 121 Chamaeleo calyptratus Duméril & Bibron, 65, 92 Chamaeleo höhnelii (Steindachner), 48 Chamaeleo jacksoni (Boulenger), 48 Chamaeleo quilensis Bocage, 48 Chamaeleo senegalensis Daudin, 48 Dispholidus typus (Smith), 49 Erix miliaris (Pallas), 129 Pituophis melanoleucus affinis Conant & Collins, 127, 129 Pseudechis porphyriacus Shaw, 64–65. See also Pseudechis porthyriacus (sic) Pseudechis porthyriacus (sic), 64–65 Scincus scincus Linnaeus, 50, 121 AMPHIBIA Order Urodela Salamandra Laurenti sp., 49 Triturus Rafinesque sp., 49 Order Anura Agalychnis callidryas Cope, 129–130 Eleutherodactylus coqui Thomas, 53 Leptodactylus albilabris (Gunther), 50, 92–93, 179 Leptodactylus syphax Bokermann, 93 Lithobates catesbeiana (Shaw), 48, 50. See also Rana catesbeiana (Shaw) Lithobates clamitans (Latreille), 32. See also Rana clamitans (Latreille) Lithobates pipiens (Schreber), 51. See also Rana pipiens (Schreber) Lithobates sphenocephala (Cope), 130–131. See also Rana sphenocephala (Cope) Rana catesbeiana (Shaw), 48, 50 Rana clamitans (Latreille), 32 Rana pipiens (Schreber), 51 Rana sphenocephala (Cope), 130–131

Species Index

251

Rana temporaria Linnaeus, 51 Telmatobius halli Noble, 94 Telmatobius montanus Lataste, 94 Telmatobius marmoratus Duméril & Bibron, 94 Telmatobius pefauri Veloso & Bibron, 94 Telmatobius peruvianus Weigmann, 94 Telmatobius zapahuirensis Veloso, Sallaberry, Navarro, Iturra, Valencia, Penna & Díaz, 94 Xenopus laevis Daudin, 53–55 OSTEICHTHYES Cottus bairdii Girard, 52, 95 Oncorhynchus keta Walbaum, 136 Oncorhynchus nerka (Walbaum), 136 Opsanus tau (Linnaeus), 94–95 INSECTA Order Plecoptera Acroneuria abnormis (Newman), 141 Capnia bifrons Newman, 140 Dinocras cephalotes Curtis, 140 Nemoura spiniloba Jewett, 138 Isogenoides zionensis Hanson, 140 Isoperla Banks sp., 140 Oconoperla innubila (Needham & Classen), 141 Perla marginata Panzer, 140 Perlinella drymo (Newman), 139 Pteronarcella badia (Hagen), 139 Taeniopteryx burksi Ricker & Ross, 140 Order Megaloptera Sialis fuliginosa Pictet, 98, 141–142 Sialis lutaria Linnaeus, 98, 141–142 Order Orthoptera Ametrus Brunner von Wattenwyl sp., 99, 145 Anurogryllus muticus (DeGeer), 2 Arachnoscelis Karny sp., 207 Balamara gydia Otte & Alexander, 143–144 Chorthippus parallelus (Zetterstedt), 96 Conocephalus nigropleurum (Bruner), 108, 144 Copiphora rhinoceros Pictet, 107, 207 Decticus verrucivorus (Linnaeus), 175 Docidocercus chlorops Nickle, 107 Ephippiger ephippiger Fiebig, 107, 143 Gryllotalpa major Saussure, 3–4, 177–181 Gryllus bimaculatus DeGeer, 142 Gryllus campestris Linnaeus, 2 Hadrogryllacris Karny sp., 99, 145 Hemiandrus pallitarsis (F. Walker), 112, 145

252

Species Index

Hemideina crassidens (Blanchard), 111–112, 144–145 Hemideina femorata Hutton, 111–112, 144–145 Hemideina ricta (Hutton), 111–112 Meconema thalassinum (DeGeer), 99, 145 Melanonotus powellorum Rentz, 107 Myopophyllum speciosum Beier, 107 Oecanthus nigricornis F. Walker, 25, 143, 208 Ornebius aperta Otte and Alexander, 144 Phaeophilacris spectrum Saussure, 112 Polysarcus denticauda (Charpentier), 72 Stenopelmatus Burmeister sp., 98–99 Tettigonia cantans Fuessly, 142, 176 Order Blattodea Arenivaga investigata Friauf & Edney, 133 Periplaneta americana Linnaeus, 69–70 Order Isoptera Macrotermes bellicosus (Smeathman), 133, 184 Macrotermes subhyalinus (Rambur), 133, 184 Pseudacanthotermes militaris (Hagen), 133 Pseudacanthotermes spiniger (Sjöstedt), 133 Reticulitermes flavipes (Kollar), 96 Zootermopsis angusticollis (Hagen), 133, 184–185, 188 Zootermopsis nevadensis (Hagen), 133, 184 Order Hemiptera (Heteroptera: Gerromorpha) Aquarius remigis, 32. See also Gerris remigis Say Gerris buenoi Kirkaldy, 32 Gerris lacustris (Linnaeus), 31 Gerris odontogaster (Zetterstedt), 32 Gerris remigis Say, 31–32. See also Aquarius remigis Limnoporus dissortis (Drake & Harris), 33 Limnoporus notabilis (Drake & Hottes), 33 Limnoporus rufoscutellatus (Latreille), 33 Rhagadotarsus spp. Breddin, 32–33 Order Hemiptera (Heteroptera: Nepomorpha) Corixa dentipes (Thomson), 106, 150 Corixa punctata (Illiger), 106 Notonecta glauca Linnaeus, 30–31, 67–68 Notonecta kirkii, 32 Notonecta lutea Müller, 32 Order Hemiptera (Heteroptera: Cimicomorpha) Corythucha hewitti Drake, 198–199 Dipetalogaster maxima (Uhler), 105 Nabis kinbergii Reuter, 128 Rhodnius prolixus Stål, 105 Triatoma guasayana Wygodzinsky & Abalos, 105 Triatoma infestans (Klug), 105 Triatoma sordida (Stål), 105

Species Index Order Hemiptera (Heteroptera: Cydnidae), 68, 97, 117 Sehirus (Canthophorus) dubius Scopoli, 24 Sehirus (Canthophorus) impressus Horváth, 24 Sehirus (Canthophorus) melanopterus (Herrich-Schäffer), 24 Sehirus (Tritomegas) bicolor (Linnaeus), 24 Order Hemiptera (Heteroptera: Pentatomidae) Acrosternum hilare (Say), 149–150 Euschistus conspersus (Uhler), 150 Holcostethus strictus (Fabricius), 150 Murgantia histrionica (Hahn), 27, 150 Nezara viridula (Linnaeus), 27, 29, 71, 75, 122, 145–150 Podiscus maculiventris (Say), 122–123 Thyanta custator accerra ( Jones), 150 Thyanta pallidovirens (Stål), 150 Order Hemiptera (Auchenorrhyncha Dictyopharidae) Dictyophara europea (Linnaeus), 151 Order Hemiptera (Auchenorrhyncha: Delphacidae) Euides speciosa (Boheman), 24 Javesella Fennah sp., 152 Laodelphas striatellus Fall, 24, 114–115, 151 Nilaparvata lugens (Stål), 24, 28, 111, 114–115, 151–152, 175 Oliarus Stål sp., 153 Prokelisia dolus Wilson, 152 Prokelisia marginata (Van Duzee), 152 Ribautodelphax Wagner sp., 152 Sogatella furcifera (Horváth), 24, 114–115, 151 Solonaima Kirkaldy sp., 153 Tachycixius Wagner sp., 153 Order Hemiptera (Auchenorrhyncha: Cicadidae) Okanagana rimosa (Say), 116–117 Tettigarcta crinita Distant, 114 Tettigarcta tomentosa White, 114–115 Order Hemiptera (Auchenorrhyncha: Membracidae) Calloconophora pinguis (Fowler), 198 Enchenopa binotata (Say), 156 Ennya chrysura Fairmaire, 158, 175 Tomogonia vittatipennis (Fairmaire), 133 Umbonia crassicornis (Amyot & Serville), 27, 156, 200 Vanduzea arquata Say, 158–159 Order Hemiptera (Auchenorrhyncha: Cicadellidae) Agallia constricta Van Duzee, 154 Agalliopsis novella (Say), 154 Amrasca devastans (Distant), 28, 154. See also Empoasca fabae (Harris) Balclutha incisa (Matsumura), 128 Circulifer tenellus (Baker), 113, 115–116 Empoasca fabae (Harris), 115 Euscelidus variegatus Kirshbaum, 24

253

254

Species Index

Euscelis lineolatus Brulle, 24 Graminella nigrifrons (Forbes), 155–156, 164, 180 Oncopsis avellanae Edwards, 154–155 Oncopsis flavicollis (Linnaeus), 75, 154–155 Oncopsis tristis (Zetterstedt), 154–155 Spissistilus festinus (Say), 155 Order Hemiptera (Sternorrhyncha: Aleyrodidae), 111 Order Hemiptera (Sternorrhyncha: Psylloidea), 105, 159 Order Hemiptera (Coleorrhyncha) Hackeriella veitchi (Hacker), 96 Order Neuoptera Brinckochrysa Tjeder sp., 109 Chrysocerca Weele sp., 109 Chrysopa carnea Stephens, 160. See also Chrysoperla carnea (Stephens) Chrysoperla agilis Henry, Brooks, Duelli & Johnson, 161 Chrysoperla carnea (Stephens), 25, 67–68, 71, 160–161 Chrysoperla downesi (Smith), 160 Chrysoperla lucasina (Lacroix), 67–68 Chrysoperla pallida Henry, Brooks, Duelli & Johnson, 161 Chrysoperla plorabunda Fitch, 160–161 Dichrostigma flavipes (Stein), 71 Hemerobius micans Olivier, 71 Meleoma (Fitch) sp., 109 Myrmeleon formicarius Linnaeus, 123 Osmylus fulvicephalus (Scopoli), 71 Semidalis aleyrodiformia (Stephens), 71 Sisyra terminalis Curtis, 71 Order Coleoptera Aphodius ater (DeGeer), 104 Aphodius granarius (Linnaeus), 104 Aphodius pectoralis LeConte, 104 Aphodius rainieri Hatch, 104 Aphodius sigmoideus Van Dyke, 104 Dineutes discolor Aubé, 33, 76 Epilachna varivestis Mulsant, 122 Eupsophulus castaneus (Horn), 161–162 Eusattus convexus LeConte, 161 Eusattus muricatus LeConte, 161 Eusattus reticulatus Say, 161 Eusattus robustus LeConte, 161 Hylotrupes bajulus (Linnaeus), 105–106 Lepidochora discoidalis (Gebien), 74 Onymacris plana plana (Peringuey), 74 Onymacris rugatipennis (Haag), 74 Phelypera distigma (Boheman), 199 Phrynocolus somalicus Wilke, 162 Physodesmia globosa (Haag), 74 Platypus quercivorus (Murayama), 104

Species Index Polychalma multicava (Latreille), 198 Psammodes striatus (Fabricius), 204–205 Xestobium rufovillosum (DeGeer), 162 Zophosis orbicularis (Deyrolle), 74 Order Mecoptera Panorpa alpina (Rambur), 162 Panorpa communis Linnaeus, 162 Panorpa germanica Linnaeus, 162 Order Diptera Anastrepha suspense Lowe, 125 Drosophila cyrtoloma Hardy, 163 Drosophila fasciculisetae Hardy, 163 Drosophila heteroneura Perkins, 163 Drosophila melanogaster (Meigen), 163 Drosophila silvestris Basden, 163 Formicosepsis hamata, 175 Lipara japonica Kanmiya, 84, 112, 164, 175 Order Hymenoptera Aphaenogaster Mayr sp., 185 Apis cerana Fabricius, 196 Apis dorsata Fabricius, 196 Apis florea Fabricius, 196 Apis laboriosa Smith, 196 Apis mellifera Linnaeus, 73–74, 112, 190–198 Apis mellifera scutellata Lepeletier, 195 Atta cephalotes (Linnaeus), 73, 98, 186–188 Atta sexdens (Linnaeus), 98, 103–104, 186 Belonogaster Saussure sp., 189 Biosteres longicaudatus Ashmead, 125 Bombus hortorum (Linnaeus), 37 Bombus terrestris (Linnaeus), 37 Camponotus ligniperda Latreille, 73, 98, 185 Cotesia rubecula (Marshall), 164 Dasymutilla foxi (Cockerell), 164 Euglossa imperialis Cockerell, 37 Formica rufa Linnaeus, 123 Iridomyrmex anceps (Roger), 132 Leptogenys chinensis Mayr, 188 Megaponera foetens (Fabricius), 133, 188–189 Melipona seminigra Friese, 192 Messor capitatus (Latreille), 186 Metapolybia aztecoides Richards, 194 Monomorium Mayr sp., 188 Novomessor albisetosus (Mayr), 185–186 Novomessor cockerelli (André), 185–186 Perga dorsalis Leach, 198 Pimpla instigator Fabricius, 101–102 Pimpla turionellae (Linnaeus), 101–102, 206

255

256

Species Index

Pogonomyrmex Mayr sp., 186 Polistes dominulus (Christ), 99, 190 Polistes fuscatus (Fabricius), 99, 189–190 Polistes metricus Say, 189 Polyrhachis Arnold sp., 98 Sympiesis sericeicornis Nees, 124–125 Tapinoma Foerster sp., 188 Vespa orientalis Fabricius, 190 Order Lepidoptera Barathra brassicae (Linnaeus), 82 Caloptillia serotinella (Ely), 117 Drepana arcuata Walker, 199 Jalmenus evagoras (Donovan), 132 Phyllonorycter blancardella Fabricius, 124–125 Phyllonorycter cydoniella (Denis & Schiffermüller), 124 Phyllonorycter malella (Gerasimov), 123–125 Plathypena scabra (Fabricius), 122 Semiothisa aemulataria (Walker), 131 Thisbe irenea (Stoll), 131–132 AR ACHNIDA Order Scorpiones Paruroctonus mesaensis (Stahnke), 5–6, 76–78, 86, 122, 133 Order Acariformes Neumania papillator Marshall, 172 Order Araneae Acanthoscurria suina Pocock, 101 Agelena Koch sp., 83 Agelenopsis aperta (Gertsch), 169 Araneus (or Araneous) sericatus Clerck, 79, 83 Argiope argentata (Fabricius), 126 Argyrodes elevatus Taczanowski, 126 Cupiennius coccineus Pickard-Cambridge, 28, 171, 207 Cupiennius getazi Simon, 28, 109–110, 171, 173–174, 207 Cupiennius salei (Keyserling), 28–30, 76–81, 110, 126, 169–172, 207 Cyrtophora cicatrosa (Stoliczka), 168 Dolomedes triton (Walckenaer), 33, 77, 79 Eupalaestrus weijenberghi (Thorell), 101 Habronattus dossenus Griswold, 22, 164–165 Habronattus pugillis Griswold, 165–166 Heteropoda venatoria (Linnaeus), 109 Hygrolycosa rubrofasciata (Ohlert), 99–100, 166–167 Isopoda immanis (Koch), 109 Leucorchestris arenicola Lawrence, 175 Lycosa tarentula fasciiventris Dufour, 99, 171–173 Micrathena sagittata (Walckenaer), 126 Nephila clavipes (Linnaeus), 36, 83, 126 Phidippus johnsoni (Peckham & Peckham), 205

Species Index Pisaura Simon sp., 83 Portia fimbriata (Doleschall), 126–127 Schizocosa ocreata (Hentz), 99–100, 167 Schizocosa rovneri Uetz & Dondale, 100, 167 Schizocosa stridulans Stratton, 100, 167–168 Scytodes pallida Doleschall, 130 Sericopelma Ausserer sp., 83 Stegodyphus lineatus Latreille, 168–169 Theridion saxatile (Koch), 200 Zygiella Pickard-Cambridge sp., 35, 83 CRUSTACEA Gecarcinus lateralis (Fréminville), 97 Heloecius cordiformis (Milne-Edwards), 137 Nephrops norvegicus (Linnaeus), 83 Ocypode jousseaumei (Nobili), 81–82, 97–98, 102–103 Ocypode platytarsus Milne-Edwards, 81–82, 97–98, 102–103 Ocypode rotundata Miers, 81–82, 97–98, 102–103 Ocypode saratan (Forskål), 81–82, 97–98, 102–103 Ocypode sinensis Dai, Song & Yang, 81–82, 97–98, 102–103 Uca longidigitum (Kingsley), 137 Uca minax LeConte, 81–82 Uca pugilator (Bosc), 14, 81–83, 97–98, 165, 205 NEMATODA Heterorhabditis megidis Poinar, Jackson & Klein, 123 Steinernema carpocapsae (Weiser), 123 Steinernema feltiae (Filipjev), 123

257

Subject Index

Abdominal rubbing, 140–141 Abdominal wagging, 169, 189. See also Tremulation Acceleration detector, 50, 81–82 Acoustic fat, 42–43 Advertisement signal, 4, 53, 87, 92–93, 95–96, 99, 107, 119, 137, 141–142, 145, 154, 159, 161, 173, 176, 179, 182, 207 Aggressive mimicry, 126–127 Aggressive signal, 32–33, 93, 97, 106, 156, 159, 175, 180 Alarm signal, 87, 89–92, 94–95, 103, 129, 133, 184–185, 197, 200 Annelida, 13 Ant-associated, 131–133, 158–159 Auditory ossicle, 7–8, 40–41, 43–48, 52 Barth’s organ, 67, 81–82 Basitarsal compound slit sensilla (BCSS), 6, 66, 76–79, 82 Bending wave, 17, 20, 25, 30, 142 Bone conduction, 7, 41–49, 52 Boundary vibration, 17–20 Brownell, Philip, 5–6, 23 Buzz pollination, 36–37 Calling song, 4, 24, 29, 32, 83, 93, 98–99, 104–107, 111, 128, 142–143, 146–159, 161, 164, 175–181, 207

Campaniform sensilla, 66–68, 75–76, 105, 143, 192 Cave-dweller, 152–153 Cerebral cortex, 59–60, 62, 65–66 Chaetognatha, 13 Chela, 14, 97–98, 137, 205 Chordotonal organ, 66–69, 74–75, 81–82, 192 Chorusing, 3, 137, 158, 177, 180–181 Cocroft, Reginald (Rex), 2 Columella (stapes), 8, 47–48, 54 Common sound, 153–154 Contact vibration, 17, 21, 32–33, 66–67, 94, 104, 106, 143 Courtship signal, 24, 29–30, 32–34, 80–81, 92, 97–99, 101, 104, 106–107, 109–111, 115, 128, 137, 142, 144, 146–148, 151, 153–154, 156, 158, 160, 163–175, 186, 200, 205, 208 Defense signal, 95, 102, 111–112, 132, 144, 187 Dialects, 139 Dicotyledonous leaf, 27–29, 81 Directional information, 33, 49, 77, 83, 121–122, 133, 140, 146, 149, 155–156, 162, 203 Distance estimation, 6, 25, 30, 33, 80, 82–83, 122, 133, 142, 155, 170, 173

259

260

Subject Index

Disturbance sound, 97, 104–105, 119–120, 133, 154, 184, 189, 199 Dorso-ventral abdominal movement, 109–112, 114, 141, 143, 162, 168, 191, 193, 196. See also Tremulation Drumming, 14, 21, 23, 30, 46–47, 62, 80, 82, 87–102, 107, 109, 111–112, 125, 127–129, 133, 136–145, 154, 156, 158, 161–162, 164–175, 183–185, 188–190, 198–199, 205, 208–209 Duet, 26, 87–88, 98, 106, 128, 135, 139–140, 145–146, 150–151, 154, 156, 159–162, 169–171 Eimer’s organ, 62–63 Farley, Roger, 5 Female abdominal vibration (FAV), 111, 115, 151–152. See also Tremulation Female choice, 144, 152, 157–158, 166–167, 169, 171–172 Flight muscles, 36–37, 112, 163–164, 192–194 Footdrumming, 46–47, 87–88, 127–129, 183, 208. See also Drumming Frequency discrimination, 31, 79–80, 83 Gas bladder, 52, 94–95 Geophone, 22, 89, 120, 178–179, 206 Gogala, Matija, 68, 97 Hair sensilla, 66, 82–84 Harmonics, 28, 69, 103, 156, 158, 164, 177–178, 194 Herbst corpuscle, 63–65 Honeycomb, 34, 112, 190–191, 196–197 Hunt, Randy, 181 Hyomandibula, 8

Longitudinal wave, 17, 19, 21, 35–36, 79 Love wave, 18–19, 23 Lyriform organ, 34, 66, 78–81, 83–85 Meissner’s corpuscle, 56–57, 61–62 Merkel’s receptor, 56–57, 63 Mid-coxal protuberance, 84 Modulatory communication, 149, 185, 191–197, 199 Monocotyledonous leaf, 28–29, 81, 169 Multimodal signal, 13, 143, 165–168, 186, 206 Narins, Peter, 206 Nearfield motion, 17, 75, 168 Nest moving, 188, 194–195 O’Connell-Rodwell, Caitlin, 202 Opercularis, 48–49 Opisthosoma, 30, 99, 109–110, 170–174 Ossiannilsson, Frej, 3, 15, 96, 113–115, 181 Pacinian corpuscle, 56–65, 69–70, 81, 84–85, 120 Pathogen alarm, 184–185 Pedipalp, 30, 80, 99, 167–174 Percussion. See Drumming Pheromones, 109, 146, 149–150, 164, 169, 172, 185, 188, 208 Pneumatic duct, 52 Predator-induced hatching, 129–131 Propagation with dispersion, 20–21, 25 P-wave (compressional), 6, 10–11, 17–18, 23, 82, 121, 175 Quadrate bone, 7–8, 48–49 Quasi-longitudinal wave, 17, 19 Queen piping, 164, 194

Infrasound, 21, 41, 112 Johnston’s organ, 33, 66–67, 74–76, 105, 192 Juvenile hormone ( JH), 195 Lagena, 7, 50–51 Lateral vibration (LV), 189–190 Lewis, Edwin (Ted), 206

Randall, Jan, 202, 208 Rapping, 82–83, 98, 102. See also Drumming Rayleigh wave, 6, 18–19, 23, 46, 48–49, 59, 77, 86, 88–90, 93, 121, 162, 175, 206 Recruitment, 103, 133, 186–188, 190, 194–195, 197–198

Subject Index Repelling vibration, 146, 186 Rivalry song, 24, 147, 149–150, 159, 180 Rodríguez, Rafael, 2 Ruffini’s corpuscle, 56–57 Sacculus (saccule), 7, 50–52, 121 Sand, 6, 17, 21–23, 33, 45–46, 74, 76, 86, 95, 104, 120–123, 128, 165, 171, 175, 206 Scolopidia, 67, 70–72, 75 Scraping, 30, 117, 190, 199 Shaking signal, 191, 193 Sibling-group signal, 200–202 Signal, 12–13, 37–39, 94, 135–136, 181–182, 208 Signal jamming, 158, 175 Slapping. See Drumming Slit sensilla, 6, 66, 76–85 Soil, 3, 7, 9, 17, 21–23, 43, 86–87, 92–93, 97–98, 103–104, 123, 125, 141–142, 165, 177–179, 204–205 Sounding, 71–72, 101–102, 125 Spacing, 87–88, 90, 98, 117, 162, 176–177, 181–182 Spider webs, 19, 34–36, 80, 83, 126–127, 168–169 Stamping. See Drumming Stance, 33, 73, 77–79, 82, 122, 149 Stapes (columella), 8, 40, 45–49, 54 Stridulation, 21, 96–107, 109, 111–112, 117, 120, 128, 132–133, 142–145, 150–151, 159, 162, 164–165, 167–168, 172, 176–177, 185–189, 207 Subgenual organ, 14–15, 66–76, 81, 84–85, 98, 102, 105, 125, 143, 181, 192, 197 S-wave, 10–11, 18 Swim bladder, 52, 71, 94. See also Gas bladder

261

Synthetic signals, 80–81, 84, 124–125, 160, 170, 175, 179–180, 194, 197 Tapping. See Drumming Tarso-pretarsal organ, 68 Temperature dependent, 110, 138 Temporal pattern, 22, 30, 95, 97, 105, 110, 124, 130, 137, 139–140, 144–147, 152, 154, 159, 167, 170–173, 181 Thumping, 91, 93, 98. See also Drumming Ticking, 154 Torsional wave, 17, 19, 35 Transducer, 25, 39, 55, 68 Transverse wave, 17, 19–20, 35–36, 79, 107 Tremble dance, 196–197 Tremulation, 21, 30, 99, 107–112, 115, 117, 128, 140–141, 143–144, 151, 155, 160–162, 164–165, 167, 170, 172, 188–189, 191, 193, 196, 199, 207–208 Trichobothria, 34, 66, 83–84, 105 Trichoid sensilla, 66–67 Tsunami, 9–11, 59–60 Tymbal, 3, 15, 96, 112–117, 151, 154–155, 159 Tympanal organ, 14–15, 70, 97, 105–106, 114, 143 Utriculus (utricle), 50–51 Velum, 71–72 Vibration dance, 191–195 Vocalization, 91, 94 Water-borne, 20, 30–34, 55, 82–83, 122 Weberian ossicles, 52 Wind, 22, 29, 31, 74, 125, 127, 129, 205–206 Wing movement, 111–112, 116, 141, 143, 154, 156, 163–164