Freshwater Fishes of North America: Volume 1: Petromyzontidae to Catostomidae 9781421412016, 9781421412023, 1421412012, 1421412020

Certain to stand among the reference books of choice for anyone interested in the continent’s aquatic ecosystems,Freshwa

736 108 98MB

English Pages 656 [665] Year 2014

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Freshwater Fishes of North America: Volume 1: Petromyzontidae to Catostomidae
 9781421412016, 9781421412023, 1421412012, 1421412020

Citation preview

Freshwater Fishes of North America

This page intentionally left blank

Freshwater Fishes of North America VOLUME 1

Petromyzontidae to Catostomidae Edited by Melvin L. Warren, Jr., and Brooks M. Burr Illustrated by Joseph R. Tomelleri

Johns Hopkins University Press BALTIMORE

© 2014 Johns Hopkins University Press Color illustrations © 2014 Joseph R. Tomelleri All rights reserved. Published 2014 Printed in China on acid-free paper 9 8 7 6 5 4 3 2 1 Johns Hopkins University Press 2715 North Charles Street Baltimore, Maryland 21218-4363 www.press.jhu.edu Library of Congress Cataloging-in-Publication Data Freshwater fishes of North America / edited by Melvin L. Warren, Jr., and Brooks M. Burr ; illustrated by Joseph R. Tomelleri. volumes cm Includes bibliographical references and index. ISBN-13: 978-1-4214-1201-6 (hardcover : alk. paper) ISBN-13: 978-1-4214-1202-3 (electronic) ISBN-10: 1-4214-1201-2 (hardcover : alk. paper) ISBN-10: 1-4214-1202-0 (electronic) 1. Freshwater fishes—North America. I. Warren, Melvin L., Jr., editor of compilation. II. Burr, Brooks M., editor of compilation. QL625.F74 2014 597.176—dc23 2013015264 A cata log record for this book is available from the British Library. Special discounts are available for bulk purchases of this book. For more information, please contact Special Sales at 410-516-6936 or [email protected]. Johns Hopkins University Press uses environmentally friendly book materials, including recycled text paper that is composed of at least 30 percent post-consumer waste, whenever possible.

Contents

List of Contributors

vii

Preface, by Melvin L. Warren, Jr., and Brooks M. Burr Acknowledgments Chapter 1

ix

xvii

Evolution and Ecology of North American Freshwater Fish Assemblages 1 Stephen T. Ross and William J. Matthews

Chapter 2

Mating Behavior of North American Freshwater Fishes Deborah A. McLennan

Chapter 3

Petromyzontidae: Lampreys

105

Ian C. Potter, Howard S. Gill, and Claude B. Renaud Chapter 4

Dasyatidae: Whiptail Stingrays

140

Michael D. Burns, Carter R. Gilbert, and Melvin L. Warren, Jr. Chapter 5

Acipenseridae: Sturgeons

160

Bernard R. Kuhajda Chapter 6

Polyodontidae: Paddlefishes

207

Bernard R. Kuhajda Chapter 7

Lepisosteidae: Gars

243

Anthony A. Echelle and Lance Grande Chapter 8

Amiidae: Bowfins

279

Brooks M. Burr and Micah G. Bennett Chapter 9

Hiodontidae: Mooneyes

299

Eric J. Hilton, William E. Bemis, and Lance Grande Chapter 10

Anguillidae: Freshwater Eels Alex Haro

313

50

vi

CONTENTS

Chapter 11

Engraulidae: Anchovies

332

Lisa J. Hopman and Carter R. Gilbert Chapter 12

Cyprinidae: Carps and Minnows

354

Nicholas J. Gidmark and Andrew M. Simons Chapter 13

Catostomidae: Suckers

451

Phillip M. Harris, Gregory Hubbard, and Michael Sandel Literature Cited

503

Index of Scientific Names General Index

636

629

Contributors

William E. Bemis Cornell University

Lance Grande The Field Museum of Natural History

Deborah A. McLennan University of Toronto

Micah G. Bennett Southern Illinois University

Alex Haro United States Geological Survey

Ian C. Potter Murdoch University

Michael D. Burns University of Hawaii at Manoa

Phillip M. Harris The University of Alabama

Claude B. Renaud Canadian Museum of Nature

Brooks M. Burr Southern Illinois University

Eric J. Hilton Virginia Institute of Marine Science

Stephen T. Ross University of New Mexico

Anthony A. Echelle Oklahoma State University

Lisa J. Hopman Southern Illinois University

Michael Sandel The University of Alabama

Nicholas J. Gidmark Brown University

Gregory Hubbard The University of Alabama

Andrew M. Simons University of Minnesota

Carter R. Gilbert Florida Museum of Natural History

Bernard R. Kuhajda The University of Alabama

Melvin L. Warren, Jr. USDA Forest Ser vice

Howard S. Gill Murdoch University

William J. Matthews University of Oklahoma

This page intentionally left blank

Preface

The North American freshwater fish fauna comprises a little more than 1,200 native species in 50 families. It is the most thoroughly studied and largest temperate fish fauna (Page & Burr 2011) in the world. In comparison, an analysis and compendium of European freshwater fishes included 546 native species in about 24 families (Kottelat & Freyhof 2007); Europe is about one-third the land area of North America. Australia has nearly 300 freshwater fishes in 35 families (Allen 1989; Allen et al. 2002) in a land area about that of the United States (minus Alaska). This number includes many marine species that enter fresh water, and highly unusual freshwater fish lineages occur there (e.g., Salamanderfish, Lepidogalaxias salamandroides; Australian Lungfish, Neoceratodus forsteri; Nurseryfish, Kurtus gulliveri). The only other temperate fish fauna that could rival North America is Asian, but reliable information on this vast area and its fishes remains poorly understood by scientists in the New World. An estimate for the country of China stands at 1,010 native species (M. Kottelat pers. comm.). Unsurprisingly, as for many plant and animal groups, the tropical regions of the world harbor freshwater fish faunas several times larger than those of temperate regions (Lundberg et al. 2000; Berra 2007). In the mid-1970s knowledge of North American freshwater fishes was confined to a few specialists, but even so for many species (and families) little was available on natural history or ecology. In 1980, a landmark volume was published that used spot distribution maps to illustrate the ranges of all freshwater fish species in the United States and Canada (Lee et al. 1980 et seq.). That volume made available to the lay public as well as specialists a level of knowledge of the North American freshwater fish fauna theretofore unknown. Shortly thereafter a physi-

cian from Forsythe, Missouri, combined his hobby of scuba diving and snorkeling with photography and revealed, even to specialists, the incredible colors of the North American native fish fauna, especially in their brightest breeding condition, as well as some of their unique and fascinating natural histories. William N. Roston eventually traveled the continent looking for clear water and fish to photograph in their natural environment (never in aquaria). A number of his photographs are used here. During this period, numerous books focused on fish faunas of individual states (e.g., Alabama, Arkansas, California, Illinois, Kansas, Mississippi, Missouri, New Mexico, Ohio, Virginia, West Virginia, Tennessee, Washington, Wisconsin) as well as Canada, making even more detail on fishes available to the public. These works allowed for the first complete identification guide to all freshwater fishes in the United States and Canada (Page & Burr 1991, revised 2011). Nevertheless, it was not until the Freshwater Fishes of Mexico (Miller et al. 2005) was published that it was possible for us to consider editing this threevolume work on the natural history, ecology, and conservation of North American freshwater fishes. We are indebted to a large community of ichthyologists, fisheries biologists, and other workers in related fields (e.g., physiology, genetics, behavior, ecology) who have investigated the details of the lives of fishes in such a way that much technical information can now be synthesized in one place and again made available to the public and other specialists. Even though our overarching goal was to synthesize as much information as possible on North American freshwater fishes, the job of gathering information is far from

x

PREFACE

complete. In editing this work and writing synthesis chapters of our own, we were struck at once by the incredible natural history and taxonomic diversity among our native freshwater fishes but also by the large and critical information gaps that remain. Unfortunately, for many species (and nearly entire families), the syntheses presented here are (or are close to being) obituaries. For many species and groups, the biological information needed to help recover them, to slow population declines, or to prevent extinction is simply unavailable. That said, the most critical component of conserving North American freshwater fishes is the prevention of habitat loss and degradation by humans, not lack of biological information. Fishes in this fauna have an incredible tenacity for life, whether we completely understand their biology or not, but we as coinhabitants of the North American continent need to provide them the opportunity to endure. We hope that this work helps stem the high rates of population decline and extinction being experienced across the North American fish fauna. We also hope this work stimulates a whole new generation of ichthyologists and fisheries researchers to further expand our knowledge and appreciation of the natural history, ecology, and conservation of the great freshwater fish fauna of North America.

marine species in a work about freshwater fishes, but the arbitrariness reflects a biologically real gray area among fishes at the interface of saltwater and freshwater systems (e.g., “coastal” Largemouth Bass, Micropterus salmoides; Gulf Pipefish, Syngnathus scovelli; Atlantic Needlefish, Strongylura marina). Each taxonomic chapter focuses on a family or in some cases two families of North American freshwater fishes with emphasis on the natural history, biology, evolution, and conservation of each genus in the family. The sequence of the families generally follows the arrangement of Nelson et al. (2004) and Nelson (2006). In volume 1, taxonomic chapters cover the Lampreys (Petromyzontidae) through the Suckers (Catostomidae) with one exception. Because of extenuating circumstances, the chapter on Herrings (Clupeidae) will be included in a subsequent volume. Taxonomic chapters in volume 2 will cover the Characins (Characidae) through the Livebearers (Poecilidae), and volume 3, the Sticklebacks through the American Soles (Achiriidae), but we acknowledge the phylogeny of Ray-finned Fishes and Spiny-rayed Fishes by Near et al. (2012, 2013) and the expansion of that work by Betancur-R. et al. (2013) as the most comprehensive and defensible to date and present those sequences herein for North American fishes we cover (Tables P.1 and P.2).

AREA AND BREADTH OF COVERAGE The area of coverage encompasses fishes in fresh waters of North America, including Canada, the coterminous United States, and Mexico, south generally to the Isthmus of Tehuantepec. For some families, authors extended the southern boundary to include fishes of the Yucatan Peninsula region. Within the covered area, all native North American fishes, emphasizing the level of genus, are included that primarily inhabit and reproduce in fresh water and that primarily inhabit marine or estuarine systems but are frequent or even permanent components of some freshwater fish assemblages. Some primarily marine fishes are included because they are naturally established and reproduce in fresh waters (e.g., Atlantic Stingray, Dasyatis sabina; Burns et al. this volume). Others are included because young and adults penetrate deeply into freshwater systems and reproduce or are suspected of reproducing in fresh waters (e.g., Hogchoker, Trinectes maculatus), and still others simply occur with such high frequency in freshwater habitats that they likely are important functionally in those ecosystems (e.g., Striped Mullet, Mugil cephalus). Admittedly, some degree of arbitrariness was unavoidable in the inclusion or exclusion of

METHODS Sources. We encouraged authors to use only peerreviewed publications in this work, but in many cases information was only available in unpublished dissertations, theses, or even reports. We clearly indicate those unpublished sources as such in the literature cited. Scientific and common names. For scientific and common names of taxa (e.g., species, genera, families, orders), we used Nelson et al. (2004), Nelson (2006), Page et al. (2013), and occasionally FishBase (2012) as guides. Our order of presentation of families follows that of Nelson (2006), but as noted previously, Near et al. (2012, 2013) presented a phylogeny of Ray-finned Fishes (Actinopterygii) and Spiny-rayed Fishes (Acanthomorpha) based on 9–10 nuclear genes and 579 fish species. Using and expanding the Near et al. (2012, 2013) data, Betancur-R. et al. (2013) analyzed relationships of 1,401 bony fish taxa using 20 nuclear and 1 mitochondrial genes. The results clearly indicate convergence on a well-resolved phylogeny for most fishes. We accept those works as definitive (Tables P.1 and P.2) but did not follow that phylogenetic sequence due to the timing of preparation for this volume. Authors give the

Table P.1. Phylogenetic sequence, clade names, orders, and family names of Ray-finned Fishes (Actinopterygii) and Spiny-rayed Fishes (Acanthopterygii) represented in North American fresh water. The sequence follows the phylogenetic trees recovered from analysis of 9–10 nuclear genes and 579 fish species (modified from Near et al. 2012, 2013). The designation at the ordinal level of incerti ordis indicates the broader relationships of the family were undefined. Clade name

Order and family (common name)

Actinopterygii Actinopteri Acipenseriformes Acipenseridae (Sturgeons) Polyodontidae (Paddlefishes) Neopterygii Holostei Amiiformes Amiidae (Bowfins) Lepisosteiformes Lepisosteidae (Gars) Teleostei Elopomorpha Anguilliformes Anguillidae (Freshwater Eels) Osteoglossocephala Osteoglossomorpha Hiodontiformes Hiodontidae (Mooneyes) Clupeocephala Otocephala Clupeiformes Clupeidae (Herrings) Engraulidae (Anchovies) Ostariophysi Otophysi Cypriniformes Catostomidae (Suckers) Cyprinidae (Carps and Minnows) Characiformes Characidae (Characins) Siluriformes Ariidae (Sea Catfishes) Heptapteridae (Seven-finned Catfishes) Ictaluridae (North American Catfishes) Euteleostei Salmoniformes Salmonidae (Trouts and Salmons) Esociformes Esocidae (Pikes and Mudminnows) Osmeriformes Osmeridae (Smelts) Neoteleostei Eurypterygii Ctenosquamata Acanthomorpha Percopsiformes Amblyopsidae (Cavefishes) Aphredoderidae (Pirate Perches) Percopsidae (Trout-Perches) Gadiformes Gadidae (Cods) (continued) xi

xii

FRESHWATER FISHES OF NORTH AMERICA

Table P.1., continued Acanthopterygii Percomorpha Syngnathiformes Syngnathidae (Pipefishes and Seahorses) Gobiiformes Gobiidae (Gobies) Eleotridae (Sleepers) Synbranchiformes Synbranchidae (Swamp Eels) Pleuronectiformes Achiridae (American Soles) Paralichthyidae (Sand Flounders) Pleuronectidae (Righteye Flounders) Ovalentaria Atheriniformes Atherinopsidae (New World Silversides) Beloniformes Belonidae (Needlefishes) Hemiramphidae (Halfbeaks) Cyprinodontiformes Rivulidae (New World Rivulines) Goodeidae (Goodeids) Profundulidae (Middle American Killifishes) Cyprinodontidae (Pupfishes) Fundulidae (Topminnows) Poeciliidae (Livebearers) incerti ordinis Cichlidae (Cichlids and Tilapias) incerti ordinis Mugilidae (Mullets) Embiotocidae (Surfperches) incerti ordinis Gobiesocidae (Clingfishes) Unnamed clade Centrarchiformes Centrarchidate (Sunfishes) Perciformes Percidae (Perches) Cottidae (Sculpins) Gasterosteidae (Sticklebacks) Unnamed clade

Unnamed clade

complete scientific and common name in the chapters on first mention of the species and thereafter use either or both. Authors were free to deviate from these primary sources for common and scientific names for newly described species and higher taxa or because of differing or new systematic evidence (or taxonomic opinion) as well as for clarity. We capitalized the common names, if available, of all species, families, orders, and higher

incerti ordinis Moronidae (Temperate Basses) incerti ordinis Sciaenidae (Drums)

taxonomic categories (e.g., Ray-finned Fishes for Actinopterygii, Eels for Anguilliformes, Freshwater Eels for Anguillidae, American Eel for Anguilla rostrata) (see Nelson et al. 2002). We encouraged authors to use the common name in lieu of scientific names after first mention because common names are descriptive and colorful and are increasingly more stable through time than scientific names. We did not capitalize colloquial, nonstandard,

Table P.2. Phylogenetic classification of Ray-finned Fishes (Actinopterygii) represented in North American freshwaters. The sequence follows the phylogenetic tree recovered from analysis of DNA sequence data for 20 nuclear and 1 mitochondrial genes for 1,401 bony fish taxa plus 4 tetrapod species and 2 chondrichthyan outgroups representing 1,093 genera and 369 families (Betancur-R. et al. 2013). The designation at the ordinal level of incerti ordinis indicates the broader relationships of the family are undefined. Class Subclass Infraclass Megacohort

Superorder

Actinopteri Chondrostei Neopterygii Holostei

Order

Family (Common Name)

Acipenseriformes

Acipenseridae (Sturgeons) Polyodontidae (Paddlefishes) Amiidae (Bowfins) Lepisosteidae (Gars)

Amiiformes Lepisosteiformes

Teleostei Elopocephali Osteoglossocephalai

Anguilliformes Hiodontiformes Clupeiformes Cypriniphysae

Cypriniformes

Cypriniphysae Cypriniphysae

Characiformes Siluriformes

Salmoniformes Esociformes Osmeriformes Percopsiformes

Gadiformes Gobiiformes Syngnathiformes Synbranchiformes Pleuronectiformes

Cichlomorphae Atherinomorphae Atherinomorphae Atherinomorphae Atherinomorphae Atherinomorphae Atherinomorphae Atherinomorphae Mugilomorphae Blennimorphae

incerti ordinis Cichliformes Atheriniformes Beloniformes Cyprinodontiformes

Mugiliformes Blenniformes incerti ordinis incerti ordinis Centrarchiformes Perciformes

xiii

Anguillidae (Freshwater Eels) Hiodontidae (Mooneyes) Clupeidae (Herrings) Engraulidae (Anchovies) Catostomidae (Suckers) Cyprinidae (Carps and Minnows) Characidae (Characins) Ariidae (Sea Catfishes) Heptapteridae (Seven-finned Catfishes) Ictaluridae (North American Catfishes) Salmonidae (Trouts and Salmons) Esocidae (Pikes) Umbridae (Mudminnows) Osmeridae (Smelts) Amblyopsidae (Cavefishes) Aphredoderidae (Pirate Perches) Percopsidae (Trout-Perches) Gadidae (Cods) Lotidae (Cuskfishes) Eleotridae (Sleepers) Gobiidae (Gobies) Syngnathidae (Pipefishes and Seahorses) Synbranchidae (Swamp Eels) Achiridae (American Soles) Paralichthyidae (Sand Flounders) Pleuronectidae (Righteye Flounders) Embiotocidae (Surfperches) Cichlidae (Cichlids and Tilapias) Atherinopsidae (New World Silversides) Belonidae (Needlefishes) Hemiramphidae (Halfbeaks) Cyprinodontidae (Pupfishes) Fundulidae (Topminnows) Poeciliidae (Livebearers) Goodeidae (Goodeids) Profundulidae (Middle American Killifishes) Mugilidae (Mullets) Gobiesocidae (Clingfishes) Moronidae (Temperate Basses) Sciaendae (Drums) Centrarchidae (Sunfishes) Elassomatidae (Pygmy Sunfishes) Percidae (Perches) Gasterosteidae (Sticklebacks) Cottidae (Sculpins)

xiv

PREFACE

or semi-technical, but informal, common names for groups of fish (e.g., brook lamprey, shad, carp, dace, minnow, pikeminnow, shiner, chub, buffalo, carpsucker, redhorse, jumprock, piranha, bullhead, madtom, salmon, trout, pickerel, splitfin, killifish, molly, mosquitofish, bass, blackbass, crappie, darter). Fossil taxa. We indicated fossil taxa by a dagger “†” placed before the genus name. In general, we followed Walker & Geissman (2009) for designation of geologic time (period, epoch, age) in millions of years ago (mya), but the original references should be consulted to determine how the geological formations were dated or how fossil dates were estimated. Abbreviations and museum acronyms. We abbreviated standard length, total length, and fork length as SL, TL, and FL, respectively. Museum acronyms followed Leviton et al. (1985) and Leviton & Gibbs (1988). Distributional maps. We provided to authors shaded maps for each genus showing the estimated native freshwater range of the genus in North America. For genera with expansive marine ranges, the freshwater and nearshore marine range is given, not the entire marine range. Although we took care to make the maps as accurate as possible, the scale of the maps and shading of the range obviated pin-point accuracy. Also, for many fishes that have been widely introduced, the native range can only be estimated from often limited historical data.

TAXONOMIC CHAPTER OR GA NI ZATION With a few exceptions (e.g., Lampreys, Petromyzontidae), each taxonomic chapter contains 13 major sections and various numbers of subsections. For some families, little to nothing may be known about certain sectional and subsectional topics, and in those cases, the paucity or lack of information is generally acknowledged. Even in a work this large and broad ranging, we came to realize early on that some important topics could not be covered adequately when our focus was largely at the level of genus. Hence, contributors did not cover the zoogeography of species within each family. The zoogeography of North American freshwater fishes would require another volume to update and reassess information previously synthesized on that topic (e.g., Hocutt & Wiley 1986; Mayden 1992; see also Ross & Matthews, this volume). Likewise, contributors did not include tools or aids in identification of species (e.g., illustrated keys) because identification is most often a species-level exercise and it is so well covered in numerous state and regional fish

books, including Canada (Scott & Crossman 1973) and Mexico (Miller et al. 2005; see also literature guide sections in each taxonomic chapter), as well as in a field guide for North American freshwater fishes north of Mexico (Page & Burr 2011). Another important, but large topic not covered in detail is the area of fishing statistics, which again is deserving of a separate synthesis (but see commercial importance sections). We describe the content of major sections and subsections in each chapter as follows. Chapter introduction. In an initial section, authors introduce the family to the reader by relating the scientific and common names of the fishes, highlighting some specialized or unusual features of the group, and for those families with a large number of marine species or those not wholly endemic to fresh waters of North America, placing the focal taxa in context of the diversity and geographic distribution of the entire family. Diversity and distribution. Contributors summarize the general diversity of the focal family, including a discussion of each genus, its native distribution, the number of species in each genus, and evidence of polytypy or phylogeographic structure. Authors also included a non-native distribution subsection if information was available outlining introductions outside the native range and, if known, the effects of the taxa as non-natives. Phylogenetic relationships. Contributors cover all phylogenetic hypotheses (those based on cladistic methodologies) for the focal family (inter- and intrafamilial) identifying, if possible, the sister group of the family and then detailing the relationships of all genera within the family. Fossil record. Authors summarize the fossil history, if any, for the focal family. Minimally the section synthesizes information on each known fossil genus in the family and the number of extinct fossil species in each genus. Ages or approximate ages are given when known. Morphology. Contributors synthesize information on morphological structures with an emphasis on diversity in morphology across family members and specialized, unusual, and unique features. In the introductory subsection, authors describe the general physiognomy of the respective family (e.g., body shape; fin shape, type, relative size, and placement; mouth size and placement; scale type, color, and pigmentation patterns). In other subsections, authors detail unusual or specialized external and internal anatomical characteristics (e.g., reproductive anatomy, sensory organs, functional biology, and ecomorphology). Genetics. Contributors synthesize genetic-based studies focusing on topics such as karyology, phylogeography,

PREFACE

infraspecific variation, and hybridization and introgression. They also present studies employing genes or gene products to determine phylogeny in the phylogenetic relationships section. Physiology. Authors highlight the incredible diversity of physiological traits exemplified by fishes in each family. When information is available, the sections include syntheses on tolerances to and effects of water temperature, dissolved oxygen, pH, salinity, and turbidity. Contributors also highlight other specialized physiological adaptations of each family (e.g., swimming performance, sensory physiology, chemical ecology, bioenergetics, metabolism). Behavior. Contributors cover non-reproductive-associated behaviors in this section. These include behavioral areas such as aggression, dominance, learning, memory, migratory and non-migratory movement, diel activity, schooling behavior, foraging behavior, alarm signaling, patch choice, and any other unique or specialized behavior. Reproduction. Authors synthesize features of the reproductive cycle, including reproductive behaviors. The section focuses on topics such as age and size at maturity; sexual dimorphism; spawning migrations, cues, and

xv

sites; pre-spawning and spawning behaviors; male and female reproductive allocation; parental care; unusual mating systems; and embryo and larval development. Ecology. Minimally contributors focus on habitat, diet (particularly diet breadth and specializations), predation, and parasitism but also when possible range across topics from autecology to the functional importance of individual taxa in communities and ecosystems. Conservation. Authors discuss imperiled fishes in the focal family and the likely reasons for their decline. Contributors incorporate the best available information and summarize the reasons for declines or anticipated declines. Commercial importance. Contributors cover economic importance and values of taxa in the focal family. This includes the importance in historic or present commercial fisheries, cultural significance, aquaculture, sport fisheries, and the aquarium trade. Literature guide. In the final section, authors point the reader to major sources of information on the family. In particular, detailed family, species, or topic-specific treatments are referenced.

This page intentionally left blank

Acknowledgments

We acknowledge Richard L. Mayden for originating the core concept of this work and getting it underway. We thank Vincent Burke, Johns Hopkins University Press, for wise counsel, steady guidance, and encouragement on diverse matters that arose in finalizing manuscripts and coordinating with the contributors. Jennifer Malat, formerly with Johns Hopkins University Press, and Sara Cleary and Courtney Bond with the Press always promptly responded to questions concerning myriad details associated with readying manuscripts for publication. Many thanks to Rob Hopkins for designing and painstakingly preparing the distribution maps and patiently revising them to the satisfaction of the editors and contributors. Our copyeditor and indexer, Maria denBoer, was simply superb in attention to myriad details of style, consistency, and clarity. Numerous other individuals worked diligently with us to prepare this volume. We appreciate Amy CommensCarson for creating and redrawing numerous figures. Amy Commens-Carson, Eryon Maynard, Elizabeth McGuire, Gordon McWhirter, Vicki Reithel, Anthony Rietl, and Daniel Warren patiently formatted, proofed, and cross-referenced numerous drafts of a large and ever expanding literature cited as well as in-text references to tables and figures. Steve Platania and Ingo Schlupp assisted in locating photographs. Nancy Smith kindly and adeptly coordinated financial contributions, and Cathy Jenkins and Brenda Marshall provided critical logistical support. We also acknowledge Ted Leininger, Project Leader, Center for Bottomland Hardwoods Research, Southern Research Station, USDA Forest Ser vice, and Katherine Smith, Aquatic Ecologist, Office of the Chief, Research and Development, USDA Forest Ser vice, for substantial and continued support of the project.

COLOR PLATES AND PHOTOGRAPHS We used color drawings by Joe Tomelleri, the premier fish artist in North America, to illustrate as many North American genera as possible. The colors are as seen on the fishes when they are first removed from the water, and many were drawn when at their peak breeding colors. Joe’s fish portraits are done in Berol Prismacolor pencils, and using those pencils Joe is renowned for precisely portraying life color, scale and fin-ray counts, and a full spread of the fins. We express our appreciation to Joe for granting us a generous licensing arrangement for use of his drawings. For contributing or providing liberal licensing agreements of copyrighted photographs to volume 1, we are grateful to Juan M. Artigas Azas; Ginny Adams; Jeffrey Basinger, Jeremy Monroe, and Dave Herasimtschuk (Freshwaters Illustrated); William Bemis; R. J. Beamish; Heiko Bleher and Natasha Khardina (Aquapress Publishers); Bill Bonner; Bowfin Anglers Group 2011; Richard T. Bryant; Brooks M. Burr; G. Burton; Ronald R. Cicerello; Andy Dolloff; Eric Engbretson, Christopher Morey, Roger Peterson, and Paul Vecsei (Engbretson Underwater Photography); Kevin Estrada (Sturgeon Slayers); A. Ferrara; Dean Fletcher; Dennis Frates; Lance Grande; Wendell R. Haag; Eric Hilton; Jan Hoover; Gerald Jennings (Calypso Photographic Library, www.calypso.org.uk); Stephen M. Kajiura; Katie May Laumann; Larry Linton (Larry Linton Fine Art); Lance Merry (www.lancemerry.com); K. Oliveira; Kyle Piller; D. E. Scott; Garold Sneegas (Aquatic Kansas Images); P. Sorensen; David B. Snyder; Todd Stailey (Tennessee Aquarium); C. Taber (Upper Colorado River Endangered Fish Recovery Program); Matt Thomas;

xviii

AC KNOW LEDG MENTS

Uland Thomas; Michael Tobler; C. Vaughn; G. Verrault; and Tim Watts, glooskapandthefrog.org/eel%20gallery .htm. Likewise, we gratefully acknowledge Noel M. Burkhead, Stephen T. Ross, and William N. Roston for granting us and the contributors blanket permission to use any of their fish photographs.

Southeastern Fishes Council

Alaska Fisheries Science Center, National Marine Fisheries Ser vice, NOAA

FINANCIAL SUPPORTERS Desert Fishes Council

North American Native Fishes Association

Center for Bottomland Hardwoods Research, Southern Research Station, USDA Forest Ser vice Museum of Southwestern Biology, University of New Mexico Oklahoma State University Roanoke College Robert C. Cashner Southern Division of the American Fisheries Society

Research and Development, Office of the Chief, USDA Forest Service

University of Iowa University of Toronto Southeast Ecological Science Center, United States Geological Survey

Freshwater Fishes of North America

This page intentionally left blank

Chapter 1

Evolution and Ecology of North American Freshwater Fish Assemblages Stephen T. Ross and William J. Matthews

Fish assemblages in North America comprise a rich array of species with attendant diversity in morphology, physiology, behavior, ecology, life history, and range of habitats. We suspect that, among the world’s largest continental fish assemblages, the North American assemblage has the distinction of being the most thoroughly studied. Even so, much remains to be learned at even the most basic levels. Indeed, anyone who has written, or read, a regional fish book should be impressed immediately by how little is known about most species, especially in regard to the availability of information across the species’ ranges. Our goal in this chapter is to synthesize information concerning aspects of the evolution and ecology of fish assemblages in North America. To keep this effort manageable, we have focused on what we believe are some key biological elements of patterns and responses. The unifying theme is to provide an understanding of what controls the kinds and numbers of species in a local fish assemblage. We include information on general origins of North American fishes, ages of assemblages, distributions, and the responses of assemblages and species to habitat characteristics (size, quality, diversity, and variation), resource acquisition, and species interactions. Following Matthews (1998), we consider a fish assemblage to include fish species found together in a single locality over a short ecological time period, i.e., those that have a reasonable probability of encountering each other within the course of feeding, resting, movements, and so forth in a given day.

Fish Diversity Fishes make up more than half of all extant vertebrates with 27,977 named species and an average of 200 new

species described annually (Eschmeyer 1998; Nelson 2006). Of the world’s fish fauna, Cohen (1970) estimated that 41.2% were essentially restricted to fresh water. This number is remarkably close to the current number of 43%, or 11,952, recognized freshwater species worldwide (Nelson 2006). This diversity is particularly surprising given that liquid fresh water only makes up 0.0142% of the water on our planet (Shiklomanov 1993). The bulk of the world’s fresh water is unavailable as fish habitat because it is frozen (78%) or is groundwater (22%) (Horn 1972; Goldman & Horne 1983). Considering the six major zoogeographic realms (reviewed by Berra 2007), the greatest freshwater fish diversity occurs in the Neotropical realm (Central and South America and tropical Mexico) with an estimated 5,000–8,000 species, followed by the Oriental realm (India and Southeast Asia) with about 3,000 species, the Ethiopian realm (Africa and southern Arabia) with about 2,850 species, the Nearctic realm (North America except tropical Mexico) with 1,061 species, the Palearctic realm (Europe and Asia north of the Himalayas) with 552 species, and the Australian realm with 500 species, including marine fishes that enter fresh water (Burr & Mayden 1992; Matthews 1998; Lundberg et al. 2000; Moyle & Cech 2004; Berra 2007). These numbers sum to more than that given in the preceding paragraph because some regions include estimates of as yet undescribed taxa, for instance, the most recent estimate raises the number of North American freshwater fish species to 1,116 (Smith et al. 2010). The North American species are included in 201 genera and 50 families, including various marine, peripheral species (Burr & Mayden 1992).

2

FRESHWATER FISHES OF NORTH AMERICA

North American freshwater fish diversity is greatest in the southeastern region of the United States, which contains more than half (560 described species) of the freshwater fish fauna (Warren et al. 2000). Western North American fish diversity is about one-third that of overall eastern diversity, although endemism tends to be greater in the west (Moyle & Herbold 1987; Burr & Mayden 1992). In the United States and southern Canada, geographic grids of 1 degree latitude and longitude contained on average ≤10 species in western areas with maximum values of 19 in Oregon, 14 in California, and 11 along the Colorado River (McAllister et al. 1986). In contrast, the same-sized grids in the southeastern United States supported ≤73 species. Another treatment of North American diversity patterns including Mexico further illustrates this pattern (G. R. Smith in Lundberg et al. 2000; Smith et al. 2010; Fig. 1.1). Regional differences in native fish species can be appreciated more readily by comparing the native species diversity that normal sampling effort (0.75–1.0 h of seining) might yield along several hundred meters of stream. This also relates more closely to the actual richness of fish assemblages. In species-rich southeastern streams, as in Mississippi or in the Ozark Plateau, capture of 25–35 species from a medium-sized stream is not unusual (STR & WJM pers. obs.), compared with 8–12 species in central Oklahoma

200-249 150-199 100-149 50-99 25-49 1-24

Figure 1.1. Geography of fish diversity in North America and the limit of Pleistocene glaciation (solid blue line). Key shows number of species/grid (reproduced with permission from G. R. Smith).

(Matthews 1998), or 45,000 dams with a height ≥15 m are capable of storing about 15% of the world’s annual river runoff (Nilsson et al. 2005). Of the 74 large North American rivers with a virgin mean annual discharge of >350 m3/s, 39 were judged to be moderately to strongly affected by impoundments or flow regulation (Dynesius & Nilsson 1994). Of the 35 rivers judged not affected, only one, the Pascagoula River, Mississippi and Alabama, is located south of Alaska or Canada. The preponderance of lentic (lakes, ponds, artificial impoundments) versus lotic (flowing water) habitats might suggest that lake fishes would dominate the diversity of freshwater fishes in North America, and at least in some regions of the world, lake fish assemblages are diverse and may support a high degree of endemism (e.g., Fryer & Iles 1972; Echelle & Kornfield 1984; Smith & Todd 1984). In present-day North America, however, the number of species unique to lakes is much lower than those in flowing waters, primarily because of the young age of large North American lakes (G. R. Smith 1981). New World Silversides (Atherinopsidae: Menidia spp.) from Mexico’s largest natural lake, Lake Chapala on the Mexican Plateau, provide an example of a small North American species flock with 12 species either restricted to the lake or also occurring in the surrounding streams (Barbour 1973; Miller et al. 2005). The Laurentian Great Lakes also support (or supported because three taxa are now extinct) perhaps nine species of ciscoes (family Salmonidae, Trouts and Salmons, subfamily Coregoninae) (Underhill 1986; Cudmore-Vokey & Crossman 2000; Etnier & Skelton 2003). These are considered an incipient species flock (Smith & Todd 1984). The known native fish fauna of the Laurentian Great Lakes, excluding the St. Lawrence River and tributaries and including extirpated or extinct taxa, comprises 126 species (Cudmore-Vokey & Crossman 2000; Etnier & Skelton 2003), but only 5% are endemic (primarily ciscoes, Coregonus spp.) (see Cudmore-Vokey & Crossman 2000). The Great Lakes fauna is derived primarily from

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

the upper Mississippi River drainage, streams of the Atlantic Coastal Plain, and the Beringian Refugium of the Yukon Valley following retreat of the Pleistocene glaciers (Underhill 1986; Mandrak & Crossman 1992). In contrast, 17.9% of the rich lotic Tennessee River fish fauna (about 229 species and subspecies) is endemic (Etnier & Starnes 1993; Warren et al. 2000). Although North American lakes presently do not support large species flocks, this has not always been the case. Examples of now extinct North American lacustrine species flocks include the fossil Sculpin fauna (Cottidae, genera Myoxocephalus and Kerocottus) of the Pliocene (5.3–2.6 mya) Glenns Ferry Formation, Idaho (G. R. Smith 1981), and the rich, extinct semionotid fauna from the Mesozoic (251–65.5 mya) Newark lakes of eastern North America (McCune et al. 1984). Recognizing the influence that species from lotic environments have on lentic assemblages, Kitchell et al. (1977) proposed the term “River Analogy” to explain the distribution of large percid fishes (e.g., Walleye, Sander vitreus). They argued that most North American and European lakes were of recent origin (i.e., Pleistocene or later, ≤2.6 mya), that lake-inhabiting fishes had a riverine ancestry, and that pool habitats in low-gradient rivers (sloughs, oxbows) were analogous to littoral lake habitats. We suggest, like Kitchell et al. (1977), that the River Analogy applies to many other groups of lake-inhabiting fishes in North America. Further, fishes of large rivers may use habitats in new lakes (or impoundments) similarly to use of their unimpounded, large-volume habitats (e.g., Blue Catfish, Ictalurus furcatus, versus Channel Catfish, I. punctatus, in Lake Texoma, Edds et al. 2002). Thus, it is often useful to divide North American fishes on the basis of lentic and lotic habitats especially at a more local scale. On a broader scale, however, this may not be the most meaningful ecological axis along which to consider fish assemblages. We suggest that an equally meaningful axis would be upland versus lowland forms.

ORIGIN AND AGE OF NORTH AMERICAN FISH FAMILIES Fishes constituting a given assemblage often are considered to have similar origins and histories of interactions; however, this may not be true (Brooks & McLennan 1991; Matthews 1998). Contemporary species assemblages may be due to the association of the species’ ancestors in that particular geographic region, or the species may have

3

evolved among different assemblages and entered the assemblage through dispersal. For example, many fishes occurring in Arkansas have affinities with faunas to the north or northeast, but many others are more associated with faunas of the southeastern or southwestern United States. Within one watershed (Piney Creek), 43 fish species had affinities with faunas to the northeast, east, and southeast (Matthews 1998). Because of this, any assemblage is likely a mixture of species that have different origins and evolutionary ages and have been interacting for widely different periods of time. This situation is now made even more extreme by the widespread and relatively rapid introduction of non-native species (Courtenay et al. 1986; Fuller et al. 1999; Gido & Brown 1999; Gido et al. 2004) and the resulting homogenization of faunas (Rahel 2000, 2002, 2010). The mass invasion resulting from the bypassing of natural barriers that have existed in many instances since the Early or Middle Triassic (235–250 mya) represents a new and unique form of global change (Olden 2006; Ricciardi 2007). A conceptual model for natural colonization of a habitat is that of a series of filters that block potential species at various levels (Fig. 1.2; Smith & Powell 1971; Tonn et al. 1990; Poff 1997; Matthews 1998; Jackson et al. 2001; Rahel 2002). Here, we use this model as the framework for understanding the factors affecting the composition of local fish assemblages. In terms of understanding the origin of specific assemblages, one must also consider the history of lineages leading to particular species, and because of the continuing role of natural selection in conjunction with environmental change, the role of speciation and extinction in affecting the structure of a specific assemblage (see branching lineages, Fig. 1.2). Teleosts likely arose in the Middle or Late Triassic (202–234 mya) of the Early Mesozoic, and based on the fossil record, representative forms of half of the extant 40 teleostean orders were present at least by the Cretaceous of the Late Mesozoic, some 68–145 mya (Nelson 2006; Helfman et al. 2009; geologic times from Walker & Geissman 2009). Earlier, beginning in the Late Paleozoic (about 306 mya), precursors to present-day continents formed a single large, dynamic land mass, Pangaea, which persisted through the Triassic (202–251 mya) before separating by the Middle Jurassic (about 161–176 mya), forming major northern (Laurasia) and southern (Gondwana) land masses. Proto–North America included western Laurasia. Continued rifting resulted in the gradual breakup of Laurasia and Gondwana into the present-day arrangement of continents (Cracraft 1974; Briggs 1987;

4

FRESHWATER FISHES OF NORTH AMERICA Ancestral Pangaean Fish Fauna

Potential North American Fish Fauna Continental Biogeographic Filter (tectonic events, glaciation, sea level changes, stream and lake development) Potential Regional Species Pool Physiological Filter (water chemistry, temperature) Physical Habitat Filter (discharge, current speed, structural habitat, habitat size)

Decreasing Temporal & Spatial Scales

Coarse Biogeographic Filter (Laurasia, Gondwana)

Figure 1.2. A conceptual model of the formation of fish assemblages through progressive loss (i.e., filtration) and addition (i.e., speciation) of lineages. The curved arrows on either side of the figure suggest the interplay between local and regional faunas (adapted in part from Smith & Powell 1971; Tonn et al. 1990; Poff 1997; Matthews 1998; Rahel 2002; Ross 2013).

Biotic Interaction Filter (competition, predation, facilitation) Human Impact Filter Local Assemblage

Hocutt 1987). Because species ancestral to most modern lineages were present before the breakup of Pangaea, the subsequent movements of tectonic plates and their associated faunas were primary factors shaping fish assemblage composition (Fig. 1.2), although in some cases phylogenies are understood too poorly or no fossil material is available to clearly establish an area of origin.

Faunal Origins North American fish families exhibit a variety of origins, including archaic groups that originated in Pangaea (Fig. 1.3). Of 50 North American fish families (Burr & Mayden 1992), half are of marine origin. In some of these groups, the radiation into fresh water occurred early, such as the Bowfins (Amiidae) in which one subfamily (Amiinae) has occupied freshwater habitats in the Northern Hemisphere since the Late Cretaceous (about 90 mya) (Grande & Bemis 1999). The second largest group has a North American origin and includes lineages originating in Pangaea or Laurasia, followed in number by groups originating in Central and South America and Eurasia (Fig. 1.3). Although the North American freshwater fish fauna includes at least 50 families, only about half could be considered as major components based on their number of

species and/or their breadth of distribution. In addition, 90% of the extant North American freshwater fish species are contained in only 15 families. These families had their origins in Eurasia (Minnows and Carps, Cyprinidae; Suckers, Catostomidae), Central America (Livebearers, Poeciliidae; Topminnows, Fundulidae), North America including Pangean-Laurasian elements (North American Catfishes, Ictaluridae; Trouts and Salmons; Goodeids, Goodeidae; Sunfishes, Centrarchidae; Perches, Percidae), the marine environment (New World Silversides; Pupfishes, Cyprinodontidae; Sculpins; Lampreys, Petromyzontidae; Herrings, Clupeidae), and South America (Cichlids, Cichlidae).

Faunal Ages The time of occupation of North America by major fish families also varies. Of 27 families or subfamilies for which age data are available (from fossils or from phylogenies calibrated by geological events, fossils, or molecular data), Lampreys (Petromyzontidae) are by far the oldest recorded extant family, dating from the Paleozoic (Fig. 1.4). Other old groups, dating from the Cretaceous Period of the Mesozoic, are Sturgeons (Acipenseridae), Bowfins (Amiidae), Gars (Lepisosteidae), Paddlefishes (Polyodonti-

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

North America- 16

Eurasia via Beringia-2 Catostomidae1,2,3,9 Cyprinidae1,3,5,9

5

Amblyopsidae19 Aphredoderidae3,9 Centrarchidae3,9 Elassomatidae3 Goodeidae3,9 Hiodontidae3,9 Ictaluridae3,9 Percopsidae3,9,19,21

Amiidae10,14,15 Lepisosteidae21,23 Polyodontidae10,12,13,16 Umbridae3,19,21

Central America- 3

Marine- 25 Achiridae3 Anguillidae9 Atherinopsidae3,9 Belonidae3 Cottidae3,9 Cyprinodontidae7 Embiotocidae3 Engraulidae3 Gasterosteidae8 Gobiesocidae3 Hemiramphidae3 Lotidae21 Mugilidae3,18 Osmeridae3 Pleuronectidae3 Sciaenidae3,9 Syngnathidae3,9

Laurasia/Pangaea

Acipenseridae13 Esocidae9,11,19,24 Percidae?3,6 Salmonidae19,22,24

Ariidae3,18 Clupeidae3 Eleotridae3,18 Gadidae3 Gobiidae3 Moronidae9 Petromyzontidae4,9 Synbranchidae18

Fundulidae2 Poeciliidae3,9 Profundulidae20

South America- 4 Characidae9 Cichlidae 3,9 Pimelodidae 9 Rivulidae17, 20

Figure 1.3. General origins of the major families of North American freshwater fishes. Families listed as North American include those of Laurasian-Pangaean origin because of the general uncertainty in determining exact locations. References: (1) Berra (2001), (2) Briggs (1986), (3) Burr & Mayden (1992), (4) Cavender (1986), (5) Cavender (1991), (6) Collette & Banarescu (1977), (7) Echelle & Echelle (1992), (8) Foster et al. (2003), (9) Gilbert (1976), (10) Grande (1984), (11) Grande (1999), (12) Grande & Bemis (1991), (13) Grande & Bemis (1996), (14) Grande & Bemis (1998), (15) Grande & Bemis (1999), (16) Grande et al. (2002), (17) Hrbek & Larson (1999), (18) Miller & Smith (1986), (19) Moyle & Cech (2004), (20) Parenti (1981), (21) Patterson (1981), (22) Smith & Stearly (1989), (23) Wiley (1976), (24) Wilson & Williams (1992).

dae), and Pikes (Esocidae). The remaining North American families all date from within the Cenozoic. Families occupying North America since the Paleogene Period of the Early Tertiary include Ictaluridae, Percopsidae (Trout-Perches), Clupeidae, Salmonidae, Moronidae (Temperate Basses), Hiodontidae (Mooneyes), Catostomidae, Centrarchidae, Aphredoderidae (Pirate Perches), Umbridae (Mudminnows), and Cyprinidae. Families appearing in the Neogene Period of the Late Tertiary include Goodeidae, Poeciliidae, Cichlidae, Percidae, Fundulidae, Cyprinodontidae, Cottidae, Gasterosteidae (Sticklebacks), Atherinopsidae, and Sciaenidae (Drums and Croakers) (Fig. 1.4). Although the earliest percid fossils in North America date only from the Pleistocene, a fossil-calibrated molecular phylogeny of the darters (Percidae, Perches) shows the separation of darters from nondarter percids occurring 19.8 mya (Carlson et al. 2009). Within the darter genus Nothonotus, the age of the most recent common ancestor dates to 18.5 mya (Near & Keck 2005). Consequently, percids likely occurred in North America at least by the Early Miocene (about 23 mya). Seventy-eight percent of the 27 major families were pres-

ent in North America by the Early Miocene (23–16 mya) and were thus affected by numerous geologic and climatic events of the Late Tertiary. In western North America a freshwater fauna dominated by teleosts first appeared by the Late Paleocene (about 59 mya), followed by the expansion of an essentially modern fauna by the Oligocene and Miocene (Minckley et al. 1986). The western fauna during the Oligocene (23–34 mya) and Eocene (34–56 mya) shared forms with an eastern fauna including Paddlefishes (Polyodontidae), Gars (Lepisosteidae), Sturgeons (Acipenseridae), Bowfins, Trouts and Salmons, Mooneyes, Suckers, North American Catfishes, Trout-Perches, and Pikes (Esocidae) (Grande 1984; Minckley et al. 1986; Grande & Lundberg 1988; Grande 1999). The Oligocene fauna included some of the earlier fauna such as Mooneyes, Trouts and Salmons, Pikes, and also cyprinids, atherinopsids, cyprinodontids, fundulids, gasterosteids, centrarchids, embiotocids (Surfperches), and cottids (Minckley et al. 1986). Of the non-teleosts, Sturgeons are the only extant western forms, but Gars, Paddlefishes, and Bowfins are now absent from the western fauna.

FRESHWATER FISHES OF NORTH AMERICA

65

145

202

251

C ar bo ni fe ro us

56

PALEOZOIC Pe rm ia n

34

Tr ia ss ic

Pa le oc en e

23

Petromyzontidae 2, 18 Amiidae 5, 8 Esocidae 6, 20 Acipenseridae 2, 20 Lepisosteidae 2, 20 Polyodontidae 5, 7, 9 Percopsidae 2, 14 Clupeidae 4 Ictaluridae 5 Salmonidae 20 Moronidae 2 Hiodontidae 5, 20 Catostomidae 3, 5 Centrarchidae 2, 17, 20 Aphredoderidae 2, 20 Umbridae 2,14 Cyprinidae 3 Goodeidae 2, 20 Poeciliidae 10, 11, 19 Cichlidae 13, 15 Percidae 1, 2, 16, 20 Fundulidae 2, 20 Cyprinodontidae 12, 20 Cottidae 2, 20 Gasterosteidae 2, 20 Atherinopsidae 2, 20 Sciaenidae 2, 20

Ju ra ss ic

Eo ce ne

5.3

O lig oc en e

2.6

MESOZOIC

M io ce ne

0.01

Pl ei st oc en e Pl io ce ne

H ol oc en e

CENOZOIC

C re ta ce ou s

6

359

299

?

Figure 1.4. The earliest representation of major fish families in North America based on the first occurrence of fossils or from calibrated molecular phylogenies. Because the earliest fossils represent a minimal age of origin, families could be much older. Within the Cenozoic, geologic ages refer to epochs; within the Mesozoic and Paleozoic, ages refer to periods. Numbers at the top of each column are the beginning age (mya) of each geologic age or period. Numbers after families indicate sources; gaps in the fossil record are not shown. References: (1) Carlson et al. (2009), (2) Cavender (1986), (3) Cavender (1991), (4) Grande (1982), (5) Grande (1984), (6) Grande (1999), (7) Grande & Bemis (1991), (8) Grande & Bemis (1996), (9) Grande et al. (2002), (10) Mateos et al. (2002), (11) Meyer & Lydeard (1993), (12) Miller (1981), (13) Murray (2001), (14) Murray & Wilson (1996), (15) Myers (1966), (16) Near & Keck (2005), (17) Near et al. (2005), (18) Nelson (2006), (19) Webb et al. (2004), (20) Wilson & Williams (1992).

The importance of the varying ages of occupation of fish groups in North America to our understanding of fish assemblages is that the forces shaping the evolution of morphology, physiology, and behavior of species making up present-day assemblages are unlikely to be found by looking only within the contemporary assemblage. Instead, selective pressures leading to various traits may date to earlier time periods and may not even include the present-day assemblage. For instance, although feeding and morphological specializations are little changed in Pikes, the community relationships have changed a great deal since the Paleocene (56–66 mya) when Pikes were part of fish assemblages including osteoglossomorphs, percopsiforms, amiids, gonorynchids (Beaked Sandfishes), lepisosteids, asineopids (now extinct), osmerids (Smelts), clupeids, cyprinoids (possibly catostomids), and

ictalurids (Wilson & Williams 1992). Thus, major adaptations of Pikes evolved before modern predator-prey systems existed (Wilson & Williams 1992).

Tertiary and Quaternary Events and North American Fish Assemblages North American fish assemblages were shaped by largescale geologic and climatic events occurring before and during the Quaternary Period (2.8 mya–present) (Fig. 1.2), and the variation in fish assemblage composition, including species richness along both north-south and east-west gradients in North America, attest to the influence of these intracontinental factors (Fig. 1.1). Middle to Late Tertiary (34–2.6 mya) changes in landform and climate were extensive throughout North America but were par-

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

Early Pliocene (about 5 mya) (Morgan & Swanberg 1985; Sahagian et al. 2002). Of importance to our understanding of fish assemblages is that these changes, along with their impacts, extended well into the Cenozoic and were thus recent enough to have affected the flora and fauna of western North America (Minckley et al. 1986). Contiguous patterns of modern fish faunas in the West generally correspond to these continental subplates; drainages that extend across zones into adjacent subplates tend to have faunas derived from several sources (Minckley et al. 1986). The present-day Colorado River fauna is a case in point. As outlined previously, the regions of the Great Basin and Colorado Plateau (Fig. 1.5) were uplifted primarily during the Early Pliocene (5.3–3.6 mya), and extensive faulting from the Miocene (23–5.3 mya) into the Pliocene resulted in the isolation of the Colorado Plateau when northtrending streams from central Arizona were interrupted. The Early Eocene (56–49 mya) uplift of the Wasatch Front and the subsequent drop of the Great Basin in the Late Oligocene (28–23 mya) isolated the upper Colorado River fauna from that of the Great Basin. The origin of the upper Colorado fauna (including the Green and Colorado River watersheds) occurred before the Miocene in streams draining the uplifted Rocky Mountains and flowing across the Colorado Plateau to interior basins in Arizona, Colorado, and New Mexico (the Miocene Bidahochi

ticularly so in western North America (Schermer et al. 1984; Dickinson 2004). An extensive geological literature documents that present-day North America west of the Basin and Range Province (essentially comprising the western halves of Washington, Oregon, California, and northern Mexico) and north into Alaska is a conglomerate of allochthonous, accreted terranes of mostly Oceanic origin (reviewed by Minckley et al. 1986). A terrane is a fragment of crustal material formed on, or broken off from, one tectonic plate and accreted to crust lying on another plate. From the Paleozoic through the Late Miocene, these terranes were joined to the North American Craton (= Laurentia) primarily by subduction under the North American Plate (Dickinson 2004). This activity also was directly or indirectly responsible for periods of intense volcanism and orogeny contributing to the formation of the Rocky Mountain Range, the Cascade Range, and later in the Middle to Late Miocene, the Sierra Nevada Range (Schermer et al. 1984; Dickinson 2004; Mulch et al. 2006, 2008; Crowley et al. 2008; Eaton 2008). Uplift of the Colorado Plateau and formation of the Basin and Range also was related to the subduction of Oceanic plates, in this case the Farallon Plate as it slid beneath the North American Plate (Minckley et al. 1986; Eaton 2008). Although uplift of the Colorado Plateau occurred at least from the Oligocene (about 25 mya), the period of most rapid uplift took place in the

Figure 1.5. The Colorado River drainage and associated geological features (based on Minckley et al. 1986 and Gross et al. 2001).

Oregon Idaho k La eI

ho da

Sacramento R.

California Bouse Formation

N W

E S

300

in Virg

KM

300

do ora Col

R.

Colorado

boundaryy of upper Colorado R. u uppe pper and and lower lo C

Colorado C olo orado rado Plateau latteau eau Li P ttl e C ol or ad Arizona o R . Gila

0

Green R.

Wa sat ch Fro nt R.

Wyoming

tns. ky M

Hualapai Formaton

Utah

Roc

White R .

Great Basin

Bonn e ville B asin

Nevada

R.

7

Bidahochi ho Formation

New M Mexico

8 FRESHWATER FISHES OF NORTH AMERICA

Basin) or northwestern Arizona (the Miocene Hualapi Basin). A middle section of the Colorado River, currently comprising the Little Colorado, Virgin, and White Rivers, drained to the southwest. A third section, including the Gila River, was incorporated into the drainage after the retreat of the Bouse Miocene–Early Pliocene Embayment (Minckley et al. 1986). The upper and lower Colorado River systems were joined 10.6–3.3 mya via headward erosion of streams of the middle and lower Colorado watersheds and through reoccupation and reversal of flow in older channels. The Colorado River reached the Gulf of California by the Pliocene (Minckley et al. 1986). Origins of the dominant components of the mainstem Colorado fish assemblage (Colorado Pikeminnow, Ptychocheilus lucius; Humpback Chub, Gila cypha; Roundtail Chub, G. robusta; Bonytail Chub, G. elegans; Speckled Dace, Rhinichthys osculus; Razorback Sucker, Xyrauchen texanus; Bluehead Sucker, Catostomus discobolus; and Flannelmouth Sucker, C. latipinnis) can be traced to these various geological events (Minckley et al. 1986; Smith et al. 2002c). Species with relationships to the northwest, including the Sacramento–San Joaquin Basin and, more closely, the Bidahochi Lake deposits to the east, include the Colorado Pikeminnow, Roundtail Chub, Humpback Chub, and Bonytail Chub (see also Smith et al. 2002c). The Speckled Dace colonized the upper Colorado River Basin from source populations to the north and west in the northern Bonneville and Snake River drainages about 3.6 mya, and divergence of lineages in the upper and lower Colorado began 1.9–1.7 mya (Oakey et al. 2004; Smith & Dowling 2008). The Flannelmouth Sucker shows relationships to the north and west, but the Bluehead Sucker is related to forms in the Bonneville Basin to the west. Origins of the distinctive Razorback Sucker are less understood, but the divergence of the Xyrauchen lineage from that of Deltistes and Chasmistes likely occurred in the Late Miocene, if not before, and suggests a relationship to the north or northwest (Miller & Smith 1981; Hoetker & Gobalet 1999). Clearly, the main-channel fish assemblage of the Colorado River is a composite of species of different evolutionary origins and ages, and most of the changes predate Pleistocene events. In addition to geomorphic changes, the composition and distribution of western fish assemblages were shaped by a general climatic trend toward increasing aridity, resulting in the drying of large lakes and shrinking or loss of streams present during the Miocene and Pliocene. In part, the uplift of major mountain ranges contributed to expanding aridity as atmospheric circulation patterns

changed and rain shadows formed on the eastern sides of the ranges (e.g., Kohn & Fremd 2008; Mulch et al. 2008). Nowhere is the pattern of increased aridity more striking than in the western Great Basin of what is now Nevada and Utah. During the Early to Middle Pleistocene (about 650,000 years ago), this area, which is now desert, was a land of abundant and large natural lakes, especially Lake Lahontan to the west and Lake Bonneville to the east (Reheis 1999; Mock et al. 2006). As a consequence of overall drying in the region, faunas were increasingly isolated, resulting in high levels of endemism and loss of species through extinction (Hubbs et al. 1974; G. R. Smith et al. 2002). Examples include subspecies and species of desert Pupfishes (Cyprinodon spp.) that now exist in isolated spring runs as relicts from once large lacustrine systems, although the divergence times of some lineages or species extend to the Late Pliocene, substantially predating the Holocene desiccation of large lakes (Hubbs et al. 1974; Miller 1981; Smith 1981; Minckley et al. 1986; G. R. Smith et al. 2002; Echelle 2008). The impact of post-Pleistocene drought on western fishes is further illustrated by work on genetic variability in the Flannelmouth Sucker, one of the ancient, endemic species of the Colorado River (Douglas et al. 2003). Genetic diversity in the Flannelmouth Sucker is surprisingly limited for such an ancient species and is consistent with the hypothesis of a major, basin-wide population crash during a documented post-Pleistocene, severe western drought. Further, populations of this species in the upper Colorado River basin are the result of migration from refugia in the lower part of the system likely within the last 10,000–11,000 years (Douglas et al. 2003; Douglas & Douglas 2010). Fish assemblages in central and eastern North America also were affected by Late Tertiary (Miocene and Pliocene) (23–2.6 mya) geologic events. One of the most species-rich areas in North America is the Central Highlands region (Fig. 1.6). Various authors have summarized information on Pliocene and Miocene drainage patterns of this area and demonstrated that biogeographic patterns of modern fish assemblages in the Central Highlands are often better explained by these early drainage patterns, especially the Old Mississippi and Teays River systems, than by Pleistocene or Holocene (2.6 mya–present) drainage patterns (Pflieger 1971; Wiley & Mayden 1985; Mayden 1987b, 1988). Abundant biogeographical and geological data support the existence of a Central Highlands province present at least from the Eocene (56 mya) (references in Mayden 1985a, 1987ab, 1988; Wiley &

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

9

Figure 1.6. The Central Highlands region of eastern North America showing current river drainages (based on Mayden 1987a, 1988).

Central Lowlands er

Eastern Highlands

Ozark Highlands

Ouachita Highlands Re d

Ri ve r

i River

as Riv

Mississipp

Arkan s

n Coastal Plai

Mayden 1985). During the Pleistocene, southward movement of pre-Wisconsinan glaciation split the Central Highlands into Eastern and Interior Highlands. This was followed by the penetration of the lowland area connecting the Eastern and Interior Highlands by the Mississippi River, now enlarged because of southward deflection and increased flow of streams that formerly drained into Hudson Bay (Missouri River) or the Laurentian stream system and the Atlantic Ocean (Ohio River) (Pflieger 1971; Mayden 1985a; Wiley & Mayden 1985). Two principal hypotheses have been proposed to explain the high diversity of the region: the Pleistocene dispersal hypothesis and the Central Highlands vicariance hypothesis (CHVH) (Mayden 1988; Strange & Burr 1997; Near & Keck 2005). In the former hypothesis, the Eastern Highlands represented a center of origin for lineages that subsequently dispersed along glacial fronts during the Pleistocene to streams of the Interior (Ozark and Ouachita) Highlands. As such, species in the Interior Highlands should be no older than the Pleistocene. The CHVH predicts that the fauna diversified in a widespread and interconnected Highlands region during the Miocene and Pliocene and was then fragmented by Pleistocene events, after most speciation events had occurred, into the Ozark and Ouachita Highlands west of the Mississippi River and the Eastern Highlands east of the Mississippi River (Fig. 1.6). Other studies, however, indicate that understanding fish diversity in the Central Highlands is more complex than first thought (Strange & Burr

1997). Phylogeographic analyses using molecular data do show some support for predictions of the CHVH in divergence times of various lineages. The darter subgenera Litocara (genus Etheostoma) and Odontopholis (genus Percina) have species in the Ozark and Eastern Highlands and both groups show deep divergences of species between the two regions that likely occurred in the Miocene (Strange & Burr 1997). Four species of the minnow genus Erimystax, which occur in the Ozark, Ouachita, and Eastern Highlands and adjoining areas, also show Miocene speciation events (Simons 2004), and divergence within the hogsuckers (genus Hypentelium) occurred prior to the Pleistocene (Berendzen et al. 2003). Not all evidence, however, supports Miocene or Pliocene ages of species. In a study of lineage divergences in the 20 species of the darter genus Nothonotus, times ranged from the Miocene (6 events) to the Pliocene (4 events) to the Pleistocene (8 events) (Near & Keck 2005). Divergences of subspecies of the Northern Studfish, Fundulus catenatus, occurred by dispersal or peripheral isolation in the Late Pleistocene or later; divergence of subspecies of the Banded Sculpin, Cottus carolinae, perhaps by peripheral isolation, also occurred in the Pleistocene (Strange & Burr 1997), as did divergence within the Gilt Darter (Percina evides) (Near et al. 2001). In summary, the rich Central Highlands ichthyofauna seems to be a product of both vicariant and dispersal events, facilitated by the region’s great age and topographic diversity, and the high fish diversity in many ways follows predictions

10

FRESHWATER FISHES OF NORTH AMERICA

of island biogeography theory and species-area relationships (Page 1983; Near & Keck 2005). Pleistocene impacts through glaciation, changes in stream patterns caused by ice dams, flow changes, stream captures, sea level lowering, alteration of land by glacial scour, and creation of new lake habitats through terminal moraines or kettle lake formation all had major effects on fish assemblages in northern North America (Pflieger 1971; Crossman & McAllister 1986; McPhail & Lindsey 1986; Matthews 1998). Previously, four major glacial advances were recognized within the Quaternary; however, the estimate of the number of glacial advances from the dawn of the Pleistocene (about 2.6 mya) is now 18–20 for the entire planet and 13–18 for North America (Davis 1983; Ehlers 1996). The last major advance (the Wisconsinan) reached its maximum extent 8,000–10,000 years ago (Ehlers 1996; Lowe & Walker 1997). In addition, there were climatic fluctuations embedded within each of the major advances. For instance, the Wisconsinan glaciation can be subdivided into three periods of advances with the last advance starting about 23,000–25,000 years ago (Ehlers 1996). The limits of glacial advance (Fig. 1.1), defined by terminal moraines or existing waterways, extended the farthest south in the central United States, reaching across Illinois, Indiana (except the south-central region), and most of Ohio nearly to the present course of the Ohio River (Frye et al. 1965; Goldthwait et al. 1965;

Figure 1.7. The relationship of the modern Red River to pre-Pleistocene drainages. Pre-Pleistocene drainages are shown in black: 1—Plains Stream; 2—Old Ouachita River; 3—Old Red River; 4— Old Mississippi River. Ancestral drainages are based on Mayden (1987a, 1988). The closed circle shows the 1997 collecting site on the modern Ouachita River referred to in Table 1.1.

1 4

Miss

Arkansas R.

issip pi R

.

Ouachita R.

2 . dR Re

3

Wayne & Zumberge 1965; Clark et al. 1996; Lowe & Walker 1997). Farther east, ice covered upper Pennsylvania and all of New York and New England (Muller 1965; Schafer & Hartshorn 1965). Except for montane glaciers, glacial penetration was less in western states, covering the upper half of most of Washington, Idaho, and Montana and all but the southwest corner of North Dakota (Flint 1971). Higher elevations along the Rocky Mountains supported extensive glaciers as far south as New Mexico (Richmond 1965). In the far west, there were large glaciers in the Sierra Nevada Range and even in the transverse ranges of southern California (Owen et al. 2003). Pleistocene events resulted in substantial changes to earlier drainages so that faunas of present-day rivers may reflect contributions from once separate drainages. For instance, the modern fish fauna of the Red River of the South and its tributaries (Fig. 1.7) likely comprise older faunas from three distinct, pre-Pleistocene river systems: the Plains Stream in the headwaters of the Red River, the Old Ouachita River (Little-Kiamichi-Ouachita system), and the Old (lower) Red River (Mayden 1985a). The amalgamation of faunas is illustrated by examining the biogeographic relationships of 13 fish species taken in a single sample from the Ouachita River, Arkansas (STR pers. obs.; Table 1.1). Five species are endemic, or largely so, to all three regions of the Central Highlands and thus would have had the potential for interaction since the Pliocene

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

11

Table 1.1. Biogeographic relationships of species from a sample of fishes (STR pers. obs.) from the Ouachita River, Arkansas, at the confluence with the Little Missouri River. (1) Found in all Central Highlands (some with disjunct populations in Central Lowlands); (2) endemic to Ouachita Highlands; (2a) Ouachita Highlands and various adjoining regions; (3) widespread, primarily lowland species with sister-species found in Central Highlands (i.e., cladogenesis likely before uplift of Central Highlands); (4) widespread but biogeographically non-informative species.

Taxa

Origin/PrePleistocene Distribution

Highland Stoneroller, Campostoma spadiceum Blacktail Shiner, Cyprinella venusta Steelcolor Shiner, Cyprinella whipplei Redfin Shiner, Lythrurus umbratilis Bigeye Shiner, Notropis boops Bullhead Minnow, Pimephales vigilax Creole Darter, Etheostoma collettei Orangebelly Darter, Etheostoma radiosum Speckled Darter, Etheostoma stigmaeum Redspot Darter, Etheostoma artesiae Mountain Madtom, Noturus eleutherus Banded Darter, Etheostoma zonale Channel Darter, Percina copelandi

or earlier. Four species are primarily restricted to the Ouachita Highlands and perhaps had a later origin. The remaining five species are widespread, generally lowland forms, some of which are sister species to forms occurring in the Central Highlands. Clearly, even in this example of one sample of fishes, evolutionary origins, ecological histories, and ages of the taxa are different with the assemblage including groups fragmented from a once intact prePleistocene Central Highlands fauna, more recent taxa endemic to the Ouachita Highlands, and components derived from generally widespread, primarily lowland, prePleistocene taxa. Such separate origins have substantial consequences for the interpretation of factors like coevolution of species’ traits (see later treatment herein). Farther west, in addition to portions of the Missouri River that originally flowed northward into Hudson Bay, the Bonneville Basin was also likely once part of the Hudson Bay drainage during the Late Miocene via the upper Snake River (G. R. Smith 1981; Crossman & McAllister 1986). The connections are reflected in the current fish faunas where, for instance, the Bonneville drainage (located primarily in Utah) contains faunal elements from the north and northeast such as Prosopium spp. (whitefishes), Catostomus spp. (Suckers), and the cyprinid genera Richardsonius and Rhinichthys (G. R. Smith 1981). Glacial advances and retreats also impacted fish assemblages directly through extirpation or displacement into

2 3 1 4 1 4 2 2 3 3 1 1 1

References Mayden (1987a); Blum et al. (2008); Cashner et al. (2010) Mayden (1987a) Mayden (1987a) Mayden (1987a) Wiley & Mayden (1985); Mayden (1987a) Mayden (1987a) Mayden (1985a) Page (1983); Mayden (1985a, 1987a) Page (1983); Simon (1997) Mayden (1985a); Piller et al. (2001) Mayden (1985a, 1987a) Page (1983); Mayden (1988) Mayden (1987a)

glacial refugia, followed by subsequent colonization of newly available habitats when glaciers retreated. At least five major glacial refugia, as well as various minor refugia, allowed the survival of organisms displaced by advancing ice (Fig. 1.8) (Flint 1971; Crossman & McAllister 1986; Cox & Moore 1993; Stamford & Taylor 2004). As a consequence, some northern fish assemblages have only been formed within the last 10,000 years and colonization of once glaciated areas is an ongoing process (Crossman & McAllister 1986; Lundberg et al. 2000). For example, species richness in formerly glaciated areas, as shown for Ontario, Canada, is related strongly to distance from glacial refugia and the time that colonization corridors have been free of ice (Mandrak 1995). In central North America, the majority of colonizations of once glaciated areas occurred via the Mississippi Refugium (Fig. 1.8), contributing species to north-central Canada, the Hudson Bay drainage, and the Arctic Archipelago (Mandrak & Crossman 1992; Matthews 1998). Across Ontario, Canada, which was totally covered by the Wisconsinan glacial advance, 77 of 91 species for which glacial refugia are resolved colonized from the Mississippi Refugium, 18 from the Atlantic Refugium, and 2 from the Missouri Refugium (Mandrak & Crossman 1992). Most of the Ontario species (86) for which refugia could be identified survived the glacial advance in a single refugium, and only 5 had multiple refugia. Of the 21 common species limited to the Great Lakes and Nelson River (Hudson Bay

12 FRESHWATER FISHES OF NORTH AMERICA

Figure 1.8. Glacial refugia during the Wisconsinan glacial advances and pathways of recolonization. Smaller refugia are indicated by closed circles. Based on data from McPhail & Lindsey (1970, 1986), Mandrak & Crossman (1992), Matthews (1998), McCusker et al. (2000), Smith et al. (2001), and Stamford & Taylor (2004).

Beringia

Nahanni Queen Charlotte Islands Banff-Jasper

Cascadia Missouri

Mississippi Atlantic

drainage) watersheds, 14 originated from the Mississippi Refugium, 1 species originated from both the Mississippi and Atlantic Refugia, 1 species originated from the Atlantic Refugium, and 1 species originated from the Atlantic, Mississippi, and Missouri refugia (Fig. 1.8; Mandrak & Crossman 1992). Whether assemblages tended to move as groups of species or as individual species is unknown, although colonization likely occurred in waves of immigrants, when passageways from various refugia became free of ice. Recolonization of New England and southeastern Canada by fishes from the Atlantic Refugium likely occurred initially via drainages in Connecticut because these drainages were the first to be fully free of ice, and a series of proglacial lakes and rivers forming within the Connecticut River Valley (extending from Connecticut through Massachusetts, New Hampshire, and Vermont) would have provided suitable habitats, dispersal routes, and subsequent access to more northern habitats. Gla-

cial Lake Connecticut occupied the area of Long Island Sound from about 19,000–15,500 years ago, and Glacial Lake Hitchcock occupied the Connecticut River Valley about 18,000–12,000 years ago (Poppe et al. 2000; Benner et al. 2009). Based on trace fossils (e.g., tracks and traces in bottom sediments made by invertebrates and fishes), early recolonizers likely included species of Salvelinus and Cottus, which were present in Glacial Lake Hitchcock by 13,700 years ago (Benner et al. 2008, 2009). The Eastern Blacknose Dace (Rhinichthys atratulus) also recolonized proglacial habitats. The species initially recolonized eastern drainages in Connecticut (including the Connecticut River drainage) through a single founding event from a single Atlantic refugium in the early stages of deglaciation. Some 9,000 years later, a second colonization occurred in the more western Housatonic River drainage of Connecticut but not in the other two more eastern drainages (Connecticut and Thames Rivers). Consequently, Eastern Blacknose

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

sheds that trended in a north-south direction, fish species also showed this pattern. Southward displacement likely was aided by a combination of increased stream discharge and changes in the character of streams in addition to cooler temperatures (Cross 1970; G. R. Smith 1981; Cross et al. 1986). For instance, species that today have a primarily northeastern or northcentral distribution, such as the Redbelly Dace (Chrosomus erythrogaster), Northern Studfish, and Rainbow Darter (Etheostoma caeruleum), have disjunct populations as far south as Mississippi (Ross 2001) and the Redbelly Dace and Creek Chub (Semotilus atromaculatus) have disjunct populations in northeastern New Mexico (Pflieger 1971). A general pattern of reduced species richness is associated with glaciated areas (Fig. 1.1); however, in some cases isolation stemming from glacial activity resulted in increased rates of speciation (e.g., Bernatchez et al. 1996). For instance, the Lake Whitefish (Coregonus clupeaformis) diverged into three genetically distinct races during isolation in the Beringia, Mississippi-Missouri, and Nahanni glacial refugia (Foote et al. 1992). In addition to coregonines, speciation in formerly glaciated areas occurred in other groups of fishes, including Smelts (Osmerus spp.), Sticklebacks (Gasterosteus spp.), whitefishes (Prosopium spp.), and chars (Salvelinus spp.) (Schluter & McPhail 1993).

Dace populations in the three major drainages of Connecticut are derived from at least two refugia and differ greatly in how long they have occupied the region (Tipton et al. 2011). In western North America, four refugia (Beringia, Cascadia = Pacific, Mississippi, and Missouri) contributed to the formation of present fish assemblages. Times of faunal movement out of these refugia differed because of an earlier retreat of ice from coastal refugia and from the Missouri Refugium of the Great Plains compared with the Mississippi Refugium (McPhail & Lindsey 1970). Fish faunas of six hydroregions of Alaska each contain immigrants from the four principal refugia, although the contribution of the southwest Cascadia Refugium decreases from south to north and that of the Beringia Refugium decreases from north to south (Fig. 1.9; Morrow 1980; Oswood et al. 2000, using data from McPhail & Lindsey 1970). These examples suggest that fish assemblages in formerly glaciated regions experienced a step-like increase in colonizers over time as passage from the various refugia became possible and that western assemblages, like their northern counterparts, often contain species from multiple refugia. Cooling associated with the Pleistocene resulted in a general southward displacement of terrestrial plants and animals outside of the areas of direct glacial impact (Pflieger 1971; Whitehead 1973; Pielou 1991). In water-

Figure 1.9. Contributions of three glacial refugia to the fish faunas of Alaskan hydroregions. Beringia—black; Cascadia—light gray; upper Mississippi—white (modified with permission from Oswood et al. 2000, Journal of the North American Benthological Society 19:405–418; copyright 2000 North American Benthological Society).

4% 3%

15%

Arctic 24%

81%

73%

Northwest 6% 23%

Yukon 71%

7%

Southcentral 63%

30% 3%

Southwest 59%

38%

13

Southeast 4% 52% 44%

14 FRESHWATER FISHES OF NORTH AMERICA

RESPONSES OF FISH ASSEMBLAGES TO LOCAL AND REGIONAL EFFECTS We have shown that regional fish faunas are products of various and complex historical events, and that fish assemblages, like those of other biota (Jablonski & Sepkoski 1996), have undergone continual cycles of breakup and rearrangement over geological time. In this section we examine how the regional environmental characteristics, the regional fauna, habitat type, and habitat quality influence local fish assemblages. In doing so, we are making the transition from the realms of biogeography and evolutionary ecology to that of community ecology (see Keddy & Weiher 2001). Fish assemblages are influenced by factors operating at the local scale (e.g., physical habitat, predators, competitors) and by regional effects including climate, elevation, geographic location, and the regional fish fauna of which they are part (Fig. 1.2). Hugueny et al. (2010) proposed a broad framework by which to view fish communities, integrating across scales from historical and biogeographic factors to interactions among species in local communities and including regional synchrony in community changes. To understand relationships of species or functional groups to environmental variables, one must recognize that the presence of a particular kind of fish in a habitat may vary spatially and temporally on annual, seasonal, or daily scales and that such variation often is compounded by changes in life history stage. Within North American freshwater fishes, the duration spent in a particular habitat can range from species that remain in the same general habitat throughout their entire lifespan, as in certain riffle-inhabiting darters such as the Orangebelly Darter (Etheostoma radiosum) (Scalet 1973) and Mottled Sculpin (Cottus bairdi) (Petty & Grossman 1996), or the extreme case of the Devils Hole Pupfish (Cyprinodon diabolis), which is restricted to Devils Hole, a 3 m × 20 m pool in the Death Valley system (Miller 1948; Deacon & Williams 1991), to those that move seasonally among habitats for purposes of reproduction, such as spawning migrations of catostomid fishes from large rivers into headwater streams (e.g., Curry & Spacie 1984), to long-distance migration hundreds or thousands of kilometers shown by diadromous fishes such as Sturgeons, Herrings, and Freshwater Eels (Anguillidae) (Heise et al. 2005; Helfman et al. 2009). Although we use it here in a general way, the term “habitat” has various meanings and the significance of

these meanings has garnered considerable debate (e.g., Ryder & Kerr 1989), generally relating to the distinction between habitat and environment. A current view is that habitat comprises the localized structured component that acts as a template for organisms, and environment is the sum of the biotic and abiotic surroundings, including habitat and other organisms (Peterson 2003). Habitat includes both static (i.e., substratum characteristics) and dynamic (i.e., water-column characteristics) components, and the extent of suitable habitat is defined by the degree of overlap between suitable dynamic and static components (i.e., suitable static habitat alone is not sufficient if suitable dynamic habitat does not also occur) (Peterson et al. 2007).

Local and Regional Environmental Effects on Assemblages Various models have been proposed relating large-scale, regional factors (e.g., geology, climate, rainfall, elevation) to the primary structure of assemblages (e.g., species presence, relative abundance) or to emergent assemblage structure (e.g., species richness, diversity, assemblage complexity, trophic relationships) (Marsh-Matthews & Matthews 2000). A recent evaluation of temporal changes in fish assemblages of three large Great Plains river basins emphasizes the importance of understanding the comprehensive and comparative impacts of broad regional factors, such as groundwater withdrawal, sedimentation, habitat fragmentation, and invasive species, if long-term changes in local and regional fish communities are to be fully appreciated (Gido et al. 2010a). Approaches relating local habitat features to fish assemblages are treated in the subsection on habitat type and quality. Three conceptual models historically used to predict emergent structure in lotic assemblages from basic ecological principles are the habitat template (Southwood 1977, 1988), landscape filters (Poff 1997), and the river continuum concept (Vannote et al. 1980) (see summary by Goldstein & Meador 2004). Frimpong & Angermeier (2010) suggested that incorporating traits of individual species, a trait-based approach, can profitably combine knowledge about the basic biology of individual species with environmental conditions to provide a robust view of success of individual species in local habitats or the composition of a local assemblage or functional groups. Other traits, in addition to those internal to a fish species, might strongly relate to the role of that species in an ecosystem. For example, McIntyre & Flecker (2010) highlighted the

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

fact that fish species differ markedly in the elemental composition of their bodies (stoichiometry) and thus, in the nutrient ratios of their waste products, making species identity in an assemblage an important factor in the effects of those fish on the ecosystem. In the habitat template model, the habitat is suggested as a template providing a predictive pattern for the evolutionary assembly of communities and life history traits thereof, much like the periodic table of elements in chemistry (Southwood 1977, 1988). An important assumption is that current organismal traits match current environmental conditions, which is not necessarily the case (see origin and age of North American fish families section). To test predictions of the habitat template model, Townsend & Hildrew (1994) used a large data set from the River Rhône drainage, France. They focused on two axes, temporal habitat heterogeneity (a measure of the frequency of disturbance) and spatial heterogeneity (a measure of the availability of refugia), in developing predictions of how species traits would respond to the habitat template. For instance, the species trait of body size should decrease in environments with low spatial heterogeneity and high temporal heterogeneity (i.e., unstable environments) and be large or small in environments with high spatial heterogeneity and low temporal heterogeneity (i.e., stable environments). Similarly, lifespan should be short in unstable environments and long or short in stable environments. Tests of these and other predictions based on 13 taxonomic groups of plants and animals from the Rhône River drainage resulted in only mixed support for the habitat template model and support for fishes was totally lacking (Resh et al. 1994). The large data set used to test the predictions might have had methodological limitations that precluded a fair test of the model (Resh et al. 1994). Studies of U.S. midwestern streams (Poff & Allan 1995) and comparisons of functional convergence between European and eastern North American fish assemblages (Lamouroux et al. 2002) offer somewhat stronger support for the habitat template model. For instance, two predictions of the habitat template model, that variable habitats should contain more resource generalists and that nonvarying habitats should contain more specialists, was supported for midwestern U.S. stream fish assemblages (Poff & Allan 1995). Variables used to characterize habitat variability included flow predictability and variation, base flow stability, and frequency of spates (Poff & Allen 1995). Ecological traits of species, including body size, longevity, fecundity, water-column position, body shape, and swim-

15

ming ability responded similarly to the physical habitat template, determined by Froude number (ratio of current velocity and water depth), in France and Virginia (Lamouroux et al. 2002). Even though this shows predictive ability of the habitat template model, the amount of explained variation was generally 0, and the premise that local richness cannot logically exceed regional richness and so should originate at 0 (indicated by the dotted line in Fig. 1.10). At a local scale, however, diversity in pools was related strongly to diversity at collecting sites (a site included three or more habitat units such as pools or riffles), indicating that the local habitats (pools) were not saturated (Fig. 1.10). The number of local introduced species was related positively to the regional number of introduced species at all regional scales, and in contrast to native species, showed no evidence of saturation. In addition, the number of native fish species did not influence the number of nonnative species, suggesting that high native fish diversity does not preclude invasion by non-native fishes. The strong influence of regional compared with local factors also is evident in lakes. Fish faunas of watersheds within the Laurentian Great Lakes were impacted by the effect of large-scale regional processes reflective of postglacial dispersal or climate but were much less related to measures of environmental similarity (e.g., lake depth, area, and pH), although such factors likely have some role in affecting species composition (Jackson & Harvey 1989). In a comparison of small lakes in Wisconsin and Finland, species richness in individual lakes was related to regional species richness, but local richness reached an asymptote, suggesting that individual lake faunas become saturated with species (Tonn et al. 1990; Fig. 1.10).

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

80 70 60 50

2

R = 0.10

40 30 20

25 20 15

10

20

30 40

50

60

2

R = 0.57

10 5

10 0 0

Site level

30

Local: Richness in Pools

Local: Richness at Sample Site

Virginia Streams Drainage level

0

70 80

Local Species Richness

30

Regional: Richness at Site

Regional: Richness in Drainage Basin

Wisconsin Lakes

7

20

10

0

6

reg

5

a i on

t en chm i r l en

17

Figure 1.10. The relationship between native fish diversity of local assemblages to regional fish diversity in Virginia streams at the drainage and local scales (based on Angermeier & Winston 1998) and Wisconsin lakes (based on Tonn et al. 1990). Dashed lines indicate a hypothetical direct relationship between regional and local diversity; for Virginia streams, solid lines indicate actual relationships between regional and local diversity; dotted lines indicate extrapolation of local diversity to 0. The closed circle and vertical line indicate the mean and 95% confidence interval of local species richness for Wisconsin lakes. See text for further explanation.

4 local saturation

3 2 1 0 0

10

200

30

40

50

60

70

Regional Species Richness

Regional factors alone, however, could not explain local species composition because biotic factors, particularly the presence of large predators, also influenced species composition. Predator composition, lake morphometry, and winter oxygen levels also affected the structure of fish assemblages in small Wisconsin lakes (Tonn & Magnuson 1982). These studies all suggest a general, but highly variable, link between regional and local species richness. In contrast, in a study of the Interior Highland region (Ozark and Ouachita Mountains, Arkansas, Oklahoma, Missouri, and Kansas), regional (river basin) fish species richness accounted only marginally for species richness at local sites over all species, and the regional-local species richness relationship was nonexistent within the Cyprinidae and Percidae. Overall, the number of species in local assemblages varied greatly within basins, suggesting a lack of strong regional-local effects and the greater influence of local physical or biotic factors on local diversity (Matthews & Robison 1998). Similarly, for midwestern stream fishes (65 sites, 13 drainages, Nebraska and Iowa south to Texas), local factors affected species richness more than the overall size of the regional species pool (MarshMatthews & Matthews 2000). Even so, in contrast to

emergent assemblage properties (i.e., species richness), primary assemblage structure (i.e., the occurrence of particular species) was influenced strongly by broad geographic factors, primarily latitude, reflective of the fact that many species have restricted north-south distributions (Conner & Suttkus 1986; Cross et al. 1986). Similarly, in Texas stream fish assemblages, regional environmental factors and the regional species pool were important in affecting species composition of local assemblages (Hoeinghaus et al. 2007); however, functional group response was influenced more strongly by local environmental and biotic factors. Regional and historic filters (see Tonn et al. 1990) clearly can have a major influence on local assemblages and in some cases, especially southeastern streams and northern lakes, the richness of local fish assemblages is affected strongly by regional diversity. Species composition also can be greatly influenced by large-scale factors such as latitude, zoogeographic region, divisions between major river basins, and stream size (e.g., Swift et al. 1986; Hitt & Angermeier 2011). Nevertheless, not all assemblages show a relationship between regional and local diversity, as evidenced by harsh midwestern streams and speciose upland streams.

18

FRESHWATER FISHES OF NORTH AMERICA

Effect of Habitat Type and Quality Current velocity, water depth, water quality, bottom type, food availability, and structure are all important factors affecting habitat selection by fishes and thus have substantial impacts on assemblage structure. Within limits set by the regional species pool, the occurrences of particular species, and thus the composition of a local assemblage, are dictated to a large degree by the type and quality of the local habitat and the surrounding environment, including riparian zones. At the level of stream reaches (i.e., lengths of streams including several riffle-pool sequences), hydrologic variables, and in particular Froude number, explained ≤50% of the variance in functional traits of stream fish assemblages in both eastern North America and Europe (Lamouroux et al. 2002). In the upper Red River, Oklahoma, predictability of fish assemblages also was based on environmental gradients; however, in this physically harsh, often saline system the conductivity gradient had the most predictive power (Taylor et al. 1993). For midwestern stream fishes, local aquatic habitat variables explained a small but significant amount of variation in species richness (14%) and assemblage complexity (15%) (Marsh-Matthews & Matthews 2000). On a finer scale, species differences also occur between shallow runs or riffles and deeper pool habitats (Schlosser 1987) with the differences more pronounced in large versus small streams (Taylor 2000). Schlosser (1987) developed a conceptual framework of processes affecting fish assemblages along a gradient of pool development, habitat volume, and habitat heterogeneity. He proposed that fish assemblages in upstream, shallow areas are driven primarily by high variability in physical factors such as droughts and floods, but assemblages in downstream areas containing environmentally complex, deep pools are less variable and driven more by biotic interactions. In the Little River, Oklahoma, patterns of faunal similarities among 74 stream sites were different between adjacent pool and riffle assemblages, suggesting that the 2 assemblages responded differently to environmental factors (Taylor 2000). Pools almost always had more species than riffles, but riffle assemblages tended to be more similar to pool assemblages in smaller streams. In an upland stream reach of the Illinois River, Oklahoma, most species shifted from backwater pools early in life to pools or riffles as they increased in size (Bart 1989). Riffles were not used to any great extent by young (only 30% of riffle species occurred in the riffles as young). In an even smaller head-

water stream in the same drainage (Gelwick 1990), more species occurred in pools (21) than in riffles (11), and only 3 species (Slender Madtom, Noturus exilis; Fantail Darter, Etheostoma flabellare; and Banded Sculpin) were exclusive to riffles. Pool and to a lesser extent riffle assemblages showed some changes longitudinally over the 6 km stream section. In apparent contrast with studies in larger streams, riffles seemed to function more as supplemental habitats or as refuges from predation for juvenile individuals of taxa found in pools. Riffle taxa shifted into pools during droughts or floods (Gelwick 1990). Thus, the use of pool and riffle habitats by fishes varies relative to stream size and hydrologic conditions. Sizes of fishes also vary among riffles, runs, and pools with larger individuals of the same species, and also larger species, occupying pools (Mahon & Portt 1985). This is part of the overall phenomenon of bigger fishes being present in deeper habitat that is demonstrated for stream fishes on a variety of scales and locations (e.g., Power 1987; Gorman 1987; Harvey & Stewart 1991; see predation subsection). These and other studies show that some species are restricted to riffle habitats. An example of a specialized riffle inhabitant is the Bayou Darter (Nothonotus rubrum) of Mississippi, which shows strong selection not only for riffle habitats but also for riffles with certain characteristics. Bayou Darters occur in shallow riffles characterized by current speeds averaging 79 cm/s and having a coarse (mean particle size 16–32 mm), firm substratum (Ross et al. 1990, 1992). Individual fish are rarely encountered outside of favorable riffle habitats, although larval stages do move downstream in the drift (Slack et al. 2004). In the winter, selection for coarse structure in riffles increases when fish are energetically constrained to seek out refuges from high current speeds and preferentially choose larger over smaller refuges (Ross et al. 1992). Further, as riffle habitats are created (in this case by headward erosion), the Bayou Darter has expanded its range into more upstream reaches (Ross et al. 2001). Even within riffles, fish species may use habitats quite differently, as shown by studies of habitat partitioning among riffle-dwelling fishes. For example, five species of darters within the genus Etheostoma (Greenside Darter, E. blennioides; Rainbow Darter; Orangethroat Darter, E. spectabile; Missouri Saddled Darter, E. tetrazonum; and Banded Darter, E. zonale) differed in their occupation of riffle habitats in streams of the Ozark upland region, primarily along a gradient of association with submerged and emergent vegetation and less so on the basis of substratum size, water depth, or current speed. Orangethroat Darters

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

typically occurred on riffles lacking vegetation, but Rainbow Darters were associated strongly with emergent vegetation. Species also segregated relative to stream size with Missouri Saddled Darters found more in riffles of large than small streams (White & Aspinwall 1984). Within a single riffle in the Roanoke River, Virginia, microhabitats of three darter species (Fantail Darter; Riverweed Darter, Etheostoma podostemone; and Roanoke Logperch, Percina roanoka) showed distinct differences in water flow and depth of microhabitats (Matthews et al. 1982a). The differences in current speeds selected in the riffles corresponded to differences in morphology, behavior, and the ability to tolerate exposure to flow for two of the species, the Fantail Darter and Roanoke Logperch (Matthews 1985). Underwater observations are used to study habitat use of fishes within a short stream reach, usually consisting of a pool, run, and riffle. For instance, in a fish assemblage in a southern Appalachian stream, benthic and water-column fishes generally used habitat non-randomly (Grossman & Ratajczak 1998). The habitat gradient along which fishes primarily differed contrasted high-velocity, erosional areas with low-velocity, depositional areas. In addition, some species shifted in microhabitat use both seasonally and ontogenetically. Large individuals of water-column species, such as the Warpaint Shiner (Luxilus coccogenis), River Chub (Nocomis micropogon), Rainbow Trout (Oncorhynchus mykiss), and Creek Chub, tended to occupy deeper microhabitats than small individuals. Large individuals of benthic species, such as the Longnose Dace (Rhinichthys cataractae) and Mottled Sculpin, occurred closer to shelter and in higher current velocities than did small individuals. In wide pools of Baron Fork, Oklahoma, small, juvenile Central Stonerollers (Campostoma anomalum pullum) are restricted to shallow pool margins, but large individuals occur in deep water, mid-pool areas (WJM pers. obs.). In a montane stream in Idaho, both Cutthroat Trout (Oncorhynchus clarkii) and Bull Trout (Salvelinus confluentus) were non-random in macrohabitat use, selecting pools over riffles (Bonneau & Scarnecchia 1998). Non-random habitat use also is well documented in lentic systems, where water depth and distance from shore, submerged aquatic vegetation, and vertical water-column position are often important axes of separation (Moyle 1973; Werner et al. 1983ab; Ross 1986; Benson & Magnuson 1992). In small Michigan lakes, species differed primarily by habitat (Werner et al. 1977). The shallow, vegetated littoral zone was used primarily by juvenile centrarchids, such as small Bluegills (Lepomis macrochirus) and various cyprinids, but larger Bluegills, Black Crappies (Pomoxis ni-

19

gromaculatus), and Largemouth Bass (Micropterus salmoides) occurred in deeper, more open areas. Fishes also segregated vertically. For example, Blackchin Shiners (Notropis heterodon) used the upper water column, but Blacknose Shiners (Notropis heterolepis) were more associated with the bottom. Similarly in Florida lakes, the fish assemblage responded strongly and positively to the location and type of submerged aquatic vegetation, and fishes were uncommon outside of the vegetated areas. In addition, Bluegills tended to increase in size with increasing water depth, although Largemouth Bass, which were concentrated just outside the vegetated areas, did not show size increases with depth (Werner et al. 1978). Fish species also segregated vertically within the water column, a pattern documented in other studies of lake fishes (Keast & Fox 1992). The use of habitats by fishes in both lentic and lotic habitats may vary temporally over a 24-h period and also seasonally (including shifts due to different life history stages). Studies of diel shifts in habitat use are more common in lakes than in streams and often show regular patterns of movement of fishes into and out of specific habitats, but studies of seasonal changes in habitat are more common in streams. For instance, all sizes of Bluegills and small Largemouth Bass in a northern lake tended to move inshore and higher up in the water column at dusk (Werner et al. 1977). In another northern lake, striking changes also occurred in abundance of fishes in particular habitats between day and night samples, and the patterns varied somewhat by month. The changes were not so much shifts in assemblage composition as shifts in relative abundances. In that study, the night sampling occurred between 2200 and 2400 h and divers used lights to locate fishes (Keast et al. 1978). Bluegills tended to move to more exposed areas to forage; in addition, individual Bluegills and Pumpkinseeds (Lepomis gibbosus) were observed motionless on the bottom in shallow water (6); latitude and longitude are from about the midpoint of the study area; spatial scale, if not stated, was estimated from map of study area. Spp. = Number of species analyzed; HD = Human disturbance; L = Low stress; P = Persistence; S = Stability; Cj = Jaccard Coefficient; PSI = Proportional Similarity Index; CV = Coefficient of Variation (proportion); Im = Morisita’s index of similarity. Stations

Potential Stressor and Categorization of Stress

Site

Habitat

Spp. (no.)

Ball Creek, NC

Small stream

1

Spring flooding; late spring to autumn drought (generally low stress)

3

Ball Creek, NC

Small stream

1

Spring flooding; late spring to autumn drought (generally low stress)

4

Coweeta Creek, NC

Medium stream

1

Spring flooding; late spring to autumn drought (generally low stress)

5

Cedar Fork Creek, OH

Medium stream

1

Annual flooding (moderate stress)

30

Undisturbed streams, Savannah River site, SC

Small stream

9

Annual variation in flow (low stress)

15a

Martis Creek, CA

Small stream

4

Periodic flooding; non-native predators (moderate stress)

7

Authors’ Conclusions Temporal persistence of resident species; moderate temporal stability of resident species based on numbers of individuals (mean CV = 0.53) Temporal persistence of resident species; temporal stability of resident species based on relative abundance; moderate to low temporal stability based on numbers of individuals (mean CV = 0.75) Temporal persistence of resident species; temporal stability of resident species based on relative abundance; moderate to low temporal stability based on numbers of individuals (mean CV = 0.62) Temporal persistence; temporal stability indicated by consistency in rank order data; high variation in numbers of individuals Temporal persistence overall; temporal stability based on rank order data and similarity analyses; moderate stability based on CV of 0.44 Temporal persistence overall; temporal stability based on rank order data; number and biomass data showed high variation (last sample in 1983)

HD

L

High P

High S

N

Y

Y

Y

Freeman et al. (1988)

N

Y

Y

Y

Freeman et al. (1988)

N

Y

Y

Y

Freeman et al. (1988)

N

Y

Y

Y

Meffe & Berra (1988)

N

Y

Y

Y

Paller (2002)

N

Y

Y

Y

Moyle & Vondracek (1985)

References

Black Creek, MS

Medium stream

5

Annual overbank flooding (low stress)

25

Piney Creek, AR

Medium stream

5

Periodic flooding (low stress)

10

Pearl River, MS

Large river

8

Periodic flooding; upstream impoundments (generally low stress)

28

French Creek, NY

9

Normal seasonal variation in flow and temperature (low stress) Annual high flows (low stress)

41

Kiamichi River, OK

Small to medium stream Medium stream

Coweeta Creek, NC

Medium stream

1

Annual variation in flow, including droughts (generally low stress)

16

Otter Creek, IN

Medium stream

1

Upstream mill dam; no other major impacts (low stress)

18

Martis Creek, CA

Small stream

4

Periodic flooding; severe spring flood in 1983; nonnative predators (generally low stress)

5

6

10

Temporal persistence; temporal stability based on numbers of individuals and rank order data Temporal persistence overall; temporal stability overall based on rank order data and similarity analyses; temporal stability at individual stations based on rank order data Temporal persistence overall; temporal stability overall based on similarity analyses; high variation in numbers of individuals (CV = 1.03) Temporal persistence overall; temporal stability based on species abundances Temporal persistence of common species overall; temporal stability overall based on rank order data and similarity analyses; stability at three individual stations and instability at three others based on rank order data Temporal persistence of common species (mean Cj = 0.79; range 0.67–1.0); temporal stability altered by drought (pre-drought, drought, and post-drought assemblages distinct) Temporal persistence (mean Cj = 0.80); low to moderate stability (PSI = 0.47); low stability based on numbers of individuals (CV = 1.37) Temporal persistence overall; low temporal stability based on relative abundance data (species abundances changed dramatically after 1983 flood)

N

Y

Y

Y

Ross et al. (1987)

N

Y

Y

Y

Ross et al. (1985); Matthews et al. (1988)

N

Y

Y

Y

Gunning & Suttkus (1991); data analyzed by Matthews (1998)

N

Y

Y

Y

Hansen & Ramm (1994)

N

Y

Y

Y

Matthews et al. (1988)

N

Y

Y

N

Grossman et al. (1998); additional analysis by STR

N

Y

Y

N

N

Y

Y

N

Whitaker (1976); Grossman et al. (1982); data reanalyzed by Matthews (1998) Strange et al. (1992)

(continued)

Table 1.2, continued Stations

Potential Stressor and Categorization of Stress

Site

Habitat

Spp. (no.)

Sagehen Creek, CA

Small stream

11

Periodic flooding; severe winters; no major human disturbances (low to moderate stress)

8

Aravaipa Creek, AZ

Medium stream

3

Flash flooding and drought (moderate to high stress)

7

Brier Creek, OK

Small stream

5

Flash flooding and drought (moderate to high stress)

10

Purgatoire River tributaries, CO

Small streams; some intermittent

5

Flash flooding and drought (high stress)

11

Wabash River, IN

Large river

29

Dam construction; positive and negative changes in water quality; urbanization; periodic flooding (moderate stress)

75

Authors’ Conclusions Temporal persistence; moderate temporal stability based on rank order data; low temporal stability based on changes in standing crop Temporal persistence of species overall; temporal stability based on rank order data; actual numbers fluctuated extensively Temporal persistence overall; low temporal stability overall based on rank order data and similarity analysis (Im = 0.40) Temporal persistence at four of five sites (fifth site had intermittent flow); low temporal and spatial stability due primarily to variation in numbers of rare species; stability greater in sites with deep pools than with only shallow riffles Moderate temporal persistence overall; low temporal stability overall based on Bray-Curtis similarity; similarity decreased with greater time between samples to about 0.25; low similarity at individual stations based on multivariate measures using abundances

HD

L

High P

High S

N

Y

Y

N

Gard & Flittner (1974)

N

N

Y

Y

Meffe & Minckley (1987)

N

N

Y

N

Ross et al. 1985; Matthews et al. (1988)

N

N

Y

N

Fausch & Bramblett (1991)

Y

N

Y

N

Pyron et al. (2006)

References

Blue River, KS

Large river

14

Disturbed streams, Savannah River site, SC

Small stream

8

Little Uchee Creek, AL

Small stream

2

Halawakee Creek, AL

Small stream

2

Wacoochee Creek, AL

Small stream

4

Bogue Chitto River, LA

Medium stream

7

a

Average over all sites.

Reservoir construction; introduction of non-native species (moderate stress) Post-thermal discharge; periodic anoxic discharge; toxic chemicals (high stress)

Increase in pine monoculture; 69% human population increase in region; 39% decline in annual flow; flashier runoff (moderate stress) Increase in pine monoculture; 69% human population increase in region; flashier runoff (moderate stress) Increase in pine monoculture; 69% human population increase in region; flashier runoff (moderate stress) Land-use changes including increases in human population, dairy farming, cattle ranching, gravel mining, road construction, and silviculture (moderate stress)

29

14a

12

15

20

95

Low to moderate persistence (mean Cj = 0.41; range 0.2–0.54) temporal stability based on relative abundances Low temporal persistence; low temporal stability overall based on rank order data and similarity analyses; mean CV = 0.59, based on numbers of individuals Low to moderate temporal persistence with rare species eliminated (mean Cj = 0.57; range 0.22–1.0); moderate temporal stability (mean Im = 0.71; range 0.22–0.96)

Y

N

N

Y

Gido et al. (2002)

Y

N

N

N

Paller (2002)

Y

N

N

N

Johnston & Maceina (2009); additional analysis by STR

Low temporal persistence with rare species eliminated (mean Cj = 0.33; range 0.22–0.40); low temporal stability (mean Im = 0.53; range 0.36–0.71) Low to moderate temporal persistence with rare species eliminated (mean Cj = 0.27; range 0.14–0.50); low temporal stability (mean Im = 0.53; range 0.24–0.88) Low to moderate temporal persistence (Cj = 66–74%); temporal stability low (27-year comparison) to high (11- and 16-year comparisons)

Y

N

N

N

Johnston & Maceina (2009); additional analysis by STR

Y

N

N

N

Johnston & Maceina (2009); additional analysis by STR

Y

N

N

N

Stewart et al. (2005); additional analysis by STR

32

FRESHWATER FISHES OF NORTH AMERICA

et al. 1988). In spite of extreme conditions, including total dewatering of some stream reaches, the fish fauna over an 18-year period showed strong persistence on a streamwide basis with abundant species remaining abundant and rare species remaining rare with only a few exceptions (Table 1.2). Stability (a quantitative measure) of the Brier Creek fish fauna showed greater variation, and the fauna at individual collection sites (i.e., at the assemblage level) was even less persistent and stable. Long-term stability (or changes) in a fish community in arid or semi-arid environments may depend substantially on the response of individual species to drought conditions. Experiments on five common fish species in Brier Creek showed species-specific responses to drought with respect to outright survival and to post-drought recovery. For example, the Blackstripe Topminnow (Fundulus notatus) and Longear Sunfish (Lepomis megalotis) had lower survival during simulated drought compared to Central Stonerollers, Bigeye Shiners (Notropis boops), and Orangethroat Darters. Orangethroat Darters that survived an experimentally imposed drought in one summer actually recovered in physical condition to match that of conspecifics in the wild and were reproductively competent (Marsh-Matthews & Matthews 2010). Importantly, in systems with strong environmental filters, assemblages may be controlled more by stochastic processes, even though the high persistence of species might suggest primacy of niche-related processes (Chase 2007). Long-term data also exist for Piney Creek, a permanent upland Ozark stream (Ross et al. 1985; Matthews 1986c; Matthews et al. 1988) that offers a more benign habitat (i.e., no dewatering and less temperature variation). Not surprisingly, Piney Creek fishes also were highly persistent; however, in contrast to Brier Creek, the fish fauna in Piney Creek also had greater faunal stability, both overall and at the assemblage level. Piney Creek had a severe flood in December 1982; however, immediately after the flood no major changes occurred in rank abundance of the 10 most abundant species (Matthews 1986c). Less common species did change in abundance so that local assemblages were altered immediately post-flood. Eight months after the flood, the overall fish fauna and the fauna at individual collecting stations had essentially recovered to pre-flood conditions, rendering the Piney Creek fish fauna stable and persistent across years and a range of flow conditions (Matthews 1986c). Although fish assemblages clearly can rebound rather quickly from major impacts (see also Detenbeck et al. 1992 and Taylor et al. 1996a), other studies indicate that

floods or droughts changed or reset assemblage structure. Later studies on Brier Creek documented that two severe droughts resulted in a substantial change in the Brier Creek fauna, which did not recover to its former state until 3– 4 years post-drought (Matthews & Marsh-Matthews unpubl. data). In Coweeta Creek, North Carolina (Table 1.2), a severe drought resulted in 3 distinct assemblages over a 10-year period corresponding to pre-drought, drought, and post-drought (Grossman & Ratajczak 1998; Grossman et al. 1998). Finally, in a 22-year study of a 150 km reach of the Pearl River, Louisiana and Mississippi, fish assemblages showed stochastic structuring resulting from droughts, hurricanes, dams, and channel modifications. Even so, in periods between major perturbations, biotic interactions were likely important in structuring assemblages (Geheber & Piller 2012). Much of the detected variation in persistence and stability of fish assemblages is perhaps related to hydrologic variability, the variation from system to system in what composes a catastrophic event, and the timing of major perturbations (Grossman & Sabo 2010). Flooding in Brier Creek when fishes are spawning can have severe impacts on larval survival, as shown when a major flood displaced downstream and killed larval cyprinids and centrarchids 100 mm TL) individuals of the two bass species and the Central Stoneroller (Power & Matthews 1983). In pools with schools of the Central Stoneroller, attached algae (mostly Rhizoclonium and Spirogyra) was much reduced in height and standing crop from grazing by these algivorous fish. In contrast, Central Stoneroller was absent or rare in pools containing large bass, and these pools had dense, tall growths of algae. This pattern, once detected, was tested across time and by bass addition-removal experiments (Power et al. 1985) and persisted throughout a year of study. The addition of Largemouth Bass to a predator-free pool containing Central Stonerollers resulted in a major change in composition and growth form of algae in the pool. Once Largemouth Bass were introduced, Central Stonerollers rapidly emigrated out of the test pools (Power et al. 1985; Power 1987) or took shelter and remained in shallow pool edges. As a result, algae grew densely over the next several weeks in the deeper parts of the pools guarded by bass, and algal growth gradually spread into shallow areas as the numbers of Central Stonerollers decreased through emigration. This was interpreted as an example of a threelevel trophic cascade with strong effects by the bass controlling ecosystem processes in these pools. The effects of the algivorous Central Stoneroller in stream ecosystems are pervasive. By removal of algae they initiate changes within pools with consequences for invertebrates, processing of particulate organic matter, and bacterial standing crops, causing a total of >20 measurable ecosystem responses (Matthews et al. 1987; Gelwick & Matthews 1992; Matthews 1998). In Brier Creek, the reciprocal distribution of large bass and Central Stoneroller is temporally persistent. In six of eight surveys (Power et al. 1985; Matthews et al. 1994), the dichotomy in pool-to-pool distribution persisted across 14 pools for >1 year, and 12 additional surveys (1995–2003) indicated the pattern was again persistent (Matthews & Marsh-Matthews 2006b). In addition, these piscivores also result in avoidance of pools by some, but not all, other small-bodied and potential prey species and thus have major effects on local assemblage structure. Across streams, the impacts of bass species on prey fishes, particularly cyprinids, are variable and depend strongly on the physical setting of a stream and identity of the potential predator. When a search for the existence of the bass-stoneroller-algae trophic cascade was extended to

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

other, larger stream systems, where Smallmouth Bass were the dominant predator, the dichotomy between bass and stonerollers broke down, and they often were found together in pools (Matthews et al. 1987). Smallmouth Bass apparently were a less controlling predator than the other bass species, and this was confirmed when Smallmouth Bass were moved to an experiment in Brier Creek (Harvey et al. 1988). In seeming contrast, two size classes of another cyprinid, the Hornyhead Chub, reduced their use of deep pools of an experimental stream in the presence of Smallmouth Bass and instead occupied shallow raceways, suggesting the importance of particular species or systems in affecting the strength of the predation response (Schlosser 1988b). In many systems (Matthews 1998) top-down predation or predator threat by piscivores can control one or more lower trophic levels in a classic HSS (Hairston-SlobodkinSmith) pattern (Slobodkin et al. 1967). It is equally clear from the above examples that few patterns fit widely across all systems and all potential predators, and each situation probably needs to be assessed for its own unique properties when we consider effects of piscivores on species occurrences in local assemblages and on food webs. The family Cyprinidae (Carps and Minnows), although composed primarily of small-bodied species, also includes taxa that are large-bodied and piscivorous, such as Ptychocheilus spp., the pikeminnows (140–180 cm TL) (Page & Burr 1991), and Creek Chub (maximum size about 300 mm TL, Ross 2001). Following the introduction of the non-native Sacramento Pikeminnow (Ptychocheilus grandis) in the Eel River, California, habitat use of native fishes shifted (Brown & Moyle 1991). Responses of native fishes generally followed the predictions of the body size, predation risk, and water depth model (Fig. 1.16). Resident fishes shifted from broad use of riverine habitats to general avoidance of deep habitats, either by shifting microhabitat use within a habitat or by shifting habitats. Changes were most extreme for the Threespine Stickleback, juvenile Sacramento Sucker (Catostomus occidentalis), and juvenile Rainbow Trout, which shifted to habitats that were shallower than the shallowest depth usually occupied by Sacramento Pikeminnow (about 50–70 cm deep). Creek Chubs become increasingly piscivorous at >80 mm SL (Fraser & Cerri 1982), and large Creek Chubs can impact smaller fishes, including juvenile Creek Chubs (e.g., Fraser & Cerri 1982; Fraser & Emmons 1984; Schlosser & Ebel 1989). Even in small streams with distinct pool-riffle habitats, the impact of Creek Chubs on

43

habitat use of juvenile conspecifics or small cyprinid species, although measurable, is less extreme than that shown by prey fishes in the presence of bass species (Fraser & Cerri 1982; Schlosser & Ebel 1989). In a replicated, experimental stream, juvenile Creek Chubs and Blacknose Dace were less numerous in treatment compartments containing adult Creek Chubs than in those without the predator (Fraser & Cerri 1982; Fraser & Emmons 1984). Responses of the prey to the predator also were mediated by time of day and habitat structure with both reducing the effect of the predator. The prey selected habitats with cover and a predator over those that lacked cover and a predator, and prey were more likely to be associated with a predator during the day than at night, when predation risk was presumably greater. Prey responses to the predator varied depending on the amount of food available to the prey. Juvenile Creek Chubs accepted greater predation risk as the potential reward (greater food density) increased (Gilliam & Fraser 1987). Later work in a natural stream (Fraser et al. 1987) generally supported the results from the experimental stream, except that predator avoidance by the prey did not vary with the amount of habitat structure (probably because there was always some structure in the natural stream) nor diurnally. An exception to the rule of small fishes in shallow habitats and large fishes in deep habitats can occur with larval fishes. In both lentic (e.g., Werner & Hall 1988) and lotic (e.g., Harvey 1991ab) systems, large predators can create predator-free zones for larval fishes by eliminating the small fishes that would prey on the larvae. In northern lakes, larval Bluegills move into the pelagic zone immediately after hatching to feed on zooplankton and remain there until about 12–14 mm SL (Werner & Hall 1988). The initial move into the pelagic zone is likely an effect of Largemouth Bass predation, in that the presence of bass forces juvenile fishes (which would prey on larval fishes) out of the pelagic and into the littoral zone, creating a predator-free space in the pelagic zone for larvae or early juveniles. Movement back into the littoral zone is likely due to the increased predation risk caused by increased pigmentation and a larger body size. Similarly, in pools of Brier Creek, Oklahoma, survival of larval centrarchids and cyprinids was low in pools that contained juvenile centrarchids and cyprinids, but significantly higher in pools with adult Largemouth Bass (Harvey 1991a). Larvae were generally in deep pools, or deeper sections of pools, where the presence of a predator had reduced or eliminated juvenile fishes. Juvenile fishes were shifted to

44 FRESHWATER FISHES OF NORTH AMERICA

shallow water habitats where predation from Largemouth Bass would be limited both by access and by risk of predation on the bass by terrestrial predators (Fig. 1.16). Likewise, Smallmouth Bass impacted larval fish survival in a larger river, Baron Fork of the Illinois River, Oklahoma (Harvey 1991b). Even with relatively high rates of larval drift, natural pools with adult bass had higher larval densities than did pools lacking adult bass, again suggesting the importance of the predator-free zone for larval survival. In addition to habitat shifts, activity periods, such as foraging time, also may be affected by the risk of predation. Longnose Dace in two Canadian streams foraged almost exclusively at night; this pattern was maintained throughout the ice-free season. Although not tested directly, the nocturnal foraging pattern, which is rare among cyprinids, was attributed primarily to increased risk from avian and fish predation during the day (Culp 1989). Shifts in habitat use as a consequence of the threat of predation also can alter the presence or strength of competitive interactions. For instance, the crowding of small fishes in the littoral zone due to the threat of bass predation may cause increased competitive interactions in streams (Gorman 1988b) and lakes (Werner et al. 1983a), an effect experimentally demonstrated in ponds (Mittelbach 1988). Life history attributes, including body size and age at maturity, can be impacted by predation pressure and thus affect the size structure and population dynamics of local fish assemblages. The Utah Chub (Gila atraria) comprises two distinct clades resulting from an Early Pleistocene divergence between the upper Snake River and the Bonneville Basin. The Utah Chub in the Bonneville Basin evolved in the presence of their primary predator, Cutthroat Trout, but those populations became predator free when Utah Chub populations became fragmented during the Late Pleistocene recession of Lake Bonneville beginning about 10,000 years ago. In contrast, Utah Chub populations in the Snake River have co-existed continually with Cutthroat Trout, producing a predator phenotype. The predator phenotype of the Utah Chub has higher juvenile growth rates, reaches a larger adult size, and is longer lived than the derived predator-free phenotype. In addition, the predator phenotype matures later and at a larger body size and has lower female reproductive effort (suggesting a tradeoff between reproductive effort and lifespan) compared with the predator-free phenotype (Johnson & Belk 1999; Johnson 2002).

Facilitation and Mutualism In our discussion of species interactions we primarily focused on symmetrical negative interactions like competition or asymmetrical (+/−) interactions like predation. In this section we show that positive interactions (+/+; +/0) among fish species are also common and of potential importance to the formation and maintenance of fish assemblages. Facilitative interactions are asymmetrical or symmetrical positive encounters among organisms that benefit one or more of the participants and do not harm either (Stachowicz 2001). For instance, the predator-free zone provided to larval fishes by large, predatory fishes is an example of facilitation involving habitat modification. Facilitative interactions also may be symmetrical (mutualistic) when both species benefit from the association (e.g., Boucher et al. 1982). The terminology of interspecific interactions is complex; for simplicity, we follow Boucher et al. (1982) in dividing mutualistic behavior into direct and indirect mutualisms. In the former, direct interaction occurs between two species. In the latter, no direct contact occurs, but each species benefits from the other’s presence. Direct mutualisms can be further subdivided into symbiotic and non-symbiotic mutualisms with the distinction between them based on the level of their physiological integration. Boucher et al. (1982), although acknowledging exceptions are frequent, considered that symbiotic mutualisms tended to be coevolved and obligate, but non-symbiotic mutualisms tend to be facultative and not co-evolved. Much of the earlier literature on species associations treated mutualism as an obligatory response. Nevertheless, various studies (Gomulkiewicz et al. 2003; Hay et al. 2004) indicate mutualisms are often context dependent and may be obligatory in one area but not another and may even change to antagonistic interactions. Facilitation and mutualism in fish assemblages can occur between fishes and other taxa, especially with foundation species (species that contribute a framework for the entire community, such as trees, grasses, or beaver [Castor spp.]; Bruno et al. 2003; Pollock et al. 2003), or only among fishes. Although we focus primarily on the latter, others provided evidence of a diff use mutualism between species of Oncorhynchus and the trees surrounding the natal streams (Hay et al. 2004; Drake et al. 2006; Gende et al. 2007). Streams that are forested support greater densities of juvenile salmon because of the input of nutrients from leaf litter and instream wood. Spawning runs of adult salmon import large quantities of marine-derived

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

nutrients (nitrogen, carbon, and phosphorus) via their post-spawning carcasses and thus subsidize growth of riparian trees. Similarly, significant transport of marine nitrogen into grape vineyards occurs via spawning migrations of anadromous Chinook Salmon and movement of carcasses onto the terrestrial landscape by scavengers (Merz & Moyle 2006).

Species Associations and the Potential for Facilitation If co-existing species always segregated into distinct use of resources (Werner et al. 1977; Ross 1986), they would overlap minimally in use of space or foods, and individuals would occur most often with conspecifics and not in mixed-species groups. Even so, mixed-species contact groups are common in North American fish assemblages with two or more species of fishes occurring in a small space or with individuals intermingled. Suggestions of mixed-species interactions among freshwater fishes are not new. Reighard (1920) described minnow and Northern Hogsucker (Hypentelium nigricans) feeding interactions in detail, and also commented that he had observed the White Sucker (Catostomus commersonii) to be much less easily startled when accompanied by a group of Logperch (Percina caprodes). Although perhaps lacking proof, Reighard in this one paper suggested both enhanced feeding and safety as benefits from mixedspecies groupings. In 1 m2 plots in a small Minnesota lake, 14 significant negative species associations existed compared with 3 significant positive associations (Moyle 1973). Six of the negative associations were between minnows and piscivorous centrarchids. Positive associations between the Mimic Shiner and Bluntnose Minnow (Pimephales notatus), the Bluntnose Minnow and White Sucker, and the Common Shiner (Luxilus cornutus) and White Sucker suggested that some fish species were attracted to others by the protection and feeding advantages that a large school of fishes offers and that mixedspecies schooling was common in this lake. Multispecies schooling occurred among four minnow species in a small Wisconsin stream, and observations in aquaria suggested that Notropis spp. are mutually responsive with individuals of one species readily following those of another (Mendelson 1975). Young of the Longnose Dace and Creek Chub often formed mixed schools in streams (Copes 1983). In Florida canals, Largemouth Bass ≤30 cm TL and similarly sized Bluegills form mixed-species foraging groups (typical group size of

45

five) that interact in hunting small poeciliids and cichlids (Annett 1998). We have already discussed studies documenting vertical habitat segregation of fishes. Here, we emphasize that species in benthic or water-column guilds often show little if any intraguild separation in habitat use (e.g., Grossman & Freeman 1987), although this is certainly not always the case (e.g., Baker & Ross 1981). The frequent lack of intraguild resource differences at least suggests that species in a guild often occur in mixed-species schools as documented in Coweeta Creek for several minnow species and Rainbow Trout (Freeman & Grossman 1992) and for six pool-dwelling minnow species in an Ozark creek (Gorman 1987, 1988ab). For these mixed-species groups, Gorman suggested both facilitation of feeding and enhanced anti-predator effects and that microhabitat use within the guild was a dynamic balance between segregation (lessening interspecific competition for resources) and associations (with potential benefits to mixed-group members). Ozark stream minnows clearly may segregate partially into distinct microhabitats, but this segregation is modified by the presence of other minnow species with slightly more positive associations (i.e., converging to common habitats when mixed species were present) than dissociations (i.e., avoidance of each other in mixedspecies pairings) (Gorman 1987, 1988ab). In another small Ozark stream, species often formed mixed schools, intermixed vertically and horizontally (McNeely 1987). In virtually all aquatic habitats, capturing >1 fish species in a single seine haul is more common than not (raising the issue of whether the sampling method is capturing fishes across more than one microhabitat, or whether species are occurring in mixed associations), but direct visual observation by snorkeling also suggests the preponderance of mixed-species groups, especially for minnows. In a total of 787 seine hauls or snorkeling observations made in diverse habitats from large rivers, Lake Texoma, Oklahoma (a large reservoir), and smaller streams, 71% included >1 minnow species. Included in these samples were small seine hauls, about 10 m2 each, in the Canadian River, Oklahoma (Matthews 1977), in which 71% had 2 species and 46% had ≥3 species present. In 20 m shoreline seine hauls in Lake Texoma (Matthews 1998), 86% had ≥2 minnow species, and 60% contained ≥3 minnow species. In 2 m2 kickset seine hauls (Roanoke River, Virginia, Matthews et al. 1982a), 55% had ≥2 minnow species present (WJM unpubl. data). The trend for mixed-species groups of minnows also was found in additional small streams of south Oklahoma and in the Arkansas Ozarks

46

FRESHWATER FISHES OF NORTH AMERICA

(Matthews 1998). Such examples of mixed-species occurrences in localized field samples do not yield information on actual interactions among species but do suggest the possibility that North American minnows could form mixed-species groups under a wide range of conditions. Thus, in North American lakes or streams it is common to find ≥2 related fish species in such close proximity to each other that information transfer or interactive benefits seem possible.

Facilitation and Mutualism in Fish Assemblages Proposed benefits from interspecific aggregation include increased foraging efficiency and enhanced detection and avoidance of predators (Boucher et al. 1982). Earlywarning benefits in mixed groups are well known in some organisms such as birds (Thompson & Thompson 1985). Fish in monospecific shoals (i.e., schools) transfer information about predators (Magurran & Higham 1988), but early warning would be effective in a mixed-species shoal only if information is transmitted rapidly (Godin et al. 1988). In mixed-species shoals, competition for food might be less intense than in a monospecific shoal while presenting a predator the impression of a large shoal. A possible cost of mixed-species shoaling is incurred if predator evasion maneuvers were impeded by heterospecifics (during an actual predator attack) (Pitcher 1986). Also in mixed-species shoals one species might facilitate access to prey by another, or feeding opportunities might be improved by observation of feeding by heterospecifics (Pitcher 1986). For example, in North American streams, if one species of Sunfish (genus Lepomis) strikes at an insect on the surface of a pool, other Sunfish species rapidly approach to feed. In multispecies assemblages, temporarily rare species might persist by virtue of hiding within schools of more abundant heterospecifics (Moyle & Li 1979). By joining mixed-species groups, a fish might enjoy the benefit of being a member of a larger group (antipredation; vigilance) yet reduce the cost of intraspecific competition to less than it would be in an equally large group of conspecifics (Allan 1986). In an artificial stream, Allan (1986) tested three sympatric European cyprinids that often formed mixed-species shoals in streams (Dace, Leuciscus leuciscus; European Minnow, Phoxinus phoxinus; and Gudgeon, Gobio gobio) and suggested that altered behaviors of the species in mixed aggregations reflected a balance between the tendency to avoid identical use of resources, yet to remain close enough to the other species to

gain anti-predation benefits from the appearance of a large shoal. Some species in mixed groups may benefit from feeding actions of others. When benthic feeding fishes like Suckers (Catostomidae) disturb substrata, they may increase the accessibility of invertebrates to smaller insectivorous fish species. Bigeye Shiners swim above schools of benthic-foraging Ozark Minnows, feeding on items that were suspended in the water column, and Hornyhead Chubs follow the benthic algivore, Central Stoneroller (Gorman 1988b). Similarly, Gilt Darters (Percina evides) commonly follow two other darters, Logperch and Blotchside Logperch (P. burtoni), feeding on items exposed when the logperches flip over stones (Greenberg 1991). In clear upland streams, minnows (e.g., Blacktail Shiners) commonly closely follow large Northern Hog Suckers, as suggested by Reighard (1920), feeding on invertebrates suspended by benthic feeding of the Sucker (Baker & Foster 1994). In this relationship, the larger, benthic-feeding species might also benefit if it received early warning of a threat if the smaller fishes took flight from a potential predator. If this is beneficial to both participants, then the Hog Sucker–minnow tandem might represent a true mutualism. In addition to potential benefits from interactions of species in feeding or general predator defense, examples of more complex interspecific groupings exist, providing positive benefits to one or both members of the association. For instance, nest associations (Wallin 1989, 1992; Johnston & Page 1992; Johnston 1994ab) are observed frequently in North American streams. In many, visiting species lay eggs in or on gravel nests or other structures that are tended by the original (often larger) builder. For example, numerous chub species (genera Nocomis, Semotilus) build large gravel nests on which other minnow species spawn. The minnow eggs apparently benefit by being in clean gravel of the nest or from protection by the guarding nest owner. Hornyhead Chub nests are piles of gravel up to 0.9 m across and 0.3 m tall, constructed by the host carrying stones some distance by mouth and forming quite prominent structures in the gravel substrata of small streams (Robison & Buchanan 1988). Nests of Semotilus also are used frequently as spawning sites by other species. Creek Chubs construct pit-ridge nests (Reighard 1910; Ross 1977a; Maurakis et al. 1990; Johnston & Page 1992) where the male initially excavates a spawning pit about 7 cm deep by shoving stones away or moving them with his mouth. Males are multiple spawners and on each successive spawning the male extends the

EVOLUTION AND ECOL OGY OF NORTH AMERICAN FRESHWATER FISH ASSEMBLAGES

pit farther downstream, covering the newly laid eggs with stones from the extended pit. As this process continues, the result is a longitudinal ridge of pebbles (about 69 cm long × 22 cm wide × 4 cm high) that covers the fertilized eggs with the active spawning pit at the downstream end. Nest associates can gain improved reproductive success by spawning over host nests. The associates do not necessarily respond to the nest substratum but to the presence and activity of the host (Johnston 1994a). In an experiment that consisted of constructing artificial gravelmound nests in two streams (to mimic nests of the Bluehead Chub, Nocomis leptocephalus, and Green Sunfish, which naturally spawned in the streams), the artificial nests were unused by any of the numerous minnow species (including Central Stoneroller; Rosyside Dace; and Greenhead Shiner, Notropis chlorocephalus) that were observed spawning over natural nests in the streams (Johnston 1994a). In contrast, Topeka Shiners, a nest associate with Sunfish (genus Lepomis), apparently also spawn in the absence of Sunfish nests, as shown by males selecting and defending spawning sites over sand substrata (Witte et al. 2009). The association between the Yellowfin Shiner (Notropis lutipinnis) and Bluehead Chub may be a true mutualism because Yellowfin Shiners failed to reproduce in the absence of chubs. Bluehead Chub eggs also may benefit because repeated, vigorous activity by spawning minnows may keep sediments from accumulating in the nest, but advantages gained by the host were not directly tested (Wallin 1989, 1992). The hypothesis that the host also gains improved reproductive success in the presence of a nest associate was supported with the nest associate, the Redfin Shiner (Lythrurus umbratilis), the Green Sunfish host, and a predator (on eggs and larvae), the Longear Sunfish (Lepomis megalotis) (Johnston 1994b). In the presence of the predator, Green Sunfish nests that had associates present showed higher larval Sunfish survival than in those without associates, clearly indicating that hosts accrue benefits from the nest associate and that the association is a true mutualism. In tests without the predator present, no difference existed in survival of Green Sunfish larvae in nests with and without associates, indicating that associates were not detrimental to the host. The benefit of the associates to the host in the presence of a predator was likely achieved through a dilution effect. Whether or not suggested mutualisms are fortuitous interactions or arise from co-evolution is difficult to discern, in part because views differ substantially on what constitutes co-evolution. Ehrlich & Raven (1964) used the term in the sense of tightly coupled species pairs with evolution

47

of a given trait in one species producing subsequent evolution of a trait in the other species of the pair, which results in selective pressure producing further modification of the trait in the first species, and so on. An alternate view to tightly coupled co-evolution of pairs is that diff use co-evolution could affect entire assemblages of species simultaneously. That is, communities could evolve by virtue of overall contacts among species without involving reciprocal evolutionary steps by all species (Inouye 2001). Other views of co-evolution include geographic mosaic models based on different co-evolutionary outcomes for populations in different communities (e.g., Thompson 1994, 1999ab, 2005; Gomulkiewicz et al. 2000; Nuismer et al. 2003). Tightly and loosely coupled co-evolution are quite different phenomena, and co-evolution of communities is controversial (Krebs 1994), resulting in a huge primary literature and numerous major summaries (e.g., Gilbert & Raven 1980; Thompson 1982, 1994, 2005; Futuyma & Slatkin 1983; Stone & Hawksworth 1986). In fish assemblages, the term “co-evolution” most often implies diff use co-evolution or merely that members have traits influenced by long-term general associations with others (e.g., MacLean & Magnuson 1977; Gorman 1987, 1988a; Wikramanayake & Moyle 1989; Matthews 1998), although Baltz & Moyle (1983) suggested that habitat segregation between two species (Rainbow Trout and Sacramento Sucker) was due to their long history of co-evolution. In the section on the origin and age of fish assemblages, we demonstrated that species comprising local assemblages are often of widely different evolutionary ages and that fish assemblages have gained and lost species over evolutionary time scales. Because of this, the likelihood of a suite of species remaining together over long periods of time (i.e., 1,000s of years) is seemingly low. Similar arguments are given by others (Matthews 1998, Chapter 9; Grossman & Freeman 1987; Grossman et al. 1987ab, 1998) who suggested from multiyear studies that habitat availability, environmental disturbance, predation, or independently evolved species traits dominate microhabitat use by individual species in streams. Grossman & Ratajczak (1998) argued that stream fish communities are not highly co-evolved systems so that interactions among species are not regulated by strong or consistent interspecific interactions. For co-evolution to be important, species should have a high encounter rate (Thompson 1982; Price 1984), interspecific interactions should not be weak or of short duration (Futuyma & Slatkin 1983; Farrell & Mitter 1993), and

48

FRESHWATER FISHES OF NORTH AMERICA

substantial periods of close association should prevail (Brown 1995). Interactions, hence selective pressures of one species on another, change in the context of the other species present in a community, affecting mechanics of diff use co-evolution (Inouye 2001). Habitat availability clearly is an important factor in species co-occurrence, or the degree of persistence of local assemblages. Certain habitats, like deep pools, can be critical to the stability of a fish community (Schlosser 1987; Fausch & Bramblett 1991; Lohr & Fausch 1997). In practice, whether fish species congregate consistently in a particular kind of habitat in response to past co-evolution of traits or more because of independently acquired traits, the likelihood of future trait modification by co-evolutionary interactions is increased if they now co-occur in regular, direct contact. If a species at one location is part of a persistent suite of other species at the scale of years or decades, the probability increases that those species will consistently influence its evolution. Conversely, a species surrounded by a frequently changing milieu of other species is less likely to evolve under the influence of predictable biotic selection pressures (Futuyma & Slatkin 1983). What is lacking in many of the debates about co-evolution is empirical information on just how long (months, years, or generations) patterns of direct contact among mobile species in realworld communities must persist to provide a consistent template within which interspecific genetic adjustments of traits, hence local co-evolution, can occur. Some coevolved relationships among organisms are assumed to be quite old (McNaughton 1984) or develop over long periods of intimate contact (e.g., 50 million years, Currie et al. 2003). Nevertheless, reviews (Thompson 1998, 1999ab, 2005) and empirical evidence (Thompson & Cunningham 2002; Palkovacs et al. 2009) show that co-evolutionary processes among species can operate at time scales of decades, not requiring long periods of geologic time (i.e., rapid co-evolution). Thus, observations of persistent contacts in ecological time (e.g., Matthews & MarshMatthews 2006b) can be pertinent to evolutionary processes for species and communities, and the duration of contact shown for pairs and triads of species in the section on persistence of assemblages could be sufficient for rapid co-evolution to occur. The biology of individual species also influences the possibility of the acquisition of co-evolved traits. Having the potential for rapid genetic modification combined with short generation times increases the probability that fish co-existing in local communities at scales of years to decades could co-evolve. Genetic adaptation to changed

temperatures apparently occurred in 1 female visit the small male sequentially, or if they see just one female near him for an extended period of time. Hence, female Sailfin

53

Mollies use a series of decision-making rules that place a genetic predisposition for a particular male trait (large size) within a larger social context. This is a critical discovery because it means that the effects of nature (genes) can be modified by nurture (in this case, learning) so that a strict preference need not reach its fullest potential in a naturally breeding population. The effect of context-dependent copying may serve ultimately to reduce the consistency of female choice and weaken the influence of sexual selection (e.g., experiments with the Guppy, Poecilia reticulata, Dugatkin & Godin 1992; Perugia’s Limia, Limia perugiae, Applebaum & Cruz 2000). This begs the question of how often females in a natural population are exposed to mate-copying situations that would counter their normal biases. These data are needed because the two dynamics, copying under conditions of limited information (males closely matched) or copying leading to reversals of female preferences, have different evolutionary implications.

Nuptial Coloration and Limits of Vision Light transmission is much more complicated in aquatic ecosystems than in air (e.g., Levine et al. 1980; Lythgoe 1980). Depending on their wavelength, photons are absorbed or scattered by water molecules, organic particles, phytoplankton, and other plants. Absorption of light on its way from the object to the eye leads to image degradation because information is missing. The scattering of photons from other sources into the visual pathway decreases contrast between the object and the background so that distant objects appear faint and blurred, as if covered by a veil or fog (the veiling effect). Scattering of light thus leads to image degradation because some information is lost (from the object) while irrelevant information is added (from the background). So, the composition of light in any given environment is dependent on the distance that light has traveled since entering the water and what is in the water. For example, chlorophylls and other organic substances absorb most light at depths of about 25 m and shift the transmission maximum to wavelengths (λ) of about 500–600 nm (greenish-yellow: majority of coastal waters, lowland ponds, rivers). Add tannins and lignins to the picture, and little light penetrates >3 m with the transmission maximum pushed to ≥600 nm (reddishbrown: swamps, marshes, tropical blackwater rivers). Because of the differential transmission in these systems, the quality (intensity and λ composition) of light may vary dramatically along different lines of sight radiating from

54

FRESHWATER FISHES OF NORTH AMERICA

the same point (i.e., downwelling, upwelling, and horizontally scattered photons). What do fishes see? In brief, most freshwater fishes have at least two types of light-sensitive pigments in their retinas. The peak spectral sensitivity of matching pigments matches the dominant background wavelength (spacelight). Objects reflecting the same wavelengths as the spacelight appear invisible, but objects reflecting other wavelengths will appear dark against the background. Light detected by offset pigments does not match the spacelight (hence the term “offset”), so objects reflecting these wavelengths appear bright against a darker background. The distribution of light-sensitive pigments in freshwater systems tends to track the light environment (e.g., shallow, clear waters—ultraviolet, blue, green, and orange-red pigments in diurnal poeciliids and cyprinodonts; greenish freshwaters—blue, green, and orangered in diurnal Cichlids, minnows, Catfishes, and salmonids: Loew & Lythgoe 1978; Crescitelli et al. 1985). In the following I focus on the role of color in the male-female courtship dialogue. Given the variety of color patterns and our focus on visual signals, the most astonishing thing about research in this area is not so much what we have discovered (although that is, of course, fascinating), but rather that we have discovered so little.

How to Be Seen: Color in Killifishes The Bluefin Killifish, Lucania goodei, inhabits a wide variety of freshwater habitats throughout (primarily) Florida (Fig. 2.3). These habitats range from clear waters in which

Figure 2.3. Male Bluefin Killifish, Lucania goodei, flare their colorful dorsal and anal fins to compete with other males for spawning sites and to court approaching females. The color of the fins and composition of photoreceptor pigments varies with water clarity; these fish occur in habitats as diverse as the clear waters of springs to the turbid, tea-stained waters of wetlands (picture courtesy the Calypso Photographic Library, www.calypso.org.uk).

all wavelengths are transmitted quite well, but short wavelengths (UV-violet) are transmitted most effectively, to turbid, tea-stained waters, an environment that favors longer wavelengths (yellow-red). Retinal design parallels these transmission differences. Bluefin Killifishes from clear springs have more UV and violet cones and higher concentrations of short wavelength-absorbing opsins (SWS1, SWS2B), and consequently fewer yellow and red cones and lower concentrations of long wavelength-absorbing opsins (Rh2, LWS), than do swamp dwellers (Fuller et al. 2003, 2004). Much of this variability is due to environmental effects, particularly the lighting conditions under which offspring are raised (Fuller & Travis 2004; Fuller et al. 2005). Male Bluefin Killifish (Fig. 2.3) compete for and guard patches of vegetation that serve as oviposition substrates. Fighting involves flaring the dorsal and anal fins before proceeding to a circle fight. Once established, a male courts an approaching female by flaring his fins and swimming in circular loops around and in front of her. Females visit several males before spawning and spawn many times over the breeding season (Fuller 2001). Males develop a red dot at the base of their caudal fins and blue pigmentation in the anterior three-quarters of their dorsal fins. The remaining part of the male’s dorsal fin may be blue, red, or yellow; the pelvic fins are either red or yellow; and the anal fins may be red, yellow, blue, red and blue, or yellow and blue. Some color combinations (yellow-blue and to a lesser extent red-blue anal fins) in this extreme polymorphism were predicted by drainage or longitude. Males with red, and to a lesser extent, yellow anal fins were more common in clear springs. Because spring fish are more sensitive to short wavelengths, the red-yellow signal appears dark against the bright blue spacelight (contrast). Males with blue anal fins were more common in swamps (Fig. 2.3). In swamps, the male’s anal (and dorsal) fin color would appear quite dull because short wavelengths are attenuated quickly, and the fish themselves have a retina set to detect yellow-red. Even if this maximizes contrast (blue fins, dark, against red spacelight, bright), the signal will not be transmitted far nor will it be detectable by a potential mate from a distance. This paradoxical result might indicate a need for up close and personal communication; i.e., perhaps the cost of attracting predators from a distance is greater than the benefit of attracting females (Fuller 2002). So, what factors are maintaining such a variety of male color morphs? Investigations using red and yellow males indicated that females do not choose mates based on color (McGhee et al. 2007), level of male aggression and fertil-

MATING BEHAVIOR OF NORTH AMERICAN FRESHWATER FISHES

55

ization success are not correlated with color (Fuller 2001; Fuller & Travis 2001), and negative frequency-dependent mating success (rare male effect) is not operating in the field (Fuller & Johnson 2009). Given the environmentally based plasticity in visual pigments, the fitness of a given color morph possibly varies across different spectral environments, such that each morph will have the most reproductive success where it is most conspicuous (Chuno et al. 2007; Gray et al. 2008). Tests of this hypothesis, which will require more complicated control of lighting conditions in the lab, should prove fascinating.

Problems with Patterns: Color in Darters Darters (Percidae, Perches) are among the most complex and flamboyantly colored of North American freshwater fishes (Fig. 2.4). Surprisingly, no correlation is known between the intensity of carotenoid-based colors (yellows, oranges, and reds) and male reproductive success (Orangethroat Darter, Etheostoma spectabile, Moerchen 1973; Pyron 1995; Striped Darter, Etheostoma virgatum, Porter et al. 2002; but see Reeves 1907 and Page 1974). This does not rule out a role for color in male-male interactions or the possibility that other components of the male signal, perhaps blue- or green-based colors or contrast independent of color, are used by females to assess male suitability. For example, females from two Rainbow Darter (Etheostoma caeruleum) populations (Prairieville Creek and Seven Mile Creek, Michigan) performed more nosedigs, a measure of female sexual motivation, near Prairieville Creek males. Prairieville Creek males were smaller but appeared to display a greater red-blue contrast than did the larger, darker Seven Mile Creek males (Fuller 2003). Females from several darter species preferred larger males (Spottail Darter, Etheostoma squamiceps, Bandoli 1997; Relict Darter, Etheostoma chienense, Piller & Burr 1999; Fantail Darter, Etheostoma flabellare, Moretz & Rogers 2004), so the Rainbow Darter females’ response was unusual, and perhaps, color related. To test this suggestion we would have to determine whether Rainbow Darter females do indeed prefer smaller males independently of color. If color patterns are not transmitting information about male condition that females can use to differentiate among conspecific suitors, perhaps the nuptial signal is acting in species recognition, a form of sexual selection involving discrimination between a conspecific and another species (Ryan & Rand 1993). Given, however, that about 25% of all darter species are introgressed, indicating hybridization occurred at some point in their evolu-

Figure 2.4. Like many species of darters (family Percidae), breeding male (top) Rainbow Darters, Etheostoma caeruleum, develop bright nuptial coloration during the spring, while females (bottom) remain comparatively cryptic. Nevertheless, even though sexual dichromatism is widespread among darter species, the actual function of male color (e.g., male-male interactions, female courtship) is not well understood (male and female, Vermilion River, Vermilion County, Illinois, May 2009; photograph by and used with permission of Uland Thomas).

tionary history, and that the dominant crosses within one genus, Nothonotus, are between egg-burying species in which the males are noticeably different in coloration (Keck & Near 2009), it is difficult to explain what role nuptial color plays in these enigmatic fishes.

Truth in Advertising: Color in Pupfishes Territorial male Pecos Pupfish (Cyprinodon pecosensis) develop a metallic blue color over their entire bodies, but females and juveniles remain cryptically colored olivebrown, black-barred fish. A range of nuptial color develops from light blue with conspicuous lateral bars to intense blue and no bars. Male color development is context dependent; males do not develop the signal in isolation or in the presence of juveniles. The signal appears at low levels during agonistic interactions with other males then intensifies dramatically in the presence of a courting female and increases again after a spawning bout, reliably signaling, “I am a successful male; a female has already chosen me” (Kodric-Brown 1996; see also Threespine Stickleback, Reisman 1968; McLennan & McPhail 1990). The intensity of blue also appears to be signaling more than just past spawning success. Relative to their dull-colored non-territorial counterparts (Kodric-Brown &

56

FRESHWATER FISHES OF NORTH AMERICA

Nicoletto 1993), territorial males are much better swimmers, a measure positively correlated with metabolic efficiency and stamina (Smit 1965; Beamish 1978; Stahlberg & Peckmann 1978), and are in better condition (i.e., heavier / unit length, Bolger & Connolly 1989). Within territorial holders, more intensely blue males are more aggressive and thus better at defending their territories from intruders (Kodric-Brown 1983; other examples of the intensity of color being correlated with dominance include the Crescent Gambusia, Gambusia hurtadoi, McAlister 1958, and Everglades Pygmy Sunfish, Elassoma evergladei, Miller 1963). Protection of the eggs is a byproduct of territory defense because males do not care for eggs and even indulge in limited egg cannibalism (Echelle 1973; Itzkowitz 1974; Loiselle 1982). Even as a byproduct, such protection is a critical component of egg survival and is important to the female. The presence of bright blue color is thus a reliable signal of three things: male quality, breeding status, and dominance. Females respond to this signal, spawning only with blue (sexually mature) males and within that subset, preferring the most intensely colored mate available (Kodric-Brown 1977, 1983). Females, however, use more than just color to identify appropriate mates. When territorial and non-territorial males were moved from lake to laboratory, fed ad libitum for a week, and then matched for size and presented with equivalent territories, females distinguished between the two before they developed full coloration. Males transmitted information about past spawning success via the intensity of their courtship and overall activity, information that was reinforced as their nuptial signal intensified (Kodric-Brown 1995).

Nuptial Color in Sticklebacks: It’s Not Just Males The nuptial male in Stickleback fishes varies from deep black (body, fins, spines, and eye bar: Brook Stickleback, Culaea inconstans, McLennan 1993) to a mosaic of redorange sides, throat and pelvic spine membranes, metallic blue-gray back, and shimmering blue-green eyes (McLennan & McPhail 1989). The function of male nuptial coloration has been so extensively investigated, at least for the Threespine Stickleback (see discussion in McLennan 2006), that for >50 years the courtship interchange appeared to be a monologue; males send, females receive. Nothing could be further from the truth. Threespine Stickleback females become bright gold with a vertical barring pattern along their backs and sides. The intensity

of the signal varies within populations; Pacific Coast females are gold all over with light bars, and Lake Ontario and New York females have gold concentrated along the lateral plates and dorsal surface with extremely dark bars (Rowland et al. 1991; DAM pers. obs.). The female nuptial signal in populations of the Brook Stickleback from central and southern Ontario consists of a bright, semi-translucent pink-gold sheen with gold concentrated in the throat and opercular regions and on top of the head, and thin, reticulated, gray bars forming a swirled pattern on the dorsum and sides of the body (McLennan 1994). Once again, differences exist among populations: fish from Nebraska develop darker vertical bars and are less intensely golden than are Ontario fish (Ward & McLennan 2006). Re-examination of previous studies indicated that female nuptial color had been documented but, in a strange kind of Orwellian double think, either ignored or described as cryptic (see photographs of courting female Blackspotted Stickleback, Gasterosteus wheatlandi, in McInerney 1969). Overall, intensification of vertical barring occurs in five species: the Threespine Stickleback, Blackspotted Stickleback, Brook Stickleback, Fourspine Stickleback (Apeletes quadracus, Rowland 1974; Blouw & Hagen 1981), and Ninespine Stickleback (Pungitius pungitius, Morris 1958; McKenzie & Keenleyside 1970). Sticklebacks respond more to bold than fine checkered patterns (Meesters 1940), so the intensification of vertical barring and swirling seen in the female signal may take advantage of a perceptual or sensory bias in the receiver’s (male) visual system. Information transferred by this signal is a reliable indicator of sex and behavioral motivation. In the Threespine Stickleback (Rowland et al. 1991) and Brook Stickleback (McLennan 1994), the origin and the intensity of female nuptial coloration is coupled tightly with ovulation, disappearing rapidly after the female spawns. The signal thus transmits the reliable message that she is a receptive female. Males respond to this message, courting nuptially colored females more intensely than gravid, but cryptically colored, alternatives (Rowland et al. 1991; McLennan 1995). From a male’s perspective the ability to differentiate between a courting female and a gravid, but non-ovulated, female provides two benefits. First, the amount of time and energy he invests courting an unreceptive partner is decreased. Second, he can identify potential nest raiders (nonbreeding females, Foster 1990) and egg stealers (neighboring male acting as sneaker). Nest raiding and egg stealing has been documented through direct observation of the phenomenon and implied through discovery of conspecific eggs in

MATING BEHAVIOR OF NORTH AMERICAN FRESHWATER FISHES

57

stomach analyses for all species of Sticklebacks (FitzGerald 1991). If the fitness of territory holders is generally determined within one season, the potential for nest destruction will be a powerful selection pressure on the male to differentiate between a courting female and threats to his clutch.

The Newest Thing: Ultraviolet Light and Courtship Ultraviolet-A radiation is composed of short λs (320– 400 nm) and is scattered easily by water molecules and solutes (reviewed by Losey et al. 1999). Short wavelengths are also high energy and thus can cause damage to biological systems. Given these two constraints, not surprisingly many biologists assumed that UV radiation was not an important part of courtship communication in fishes. The discovery that many teleosts have retinal pigments that absorb maximally around 360 nm led Macías Garcia & de Perera (2002) to question this assumption, specifically with regards to the courtship of the Darkedged Splitfin (Fig. 2.5), a viviparous toothcarp (Goodeidae) inhabiting shallow, clear waters across the Mexican High Plateau. A male’s dorsal and anal fins are edged in black and covered with yellow to orange markings (Fig. 2.5). He displays these fins vigorously during courtship, oriented either laterally to the female or dancing in figure eights in front of her. Females prefer deeper-bodied males, which have larger and more colorful fins (Macías Garcia 1991). Males also perform a seemingly paradoxical courtship display, which involves tucking the dorsal and anal fins to the side away from the female. This action de-emphasizes the fins that first attracted the female, while highlighting the male’s dull-silver sides. Under normal laboratory lighting, females pay little attention to this display, but under natural lighting (with UV radiation), a stripe along the male’s side and head glows, attracting the female. This attraction disappears when females are used as the stimulus fish. The UV-based signal thus appears to carry information about sex and possibly species that reinforces the message being transmitted by body shape, fin size, and color. Is this simply a case of redundant coding increasing the probability that the receiver will get the message? Yes and no. The yes is obvious; the UV signal says minimally, “I am a male.” No, because UV light is transmitted over such short distances that it likely comes into play only when the courting partners are close together (Fuller 2002). Such short-distance communication may be selectively advantageous because the garter snakes Thamno-

Figure 2.5. Male Darkedged Splitfins, Girardichthys multiradiatus (35 mm TL), not only vigorously display their colored dorsal and anal fins during courtship but also tuck those fins away to flash an ultraviolet stripe along their sides and heads. Females, which can detect the flash in natural light (i.e., light with ultraviolet-A radiation), use it to assess males as spawning partners (male, Laguna de Zempoala, Mexico, 16 May 1999; photograph by and used with permission of Shane A. Webb).

phis melanogaster and T. eques, which eat Darkedged Splitfins, can detect UV radiation. The snakes attack both sexes equally, but capture males with larger fins more often, possibly because of the drag resulting from the increased surface area of the erect fins (Fig. 2.5). So, dominant, large-finned males exist in a delicate balance between being preferred by females (benefit) and being more easily captured by predators (cost). When the costs become too high, selection shifts in favor of avoiding predation, so in areas of sympatry with garter snakes, males are slimmer bodied (and smaller finned) than their relatives living without garter snakes (Macías Garcia et al. 1994, 1998). If the flash from the UV stripe attracts the snake to the courting pair, then another cost is added to the balance sheet for dominant, courting males such that males in snake-infested waters should have reduced UV-reflecting areas. On the other hand, if the UV signal is transmitted only over short distances, then no differences should exist in the signal area of similar-sized males from different habitats. UV radiation also plays an important role in swordtail courtship. To our eye, the male sword is a silvery, semitransparent extension on the caudal fin. Under UV light, however, the extensions on the tail of the Panuco Swordtail, Delicate Swordtail, Mountain Swordtail (Xiphophorus nezahualcoytl), and Barred Swordtail (Xiphophorus multilineatus) (Fig. 2.6), like the Darkedged Splitfin’s lateral stripe, flash as the male swims broadside or in a figure eight around the female (Cummings et al. 2003). Interestingly, UV reflectance is not limited to the sword in Panuco

58 FRESHWATER FISHES OF NORTH AMERICA

A

B

Figure 2.6. Ultraviolet radiation also plays an important role in swordtail courtship. The breeding male’s tail extension, which affects the female’s perception of his body size (Fig. 2.1), reflects a UV flash as he courts the female (top, male Mountain Swordtail, Xiphophorus nezahualcoytl, Rio Santa Anita, Rio Panuco drainage, 1 May 2006; bottom, male Barred Swordtail, Xiphophorus multilineatus, Rio Coy, Rio Panuco drainage, 23 April 2006; photograph by and used with permission of Juan M. Artigas Azas).

Swordtails; it is distributed across the male’s body. Large males have more reflective areas on their bodies. Medium males compensate for their UV-challenged status by being more active, engaging in behaviors that enhance the UV flash. For their part, females prefer more active medium males and less active large males. Thus, each type of male possibly adopts the behavioral strategy that will best highlight his UV ornamentation (Cummings et al. 2006). The discovery of UV-based male traits provided an explanation for the puzzling observation that, unlike females in related species, female Panuco Swordtails do not prefer sworded males; indeed, they showed a trend, albeit insignificant, of being more attracted to swordless males (Rosenthal et al. 2002a). The discovery of the interaction between UV radiation and swords prompted a rerun of the mate choice trials under natural light. When UV light was available, female Panuco Swordtails settled into the

plesiomorphic pattern of preferring sworded males (Cummings et al. 2003). So, the initial experiment revealed more about sensory biases introduced into experimental design by human observers than it did about preference in Panuco Swordtails. Unlike the snake-challenged Darkedged Splitfins, the Panuco Swordtail and the other UV light-reflecting Swordtails contend with predation by Mexican Tetras, which are not sensitive to UV radiation. Such insensitivity provides the swordtails with a private channel for exchanging information: the UV λs reflected from the sword are conspicuous against the background of sidewelling green-yellow light, emphasizing the elaborate trait to conspecifics without attracting the attention of predators. Interestingly, the only species that does not reflect UV radiation from its sword, the Highland Swordtail (Xiphophorus malinche), is not sympatric with the Mexican Tetra (Cummings et al. 2003). Presumably the costs of detecting UV radiation (damage to the retina) became an important factor when the benefit of that detection (predator-free communication) was no longer significant. In summary, the evolution of spectral sensitivity is affected by the transmission properties of the environment (Endler & McLellan 1988) that in turn sets constraints on the parts of the spectrum that are available for building nuptial coloration. Given enough underlying variability in both the receiver and the sender, evolution may eventually produce signal divergence correlated with habitat types or geographical distributions. Few researchers, however, have examined the relationship between receiver response, geographic variability in sender color, and geographic variability in the spectral properties of habitats occupied across the entire range of a species. Ultimately we need these data if we really want to understand the contribution of mating system evolution to the production of freshwater fish diversity. For example, melanic Threespine Sticklebacks tend to be restricted primarily to heavily tannin-stained waters (Reimchen 1989). The general (but not absolute) absence of red males from these habitats makes sense biologically because a signal reflecting photons matching the surrounding spacelight should appear invisible from a distance (McDonald & Hawryshyn 1995; McDonald et al. 1995). In some populations, the change in the male color signal in the novel environment has been matched by a reduction in the female’s attraction to red (Boughman 2001). Are we watching speciation in progress? In another example, interpopulation differences in number of speckles on male Jeweled Splitfins, Xenotoca

MATING BEHAVIOR OF NORTH AMERICAN FRESHWATER FISHES

variata, are not correlated with either habitat clarity or variation in female preference for speckles. In fact, the preference for speckles is only expressed in clear water, implying that females living in turbid habitats may be using non-visual cues to choose mates (Moyaho et al. 2004). Once again, are we documenting speciation in progress? To really attack these and other questions, we need to begin videotaping mate choice trials in murky waters, waters illuminated under low light intensities, or mosaics of light and shadow, in other words, under the range of conditions that has formed the environment of the fish, not the human observer.

SOUND Sound is a more effective long-distance signal in water than light, electricity, or odor, so it seems intuitive that acoustical cues should play a role in fish communication. This intuition is correct; fishes from >50 families sing to one another in social contexts (Fish & Mowbray 1970; Myrberg 1981). Sound is produced by variations on two general themes (reviewed by Tavolga 1971; Fine et al. 1977; Hawkins 1993): specialized (sonic) muscles insert either directly or via modified skeletal elements on the swim bladder and strike it like a drum (harmonic, low frequency, 40–300 Hz); and processes such as rubbing or grinding bone on bone or snapping tendons over bones produce stridulations (non-harmonic, wide frequency range 100–8,000 Hz). Given that many fishes produce sounds, can they also hear them? Most fishes can detect only low-frequency sounds by direct transmission through the body to the inner ear (≤400 Hz; e.g., Trout-Perch, Percopsis omiscomaycus; Ninespine Stickleback; Northern Pike, Esox lucius; Spoonhead Sculpin, Cottus ricei; Burbot, Lota lota; Broad Whitefish, Coregonus nasus; Mann et al. 2007). Higher frequencies must be amplified to be transmitted effectively (Hawkins 1993). Amplification is provided by gas-filled structures because gas pulsates when a sound wave passes through it, amplifying and reradiating the sound in all directions. Fishes with the greatest hearing sensitivity (hearing specialists) are characterized by some method for channeling as much of this amplified sound as possible directly to the sacculus (e.g., ostariophysans: the Weberian apparatus, modified vertebrae, found in the Lake Chub, Couesius plumbeus, and the Longnose Sucker, Catostomus catostomus, can detect sounds from 100 to 1,600 Hz; Mann et al. 2007; Squirrelfishes, Holocentridae: projections from the swim bladder contact the

59

sacculus directly; gouramies, Anabantoidei: the sacculus projects into suprabranchial air chambers; see discussions by Kenyon et al. 1998; Yan et al. 2000). Oddly, many hearing specialists do not produce communicatory sounds, indicating that sound detection and sound production are not necessarily coupled evolutionarily. The use of sound during courtship creates two problems not encountered with visual cues. First, the ability to detect sound does not require that the receiver be facing the sender (along a line of sight), so the cue is broadcast to many more receivers. Given this public display, selection should favor sounds that are unambiguously speciesspecific, and if possible, occur outside the detection range of potential predators. Second, frequency, which is an important component of visually based signals, appears to be the least effective component of an acoustical cue because it is extremely sensitive to disturbances in the medium (water) and thus is rapidly degraded. This does not mean that frequency is irrelevant, only that its role is influenced strongly by environmental conditions (e.g., frequency and individual recognition in Elephantfishes, Mormyridae, Marvit & Crawford 2000). Behavioral analyses of song structure and receiver responses indicate that fishes tend to code courtship-based acoustical information in the temporal domain, e.g., the duration, interval between, and repetition rate of the pulses within the sound unit (reviewed by Kihslinger & Klimley 2002). Neurological studies, although rare, complement these observations. The temporal resolution ability of tested fishes falls within the range of other vertebrates and is sufficient for species to detect individual pulses within songs. These capabilities occur in vocal and non-vocal species, indicating that the ability to glean information from pulse duration and interpulse interval predates the appearance of a vocal repertoire. As such, these abilities are properties of the neuralsensory network that collects and processes sound (Wysocki & Ladich 2002). The focus of most sonic research has been on marine species. Surprisingly little is known about the role of acoustic signals in the breeding systems of North American freshwater fishes. For example, Anderson et al. (2008) dropped a hydrophone into the Hudson River, New York, and discovered a cacophony of sounds, 90% of which were produced between 1500 and 600 h (dusk to dawn). Although they recorded 62 different sounds, they could associate only 4 sounds to fish species, leaving a wide range of purrs, rattles, burps, claps, honks, and drums in the yet-to-be-identified category. Here, I summarize the extent of our knowledge about fishes that sing to each other

60

FRESHWATER FISHES OF NORTH AMERICA

in fresh waters. The take home message is simple: we may not know much yet, but what we do know is fascinating.

Territorial and Courtship Sounds in Pupfishes The Cuatro Cienegas Pupfish (Cyprinodon bifasciatus) (Fig. 2.7) inhabits relatively shallow, spring-fed pools in Cuatro Cienegas Basin, Coahuila, Mexico. Simultaneous videography and acoustic recordings of territorial males revealed that the majority of sounds a male produced were short (53 ms), low-frequency (391.9 Hz) calls emitted while pursuing an intruder, regardless of the intruder’s sex or species. Having dispensed with the intruder, the male returned to patrol his territory and began producing slightly longer (57 ms) and higher-frequency (396.8 Hz) sounds, which continued once a female appeared and he began to follow her. Finally, even longer (61 ms) and higherfrequency (523.4 Hz) calls were occasionally heard immediately after the male and female had spawned. Because of the proximity of the partners, which of the two was producing the post-spawning call was not determined nor was the function of the vocalization (Johnson 2000). The pulses of the Sheepshead Minnow (Cyrpinodon variegatus) are the same length and duration as those of the Cuatro Cienegas Pupfish; however, Sheepshead Minnow males often (44.2%) produce a more complex song (≥3 pulses/ call) and always sing at a much higher frequency (1,170 Hz

Figure 2.7. Breeding, territorial males of the Cuatro Cienegas Pupfish, Cyprinodon bifasciatus, a species endemic to thermal springs of the Cuatro Ciénegas Basin, Coahuila, Mexico, produce short, low-frequency calls while aggressively pursuing conspecific and heterospecific intruders. Calls directed at ovulated females are slightly longer and of higher frequency (male, Poza de la Becerra, Cuatro Ciénegas, 19 November 2006; photograph by and used with permission of Juan M. Artigas Azas).

versus 408.7 Hz) (Nicoletto & Linscomb 2008). The Sheepshead Minnow hybridizes quite easily with congenerics. Could those mating mistakes be a result, at least in part, of similarities between calls or, alternatively, of differences that might make males of Sheepshead Minnows more attractive to congeneric females (e.g., the increased call complexity)? Do any other cyprinodontiforms call? Unpublished research indicates that the Plains Killifish (Fundulus zebrinus), Northern Plains Killifish (Fundulus kansae), Golden Topminnow (Fundulus chrysotus), Gulf Killifish (Fundulus grandis), and Bayou Killifish (Fundulus pulvereus) produce low-frequency sounds during courtship (Drewry 1962), and the Bluefin Killifish (Fig. 2.3) emits low-frequency thumps (Foster 1967). Thus calling is possibly more widespread within the larger cyprinodontiform clade than currently is documented.

Territorial and Courtship Sounds in Darters Male Blackfin Darters (Etheostoma nigripinne) and Fringed Darters (Etheostoma crossopterum, Fig. 2.8) produce two types of sound, drums and knocks, during the breeding season. Some Blackfin Darter males (27%) add purrs to their repertoire. These three sounds differ in frequency and time- and structure-related parameters: pulse present or absent, pulse number and rate, and call duration. Of the three sound types, drums carry the most intra- and interspecific information. For example, drums are longer compared with knocks in both species; however, Blackfin Darters drum at a higher dominant frequency (138.7 Hz) than do Fringed Darters (89.1 Hz). The shape of the sound

Figure 2.8. Breeding male Fringed Darters, Etheostoma crossopterum, produce sounds of two types, drums and knocks. The sounds differ when males are courting females (lowerfrequency drums) than when they interact with other breeding males. The mechanism of sound production is unknown; species of Etheostoma lack a swim bladder, the site of sound production in many fishes (breeding male, Mill Creek, Union County, Illinois, 19 April 2008; photograph by and used with permission of Uland Thomas).

MATING BEHAVIOR OF NORTH AMERICAN FRESHWATER FISHES

is context dependent; males from both species produce lower-frequency drums when courting females than when interacting aggressively with conspecific males. Hybrids between the two species produce predominantly drums and do not appear to have any context dependency in their songs. Although the mechanism of sound production is unknown in these darters, they, like other fishes lacking a swim bladder, may produce sounds by the contraction of specialized muscles (Johnston & Johnson 2000a).

Territorial and Courtship Sounds in Minnows Sound production has been described in seven species of Cyprinella (Cyprinidae, Carps and Minnows) (Fig. 2.9A). Satinfin Shiner (Cyprinella analostana) males produce sounds during both territorial establishment and courtship (Stout & Winn 1958; Stout 1959, 1960, 1963; Winn & Stout 1960). In the initial stages of territory establishment, competing males produced low-frequency (1,382 Hz), single knocks during lateral threat displays, increasing the frequency of the knocking sounds as the interaction escalated through parallel swim to circle fighting. Territorial males discriminate between an intruding male and female conspecific, reacting to the male with a series of knocks and to the female with spaced, single knocks. Intruders respond to the sounds by moving away, indicating that

A

B

61

the sounds serve a threat function. Females always find knocks threatening and will leave the area if they hear them. This may explain why breeding males do not use the more intense form of the signal if a female approaches when she is not ready to breed. The male is possibly warning her to beware, but stay in the area until she is ready to spawn. When a gravid female approaches a courting male, he swims toward her, and awaits her reaction. If she does not flee, he approaches more closely, swims in a circle around her, all fins flared, purring. Purrs are lower frequency (932 Hz) and shorter lived than knocks. A female responds to a purring male by decreasing her activity, thus facilitating the courtship circle display (Stout 1975). Male Whitetail Shiners (Cyprinella galactura), Ocmulgee Shiners (Cyprinella callisema), Tricolor Shiners (Cyprinella trichroistia; Fig. 2.9A), Tallapoosa Shiners (Cyprinella gibbsi), and Edwards Plateau Shiners (Cyprinella lepida) also produce distinct sounds during aggressive and courtship interactions. All species produce very low-frequency pulse bursts, Tricolor Shiners and Tallapoosa Shiners chirp and rattle when stationary in front of their crevices, and Ocmulgee Shiners and Whitetail Shiners knock and short knock (Whitetail Shiners only) when aggression escalates from chasing to parallel swimming and lateral displays (Phillips & Johnston 2008a, 2009; Phillips

Figure 2.9. (A) In addition to developing head tubercles and displaying bright breeding colors, territorial breeding males of many species of Cyprinella (Cyprinidae), like the Tricolor Shiner, Cyprinella trichroistia, produce sounds (low-frequency pulse bursts, chirps, and rattles) as they protect their spawning crevices from interlopers and court females. (B) Breeding male Bluntnose Minnows, Pimephales notatus, which nest in cavities, do not develop bright colors but they do produce bursts of low-frequency sounds during aggressive encounters with other males; note also the large breeding tubercles on the head (photographs by and used with permission of: Noel M. Burkhead, A, breeding male, Cochran Creek, Dawson County, Georgia, 5 May 2010; Matt Thomas, B, breeding male, Bullskin Creek, Kentucky River drainage, 8 May 2007).

62 FRESHWATER FISHES OF NORTH AMERICA

et al. 2010). In general, the rate of calling is more intense at the later stages of agonistic and courtship interactions than at the beginning for all species. Significant geographical variability exists in pulse parameters (rate, duration, and interval) across populations of Whitetail Shiners in the Ozark and Appalachian Mountains. This variability was organized by distance for courtship-based calls; nonadjacent populations were more divergent than were adjacent ones (Phillips & Johnston 2008b). Low-frequency pulses are often produced by muscles associated with the gas bladder (Demski et al. 1973; Fine et al. 1977). If the underlying mechanism of courtship pulse production, and hence the signal itself, is heritable, then one or more parameters of the call might carry information important to the evolution of assortative mating based on group (population) identity (Phillips & Johnston 2008b). This, in turn, establishes the potential for allopatric speciation, adding freshwater fishes to the long list of avian and anuran species for which an interaction between sound signal variability and geographic location, mate choice, and speciation is known. The Red Shiner (Cyprinella lutrensis) and Blacktail Shiner (Cyprinella venusta), also sing, but their songs are not described in the same detail as that of the Satinfin Shiner (Delco 1960; see also courtship knocking sounds in Spotfin Shiner, Cyprinella spiloptera, and Pearl Dace, Margariscus margarita, Winn & Stout 1960). Both male and female Red Shiners and male Blacktail Shiners could differentiate between conspecific and heterospecific vocalizations. Those vocalizations, however, appeared to be produced only by females, something so unusual that Stout (1975) argued it was an artifact of experimental design. To date, that possibility is untested. Species of Cyprinella are not the only cyprinids to sing during the breeding season. Male Bluntnose Minnows (Pimephales notatus; Fig. 2.9B) also produce complex bursts of relatively low-frequency multiple pulses differing in duration and pulse interval (Johnston & Johnson 2000b). These calls are shorter in duration but at about the same dominant frequency as the songs of Ornate Minnows (Codoma ornata), which are composed of bursts of low-frequency, non-harmonic pulses. Song length depends on the behavioral context: male-male circle threat (high level of aggression) = male courtship pass over breeding site > male-male lateral threat display > chasing (low level of aggression; Johnston & Vives 2003). Fry raised in isolation from adults (which mirrors the natural situation since Ornate Minnows do not provide any parental care) produced sounds appropriate to the behavioral context,

indicating that sound production is innate in this species (Johnston & Buchanan 2007).

Territorial and Courtship Sounds in Sunfishes Males from at least six species of Sunfishes (Centrarchidae, genus Lepomis) grunt or pop during courtship (Gerald 1971). The grunt is a low-frequency sound ( MLL (P. monacha-2 lucida triploid: 12%) > P. lucida (0%) (Thibault 1974). Fry avoidance behaviors roughly parallel the changes in cannibalistic tendencies, although the changes are not as striking

101

and are confounded by fry size at birth (Lima &Vrijenhoek 1996). Cannibalism, particularly of your own offspring, is difficult to understand within an evolutionary framework (review by FitzGerald 1992). Is it just a mistaken byproduct of predatory behavior (Lima & Vrijenhoek 1996), a way to hedge bets by sacrificing some current reproductive success (using offspring as food to enhance condition) for potential success in the future (Rohwer 1978; Sargent 1992), an artifact of laboratory conditions (Schenck & Vrijenhoek 1989; Weeks et al. 1992), or simply maladaptive? The ability to manipulate both the genetic background and environmental parameters in the Poeciliopsis clones provides an elegant system for testing these hypotheses as well as for studying the evolution of fry counterstrategies to parental cannibalism in what may well be a parental-offspring evolutionary arms race.

Hybridization: A Summary Hybridization appears to happen quite often in North American freshwater fishes, but most species manage to maintain their own integrity (Whitmore 1983). In areas of widespread hybridization, introgressive swamping is generally rare because the hybrids (F1s, F2s, and backcrosses) display an abnormal sex ratio, have reduced fertility or viability, or are at a selective disadvantage (e.g., Dowling & Moore 1984, 1985ab; Allendorf & Waples 1996; Hawkins & Foote 1998). In some instances, however, gene flow from one species to another occurs without the loss of species integrity. Limited introgression may increase genetic variability in a species and thus increase the scope for selection and drift in combination with geological factors to produce distinct lineages (Anderson 1953; Lewontin & Birch 1966; Barton & Hewitt 1985; Slatkin 1987; Rhymer & Simberloff 1996). So hybridization may be an essential part of species diversification in some lineages (Verspoor & Hammar 1991; Dowling & DeMarais 1993). The implications of such a mechanism are not trivial because it will affect how we measure the cost of the initial mating mistake to the erring fishes. Introgression also may be a negative force in some interactions. For example, Rhymer and Simberloff (1996) wrote that introgression was thought to play a contributing role in 3 of 24 extinctions of animal species in North America. Those three species were all fishes: the Tecopa Pupfish (Cyprinodon nevadensis calidae), Amistad Gambusia (Gambusia amistadensis), and Longjaw Cisco (Coregonus alpenae) (McMillan & Wilcove 1994).

102

FRESHWATER FISHES OF NORTH AMERICA

CONCLUSION Mate recognition is always a complicated process, involving information transfer between prospective mates and from the courting pair to interested bystanders (other conspecifics and predators). Tinbergen (1948, 1951, 1952) was the first ethologist to delineate the complexity of this interaction. His painstaking descriptions of stimulus (signal)-response (countersignal) chains in the courtship of Sticklebacks (among other animals) highlighted not only the sophistication of information transfer in these fish, but the importance of that information for mate recognition and selection and for synchronization between partners. In North American freshwater fishes these chains are forged with visual, chemical, tactile, near touch, and acoustic links (the importance of communication based on electrical fields has not yet been explored in any detail), creating a multidimensional mate recognition template. New researchers interested in studying the evolution of mating behavior in North American freshwater fishes can expect to encounter many joys and frustrations in their search for a viable system. On the positive side, generations of ichthyologists have created a large database comprising meticulous descriptions of behavior and life history traits. Many of those species, however, have not been subjected to rigorous experimental examination of the evolutionary forces involved in shaping mating behaviors, so the field is relatively wide open for newcomers. Enhanced video technology allowing us to peer voyeuristically on spawning fishes in the field is revealing unexpected information. How many incorrect hypotheses were erected before it was discovered that the bizarre jugular position of the urogenital opening in the Pirate Perch (Aphredoderus sayanus) allows individuals to deposit gametes deep within channels through underwater root masses, where developing embryos are protected from fast-flowing currents, predators, and siltation (Fletcher et al. 2004)? On the negative side, most experimental attention has focused on only a few members of a clade (centrarchids, gasterosteids, and poeciliids are the most obvious species-ist clades) or on only one sex. In systems for which we have adequate observational and experimental data, we tend to know little about the mechanisms underlying behavioral expression. Is it genetically controlled to any extent? What is the role of hormonal control of its development and expression? How do the sensory and neural systems involved in detecting

and processing the behavioral signal operate? What effects have the transmission properties of the environment had on shaping that signal? What role does learning play, if any, in the system? As a result we have a welldeveloped database about the evolution of mating behaviors for only a small fraction of North American freshwater fish diversity (for discussion of this problem vis-à-vis imperiled fishes, see Johnston 1999). Behavior is a critical component in the production of that diversity. All nonallopatric modes of speciation are ultimately dependent on individuals maintaining genetic distinctiveness by mating assortatively. Although vicariant speciation does not require the evolution of unique mating behaviors, a population exposed to a new selective regime in a novel habitat may respond with adaptive shifts in ecology and morphology and in mate-recognition characters, forming a unique mate-recognition system (Paterson 1985). The diversification of mating behaviors is thus linked causally to the production of diversity. The corollary is that the retention of plesiomorphic mating characters may be responsible for the loss of diversity under some conditions. For example, introductions of non-native species can intensify competition for limited spawning sites or increase the probability of interspecific mating mistakes and subsequent hybridization. Even if introgression is limited, such hybridization may wreak havoc on species with limited distributions and low population diversity (Hubbs 1955). Models simulating the impact of introgressive hybridization have demonstrated that extinction rates for parental taxa can be quite rapid if one of the taxa is rare, and if assortative mating is weak initially, or weakened due to habitat degradation (Epifanio & Philipp 2001). Habitat degradation includes more than just loss of spawning sites, forcing heterospecifics into closer and closer associations, and thus increasing the probability of hybridization either via chance (e.g., nest associates) or actual mating mistakes. The efficiency of courtship communication is directly dependent on the transmission properties of the medium. Anthropogenic intervention can disrupt breeding systems by operating on those transmission parameters. For example, noise from boats or hydroelectric dams contains low-frequency components, overlapping the transmission and detection frequencies of many freshwater fishes. This interference may damage a fish’s hearing abilities in the long term (something akin to attending too many rock concerts in quick succession), and disrupt courtship communication in the short term (e.g., Scholik & Yan 2002ab). The olfactory system is exposed directly to the environment and

MATING BEHAVIOR OF NORTH AMERICAN FRESHWATER FISHES

thus extremely susceptible to damage by dissolved pollutants (e.g., copper, lead, mercury, nickel, zinc, silver, cadmium) (Saucier & Astic 1995; Beyers & Farmer 2001 and references therein; Scott et al. 2003; Sloman et al. 2003) and organophosphates (Moore & Waring 1996b, 2001; Waring & Moore 1997). Of direct relevance to North America is the demonstration that relatively minor acidification of holding waters (from pH 7.0 to 6.0), eliminates the response to alarm cues in the Fathead Minnow, Finescale Dace (Brown et al. 2002a), Pumpkinseed (Leduc et al. 2003), Rainbow Trout (Leduc et al. 2004, 2008; Scott et al. 2003), Brook Trout (Salvelinus fontinalis, Leduc et al. 2004), Iowa Darters (Etheostoma exile, McPherson et al. 2004), and Atlantic Salmon (Leduc et al. 2006, 2009), including the fish’s ability to learn the scent of a novel predator via coupling the predator’s scent with a conspecific alarm cue (Leduc et al. 2004, 2007b). Alarm-based anti-predator reactions are triggered by the hydroxylamine group on the hypoxanthine-3-N-oxide molecule at least in minnows (G. E. Brown et al. 2000). Acidification alters the chemical structure of that trigger (6,8 dioxypurine), eliminating its biological functionality (Brown et al. 2002a). Because acidification damages both the receiver (sensory system, Moore 1994b) and the sender (the cue) its effects can sweep rapidly through the system. The effects of anthropogenic interference are often complex, affecting mating systems in unpredictable ways. For example, under turbid conditions caused by enhanced phytoplankton growth, female Threespine Sticklebacks paid more attention to courtship intensity (Engström-öst & Candolin 2007) and olfactory cues (Heuschele et al. 2009) than to color signals. Under normal clear water conditions, males adjust their courtship and color intensities based on input from competitors; in general males in good condition dominate their poor-condition neighbors (Candolin 1999, 2000ab). In turbid waters, male-male interactions decreased, relaxing the social control of signal production and allowing poor-condition males to ramp up their color and courtship over the short term. Unfortunately these males generally do not have the energy reserves required to be effective fathers, which makes them a bad choice for any female (Wong et al. 2007; Candolin 2009), at least in normal, clear water. Eutrophic waters, however, decreased territorial interactions during egg and fry guarding (because males are not as visible) and increased oxygen supply to the developing embryos (decreasing the need for pectoral fanning), both of which increased hatching success. In other words, the same

103

environmental perturbation that decreased the honesty of visual signals and thus the effectiveness of mate choice increased the reproductive success of all parental males (Candolin et al. 2008). Equally fascinating, the acidification associated with eutrophication increased the efficacy of olfactory communication (Heuschele & Candolin 2007), which may compensate, in part, for the decreased reliability of courtship intensity cues, assuming that poorcondition males cannot manipulate their scent as well as their behavior and color. Just exactly what these changes will mean to population structure in the long run is something that only time can tell us. Overall, many fish species appear to be disappearing because of anthropogenic changes (Miller 1961), and many of those changes have a direct effect on the breeding system (see extensive review of the effects of pollution on the reproductive behavior of fishes by Jones & Reynolds 1997). Loss of these species involves more than just adding a name to the IUCN Red List. It involves the loss of all the unique behaviors belonging to that species, which might include tail waggling, popping, purring, zigzagging, jaw locking, gonopodial nibbling, probing, quivering, and broadside swimming. It is thus more crucial than ever that videotapes and sonograms of the behavioral repertoire of as many fishes as possible be deposited in a central repository. We have a small window of opportunity for building such a database because the attention of the general public and funding agencies is drawn to the efforts of systematists to inventory all of the diversity on this planet. Although these inventories often involve collaborations between different systematists, to my knowledge no one studying behavior is participating at any level. Ethological ichthyologists, indeed all ethologists, need to be more proactive about this, both in attaching themselves to ongoing inventories and in packaging their own gold mine of data for storage and public access. In my experience people who would normally not give much thought to conservation issues become far more interested when they see animals interacting with one another, particularly if that interaction involves courtship and caring for offspring. Of all the animals in North America, the freshwater fishes provide us with the most diverse, bizarre, and entrancing examples of such interactions. We are living in the information age and behavior is fundamentally the transmission and interpretation of information. It is not so surprising then that behavior crosses interspecific boundaries and speaks to us all.

104

FRESHWATER FISHES OF NORTH AMERICA

Acknowledgments I am grateful to Rick Mayden for originally asking me to write this chapter. As researchers, we rarely get an opportunity to stand back and look at the big picture that surrounds our individual interests. I am also grateful to M. Ryan for providing me with a quiet space in his laboratory to work

away from the distractions of the normal office routine and to J. Bull and the entire Ryan lab for their patience and encouragement during that time. The manuscript benefited greatly by comments from D. Brooks and H. Greene. Finally, I am eternally grateful to Mel Warren for tracking down all the photographs used in my chapter. This research was funded by an NSERC Discovery Grant.

Chapter 3

Petromyzontidae: Lampreys Ian C. Potter, Howard S. Gill, and Claude B. Renaud

Lampreys (Petromyzontiformes) and Hagfishes (Myxiniformes), which are both scaleless and eel-like in body form (Fig. 3.1), are the sole surviving representatives of the agnathan ( jawless) stage in chordate evolution (Hardisty 1982, 2006; Forey & Janvier 1993). The Lampreys comprise 38 species in 10 genera (Potter & Gill 2003), together with the Drin Brook Lamprey (Eudontomyzon stankokaramani), which Holčík & Šorić (2004) subsequently recognized as a valid species, and the Western Transcaucasian Brook Lamprey (Lethenteron ninae) and the Epirus Brook Lamprey (Eudontomyzon graecus) (described by Naseka et al. 2009 and Renaud & Economidis 2010, respectively). The larva of the Lamprey, which is blind and toothless (Fig. 3.1a), lives in the soft substrates of streams and rivers, where it feeds on microorganisms and detritus (Hardisty & Potter 1971a; Moore & Mallatt 1980). The burrowing behavior of the larva led to it being termed an ammocoete, which means “embedded in sand.” After a number of years, the ammocoetes of all species undergo a radical metamorphosis (Potter 1980a; Youson 1980). During this process, the Lamprey develops eyes and a suctorial disc that bears curvilinear (alate) rows of teeth, prominent infra- and supraoral tooth-bearing laminae, and a protrusible tooth-bearing and tongue-like piston (Fig. 3.2). At the completion of metamorphosis, 18 of the Lamprey species move to the wider areas of rivers or to lakes or the sea, where, depending on the species, they spend from a few months to >3 years. In this phase of the lifecycle, they use their suctorial disc to attach to their hosts, which are mainly actinopterygian (Ray-finned) fishes (Applegate 1950; Hardisty & Potter 1971b; R. J. Beamish 1980; Halliday 1991; Renaud et al. 2009a). The teeth on the suctorial

disc help Lampreys maintain their position on the host, while the teeth on the tongue-like piston are used to penetrate the host tissue (R. J. Beamish 1980; King 1980; Potter & Hilliard 1987; Renaud et al. 2009a). After a period of substantial growth, parasitic species cease feeding and migrate to their spawning areas in rivers and streams. In contrast to the 18 parasitic species, the other 23 species reach maturity within a year of completing metamorphosis and do not feed as adults; they are thus termed “nonparasitic species” (Hardisty & Potter 1971c). Nevertheless, these nonparasitic species still develop the armory used by parasitic species for feeding as adults (Bird & Potter 1979a; Holmes et al. 1999). Because the morphology of the metamorphosed individuals of many nonparasitic species closely resembles that of certain parasitic species, each of those nonparasitic species is believed to have evolved from a given parasitic species and thus together they constitute paired species (Hardisty & Potter 1971c; Potter 1980b; Hardisty 2006; Docker 2009). The two main phenotypic features that distinguish the nonparasitic member from the parasitic member of each paired species are a far smaller adult body size and much lower absolute fecundity. All Lamprey species are, however, semelparous (i.e., die after a single spawning season). In marked contrast to Lampreys, which have a larval phase spent in fresh waters, Hagfishes do not have such a phase and spend their entire life in marine environments (Hardisty et al. 1989; Hardisty 2006). Moreover, unlike Lampreys, which are iono- and osmoregulators, the Hagfishes are iono- and osmoconformers and cannot tolerate fresh water and low salinities (Bartels & Potter 2004; Wright 2007). Indeed, the extant Hagfishes are unique

106

FRESHWATER FISHES OF NORTH AMERICA

Figure 3.1. Lateral views of (a) larval Lamprey (ammocoete), (b) adult Lamprey, and (c) Hagfish (reproduced by permission of the Royal Society of Edinburgh from Transactions of the Royal Society of Edinburgh: Earth Sciences volume 80 (1989), pp. 241–254).

Figure 3.2. Oral disc of the Ohio Lamprey, Ichthyomyzon bdellium, showing the different fields and types of teeth and laminae and their nomenclature. Note the alate rows comprising an inner circumoral and an outer marginal, and the intervening intermediate rows of disc teeth. MA, median anterior tooth row; MG, marginal teeth; AF, anterior field; AC, anterior circumoral teeth; SO, supraoral tooth plate; LF, lateral field; IT, intermediate disc teeth; LC, lateral circumoral teeth; LL, longitudinal lingual laminae; TL, transverse lingual lamina; IO, infraoral lamina; PC, posterior circumoral teeth; PF, posterior field (reprinted from The Biology of Lampreys, Vol 1, M. W. Hardisty and I. C. Potter, eds., C. L. Hubbs and I. C. Potter, Distribution, Phylogeny and Taxonomy, pages 1–65, 1971).

among vertebrates in that their sodium and chloride serum concentrations approximate those of their marine environment with the result that their internal milieu is iso-osmotic with that of their environment. These characteristics imply that Hagfishes have always lived exclusively in marine environments (Lutz 1975). It was thus surprising when a Hagfish, †Myxineidus gononorum, was found in reputedly freshwater deposits of the Late Carboniferous (about 318–299 mya) in Allier, France (Poplin et al. 2001). Because adult Lampreys and Hagfishes share many features, such as an eel-like shape, a tongue-like piston that bears teeth, and gills that are located in pouches, and since both groups lack scales, bone, and paired fins, many taxonomists and comparative anatomists have considered Lampreys and Hagfishes to be closely related. Lampreys and Hagfishes were thus both placed in the class Cyclostomata or Marsipobranchii, which recognizes their possession of round mouths and pouch-like gill sacs, respectively (Hardisty 1979). The roots of the ordinal name Petromyzontiformes mean “stone,” “to suck,” referring to the attachment by adult Lampreys to rocks during nest building and mating. After comparing numerous morphological characteristics in extant and fossil agnathans and gnathostomes ( jawed vertebrates), and the physiological characteristics of the living Lampreys, Hagfishes, and gnathostomes, several workers questioned whether the two groups of living agnathans are monophyletic (e.g., Hardisty 1979, 1982; Janvier 1981; Forey & Janvier 1993; Forey 1995; Near 2009). Indeed, these comparisons led to the view that Lampreys are more closely related to the gnathostomes than to the Hagfishes. Although a few molecular analyses likewise did not support the monophyly of Lampreys and Hagfishes (Suzuki et al. 1995; Rasmussen et al. 1998), the

PETROMYZONTIDAE: LAMPREYS

majority of such analyses supported monophyly of the Cyclostomata (e.g., Stock & Whitt 1992; Mallatt & Sullivan 1998; Kuraku et al. 1999; Delarbre et al. 2002; Takezaki et al. 2003; Blair & Hedges 2005; Kuraku & Kuratani 2006). In view of marked differences between the evolutionary implications of morphological and molecular data, Near (2009) subjected a combination of phenotypic and molecular data sets for Hagfishes, Lampreys, and gnathostomes to phylogenetic analyses. He showed that the implication that Lampreys and Hagfishes were monophyletic depended on the type of analysis used (i.e., maximum parsimony versus Bayesian), and suggested that strong support for the monophyly of the cyclostomes inferred from molecular data sets should be treated with measured skepticism. While the precise relationships of Hagfishes, Lampreys, and gnathostomes was still hotly debated as recently as 2009 (Janvier 2009; Nicholls 2009), the results of microRNA studies appear to leave little doubt that the cyclostomes are monophyletic (Heimberg et al. 2010; Janvier 2010). Irrespective of arguments regarding the closeness of the relationship between Lampreys and Hagfishes, the discovery of a well-preserved Lamprey fossil in Devonian deposits (416–359 mya) in South Africa implies that these two taxa have been separated for >360 million years (Gess et al. 2006). Indeed, analyses of cDNA data led Kuraku & Kuratani (2006) to conclude that the two groups of cyclostomes diverged between 470 and 390 mya.

DIVERSITY AND DISTRIBUTION The extant Lampreys, which comprise three families, have an essentially anti-tropical distribution (Hubbs & Potter 1971). The 37 species of the Petromyzontidae, the Northern Hemisphere Lampreys, live in the cooler waters of North America, Europe, and Asia (see Figs. 3.3–3.8 for the distribution of this family in North America), but the 3 species of the Mordaciidae, the Southern Top-eyed Lampreys, and the sole species of the Geotriidae, the Southern Striped Lamprey, are restricted to temperate regions of the Southern Hemisphere (Table 3.1). The anti-tropical distribution of Lampreys is related, at least in part, to the inability of ammocoetes to tolerate high temperatures. This is exemplified by the fact that the ultimate incipient lethal temperatures of the Sea Lamprey (Petromyzon marinus) from North America, the European Brook Lamprey (Lampetra planeri) from Europe, and the Pouched Lamprey (Geotria australis) from Australia are

107

31.4, 29.4, and 28.3°C, respectively, and thus below the temperatures often recorded in tropical rivers (Potter & Beamish 1975; Macey & Potter 1978). Any description of the diversity and the distribution of Lampreys (Figs. 3.3–3.8) needs to recognize that the biological characteristics and ecological requirements of their ammocoetes and adults differ markedly. This is particularly the case with anadromous parasitic species, where the essentially sedentary and burrowing ammocoete is confined to fresh water and the adult, during its feeding phase, occupies the marine environment in which it can become widely distributed and thus move far outside the geographical range of its larval phase. Although many nonparasitic species are sympatric with their presumed parasitic ancestor, this is not always the case. For example, the nearest points in the distributions of the nonparasitic American Brook Lamprey (Lethenteron appendix) and its presumed ancestor, the Arctic Lamprey (Lethenteron camtschaticum), are >2,000 km apart (Vladykov & Kott 1978), which accounts for the highly disjunct distribution of Lethenteron (Fig. 3.3). The Northern Hemisphere family Petromyzontidae contains 8 genera and 37 species (Table 3.1). In North America, this family is represented by 6 genera and 23 species of which 11 are parasitic and 12 are nonparasitic, the latter lifecycle category sometimes being referred to as brook Lampreys (Table 3.2). Four of the parasitic species undergo an anadromous migration, moving after metamorphosis from their larval habitats in fresh water to marine environments, where they feed parasitically before ultimately returning to rivers and streams to spawn. In contrast, the other seven parasitic species are confined to fresh water for the whole of their lifecycle (Table 3.1).

Petromyzontinae Two of the Lamprey genera found in North America (Petromyzon and Ichthyomyzon) are members of one of two major clades within the Petromyzontidae, but four genera (Tetrapleurodon, Entosphenus, Lethenteron, and Lampetra) belong to the other major clade (see phylogenetic relationships section). The first clade constitutes the subfamily Petromyzontinae and the second the subfamily Lampetrinae (Table 3.1; Fig. 3.9). Petromyzon marinus, the sole representative of its genus, has an anadromous form that is widely distributed along the western and eastern seaboards of the North Atlantic Ocean, ranging in rivers in North America from Newfoundland in the north to Florida in the south (Fig. 3.4). Molecular studies strongly

Table 3.1. Classification, common names, lifecycles, and ranges of the species in the three extant families of Lampreys. In the lifecycle column, nonparasitic species are noted that can be unambiguously paired with a particular parasitic species. Asterisks denote species found in North America. Classification

Common Name

Lifecycle

Range

Order Petromyzontiformes

Lampreys

Short-headed Lamprey

Anadromous, parasitic

Mordacia praecox

Precocious Lamprey

Mordacia lapicida

Chilean Lamprey

Freshwater, nonparasitic derivative of M. mordax Anadromous, parasitic

Drainages of southeastern Australia Drainages of southeastern Australia Drainages of Chile

Family Mordaciidae Genus Mordacia Mordacia mordax

Family Geotriidae Genus Geotria Geotria australis

Southern Top-eyed Lampreys

Southern Striped Lamprey Anadromous, parasitic

Drainages of southern Australia, New Zealand, Chile, and Argentina

Caspian Lamprey

Anadromous, parasitic

Caspian Sea drainages

Sea Lamprey

Anadromous and freshwater, parasitic

Drainages of North Atlantic, European Arctic, and Mediterranean Oceans/Seas

Silver Lamprey

Freshwater, parasitic

Ichthyomyzon fossor*

Northern Brook Lamprey

Ichthyomyzon castaneus*

Chestnut Lamprey

Freshwater, nonparasitic derivative of I. unicuspis Freshwater, parasitic

Hudson Bay, Great Lakes, St. Lawrence River, and Mississippi River drainages As for I. unicuspis

Ichthyomyzon gagei*

Southern Brook Lamprey

Ichthyomyzon bdellium* Ichthyomyzon greeleyi*

Ohio Lamprey

Family Petromyzontidae Subfamily Petromyzontinae Genus Caspiomyzon Caspiomyzon wagneri Genus Petromyzon Petromyzon marinus*

Genus Ichthyomyzon Ichthyomyzon unicuspis*

Subfamily Lampetrinae Genus Tetrapleurodon Tetrapleurodon spadiceus* Tetrapleurodon geminis* Genus Entosphenus Entosphenus tridentatus*

Pouched Lamprey

Northern Hemisphere Lampreys

Freshwater, nonparasitic derivative of I. castaneus Freshwater, parasitic

Hudson Bay, Great Lakes, St. Lawrence River, and Gulf of Mexico drainages Gulf of Mexico drainages Ohio River drainages

Mountain Brook Lamprey

Freshwater, nonparasitic derivative of I. bdellium

As for I. bdellium

Mexican Lamprey

Freshwater, parasitic

Mexican Brook Lamprey

Freshwater, nonparasitic derivative of T. spadiceus

Celio, Duero, Zula, and Lerma Rivers, and Lake Chapala, Mexico Celio and Duero Rivers, and Rio Grande de Morelia drainage, Mexico

Pacific Lamprey

Anadromous and freshwater, parasitic 108

Drainages of western Canada, the United States, Mexico, and Japan

Table 3.1, continued Classification

Common Name

Lifecycle

Range

Entosphenus minimus*

Miller Lake Lamprey

Freshwater, parasitic

Entosphenus similis*

Klamath Lamprey

Freshwater, parasitic

Entosphenus macrostomus*

Vancouver Lamprey

Freshwater, parasitic

Entosphenus folletti*

Northern California Brook Lamprey Kern Brook Lamprey

Freshwater, nonparasitic

Pit-Klamath Brook Lamprey

Freshwater, nonparasitic

Upper Klamath River drainage, Oregon Klamath River drainage, Oregon, and California Lake Cowichan drainage, Vancouver Island, British Columbia Klamath River drainage, California Friant-Kern Canal and Merced River, California Klamath River drainage, Oregon, and Pit River, California

Arctic Lamprey

Anadromous and freshwater, parasitic Freshwater, nonparasitic derivative of L. camtschaticum Freshwater, nonparasitic derivative of L. camtschaticum Freshwater, nonparasitic derivative of L. camtschaticum

Entosphenus hubbsi* Entosphenus lethophagus* Genus Lethenteron Lethenteron camtschaticum* Lethenteron alaskense*

Freshwater, nonparasitic

Lethenteron appendix*

American Brook Lamprey

Lethenteron reissneri

Far Eastern Brook Lamprey

Lethenteron kessleri

Siberian Brook Lamprey

Freshwater, nonparasitic derivative of L. camtschaticum

Lethenteron ninae

Western Transcaucasian Brook Lamprey

Freshwater, nonparasitic

Drainages of Arctic and North Pacific Oceans Drainages of Brooks and Chatanika Rivers, Alaska, and Mackenzie River, Canada Great Lakes drainages and eastern United States, St. Lawrence River, and Mississippi River drainages Drainages of Amur River, Sakhalin Island, and Kamchatka Peninsula, Russia, and in South Korea and Japan Drainages between Ob and Anadyr Rivers, and of Sakhalin Island, Russia, and Hokkaido Island, Japan Drainages of the Black Sea

Carpathian Lamprey

Freshwater, parasitic

Danube River drainage

Ukrainian Brook Lamprey

Drainages of Baltic, Azov, Black, Adriatic, and Aegean Seas Drainages of Adriatic Sea

Korean Lamprey

Freshwater, nonparasitic derivative of E. danfordi Freshwater, nonparasitic derivative of E. danfordi Freshwater, parasitic

Macedonia Brook Lamprey

Freshwater, nonparasitic

Yalu River drainage, China and North Korea Strymon River drainage, Greece

Epirus Brook Lamprey

Freshwater, nonparasitic

Loúros River drainage, Greece

Western River Lamprey

Anadromous and possibly freshwater, parasitic Freshwater, nonparasitic derivative of L. ayresii

Drainages of North American Pacific Coast

Genus Eudontomyzon Eudontomyzon danfordi Eudontomyzon mariae Eudontomyzon stankokaramani Eudontomyzon morii Eudontomyzon hellenicus Eudontomyzon graecus Genus Lampetra Lampetra ayresii*

Lampetra pacifica*

Alaskan Brook Lamprey

Drin Brook Lamprey

Pacific Brook Lamprey

109

Drainages of Columbia River, Oregon, and Sacramento–San Joaquin Rivers, California (continued)

110

FRESHWATER FISHES OF NORTH AMERICA

Table 3.1, continued Classification

Common Name

Lifecycle

Range

Lampetra richardsoni*

Western Brook Lamprey

Freshwater, nonparasitic derivative of L. ayresii

Lampetra aepyptera*

Least Brook Lamprey

Freshwater, nonparasitic

Lampetra fluviatilis

European River Lamprey

Lampetra planeri

European Brook Lamprey

Lampetra zanandreai Lampetra lanceolata

Po Brook Lamprey Turkish Brook Lamprey

Anadromous and freshwater, parasitic Freshwater, nonparasitic derivative of L. fluviatilis Freshwater, nonparasitic Freshwater, nonparasitic derivative of L. fluviatilis

Drainages of Pacific Ocean, British Columbia, Alaska, Washington, and Oregon Drainages of northwestern Atlantic Ocean and Gulf of Mexico, United States Drainages of northeastern Atlantic Ocean As for L. fluviatilis, plus Danube and Volga River drainages Drainages of Adriatic Sea Iyidere River, Turkey

Figure 3.5. Geographic range of Ichthyomyzon.

Figure 3.3. Geographic range of Lethenteron in North America.

Lethenteron

Figure 3.4. Geographic range of Petromyzon in North America. Ichthyomyzon

Figure 3.6. Geographic range of Tetrapleurodon.

Tetrapleurodon

Petromyzon

indicate that the populations of the anadromous Sea Lampreys on either side of the Atlantic Ocean do not mix (Rodríguez-Muñoz et al. 2004). The landlocked derivative of the anadromous P. marinus, which is likewise parasitic, is abundant in Lakes Oneida and Cayuga and the Laurentian Great Lakes and their tributaries. The genus Ichthyomyzon is confined to the fresh waters of eastern

North America (Table 3.1; Fig. 3.5) that are tributary to the Gulf of Mexico, St. Lawrence River, and Hudson Bay (Hubbs & Trautman 1937; Hubbs & Potter 1971). This genus contains three parasitic species, the Silver Lamprey (Ichthyomyzon unicuspis), Chestnut Lamprey (Ichthyomyzon castaneus), and Ohio Lamprey (Ichthyomyzon bdellium), and their respective derived nonparasitic species,

PETROMYZONTIDAE: LAMPREYS

111

Figure 3.8. Geographic range of Lampetra in North America.

Figure 3.7. Geographic range of Entosphenus in North America.

Lampetra

Entosphenus

Plate 3.1. Sea Lamprey, Petromyzon marinus

the Northern Brook Lamprey (Ichthyomyzon fossor), Southern Brook Lamprey (Ichthyomyzon gagei), and Mountain Brook Lamprey (Ichthyomyzon greeleyi, formerly I. hubbsi) (Table 3.1).

Lampetrinae Among the four genera found in North America that belong to the second clade of petromyzontids, the genus Tetrapleurodon is endemic to the fresh waters of the Mexican Plateau (Fig. 3.6) and contains one parasitic species, the Mexican Lamprey (Tetrapleurodon spadiceus), and its nonparasitic derivative, the Mexican Brook Lamprey (Tetrapleurodon geminis) (Table 3.1). The genus Entosphenus comprises seven species that includes the Pacific Lamprey (Entosphenus tridentatus), a large and anadromous parasitic species, whose distribution in fresh water extends widely in a broad arc along the eastern seaboard of the Pacific Ocean from the northern part of Mexico in the south to Alaska in the north (Fig. 3.7) and then into northern Japan in the Western Pacific Ocean (Potter & Gill 2003). Three other species of Entosphenus, the Miller Lake Lamprey (Entosphenus minimus), Klamath Lamprey (Entosphenus similis), and Vancouver Lamprey (Entosphenus macrostomus), are also parasitic but are far smaller than E. tridentatus and confined for their entire lifecycle to

single drainages of the Pacific Ocean. The three nonparasitic species of Entosphenus, the Northern California Brook Lamprey (Entosphenus folletti), Kern Brook Lamprey (Entosphenus hubbsi), and Pit-Klamath Brook Lamprey (Entosphenus lethophagus), occur in drainages within a restricted area in the southern part of the range of E. tridentatus (Table 3.1). Three of the seven species of Lethenteron are found in North America (Potter & Gill 2003; Fig. 3.3). These include the parasitic species, the Arctic Lamprey (Lethenteron camtschaticum), which occurs in drainages of the Arctic and North Pacific Oceans and contains both anadromous and freshwater forms. The Alaskan Brook Lamprey (Lethenteron alaskense), one of the nonparasitic derivatives of L. camtschaticum, is restricted to rivers in Alaska and northwestern Canada, and the other nonparasitic species, L. appendix, occurs only in drainages of the eastern seaboard of North America and the Gulf of Mexico (Fig. 3.3). The seven species of Lampetra contain four that are endemic to North America, two that are endemic to Europe, and one that is endemic to Asia. In North America, this genus has a disjunct distribution similar to that of Lethenteron (Fig. 3.8). The anadromous parasitic Western River Lamprey (Lampetra ayresii) is found in the drainages of the North American Pacific Coast, encompassing those in

112

FRESHWATER FISHES OF NORTH AMERICA

Table 3.2. Lifecycle characteristics of Lampreys found in North America. Length and weight are for adults. General Characteristics

Characteristics of Lifecycle Categories

Total number of species 23 Lifecycle mode Anadromous and parasitic, freshwater and parasitic, or freshwater and nonparasitic Spawning Semelparous Duration: May to August (except Tetrapleurodon spp.—November to January) Habitat: shallows of rivers, streams, or lakes Eggs: about 1 mm in diameter, buried in gravel or sand, in shallow water (typically 2,000 km (see diversity and distribution section). The question of which selection pressures may have led to the evolution of nonparasitic species has been addressed in several reviews (e.g., Salewski 2003; Hardisty 2006; Docker 2009). Nonparasitic species possibly arose when barriers brought about by advances and retreats of glaciers 10,000–15,000 years ago prevented the migration of parasitic species from and into river systems (Docker 2009). This relatively recent time period could account for those paired species in which the dentitional characters of the nonparasitic and parasitic species are particularly

117

similar. A decline in the abundance of suitable hosts may also have been a factor in the selection for a nonparasitic mode of life. Assuming near equal rates of evolutionary change, variation in the extent of degeneration in the dentition of nonparasitic species indicates that nonparasitic species have evolved at different times. Indeed, Docker (2009) proposed that the extent of differences between the members of species pairs reflects the following five sequential stages in the speciation of nonparasitic Lampreys: 1. Parasitic species with no nonparasitic counterparts 2. Polymorphic population producing both parasitic and nonparasitic forms 3. Paired species without fixed morphological or genetic differences 4. Paired species with fixed genetic differences 5. Nonparasitic species that have been isolated for a long period from their parasitic ancestor, which may no longer be represented in the contemporary fauna Docker has also recognized that some typically anadromous parasitic species are represented in fresh water by parasitic praecox (premature) forms that are far smaller than the anadromous form and thus presumably have a shorter adult trophic phase and that in the past may have constituted an important intermediate stage in the evolution of certain nonparasitic species. Hardisty (2006) and Docker (2009) present excellent reviews and more detailed discussions of the ways that nonparasitic species may have evolved.

FOSSIL RECORD The first recorded fossil Lamprey, †Mayomyzon pieckoensis (Fig. 3.12), was described from the Upper Carboniferous (about 280 mya) deposits of Mazon Creek, Illinois (Bardack & Zangerl 1971). The excellently preserved fossils possessed an annular cartilage in the same position as in extant Lampreys, where it plays a crucial role in maintaining the structural integrity of the suctorial disc of adults (Lanzing 1958). †Mayomyzon pieckoensis also possessed welldeveloped eyes, which in living Lampreys develop after the completion of metamorphosis. Although this fossil Lamprey possessed these important adult characteristics, its largest specimen was only 60 mm TL, which is less than the minimum length at which any extant parasitic species enters metamorphosis. Further, †M. pieckoensis possessed a relatively smaller oral disc than contemporary Lampreys

118

FRESHWATER FISHES OF NORTH AMERICA

the petromyzontiform lineage extended to the Lower Cambrian (about 520 mya). Analyses of more recent discoveries of >500 fossils of †H. ercaicunensis indicate, however, that this species is a stem-group craniate and not a Lamprey (Shu et al. 2003).

MORPHOLOGY

Figure 3.12. †Mayomyzon pieckoensis, the first fossil Lamprey to be described (reprinted from The Biology of Lampreys, Vol 1, M. W. Hardisty and I. C. Potter, eds., D. Bardack and R. Zangerl, Lampreys in the Fossil Record, pages 67–84, 1971).

and did not contain the teeth that are so distinctive of the adults of living Lampreys. These features suggest that †M. pieckoensis was a scavenger rather than a parasite (Hardisty 1979). The fossil Lamprey †Hardistiella montanensis (from about 325 mya) was found in Carboniferous deposits in North America, as was †Pipiscius zangerli (from about 310 mya), which is possibly a Lamprey (Bardack & Richardson 1977; Janvier & Lund 1983; Janvier et al. 2004). The fossils of †H. montanensis contained a structure that could have been a complex sucking device, but the fossils showed no evidence of having piston, tectal, or annular cartilages. The disc of †P. zangerli bears a ring of simple tooth plates around the oral aperture. Like the fossils of †M. pieckoensis, those of †H. montanensis are small (i.e., 300 mm TL, and among freshwater species those of Ichthyomyzon typically reach 200–300 mm TL, and the maximum size of E. minimus is 150 mm TL. Because each nonparasitic species does not feed after completing its larval phase and undergoes some shrinkage during sexual maturation, the maximum length of its adult is slightly less than that of its ammocoete.

Eye Brain Notochord Pharynx Oral hood Cirrhi

Endostyle Velum

Feeding Mechanisms of Ammocoetes

Figure 3.14. Cleared and stained larval Lamprey (ammocoete): whole animal (top) and anterior region (bottom).

The ability of ammocoetes to ingest detritus, bacteria, diatoms, and other microorganisms is facilitated by their possession of a large oral hood that helps direct food toward the pharynx (Fig. 3.14). A ring of oral cirrhi at the entrance of the pharynx acts as a sieve, preventing large particles from passing into the branchial chamber. The muscular actions of the velum and branchial chamber move food and water into the pharynx, where the food is trapped on mucous strands and exposed to digestive enzymes produced by the prominent endostyle that is located at the

base of the pharyngeal chamber (Moore & Mallatt 1980; Youson 1981). The food is then passed backward into the simple intestine where further digestion and then assimilation occur (Cake et al. 1992). The water drawn into the pharynx is passed out over the well-developed gills, which have a large surface area that facilitates the extraction of the oxygen required for metabolism (Lewis & Potter 1982). This unidirectional water flow thus has a feeding and respiratory function (Randall 1972).

120

FRESHWATER FISHES OF NORTH AMERICA

Feeding Mechanisms of the Adults of Parasitic Species The ability of the adults of parasitic species to attach themselves to, and feed on, their hosts is facilitated by their possession of a large suctorial disc and tongue-like piston, which both possess prominent teeth (Figs. 3.1, 3.2, 3.15, 3.16). The adults of some parasitic species feed on the blood of their hosts, and those of the other species ingest mainly flesh or both blood and flesh (Renaud et al. 2009a) (see ecology section). Although these interspecific dietary differences are reflected in conspicuous differences in the dentition and other feeding structures, the fundamental components of that apparatus are the same in all Lampreys (Potter & Hilliard 1987; Potter & Gill 2003). The feeding apparatus provides, however, not only an example of an extreme form of specialization for a unique form of feeding but also many of the major characters used in Lamprey taxonomy, and the suctorial disc even plays a pivotal role in nest building and mating (Hardisty & Potter 1971b; Hubbs & Potter 1971). After the adult Lamprey has attached itself to a host by means of its suctorial disc, the teeth on that disc and the supraoral and infraoral laminae (Fig. 3.2) immediately become embedded in the host tissue, thereby enhancing the strength of the attachment. The piston then rocks backward and forward with the result that the teeth on

Figure 3.15. (a) Oral disc, and (b) dorsal and lateral views of the anterior region of an anadromous Sea Lamprey, Petromyzon marinus, which has just commenced its parasitic phase.

Figure 3.16. Oral discs of (a) the Silver Lamprey, Ichthyomyzon unicuspis, which feeds on blood as an adult, and (b) the Arctic Lamprey, Lethenteron camtschaticum, and (c) the Western River Lamprey, Lampetra ayresii, which both feed on flesh as adults. Permission granted by Fisheries and Oceans Canada; reproduced from V. D. Vladykov and E. Kott. 1979. List of Northern Hemisphere Lampreys (Petromyzonidae) and their distribution. Fisheries and Oceans Miscellaneous Special Publication 42. Reproduced with the permission of © Her Majesty the Queen in Right of Canada, 2010.

PETROMYZONTIDAE: LAMPREYS

its single transverse and two longitudinal laminae destroy the host tissue through rasping, gouging, or both (Potter & Hilliard 1987; Potter & Gill 2003; Renaud et al. 2009a). Lamphredin, a substance with anticoagulant and lytic properties, is secreted onto the host tissue, preventing blood coagulation and aiding in the digestion of the host tissue (Lennon 1954; Baxter 1956; Renaud et al. 2009a). In Northern Hemisphere Lampreys and G. australis, this substance is secreted by bean-shaped buccal glands located in the basilaris muscle to either side of the piston and immediately below the eyes. The movement of the piston and its tooth-bearing laminae then help transfer the food backward into the esophagus and eventually the intestine. The presence of velar tentacles at the entry to the water tube, which is located at the anterior end of the esophagus and leads to the branchial pouches, prevents food from entering those pouches (Fig. 3.17). Blood feeders, such as P. marinus and species of Ichthyomyzon, possess alate rows of teeth throughout the anterior, lateral, and posterior fields of their suctorial disc, a small and simple supraoral lamina, a W-shaped transverse lingual lamina and two longitudinal lingual laminae that bear fine, sharp teeth (Figs 3.2, 3.15, 3.16a). Blood-feeding Lampreys attach themselves to locations where the vascular supply of the host is well developed, sometimes remaining on the same region for a long period (Lennon 1954). They typically produce a small hole in their hosts with the wound sometimes becoming extended as the Lamprey moves its position (Figs. 3.18, 3.19; Potter & Beamish 1977; King 1980). Their two buccal glands are large, but their velar tentacles are small and few in number (Renaud et al. 2009a).

Figure 3.17. Sagittal section through the head of a Lamprey (reprinted from The Biology of Lampreys, Vol 1, M. W. Hardisty and I. C. Potter, eds., M. W. Hardisty and I. C. Potter, The General Biology of Adult Lampreys, pages 127–206, 1971).

121

In contrast to blood-feeding species, those Lampreys that feed on flesh, such as L. ayresii and L. camtschaticum, remove large chunks of host tissue (Fig. 3.20) and possess a larger and more complex supraoral lamina, and their alate tooth rows are confined to the anterior field of the suctorial disc (Fig. 3.16bc). Flesh-feeding Lampreys also possess smaller buccal glands but larger and more numerous velar tentacles. Further, their transverse lingual lamina is U-shaped and has a greatly enlarged central cusp (Fig. 3.16bc; Potter & Hilliard 1987; Gill et al. 2003; Potter & Gill 2003; Renaud et al. 2009a). The presence of far less dentition on the suctorial disc of flesh feeders than blood feeders enables such Lampreys to alter their position readily after they have removed flesh from one location. During feeding, the lateral cusps of the large blade-like supraoral

A B

Figure 3.18. (A) Landlocked Sea Lampreys, Petromyzon marinus, attached to a White Sucker, Catostomus commersonii, and (B) the round and prominent, recently formed wound and a healing wound ( just above the pelvic fins) that were produced by anadromous P. marinus on an Atlantic Salmon, Salmo salar (reproduced with permission of John Wiley & Sons from Potter, I. C., and F. W. H. Beamish. 1977. The freshwater biology of adult anadromous Sea Lampreys Petromyzon marinus. Journal of Zoology, London 181:113–130).

122 FRESHWATER FISHES OF NORTH AMERICA

A

C

B

Figure 3.19. (A) Chestnut Lamprey, Ichthyomyzon castaneus, attached to the head of a Common Carp (Cyprinus carpio) in a clear stream in the Ozark region of Missouri (photograph by and used with permission of W. N. Roston). (B) An unidentified parasitic Lamprey attached to the dorsum of a breeding male Striped Shiner, Luxilus chrysocephalus, in the Ocoee River, Polk County, Tennessee (photograph by and used with permission of Jeremy Monroe of Freshwaters Illustrated). (C) Lamprey wound on the nape of a Largemouth Bass, Micropterus salmoides, in the Sipsey Fork Black Warrior River drainage, Alabama. The wound is presumably from a Chestnut Lamprey, the only parasitic species known in Mobile Basin (Boschung & Mayden 2004) (photograph courtesy of Andy Dolloff of USDA Forest Service).

remain embedded in the host and interact with the teeth on the lingual laminae to cut away flesh (Potter & Hilliard 1987; Potter & Gill 2003; Renaud et al. 2009a). The insertion and subsequent retraction of the large central cusp on the transverse lingual lamina play a major role in gouging out flesh and passing it backward into the oral cavity. Species such as T. spadiceus, which feed on substantial amounts of blood and flesh, possess features that are more similar to those of flesh feeders than blood feeders, such as their possession of a wide supraoral lamina and numerous large velar tentacles (Gill et al. 2003; Renaud et al. 2009a). The possession of larger buccal glands by blood feeders than flesh feeders presumably reflects a greater need to pre-

vent the blood from clotting and thereby ensure a constant supply of food (Renaud et al. 2009a). The larger and generally more numerous velar tentacles in flesh feeders than blood feeders is related to flesh feeders having a greater requirement to prevent solid material from entering the branchial chamber, where it could clog the gills and restrict respiration (Potter & Gill 2003; Renaud et al. 2009a). The efficiency of the suctorial mechanism in enabling adult Lampreys to attach to their hosts, and for some species to remain on that host for many hours (Lennon 1954), relies on the well-developed annularis muscles and the strong annular cartilage to which they are attached (Rovainen 1982). Suction is enhanced by the

PETROMYZONTIDAE: LAMPREYS

mucus that is secreted by the numerous fimbriae that line the rim of the oral disc (Figs 3.2, 3.15, 3.16; Lethbridge & Potter 1979; Khidir & Renaud 2003). The imprints made by the various dentitional components of an adult of E. tridentatus on the operculum of a Chum Salmon (Oncorhynchus keta) to which it had attached (Fig. 3.21) attests to the pressure that Lamprey dentition imposes on its hosts. The evolution of a suctorial disc was accompanied by the development of structures and mechanisms that enabled the respiratory flow to become tidal (i.e., allow water to pass directly into and out of the branchial chamber via the branchiopores). This contrasts with the situation in ammocoetes and all other fishes in which the respiratory flow of water is unidirectional, passing through the oral aperture or mouth into the pharynx and over the gills and out through the vertical gill slits (Randall 1972).

123

Figure 3.20. The devastating results of an attack by the Western River Lamprey, Lampetra ayresii, on a Pacific Herring, Clupea pallasii, of 20 cm TL (photograph by and used with permission of R. J. Beamish).

Metamorphosis From the previous description of larval and adult stages, the metamorphosis of the parasitic species of Lampreys clearly involves radical changes in morphology (Potter 1980a; Youson 1980, 1988, 2003). These changes include a loss of larval structures, such as the oral hood and endostyle, which are adaptations for microphagous feeding, and the development of structures such as well-developed fins, eyes, suctorial disc, and teeth (Figs. 3.1 and 3.2) that facilitate a more active and predatory-parasitic lifestyle. Although each nonparasitic species, which during metamorphosis is often indistinguishable from that of its parasitic ancestor, does not feed after the commencement of metamorphosis, it undergoes the same changes as parasitic species, reflecting its origins from such a species. Adult characteristics, however, such as teeth and longitudinal folds in the intestine, do not become as well developed as in their parasitic ancestor, and their eyes and oral disc are relatively smaller (Bird & Potter 1979ab; Hilliard et al. 1983; Docker 2009). Further, unlike the situation in parasitic species, the gonads of nonparasitic species start to develop rapidly before the end of metamorphosis, and sexual maturity is attained within a year of the completion of the larval phase (see phylogenetic relationships section). Moreover, the maturation of the gonads is accompanied by the degeneration of the intestine as occurs in parasitic species as their gonads mature during the upstream migration (Hilliard et al. 1983).

Figure 3.21. The marks made by the various dentitional components of a feeding-phase Pacific Lamprey, Entosphenus tridentatus, on the operculum of Chum Salmon, Oncorhynchus keta, to which this Lamprey had attached. Note that the operculum has been oriented so that the marks made by the infraoral lamina are located at the bottom of the figure (permission granted by the Canadian Museum of Nature).

GE NE TICS

Karyology The Northern Hemisphere Lampreys have an exceptionally large number of small chromosomes. Thus, 8 species from 4 genera (Petromyzon, Ichthyomyzon, Lethenteron, and Lampetra) have modal diploid numbers of about 164 with most chromosomes being minute and acrocentric (Howell & Denton 1969; Howell & Duckett 1971; Robinson et al. 1974; Potter & Robinson 1981). Because cephalochordates have 36–38 chromosomes that are likewise small and have their centromere located near or at the end of the chromosomes (Wang et al. 2003), the high diploid number in Lampreys probably arose through polyploidy, a view proposed by Ohno et al. (1968) on the basis of chromosomal data for one species of Northern Hemisphere Lamprey. Geotria australis, the sole representative of the Southern Hemisphere Geotriidae, has an even slightly

124 FRESHWATER FISHES OF NORTH AMERICA

higher diploid number (about 180) with most of the chromosomes also being acrocentric (Robinson & Potter 1981). In contrast, the parasitic M. mordax and its nonparasitic derivative M. praecox, representing the Mordaciidae, the other Southern Hemisphere family, both have a far lower diploid number of 76, and their chromosomes are metacentric or sub-metacentric (Potter et al. 1968b; Robinson & Potter 1969), suggesting that extensive centric fusions have taken place during the evolution of Mordacia. The possibility that such centric fusions occurred within a chromosomal complement similar to that of other Lampreys is consistent with the similarity in the nuclear DNA contents of M. mordax and other Lampreys (Robinson et al. 1975).

Lamprey Genetics in Studies of Craniate Evolution Ohno (1970, 1999) proposed that the diversity, success, and increased complexity of the vertebrates (compared with urochordates and cephalochordates) were related, at least in part, to genome duplications that had occurred during their evolution. He argued that by increasing the number of copies of a gene, the chances of an individual gaining a beneficial mutation would be increased (i.e., natural selection would have more material on which to act). The majority of studies aimed at elucidating the timing of these duplications and the ways that major morphological novelties have arisen within the vertebrates have compared specific genes and their morphological expression in the urochordates, cephalochordates, or both with those in gnathostomes (reviewed by Holland 1999, 2003; Shimeld & Holland 2000). The results of studies, which have included Lampreys, Hagfishes, or both, suggested that at least one and possibly two genome duplications took place in a common ancestor of the agnathans and gnathostomes (Sharman & Holland 1998; Neidert et al. 2000; Kuraku et al. 2008; Putnam et al. 2008). These genomic duplications were followed by further lineagespecific genomic modifications. The characterization of some of the resulting gene families, their link to developmental pathways, and their phenotypic characteristics are increasing our understanding of the evolution of the vertebrates. For example, studies of the Hox, Sox, Pax, and Dlx families of genes are providing insights into the development of the neural crest, placodes, endoskeleton, and brain, which are innovations that define the vertebrates (Neidert et al. 2001; Ota et al. 2007; Sauka-Spengler & Bronner-Fraser 2008). Likewise, those of the opsin (Lamb

et al. 2007) and Ikaros families (Haire et al. 2000; Rolff 2007) are extending our knowledge on the evolution of the visual systems and the acquired-adaptive immune response, respectively. Indeed, Rolff (2007) suggested that genome duplications led to an imbalance in the genome, a probable underlying cause of cancer, and that this in turn probably resulted in the development of the acquiredadaptive immune system in vertebrates. The nuclear genome of P. marinus undergoes a dramatic remodeling during embryonic development, resulting in the elimination of hundreds of millions of base pairs (and at least one transcribed locus) from many somatic cell lineages (Smith et al. 2009a). This corresponds to a reduction in genome size of >20% between germline (sperm) and soma (blood), which is far greater than has thus far been found in gnathostomes. In the context of the phylogeny of the early vertebrates, it is interesting that the genomes of the other extant agnathan group, the Hagfishes, also undergo similar large-scale rearrangements, involving the removal of repetitive sequences and chromosomes from their germlines (Goto et al. 1998; Kubota et al. 2001; Kojima et al. 2010). Lampreys and Hagfishes thus represent ideal candidates for studies aimed at increasing our understanding of the mechanisms that regulate remodeling of the vertebrate genome and providing an insight into the factors that promote stability and change in that genome (Smith et al. 2009a).

Gene Order The organization of the genes in the mitochondrial genomes of the Northern Hemisphere Lampreys P. marinus and L. fluviatilis differs from those of other vertebrates, including the Inshore Hagfish, Eptatretus burgeri (Delarbre et al. 2000, 2002). Although this raises the distinct possibility that the gene order in extant Lampreys is unique, the validity of such a generalization requires confirmation that those in the two Southern Hemisphere families of Lampreys (Geotriidae and Mordaciidae) are similar to that of their Northern Hemisphere counterparts.

PHYSIOLOGY Descriptions of aspects of the physiology of Lampreys are provided in other sections of this chapter when they help account for the behaviors of a particular stage in the lifecycle of Lampreys. In this section, the focus is on respiration, osmoregulation, and vision.

PETROMYZONTIDAE: LAMPREYS

Respiration In a respirometer, the oxygen consumption by larval L. planeri in chambers supplied with a glass bead substrate into which the ammocoetes burrowed was compared with that in chambers without the substrate (Potter & Rogers 1972). The rate of oxygen consumption was greater when ammocoetes were not burrowed than when burrowed, which is the typical, and thus presumably less stressful, situation for ammocoetes (Potter & Rogers 1972; Wilkie et al. 2001). As would be expected with a poikilotherm, the routine (resting) rate of oxygen consumption of burrowed ammocoetes is related to temperature with the mean rates in larvae of I. greeleyi, e.g., ranging from 8.1 μl g−1 h−1 at 3.5°C to 90.1 μl g−1 h−1 at 22.5°C, which represents a Q10 of 3.6 (i.e., rate of change in consumption for each 10°C change) (Hill & Potter 1970). Because the ammocoetes exhibited little or no movement when burrowed, the above routine rates of oxygen consumption were considered to approximate standard rates of oxygen consumption. This conclusion is consistent with the rate of oxygen consumption by the burrowed ammocoetes of I. greeleyi at 15°C not being significantly different from the standard rate of oxygen consumption recorded at the same temperature for ammocoetes of P. marinus (i.e., by extrapolating oxygen consumption at different swimming speeds to that at zero swimming speed) (Holmes & Lin 1994). Following vigorous exercise, the rate of oxygen consumption in ammocoetes of P. marinus at 15°C increased by 5– 6 times, and although subsequently declining progressively, remained elevated for the next 3 h (Wilkie et al. 2001). When larval P. marinus are swimming at the maximum rate they can maintain for a prolonged period at a series of temperatures, the relationship between rate of oxygen consumption and temperature is unimodal with the metabolic rate being greater at 15 and 20°C than at 7, 10, and 25°C (Holmes & Lin 1994). The estimated maximum scope for activity (i.e., the difference between standard and active metabolic rates) of larval P. marinus in summer is 19°C, which is close to the estimated preferred summer temperature of 20.8°C (Holmes & Lin 1994). In laboratory trials, the ammocoetes of I. greeleyi tolerated greatly reduced oxygen tensions, a particularly valuable trait for an animal that lives an essentially sedentary life in burrows where oxygen tensions are often likely to be low (Potter et al. 1970). For example, larval I. greeleyi can tolerate for ≥4 days oxygen tensions of only about 8 mm Hg at 5°C and about 18 mm Hg at 22.5°C. Further,

125

these ammocoetes do not emerge from their burrows until oxygen tensions have declined to near lethal levels. The ability of ammocoetes to tolerate low oxygen tensions is a product of their relatively low metabolic rate and the high affinity of their blood for oxygen (Potter et al. 1970; Potter & Rogers 1972; Bird et al. 1976). Ammocoetes can use anaerobic metabolism of muscle glycogen to help fuel vigorous muscle activity (Paton et al. 2001; Wilkie et al. 2001). Yet, such activity can only be undertaken in bursts due to the rapid and substantial acidosis produced by the hydrolysis of adenosine triphosphate (ATP) (Hochachka & Mommsen 1983). Glycogen is replenished, however, within 0.5 h of the cessation of exercise and lactate returns to resting levels within an additional 0.5–1.5 h. The ability to recover rapidly from vigorous anaerobic metabolism is invaluable to an animal that relies mainly on such metabolism at those times when it has repeatedly to undertake the energy-demanding task of burrowing (e.g., when the substrate in which it lives is disturbed, such as through scouring during floods). The standard rate of oxygen consumption in the parasitic species L. fluviatilis changed little during the early stage of metamorphosis but then doubled toward the end of that radical change in morphology and physiology (Lewis & Potter 1977). These trends were paralleled during the metamorphosis of the derivative nonparasitic species with the rate of oxygen consumption subsequently remaining high in mature adults. In the case of adult landlocked P. marinus, the standard rates of oxygen consumption, derived from swimming-chamber measurements, ranged from 53 mg kg−1 h−1 at 5°C to 114 mg kg−1 h−1 at 20°C, and active oxygen consumption at 10°C was 475 mg kg−1 h−1 (F. W. H. Beamish 1973). These rates, which are comparable with those recorded for salmonids (Trouts and Salmons) of similar weight at a similar temperature, are facilitated in part by the possession of gills with a large surface area (Lewis & Potter 1976ab). Further, because the oxygen dissociation curve of whole blood of adult Lampreys shifts well to the right of that of the ammocoete measured under the same conditions, the oxygen delivery pressure to the tissues is far greater in adults, which would be of benefit to this more active stage in the lifecycle (Bird et al. 1976; Macey & Potter 1982). The shift in the oxygen dissociation curve between ammocoete and adult reflects the change from larval to adult hemoglobins that occurs during metamorphosis (Manwell 1963; Potter & Nicol 1968) and parallels the types of change that take place in amphibians during their metamorphosis and in mammals at about the time of birth (Maclean & Jurd 1972). The shift in

126

FRESHWATER FISHES OF NORTH AMERICA

the properties of the blood of Lampreys during metamorphosis is coincident with the site of hemopoiesis changing from the intestinal typhlosole and nephric fold in the ammocoete to the fat column above the nerve cord in the adult (Percy & Potter 1976). As with the ammocoete, the adult Lamprey uses anaerobic metabolism of muscle glycogen to help fuel vigorous exercise, and its glycogen levels can likewise recover rapidly (Boutilier et al. 1993; Mesa et al. 2003). Although glycogen does not recover as rapidly as in ammocoetes, recovery is still faster than is typically the case in teleosts (Boutilier et al. 1993). This rapid recovery would be particularly useful to an animal that swims in bursts, especially when faced with high water velocities and obstacles (Beamish 1974; Mesa et al. 2003; Dauble et al. 2006), and because of its relatively poor swimming performance, depends on anaerobic metabolism of glycogen to help fuel such spurts. The excretion of metabolic acid is the primary means of correcting the extracellular acidosis that follows vigorous exercise (Wilkie et al. 1998). The inferior swimming ability of adult Lampreys, at least compared with many teleosts, is probably related to the absence of a hydrostatic organ and paired fins and the use of a tidal rather than a unidirectional flow of water for respiration (Randall 1972; Beamish 1974). Further, because it generates far greater lateral forces and drag, the sinusoidal mode of locomotion used by Lampreys is less efficient than that of those teleosts in which propulsion is achieved mainly by caudal fin thrust (i.e., subcarangiform, carangiform, or thunniform swimming modes) (Helfman et al. 2009). The adult Lamprey overcomes its limitations in swimming performance by swimming in short bursts and using its oral disc at intervals to attach to rocks or other hard structures to avoid the need to keep swimming to maintain position (Quintella et al. 2009). When considering aerobic respiration, it is important to recognize that Lamprey hemoglobins, like those of Hagfishes but unlike those of gnathostomes, do not form stable tetramers, dissociating to monomers on oxygenation and associating to dimers or higher oligomers on deoxygenation (Nikinmaa 2001). Monomers have a greater affinity for oxygen than the dimers or oligomers, and are characterized by an oxygen dissociation curve that is parabolic rather than sigmoidal and is shifted farther to the left (i.e., has a lower P50). It is also relevant that, during hypoxia, the red blood cells swell, which leads to a reduction in the mean cellular hemoglobin concentration (MCHC) and thus a trend for dimeric and oligomeric hemoglobins to be converted to monomeric hemoglobins. The increase in the Bohr effect (i.e.,

shift of the oxygen dissociation curve to the left) that accompanies hypoxia also contributes to an increase in oxygen affinity (Nikinmaa 2001). These characteristics of their hemoglobins provide Lampreys with the dual ability of maximizing the uptake of oxygen under hypoxic conditions and offloading oxygen during vigorous exercise, such as swimming and burrowing, when oxygen is abundant.

Osmoregulation One of the most conspicuous physiological differences between the two groups of living agnathans is that Lampreys are iono- and osmoregulators, but Hagfishes are iono- and osmoconformers (F. W. H. Beamish 1980a; Bartels & Potter 2004; Wright 2007). Moreover, the anadromous parasitic species can osmoregulate efficiently in fresh water, when they are ammocoetes or migrating upstream as adults and in sea water during their marine trophic phase. The efficient osmoregulatory mechanisms evolved by Lampreys enable the sodium and chloride concentrations in their internal milieu to be maintained at levels well above those of fresh water when the Lamprey is in rivers and streams and well below those of full-strength sea water when it is in marine environments (Morris 1972; Beamish et al. 1978). As the osmolality of the serum is far lower than that of sea water, Lampreys probably originally lived in fresh water and only later evolved the marine phase that is characteristic of contemporary anadromous species (Lutz 1975; Hardisty et al. 1989). When in fresh water, Lampreys, like teleosts, are faced with an osmotic influx of water and an efflux of ions across the body surface (i.e., skin and gills). This problem is overcome by an active uptake of monovalent ions across the gills and the excretion of a copious supply of urine from the kidneys and in the ammocoete also through intestinal resorption of ions from its food (Fig. 3.22). Studies of ion exchange mechanisms in teleosts and in the organs of other vertebrates indicate that the intercalated mitochondria-rich cells in the gills of Lampreys when in fresh water are responsible for the uptake of chloride and that the secretion of hydrogen by these cells facilitates the uptake of sodium by the pavement cells or the intercalated mitochondria-rich cells, which are also located in the gills (Bartels & Potter 2004; Bartels et al. 2009). When in sea water, adult Lampreys are confronted with the reverse situation to that in fresh water (i.e., they lose water to the environment through osmosis) (Fig. 3.22). Although this loss is compensated for by the swallowing of sea water and by the gut absorbing sodium and chloride,

PETROMYZONTIDAE: LAMPREYS

the Lamprey then contains an excess of monovalent ions. This is overcome by excreting these ions via the gills as a near hypertonic solution (Fig. 3.22). This excretion is mediated through the chloride cells, which develop in the gills during metamorphosis and use a secondarily active transcellular transport of chloride to provide the driving force for the passive outward movement of sodium through leaky paracellular pathways between these chloride cells (Bartels & Potter 2004). The above hypotheses on the osmoregulatory mechanisms used by Lampreys in fresh and sea water are supported by results of cytochemical studies involving ammocoetes and metamorphosed Lampreys (Reis-Santos et al. 2008). Proteins crucial for sodium and chloride uptake in fresh water (carbonic anhydrase and a vacuolar-type H+ -ATPase) are co-localized in cells, whose distribution in the gill epithelium corresponds to those of the intercalated mitochondria-rich cells, but Na+/K+ -ATPase, a marker for chloride cells in gill epithelia, occurs in groups of cells in the interlamellar region of the gill filaments of metamorphosed individuals, and thus where the chloride cells are located. Further, H+ -ATPase expression is negatively correlated with the external salinity, which is consistent with the observation that intercalated mitochondria-rich cells are no longer present when the metamorphosed Lamprey has entered sea water (Bartels & Potter 2004). In laboratory experiments, ammocoetes of P. marinus cannot osmoregulate when the osmolality of the environment exceeds that of their own sera (about 225 mosmol/kg) and about 50% of ammocoetes die within 24 h of transfer to water of 350 mosmol/kg (Beamish et al. 1978). In contrast, fully metamorphosed individuals of anadromous P. marinus can readily acclimate to full-strength sea water and then can maintain their serum osmolality at 260 mosmol/kg. This ability to switch rapidly from hyper-osmotic regulation in fresh water to hypo-osmotic regulation in full-strength sea water is so effective that >80% of fully metamorphosed P. marinus can survive direct transfer to full-strength sea water (Potter & Beamish 1977). After P. marinus re-enters fresh water on its spawning run, its intestine degenerates, and the chloride cells become covered by the flanges of adjacent pavement cells and undergo apoptosis (Bartels & Potter 2004). Consequently, the animal can no longer osmoregulate in hypertonic environments.

Vision Like most other vertebrates, the eyes of Lampreys are well developed and possess a retina that is specialized for acute

127

Figure 3.22. The osmoregulatory mechanisms employed by anadromous Lampreys during the freshwater (top) and seawater (bottom) phases in their lifecycles (reproduced with permission from H. Bartels and I. C. Potter. 2004. Cellular composition and ultrastructure of the gill epithelium of larval and adult lampreys: Implications for osmoregulation in fresh and seawater. The Journal of Experimental Biology 207:3447–3462).

vision (i.e., capable of resolving fine detail). Because the eyes of Northern Hemisphere Lampreys possess two photoreceptor types, those Lampreys may have the potential to discriminate prey on the basis of contrast, color, or both (Collin 2007; Collin et al. 2009). In contrast to the eyes of Northern Hemisphere Lampreys, those of M. mordax possess a large and single type of photoreceptor, but those of G. australis contain five different types. Further, the eyes of M. mordax are unique among Lampreys in possessing a tapetum, which would reflect light back toward the photoreceptors, maximizing their capture of light and increasing sensitivity in low light intensities. The eye of G. australis is also unique among Lampreys in that it possesses an irideal flap that would reduce the amount of intraocular flare. These interspecific differences are assumed to reflect adaptations to the different modes of

128

FRESHWATER FISHES OF NORTH AMERICA

life of these two species. Thus, the characteristics of the eyes of adult M. mordax are of particular benefit to this species because they are nocturnally active, but those of G. australis would be of special value to this species during its adult trophic phase, which is spent in the bright surface waters of the Southern Ocean (Potter & Gill 2003). Although adult Lampreys have well-developed eyes, the dermal photoreceptors that were developed in the ammocoete continue to mediate light avoidance during adult life (Binder & McDonald 2008a).

REPRODUCTION Lamprey reproduction is a highly synchronized process that is initiated and mediated by a complex neuroendocrine coordination and integration of environmental cues and hormonal mechanisms (Fig. 3.23; Sower 2003). Temperature is a particularly important environmental cue for reproduction. For example, on the basis of a study of L. planeri during 14 successive spawning seasons (Hardisty 1961), the spawning of this nonparasitic species was dependent on the attainment of water temperatures of 10–11°C. The hypothalamus plays a major role in controlling reproduction in Lampreys through responding to external and internal cues by the timed release of the decapeptide gonadotropin-releasing hormone (GnRH). As in all vertebrates, GnRH acts on the pituitary to regulate the pituitary-gonadal axis (Fig.

3.23). The pituitary gland responds to GnRH by secreting gonadotropins, which are the major hormones influencing steroidogenesis and gametogenesis; gonadotropins also regulate activities such as spawning behavior (Sower 2003; Sower et al. 2009).

Sexual Dimorphism in Mature Adults Petromyzontid Lampreys only start to develop sexual dimorphism as they become mature. The mature male possesses a larger suctorial disc and a urogenital papilla and in some species, a small gular pouch. The mature females develop an anal fin-like fold (see Fig. 14 in Hardisty & Potter 1971b; Monette & Renaud 2005).

Upstream Spawning Migration of Parasitic Lampreys Following the completion of their adult trophic phase, the parasitic species of Lampreys enter rivers and migrate upstream at night to their spawning areas in the headwaters or other areas where the water is relatively shallow (Hardisty & Potter 1971b) (Fig. 3.24). Adults of landlocked P. marinus select rivers for their spawning run that contain an extremely potent pheromone, which is released into those rivers by ammocoetes and is not bound by natural organic matter in ways that reduce its natural biological potency (e.g., Sorensen & Vrieze 2003; Fine et al. 2004; Wagner et al. 2009; Fine & Sorensen 2010). This phero-

Lampreys Control of Reproduction

INTERNAL FACTORS

C

STRESS

HIGHER BRAIN CENTER HYPOTHALAMUS

-/+

GnRH-I GnRH-II

PITUITARY Ir-GTH

GONADS

Steroiogenesis and Gametogenesis OTHER: Reproductive Behavior, Sex differentiation

EXTERNAL FACTORS

Figure 3.23. Schematic diagram of the hypothalamic-pituitary-gonadal axis in the control of reproduction in the Sea Lamprey, Petromyzon marinus (reprinted from Journal of Great Lakes Research 29, Suppl. 1, S. A. Sower, The endocrinology of reproduction in lampreys and applications for male lamprey sterilization, 50– 65, 2003, with permission from Elsevier).

PETROMYZONTIDAE: LAMPREYS

A

129

Figure 3.24. (A) Chestnut Lamprey, Ichthyomyzon castaneus, a parasitic species, in the midst of moving to its spawning site in Big Creek, Wayne County, Missouri, 10 May 1987 (photograph by B. M. Burr). (B) Pacific Lamprey, Entosphenus tridentatus, in its freshwater phase, rests in a swift current as it moves to spawning areas in the Smith River, Douglas County, Oregon (photograph by and used with permission of Jeremy Monroe of Freshwaters Illustrated).

B

monal attractant comprises a mixture of at least three sulfated steroids, one of which, petromyzonol sulfate, is a Lamprey-specific bile acid derivative (Sorensen et al. 2005; Sorensen & Hoye 2007). These compounds, which induce olfactory and behavioral activity at sub-picomolar concentrations, apparently a record among fishes (Fine & Sorensen 2010), do not appear to be species specific, and their presence in a river presumably indicates to adult Lampreys that suitable Lamprey habitats are present in that river. Indeed, the size of the spawning migrations of landlocked P. marinus in the various tributaries of the Great Lakes is positively correlated with the number of larvae resident in those tributaries and thus with the amount of pheromone discharging into the lake (Wagner et al. 2009). Moreover, migrating Sea Lampreys do not enter waters that lack the larval pheromone.

Analysis of the sequence of part of the mtDNA control region of anadromous P. marinus from 11 rivers on the Atlantic Coast of North America provided overwhelming evidence that, although Lampreys enter rivers in response to the presence of the pheromone, they do not tend to home into their natal streams (Waldman et al. 2008a). Such a lack of homing, which is consistent with the conclusions from other molecular studies (Rodríguez-Muñoz et al. 2004; Bryan et al. 2005), contrasts with the homing exhibited by many anadromous teleosts. Homing would not be beneficial to parasitic Lampreys because their movements during the adult trophic phase are directed largely by their hosts. This leads to wide dispersal of the adults and consequently a reduction in their likelihood of being close to their natal streams at the end of their parasitic phase when they are

130

FRESHWATER FISHES OF NORTH AMERICA

preparing to embark on their spawning migration (Waldman et al. 2008a). Landlocked Sea Lampreys enter rivers when water temperatures in rivers exceed those of their lacustrine environment and subsequent movements occur when water temperatures are >4.5°C and reach their maxima at 10–18.5°C (Applegate 1950). Lampreys do not feed during their upstream migration, and thus their energy requirements must be met largely by the lipid stored during the preceding adult feeding phase (Beamish et al. 1979). During the upstream migration, the intestine degenerates (Youson 1981), and in anadromous species, this and the loss of functional chloride cells in the gills result in their inability to osmoregulate in a hypertonic environment (F. W. H. Beamish 1980a; Bartels & Potter 2004). The time of onset and duration of the spawning migration varies among species. The upstream migration of the anadromous Sea Lamprey in rivers on the Atlantic Coast of North America commences in May and culminates in spawning in July, and thus lasts for only about 8 weeks (Bigelow & Schroeder 1953b; Beamish et al. 1979). The timing of the spawning run of the landlocked form of P. marinus is similar (Applegate 1950). In contrast, the spawning migration of L. ayresii, which lasts for about 6 months, commences in September to February and culminates in spawning between April and June (R. J. Beamish 1980). The duration of the spawning run of E. tridentatus is even longer than that of L. ayresii with the immature adults entering rivers in April to August and not reaching maturity and spawning until the following April to July (R. J. Beamish 1980). The adults of the landlocked form of P. marinus typically migrate upstream at night and use tactile and possibly hydraulic cues, before sunrise, to seek resting places, such as under rocks or the river bank, a behavior that breaks down, however, if suitable refuge cannot be found (Hardisty & Potter 1971b; Dauble et al. 2006; Binder & McDonald 2007). Further, the typical diel pattern of activity exhibited by P. marinus during its spawning run is modified by temperature extremes; the animals become inactive at night when temperatures are 20°C (Binder & McDonald 2008b). Temperature-induced changes in diel activity appear to represent an adaptive behavior that increases the probability that Lampreys will reach their spawning grounds within the narrow thermal range required for successful embryonic development (Binder & McDonald 2008b). Experimentally blinded individuals of maturing P. marinus showed a similar propensity to migrate upstream and

at the same rate as control individuals, indicating that vision does not play a role in the behavior of this species during its spawning migration (Binder & McDonald 2007). Indeed, the marked diel pattern of locomotory activity in Lampreys is largely controlled by the synthesis and secretion of melatonin during nighttime by the pineal gland, which forms an essential component of the photoneuroendocrine system that allows Lampreys to measure and keep the time (Korf et al. 1998). Melatonin biosynthesis is regulated by signals from photoreceptors perceiving and transmitting environmental light stimuli and by endogenous oscillators that generate a circadian rhythm that does not depend on any environmental time cue. The over-riding role of the pineal gland in regulating locomotor activity in adult Lampreys is emphasized by the following results obtained from experiments with L. camtschaticum. Light-dark entrainment continued in 73% of Lampreys after their eyes had been removed but was completely abolished by subsequent pinealectomy and abolished in most animals that were pinealectomized but had intact eyes (Morita et al. 1992). Further, free-run locomotor activity in constant darkness and animals that had been entrained on a light-dark cycle was abolished by pinealectomy, demonstrating that the pineal plays a crucial role in circadian locomotor activities in Lampreys. The ability of upstream migrating Lampreys to avoid light during the day is facilitated by the presence in their tails of well-developed photoreceptors, which form part of the lateral-line system (Deliagina et al. 1995; Binder & McDonald 2007, 2008a). When the anadromous P. marinus encounters rapid flow during its upstream migration, it alternates short movements of about 67 s duration and periods of rest of about 99 s (Quintella et al. 2009). Further, when this species is faced with difficult passage areas on its spawning run, it spends nearly half of its time attached to the substrate by its suctorial disc, emphasizing that such a use of the disc enables energy to be conserved during this non-trophic period of the lifecycle (Quintella et al. 2009). The value of the suctorial disc to Lampreys during their spawning run is also demonstrated by the experimental results obtained when upstream migrants of E. tridentatus were presented with a 1.4 m vertical weir, thus simulating the type of impediment that Lampreys sometimes encounter on their spawning run (Kemp et al. 2009). The Lampreys used their suctorial disc to gain purchase by attaching to the surface of the weir and launching bursts of modified anguilliform motion. These bursts of ascent, which typically lasted for one-fifth of the total ascent time,

PETROMYZONTIDAE: LAMPREYS

were interspersed with stationary attachment using the suctorial disc. The ratio of time spent actively climbing to time spent resting decreased with distance traveled, presumably due to increased fatigue. The ability to successfully complete climbs increased with learning experience, whereby Lampreys adopted more efficient approaches or used easier routes. Adult Lampreys can even circumnavigate obstacles to their upstream migration, such as those produced by dams, weirs, and waterfalls, by moving overland through moist areas, such as those provided in grass and low-lying vegetation, and re-entering the river above the obstacle (Potter et al. 1983). The capacity of upstream migrating Lampreys to survive in moist environments outside the river is facilitated by their ability to extract substantial volumes of oxygen from their environment. In laboratory experiments, the branchial region was responsible for 87% of oxygen uptake and 80% of carbon dioxide excretion by adults of G. australis in a moist environment; the gills of these animals presumably retain their integrity in air, continuing to facilitate gas exchange (Potter et al. 1997). Although adult Lampreys are relatively poor swimmers (Beamish 1974; Dauble et al. 2006), some species cover considerable distances during their spawning run. For example, E. tridentatus was caught 400 km upstream from the mouth of the Skeena River (R. J. Beamish 1980). In the case of anadromous P. marinus, the energy cost of migrating from the estuary of the St. John River, New Brunswick, to its redds 140 km upstream was estimated to be 300 kcal for males and 260 kcal for females (Beamish 1979). Remarkably, however, this amount of energy use was not as great as that estimated for nest construction and vigorous spawning activities. The sex ratio of upstream migrant and spawning aggregations of well-established populations of Lamprey species typically ranges from close to parity to a variable excess of males with the proportion of males being greatest in large spawning aggregations (Hardisty & Potter 1971b; Beamish & Potter 1975; F. W. H. Beamish 1980b; Beamish & Austin 1985).

Upstream Spawning Migration of Nonparasitic Lampreys As they approach maturity, the adults of nonparasitic species emerge from the substrate in which they were burrowed during metamorphosis and undergo a short upstream migration and then spawn and die. The body cavity of females of many nonparasitic species that are about to spawn becomes

131

Figure 3.25. Southern Brook Lamprey, Ichthyomyzon gagei, showing the abdomen distended with mature eggs (photograph by and used with permission of G. Adams).

distended through the presence of the large and fully developed eggs that can sometimes be seen through the body wall (Fig. 3.25). Because nonparasitic Lampreys do not feed after the completion of larval life, the energy they require during metamorphosis, gonadal maturation, and the activities associated with spawning is derived largely from the substantial lipid reserves that are laid down, particularly during the later stages of larval life. In the period between the onset of metamorphosis and spawning in I. gagei, lipid and protein were catabolized to such an extent that these two components at spawning had declined to about 7 and 45%, respectively, of their levels at the commencement of metamorphosis (Beamish & Legrow 1983).

Spawning of Lampreys The spawning behavior of all Lampreys is essentially the same, irrespective of whether they are parasitic or nonparasitic (e.g., Hardisty & Potter 1971b; Mundahl & Sagan 2005; Kucheryavyi et al. 2007). As spawning time approaches, Lampreys become far more active during the day and commence nesting activity. Although spawning typically occurs in late spring or summer, the precise time varies among species according to the critical water temperatures required for spawning activity by those species. For example, in North America, the critical temperature of 22°C for I. unicuspis is greater than the 15.5°C for landlocked P. marinus and the 10.6–11.1°C for L. richardsoni (see Hardisty & Potter 1971b). When spermiating, the males of the landlocked Sea Lamprey release a pheromone (3-keto petromyzonol sulfate) through the gills, which lures ovulating but not pre-ovulating females over hundreds of meters (Siefkes et al. 2005; Johnson et al. 2009). Most Lamprey species spawn on open gravel substrate in shallow water (Cochran & Gripentrog 1992). The species of Ichthyomyzon are atypical, however, in that they

132

FRESHWATER FISHES OF NORTH AMERICA

Figure 3.27. The pairing act in Lampreys (reprinted from The Biology of Lampreys, Vol 1, M. W. Hardisty and I. C. Potter, eds., M. W. Hardisty and I. C. Potter, The General Biology of Adult Lampreys, pages 127–206, 1971).

often spawn under cover (e.g., boulders, woody debris, vegetation) where current is reduced (Cochran & Gripentrog 1992). This allows these species to reproduce in streams or rivers where the current is too strong in open water areas to allow successful spawning, and it decreases risk from visual predators. Nest construction initially involves the use of the suctorial disc to move stones and thus create a depression. The development of the nest cavity is enhanced by the vigorous activity of the Lampreys that suspends particles in the water column that are then swept downstream (Fig. 3.26). During spawning, which occurs in groups or pairs (e.g., Mundahl & Sagan 2005; Kucheryavyi et al. 2007), the female uses her suctorial disc to attach herself to the larger stones, which are located at the front of the nest facing the direction of water flow and often just upstream of riffles (Fig. 3.26). The male attaches himself to the head of the female and then coils around the female until his urogenital papilla and the cloaca of the female are apposed. The coiling action of the male and the vibrations of the female result in simultaneous emission of eggs and sperm (Fig. 3.27). The fertil-

ized eggs become buried in the substrate, a process facilitated by the disturbance caused to the substrate by the vigor of the spawning activities. Lethenteron camtschaticum tends to breed in groups where the current is appreciable (>0.5 m/s) and in pairs when they are in inshore regions of the river where the flow is slower (Kucheryavyi et al. 2007). Further, when L. camtschaticum spawns in pairs, the males exhibit agonistic behavior by fighting back other males.

Fecundity The mature eggs of Lampreys are about 1 mm in diameter, irrespective of whether the species is parasitic or nonparasitic (Hardisty 1964). The fecundity of Lamprey species is related broadly to the body size of their mature adults. Thus, the mean fecundity of large anadromous species, such as P. marinus, is about 170,000, compared with 8,000 to 19,000 in smaller species (e.g., the anadromous L. fluviatilis and the freshwater parasitic species I. unicuspis and T. spadiceus) and is typically 65% (Mundahl et al. 2005). The species composition of algae ingested by larval P. marinus and La. appendix was similar to and comparable with those in the sediment and water in the vicinity of the ammocoete’s burrow, except in the case of filamentous and epipsammic forms, which were not ingested because of their large size (Moore & Beamish 1973). The times taken for complete evacuation of algal cells from the gut at 16 and 2.5°C were 54 and 70 h, respectively, and 45 and 90% survived passage through the gut in summer and winter, respectively. Ammocoetes of I. fossor in oligotrophic streams selectively fed mainly on biofilm fragments that were suspended in the water column (i.e., in the seston) (Yap & Bowen 2003). The growth and condition of larval Lampreys is related directly to the productivity of their habitats and inversely to the density of the ammocoetes, the latter probably being due in part to the suppressing effects of the release of

133

some chemical or biological agent into the surrounding water (Morman 1987; Morkert et al. 1998; RodríguezMuñoz et al. 2003; Yap & Bowen 2003). Further, in the nonparasitic species I. gagei and L. aepyptera the sex ratios of ammocoetes vary greatly among populations of these species, and the percentage of males increases with larval density (Beamish 1993; Docker & Beamish 1994). This relationship with density and with other environmental variables in the case of I. gagei suggests that sex determination is influenced by environmental factors (Beamish 1993). Because Lampreys do not possess the hard structures whose annuli (rings) are typically used for aging gnathostomatous fishes (e.g., otoliths, vertebrae, fin rays, scales), most estimates of the age compositions of individuals in larval assemblages of single species have relied on tracing modal progressions in length-frequency distributions through sequential samples. These estimates provided overwhelming evidence that the larval phase is protracted and typically lasts for 3–7 years (e.g., Hardisty & Potter 1971a; Potter & Bailey 1972; Beamish & Potter 1975; Beamish & Austin 1985). In the case of the ammocoetes of I. greeleyi, oxytetracycline marking indicated that the bands or annuli in their statoliths, a structure analogous to teleost otoliths, were formed each year and could thus be used for aging individuals. Counts of annuli indicated that the larval period of this species is on average relatively long, lasting for 4.2 or 5.2 years (Medland & Beamish 1987). Another elegant study emphasized that the number of annuli in the statoliths of larval landlocked Sea Lampreys often did not accurately estimate the ages of ammocoetes (Dawson et al. 2009). This conclusion was based on comparisons between the known ages of ammocoetes and that estimated from the number of annuli on their statoliths and comparisons involving microsatellite data of the ages determined from statoliths with those based on parental assignment. Dawson et al. (2009) concluded that a combination of bias-corrected ages derived from statolith annuli and length-frequency data substantially increased the accuracy and precision of estimating the ages of ammocoetes. In any assessment of the age composition of larval Lamprey assemblages, however, it should be recognized that, on average and particularly among nonparasitic species, females metamorphose at a larger size and thus at a presumably older age than males (Hardisty 1965; Purvis 1970; Beamish & Austin 1985; Docker & Beamish 1994). Some of the characteristics exhibited during larval life that influence the timing of metamorphosis were revealed

134 FRESHWATER FISHES OF NORTH AMERICA

in a series of studies on P. marinus. In one of those studies, adults of a landlocked population of the Sea Lamprey were isolated in 1960 in a tributary of the Great Lakes in which they then spawned (Manion & Smith 1978). Tracing of the lengths of the resultant progeny in subsequent years showed that the growth of these ammocoetes was markedly asymptotic and that metamorphosis was size dependent and did not occur in that tributary until the ammocoetes were ≥5 years old (Potter 1980a). Notably, some ammocoetes had not even entered metamorphosis at 12 years old when the study was terminated. Although the regime in this study was in some ways artificial, variability in the age at metamorphosis is almost certainly a characteristic of all Lamprey populations, even if it is less extreme in established populations in reasonably productive rivers. Petromyzon marinus typically metamorphoses only when a particular condition factor is attained, which in turn, reflects the accumulation of large lipid stores by the body (Lowe et al. 1973; Youson et al. 1979, 1993). Most of these energy reserves, which are required during subsequent months when Lampreys do not feed, are mainly accumulated during a period at the end of larval life when ammocoetes do not increase in length. Nevertheless, when conditions are particularly favorable for growth (i.e., abundant food and low ammocoete densities), individuals in landlocked populations of P. marinus do not undergo an arrested growth phase at the end of larval life and can enter metamorphosis after only 2 years of life (Morkert et al. 1998), presumably having accumulated the requisite lipid deposits. Metamorphosis in Northern Hemisphere species becomes morphologically detectable between early and midsummer (e.g., Potter & Bailey 1972; R. J. Beamish 1980; Youson 1980; McGree et al. 2008). The subsequent internal and external changes involved in the transition from ammocoete to adult are highly synchronized (Youson et al. 1977; Potter et al. 1978; Youson & Connelly 1978; Bird & Potter 1979ab; Hilliard et al. 1983; Holmes et al. 1999; Youson 2004) and at least in the case of P. marinus, are stimulated by a rise in water temperature in spring (Youson 2003). In the laboratory, density, starvation, and photoperiod did not stimulate metamorphosis (Youson 2003). The concentration of thyroid hormones in the serum of ammocoetes peaks just before metamorphosis and then declines sharply as metamorphosis is initiated (Lintlop & Youson 1983). This explains why metamorphosis can be induced precociously in Lampreys by the administration of most goitrogens, which inhibit the synthesis of thyroid

hormones (Youson 2003). The synchrony of external and internal metamorphic changes implies that the environmental cues and endocrine factors required to initiate metamorphosis are tightly integrated (Youson et al. 1977; Potter et al. 1978; Youson & Potter 1979). Although Lampreys remain burrowed during metamorphosis, they tend, toward the end of metamorphosis, to move out into faster-flowing areas where the substrate is coarser and oxygen in the interstitial spaces is more likely to be continually replenished (Potter 1980a). When fully metamorphosed, the resultant young adults undertake a largely passive downstream movement, which is highly synchronized, occurs at night, and is initiated by increases in freshwater discharge (Applegate & Brynildson 1952; Potter 1980a). In the landlocked and anadromous forms of the Sea Lamprey, this downstream migration typically peaks in autumn and during flood conditions in the following spring (Hardisty & Potter 1971b; Potter & Beamish 1977).

Ecology of Feeding Adults In the lacustrine extensions of the large St. John River system, eastern Canada, the young adults of the anadromous form of P. marinus that did not migrate downstream in the autumn are prevented from undergoing such a movement during winter and spring because these water bodies are frozen during that period. Thus, in this river, many fully metamorphosed P. marinus do not become active until the ice melts in May, but then immediately start feeding on the Alewife (Alosa pseudoharengus), American Shad (Alosa sapidissima), and White Sucker (Catostomus commersonii) (Fig. 3.18a; Potter & Beamish 1977). This enables their lipid reserves and hemoglobin concentrations, which had become greatly depleted during the preceding nontrophic months, to be replenished prior to their migration to the sea (Potter & Beamish 1978; Beamish et al. 1979). After feeding for only 1–2 months at sea, some young adult Sea Lampreys are transported far back into the St. John River by upstream-migrating Atlantic Salmon (Salmo salar) to which they have attached (Fig. 3.18b; Potter & Beamish 1977). During its marine phase, which is estimated to last for 2–2.5 years, the anadromous Sea Lamprey feeds initially on small benthic marine fish species, such as redfishes (Sebastes spp.) and the Silver Hake (Merluccius bilinearis). As it increases in size, it then targets large and wideranging pelagic fishes, such as the Atlantic Mackerel (Scomber scombrus), Atlantic Salmon, Swordfish (Xiphias

PETROMYZONTIDAE: LAMPREYS

gladius), Bluefin Tuna (Thunnus thynnus), and Basking Shark (Cetorhinus maximus) and even marine mammals such as the North Atlantic Right Whale (Eubalaena glacialis). This results in this Lamprey species becoming widely distributed in the marine environment (F. W. H. Beamish 1980b; Halliday 1991; Nichols & Hamilton 2004). The ability of P. marinus to feed on sharks is facilitated by their possession of efficient mechanisms for rapidly excreting the large volumes of urea that are ingested when feeding on elasmobranch blood (Jensen & Schwartz 1994; Wilkie et al. 2004). The landlocked P. marinus feeds parasitically on a number of fish species for 12–20 months, with growth in weight being pronounced and linear between June and September and even increasing in October, after which most individuals were at or approaching the size when feeding ceases and the spawning migration commences in the following spring (Applegate 1950; Smith & Tibbles 1980; Bergstedt & Swink 1995). The above period is shorter than the 18–30 months proposed for the parasitic phase in anadromous P. marinus (F. W. H. Beamish 1980b; Halliday 1991), which is consistent with the far smaller maximum size of the landlocked form. The destructive effects of landlocked P. marinus on its hosts were an important contribution to the catastrophic decline that occurred in the abundance of the commercially important Lake Trout (Salvelinus namaycush) in the upper Great Lakes during the 1940s and 1950s (Smith 1971; Smith & Tibbles 1980). Given that the Lake Trout occupies deep waters, the feeding adults of the landlocked Sea Lamprey apparently prefer relatively low water temperatures. The landlocked Sea Lamprey selectively attacks large Lake Trout but feeds less frequently on and is less likely to kill the more agile Seneca strain than other strains of this species (Schneider et al. 1996; Madenjian et al. 2004) (see commercial importance section). Like anadromous P. marinus, the young adults of E. tridentatus begin feeding in either fresh or salt water, sometimes as early as mid-October, only 3–4 months after the onset of metamorphosis (R. J. Beamish 1980). After entering the sea, E. tridentatus moves into waters >20 m deep and targets species such as the Sockeye Salmon (Oncorhynchus nerka) and Pink Salmon (Oncorhynchus gorbuscha), which aggregate in these waters before entering rivers on their spawning runs. When L. ayresii enters the sea, nearly a year after the commencement of metamorphosis, it attacks and removes large amounts of flesh from the Pacific Herring (Clupea pallasii) and Oncorhynchus spp. (R. J. Beamish 1980). The adult parasitic phase of the Pa-

135

cific Lamprey can last for 3.5 years, but that of the Western River Lamprey is typically completed within 4 months. Entosphenus macrostomus, the landlocked derivative of E. tridentatus, feeds predominantly on freshwater salmonids and juveniles of the anadromous Coho Salmon (Oncorhynchus kisutch) (Beamish 2001). The three parasitic species of Ichthyomyzon, which are all confined to fresh water and attain relatively similar body sizes, feed on a wide range of actinopterygian fishes (Hubbs & Trautman 1937; Hardisty & Potter 1971b; Cochran & Jenkins 1994; Renaud 2002; Cochran et al. 2003; Cochran & Lyons 2004). The hosts of I. unicuspis include the Paddlefish (Polyodon spathula), Lake Sturgeon (Acipenser fulvescens), Northern Pike (Esox lucius), and Muskellunge (Esox masquinongy). Ichthyomyzon unicuspis and I. castaneus, which attack similar hosts, feed below the ice during winter (Cochran et al. 2003). The parasitic phase of I. unicuspis lasts for about a year with growth occurring mainly in summer (Cochran & Marks 1995). The adults of Ichthyomyzon species typically forage at night, which would enhance their foraging efficiency because they would be less visible than during the day and thus could approach the host more effectively, especially if the host is quiescent at night (Cochran 1986a). Adult Lampreys possess a well-developed olfactory organ and prominent olfactory lobes in the brain, which strongly suggests that they use olfaction when seeking prey (Kleerekoper 1972). This conclusion is consistent with a laboratory study involving adult P. marinus, in which the introduction of water in which Lake Trout had been held into their holding tank immediately elicited vigorous activity (Kleerekoper 1972). Under natural conditions in the wild, the adult Lamprey presumably uses olfaction (a relatively long-distance sense) to locate its prey, after which it probably uses its eyes to target the prey more precisely and ensure that it attaches to an appropriate region of the host’s body (see physiology section; Farmer 1980).

Blood versus Flesh Feeding Analysis of the gut contents of adults of the landlocked P. marinus, which had fed on Lake Trout and Rainbow Trout (Oncorhynchus mykiss) whose red blood cells had been labeled with chromium-51, revealed that blood constituted >98% of the food ingested by this species (Farmer et al. 1975). Inspection of gut contents and hematological tests showed that the food of I. bdellium, I. castaneus, and I. unicuspis also consisted predominantly of blood (Renaud et al. 2009a). In marked contrast to P. marinus, which

136

FRESHWATER FISHES OF NORTH AMERICA

produces a small and often well-defined hole through which body fluids are then extracted (Fig. 3.18), L. camtschaticum and L. ayresii remove large amounts of flesh from their teleost prey (Fig. 3.20; Nikolskii 1956; R. J. Beamish 1980). Tetrapleurodon spadiceus represents an intermediate feeding mode in that it ingests both blood and flesh (Álvarez del Villar 1966; Cochran et al. 1996). Petromyzon marinus tends to attach to the ventral surface and near the pectoral fins of its host, where scales are reduced, the musculature is thin, and a well-developed vascular supply is nearby (Fig. 3.18), thus optimizing the potential to obtain a plentiful supply of blood (Davis 1967; Potter & Beamish 1977; King 1980; Cochran 1986b). In contrast, the adults of Ichthyomyzon, which also mainly ingest blood, tend to attach to the dorsal surface of their hosts in shallow waters (Cochran 1986b; Renaud 2002), where pelagic species are largely absent. Attacks by I. unicuspis on the Paddlefish, however, occur in deep water, and the attachment is ventral (Cochran 1986b). Ichthyomyzon unicuspis also attaches within the branchial cavity of the Paddlefish, which would provide access to blood under pressure, especially from that in the ventral aorta, and protection from dislodgement during breaching by the Paddlefish (Cochran & Lyons 2010). Small, recently metamorphosed P. marinus often feed on small hosts, whose scales and skin are thinner than those of larger individuals of the same species, but large P. marinus and other blood-feeding Lampreys tend to attack large hosts (Cochran & Jenkins 1994; Cochran et al. 2003). The latter tendency is advantageous because a large host is more likely than a small host to survive an attack because it involves the destruction of a relatively smaller part of the body surface. The ability of fishes to recover from attacks by blood-feeding Lampreys is illustrated by the observation that 27% of Muskellunge that had been recently attacked by I. unicuspis had healed wounds attributable to earlier Lamprey attacks (Renaud 2002). Lampetra ayresii and its European counterpart, L. fluviatilis, remove large chunks of flesh from the dorsal surface of their hosts, which is the main site of attachment of flesh-feeding species (Fig. 3.20; Bahr 1933; R. J. Beamish 1980; Cochran 1986b). In contrast to blood-feeding Lampreys, flesh-feeding species typically attack small, schooling teleosts, which results in the relatively rapid death of those hosts (Cochran & Jenkins 1994). The abundant pool of prey provided by schooling fish enables flesh-feeding Lampreys to readily find another prey once they have killed their current host.

Parasitism, Commensalism, and Predation The parasites of P. marinus, I. castaneus, E. tridentatus, L. camtschaticum, L. appendix, and L. richardsoni have been studied in various North American localities (reviewed by Appy & Anderson 1981) with the majority of information being provided by the landlocked form of P. marinus in the Great Lakes. Parasites include species belonging to Bacteria (Pseudomonas and Aeromonas spp.), Fungi (Saprolegnia spp.), Protozoa (Ichthyophthirius and Trichodina spp.), Platyhelminthes (Diplostomulum, Diplostomum, Podocotyle, Nanophyetus, Eubothrium, Triaenophorus, Phyllobothrium, and Proteocephalus spp.; and larva of monorchiid spp.), Acanthocephala (Neoechinorhynchus and Metaechinorhynchus spp.), Aschelminthes (Truttaedacnitis, Cystidicola, and Eustrongylides spp.), Annelida (Pisicola spp.), and Arthropoda (Ergasilus spp.). Although no petromyzontid from North America is documented as host to freshwater unionid mussels, the glochidial stage of the unionid mollusk Anodontoides ferrussacianus was found on the gills of landlocked P. marinus; however, the metamorphosis of this parasite was not observed (Wilson & Ronald 1967). The eggs and early larval stages of Lampreys are preyed on by various fishes, including logperches (Percidae), Eels (Anguillidae), minnows (Cyprinidae), Sculpins (Cottidae), Sticklebacks (Gasterosteidae), and trouts (Salmonidae) (Hardisty & Potter 1971a). Adult Lampreys are preyed on in fresh waters by fishes, snakes, birds, and mammals and in the sea by fishes and mammals (Renaud 1997; Cochran 2009).

CONSERVATION Renaud (1997) and Jelks et al. (2008) reviewed the conservation status of Northern Hemisphere Lampreys and North American Lampreys (Table 3.3), respectively. Fifteen of the 23 North American species (65%) according to Renaud (1997) and 10 of 23 (43%) according to Jelks et al. (2008) were considered at some level of risk (Vulnerable, Threatened, or Endangered) in at least part of their North American range. This difference in percentage values is mainly due to differences in the scale of examination in the two studies. For example, Jelks et al. (2008) did not consider the six species of Ichthyomyzon at risk at the continental level, but Renaud (1997) regarded them as at risk at the national or subnational level (states and provinces). Several authors have updated the information on the conservation situation for Lampreys in Canada, California,

PETROMYZONTIDAE: LAMPREYS

137

Table 3.3. Conservation status of Lampreys in North America according to Jelks et al. (2008). Status

Listing Criteria1

NatureServe Global Rank2

Threatened Vulnerable Threatened Endangered Threatened Vulnerable Threatened

1, 2, 4, 5 1, 5 5 1, 2, 5 1, 5 1, 2 1, 5

G1G2 G3G4 G1 G1 G3G4Q G5 G5T1

Goose Lake population Lampetra ayresii Lampetra richardsoni L. richardsoni

Vulnerable Not at Risk Endangered

1, 4 Not Applicable 1, 5

G4 G4G5 G4G5T1Q

Morrison Creek population Tetrapleurodon geminis Tetrapleurodon spadiceus

Threatened Endangered

1, 5 1, 2, 5

Not Applicable Not Applicable

Taxon Entosphenus hubbsi Entosphenus lethophagus Entosphenus macrostomus Entosphenus minimus Entosphenus similis Entosphenus tridentatus E. tridentatus

1 = habitat destruction; 2 = over-exploitation of the species or its host or intentional eradication; 4 = predation by nonindigenous species; 5 = restricted range 2 G1 = critically imperiled; G2 = imperiled; G3 = vulnerable; G4 = apparently secure; G5 = secure; T1 = critically imperiled infraspecific taxon; Q = questionable taxonomy 1

Kansas, and Oregon (Close et al. 2002; Haslouer et al. 2005; Moyle et al. 2009; Renaud et al. 2009b). Summaries of the conservation status of Lampreys in the United States and Canada are updated on an irregular basis on the NatureServe website (NatureServe 2010) and also on a website that includes Mexico, maintained jointly by the U.S. Geological Survey and American Fisheries Society (USGS 2010) (taken from Jelks et al. 2008). According to NatureServe (2010; Table 3.3), five West Coast endemic taxa (species or infraspecific rank) are either Critically Imperiled (E. macrostomus, British Columbia; E. minimus, Oregon; Goose Lake population of E. tridentatus, Oregon and California; and Morrison Creek population of L. richardsoni, British Columbia) or Critically Imperiled / Imperiled (E. hubbsi, California). In Mexico, T. spadiceus is considered Endangered (Cochran et al. 1996; Jelks et al. 2008). Although the discovery of populations of E. minimus (Lorion et al. 2000), a species assumed to be extinct, is encouraging, some species of Lampreys have clearly suffered from deleterious anthropogenic effects and further work on their conservation status is likely to reveal that more populations are at risk. Since 1989, the situation for E. hubbsi, E. macrostomus, and the Goose Lake population of E. tridentatus has deteriorated (Jelks et al. 2008). The major threats remain habitat degradation and the effects of stream regulation, and

restricted ranges place West Coast endemics at risk (Table 3.3). In the Laurentian Great Lakes, the continued use of lampricides, as part of the Sea Lamprey Control Program, is affecting the non-targeted native species of Lampreys, and particularly I. unicuspis, and is thus also a cause for concern (McLaughlin et al. 2003). The susceptibility of Lampreys to modifications to rivers is well illustrated by the fact that the anadromous, parasitic E. tridentatus can no longer enter Elsie Lake, British Columbia, because the construction of a dam now prevents the downstream and upstream movement of this species (Beamish & Northcote 1989). Although the individuals of the Pacific Lamprey that became landlocked by this barrier fed on resident salmonids in the lake, they did not subsequently reach maturity and spawn. Thus, landlocking of a population of anadromous species of Lamprey does not necessarily lead to the development of a viable population of the landlocked form of that species.

COMMERCIAL IMPORTANCE Although adult Lampreys are fished commercially in numerous rivers in Europe, their potential as a food source or delicacy has not been exploited widely in North America, and as a consequence, the declines in abundances of

138

FRESHWATER FISHES OF NORTH AMERICA

the parasitic species have often not been fully appreciated. Native Americans, however, value the Pacific Lamprey so highly that this species is a cultural icon. This led to the initiation of research aimed at developing methods for restoring the populations of E. tridentatus at least in the Columbia River basin (Close et al. 2002). The over-riding economic importance of Lampreys in North America resides in the detrimental effect that they have had on fish populations. A dramatic example of such an effect is provided by the deaths caused to an estimated 60 million juvenile fishes in the Strait of Georgia in 1975 by the attacks of an estimated 667,000 Western River Lampreys (R. J. Beamish 1980). Even more impressive, however, is the fact that, in 1990 and 1991, the same species, which normally attacks the Pacific Herring, killed a minimum of 20 and 18 million Chinook Salmon, Oncorhynchus tshawytscha, respectively, and a minimum of 2 and 10 million Coho Salmon, respectively (Beamish & Neville 1995). Overall, in 1991 the Western River Lamprey killed about 65 and 25% of the total Canadian hatchery and wild production of Coho Salmon and Chinook Salmon, respectively, in the Strait of Georgia. Another striking example of the massive influence of Lampreys on host populations is provided by the destruction caused to fish stocks, and particularly those of the Lake Trout, by the landlocked form of P. marinus when it invaded the upper Great Lakes (Smith 1971; Smith & Tibbles 1980; Christie & Goddard 2003). Although the first record of the Sea Lamprey in the Lake Ontario basin dates from 1835, mtDNA analyses strongly indicate that this landlocked form of P. marinus is a natural post-Pleistocene colonizer of this large water body (Waldman et al. 2004, 2009); this conclusion, however, has been questioned, (Eshenroder 2009). By the end of the 1800s, records of attacks by Sea Lampreys on Lake Ontario fishes and particularly Lake Trout were numerous. The Sea Lamprey entered Lake Erie from Lake Ontario in 1919 following the deepening of the Welland Canal between those two lakes, which provided a bypass to the barrier posed by Niagara Falls. Petromyzon marinus spread rapidly through the Great Lakes, bringing about rapid declines in, e.g., Lake Trout in Lake Huron between 1935 and 1945, in Lake Michigan between 1945 and 1950, and in Lake Superior between 1950 and 1960. The extreme effects of the landlocked Sea Lamprey on fish populations led to a substantial investment in research, and this resulted, through the detailed studies of Vernon Applegate and his colleagues, in a sound knowledge of the biology of this species in the Great Lakes (Applegate 1950). Since

that time, various measures have been introduced to control P. marinus in the Great Lakes. When mechanical and electrical barriers proved ineffective at preventing or killing Lampreys during their migration to spawning areas, attention was turned to using chemical treatments to kill ammocoetes. After testing >6,000 chemicals, 3-trifluoromethyl-4-nitrophenol (TFM) was selected for use. This compound was more toxic to ammocoetes than other aquatic organisms, could be handled readily in the field, was effective at low concentrations, and was relatively inexpensive. TFM was first used in the late 1950s and in the 1960s. A second larvicide, the molluskicide Bayer 73 (2', 5-dichloro-4'-nitrosalicylanilide) enhanced the effects of TFM and was effective on its own when applied directly to the substrate surface. The use of chemical larvicides was successful and led to a substantial reduction in the number of spawning Lampreys. Recognition of the importance of developing more natural and integrated methods of Lamprey control led, e.g., to the testing of sterile male release techniques and the possible application of pheromones to influence migratory and spawning success (Bergstedt et al. 2003; Twohey et al. 2003ab; Johnson et al. 2009). The great relevance of the landlocked Sea Lamprey to the survival of fisheries in the upper Great Lakes is recognized by a huge investment in research and control measures, and the establishment of an Integrated Management of Sea Lamprey (IMSL) plan by the Great Lakes Fishery Commission. For further details of the Sea Lamprey in the upper Great Lakes, the reader is referred to the numerous papers that were presented on this topic at a Sea Lamprey International Symposium (SLIS II) and that were published in a special edition of the Journal of Great Lakes Research (2003, vol. 29, Supplement 1).

LITERATURE GUIDE The four volumes of The Biology of Lampreys, edited by Hardisty & Potter (vol. 1, 1971; vol. 2, 1972; vol. 3, 1981; vols. 4 a & b, 1982, Academic Press, London), provide accounts of a wide range of aspects of the biology of Lampreys. The Proceedings of the Sea Lamprey International Symposia I and II, which were published in special issues of the Canadian Journal of Fisheries and Aquatic Sciences (1980, vol. 37) and the Journal of Great Lakes Research (2003, vol. 29), provided further information on the biology of Lampreys, and detailed accounts of the invasion, effects, and control of the landlocked Sea Lamprey in the

PETROMYZONTIDAE: LAMPREYS

Great Lakes of North America. A broad, innovative, and easily approachable account of Lampreys was provided by the doyen of Lamprey biology, Martin Hardisty, in his final work entitled Lampreys: Life without Jaws (2006). Papers relating to contemporary views on the systematics of Lampreys and of the relationships of the cyclostomes and gnathostomes include Gill et al. (2003), Janvier (2009), Lang et al. (2009), and Near (2009). Several papers on Lampreys, including Margaret Docker’s excellent account of paired species, were published in the Proceedings of the Biology, Management, and Conservation of Lampreys in North America Symposium in 2009 by the American Fisheries Society edited by Brown et al. (2009). In the Food and Agriculture Organization of the United Nations Spe-

139

cies Catalogue entitled Lampreys of the World, Renaud (2011) gives an overview of the biology of all species.

Acknowledgments We are particularly grateful to Professors Helmut Bartels, Max Cake, Phil Cochran, Shaun Collin, Margaret Docker, Peter Holland, and Mike Wilkie for providing very helpful comments and criticisms of those parts of this chapter that covered their areas of expertise. Gratitude is also expressed to Dr. David Bird for providing us with Fig. 3.15, D. Naughton for Fig. 3.21, and to Gordon Thomson and Steeg Hoeksema for producing other figures. We also thank Amy Commens-Carson for redrawing several figures.

Chapter 4

Dasyatidae: Whiptail Stingrays Michael D. Burns, Carter R. Gilbert, and Melvin L. Warren, Jr.

The Whiptail Stingrays (Dasyatidae) are members of the order Myliobatiformes (Stingrays), a group of 10 families that are included, together with 3 other orders, in the subdivision Batoidea of the class Chondrichthyes (Cartilaginous Fishes) (McEachran & Aschlimann 2004; Nelson 2006). Whiptail Stingrays are members of the superfamily Dasyatoidea, which also includes the families Potamotrygonidae (River Stingrays), Gymnuridae (Butterfly Rays), and Myliobatidae (Eagle Rays). The Stingrays are cartilaginous and have depressed bodies with greatly expanded pectoral fins forming a disc. The slender tail bears a serrated, venomous spine capable of inflicting serious wounds and causing excruciating pain. Whiptail Stingrays (and other Stingrays) give birth to live young (ovoviviparity) with the young closely resembling the adults. The family name is from the Greek roots dasyatis, meaning “shaggy, rough,” likely in reference to the hardened portion of the dorsal surface that contains tubercles, thorns, or prickles in some species, and batis, meaning “skate,” in reference to rajiform fishes. The hypothesized sister-family to Dasyatidae, the River Stingrays (Potamotrygonidae), contains 3 genera, Paratrygon, Plesiotrygon, and Potamotrygon with a total of 20 species across the 3 genera. Additional species may be added to the potamotrygonids pending further research because species within two dasyatid genera (Taeniura and Himantura) are hypothesized as being more closely related to River Stingrays. River Stingrays occur in South American fresh waters and represent the sole freshwater family of Stingrays. The River Stingrays have a reduced rectal gland weight and reduced urea, both important in osmoregulation in sea water. The River Stingrays are hypothesized to have arisen in the

Late Cretaceous (99.6– 65.5 mya) or Early Tertiary (65 mya) period (Grande 1980).

DIVERSITY AND DISTRIBUTION Dasyatis sabina, the Atlantic Stingray, is a member of the largest genus in the family Dasyatidae and is the sole North American freshwater species. The Dasyatidae contain about 68 species in 6 genera: Dasyatis (≥38 species), Himantura (≥23 species), Pastinachus (1 species), Pteroplatytrygon (1 species), Taeniura (3 species), and Urogymnus (2 species) (Compagno 1999a, 2005; Nelson 2006). Of these, 5 genera and 20 species occur in the Atlantic Ocean (Western and Eastern combined, including the Mediterranean Sea) and the Eastern Pacific Ocean. Pteroplatytrygon violacea, the Pelagic Stingray (which has a worldwide distribution in tropical and temperate seas), an oceanic species, occurs in all three regions, and Dasyatis centroura (Roughtail Stingray) is found in both the Western and Eastern Atlantic Ocean. Adjusting the overall totals to account for the extended distributions of these 2 species, 3 species are limited to the Eastern Pacific (4 total), 6 restricted to the Western Atlantic (8 total), and 9 confined to the Eastern Atlantic (11 total). The genus Dasyatis with ≥38 species is circumglobal in tropical and warm temperate regions, occurring in the Atlantic, Pacific, and Indian Oceans (Table 4.1). A few species of Dasyatis, Himantura, and the Cowtail Stingray (Pastinachus sephen) occur permanently or sporadically in tropical or subtropical freshwater lakes and rivers (Thorson & Watson 1975; Taniuchi 1979; Compagno & Roberts 1982, 1984; Otake 1991; Nelson 2006). The Atlantic Stingray occurs in coastal

DASYATIDAE: WHIPTAIL STINGRAYS

141

Plate 4.1. Atlantic Stingray, Dasyatis sabina

and brackish waters (occasionally in fresh water) in the Atlantic Ocean from Chesapeake Bay south to Florida and in the Gulf of Mexico from Florida to the Campeche Gulf and the tip of the Yucatan Peninsula (Ross & Burgess 1980; Johnson & Snelson 1996). Although most records are from shallow inshore waters (typically 183 m) (identified by W. C. Schroeder, Harvard Museum of Comparative Zoology, MCZ 51812). No confirmed records exist from the Bahamas (Böhlke & Chaplin 1968), Cuba (contra Duarte-Bello & Buesa 1973), or other insular areas. Alleged occurrences in the Caribbean Sea and southward (Meek & Hildebrand 1923) are based on misidentifications or unverified supposition of occurrence (Bigelow & Schroeder 1953a; McEachran & de Carvalho 2003). The Atlantic Stingray is the only Stingray in North America known to enter fresh water. A reproducing population occurs within fresh water in the St. Johns River system, Florida (Fig. 4.1) (McLane 1955; Tagatz 1968; Johnson & Snelson 1996) (see physiology section).

PHYLOGE NE TIC RELATIONSHIPS

Higher Relationships The class Chondrichthyes contains two monophyletic evolutionary lineages, the subclasses Holocephali and Elasmobranchii (Lund & Grogan 1997). The Elasmobranchii contains three infraclasses: †Cladoselachimorphi, †Xenacanthimorpha, and Euselachii. Euselachii is the only infraclass with extant species (Nelson 2006 and references cited therein). Within Euselachii, two conflicting phylogenetic hypotheses exist. The first (outlined by Nelson 2006), coined the hypnosqualean hypothesis, places the batoids (Rays) as sister to the Pristiophoriformes (Saw Sharks) and the two are sister to Squantiformes (Angel Sharks), all ultimately sharing a common ancestor with Squaliformes (Dogfish Sharks) (see morphological works by de Carvalho 1996; Shirai 1996). The second hypothesis, summarized in and accepted by

142 FRESHWATER FISHES OF NORTH AMERICA

Table 4.1. Life history information for the genus Dasyatis compiled from numerous sources but mostly for Dasyatis sabina (see text for citations; DW = disc width). Number of extant species 1 or 2 degree freshwater Maximum size recorded in length Maximum age Age and size at first reproduction Iteroparous or semelparous Fecundity estimates Egg deposition sites Clutch size Range of nesting and spawning dates and temperatures Habitat of spawning sites; average water depth Incubation period; larval type at hatching Mean size at hatching Parental care Major dietary items General year-round habitat Migratory or diadromous Imperilment status

About 38 2 Females, 37 cm DW; males, 32.6 cm DW About 9 years Probably age 1+; females, 22–24 cm DW; males, 20.5–25.0 cm DW Iteroparous 2.3 young/female Ovoviviparous Not applicable Fertilization occurs between early March and mid-April at about 25°C, but mating can be protracted (August to mid-April) Similar to general habitat About 3.5–4 months from fertilization to parturition; ovoviviparous so larval type not applicable 96.5 ± 8.0 mm DW Female carries embryos until parturition (about late July) Small benthic crustaceans, polychaete worms, and other small invertebrates 1–2 m deep water; soft, mud bottom mixed slightly with sand toward the littoral zone Little migration known in the freshwater populations Currently stable over historic range

Figure 4.1. Geographic range of the Atlantic Stingray, Dasyatis sabina, in North America.

Dasyatis sabina

Nelson (2006) and used here, is coined the selachianbatoid hypothesis, which considers Squaliformes monophyletic without inclusion of the batoids (see Schwartz & Maddock 2002; Douady et al. 2003; Maisey et al. 2004; Naylor et al. 2005).

Relationships within Dasyatidae Of the six genera within the dasyatids, Lovejoy (1996) and McEachran et al. (1996) placed the genera Taeniura and amphi-American Himantura within the potamotrygonid clade; however, Nelson (2006) placed them within the dasyatid clade, as is done here. Pteroplatytrygon violacea is

sometimes placed in Dasyatis (Nishidia 1990; Lovejoy 1996; Rosenberger 2001b; see also Compagno 1999b), which would render it the only known species of Dasyatis to be mostly pelagic and to solely use an oscillatory-based pectoral motion. Other species of Dasyatis use an undulatory-based pectoral motion (Rosenberger 2001a) (see morphology section). To date, no synapomorphies are known that unite species within Dasyatis or species within Himantura (Lovejoy 1996; McEachran 1996; Rosenberger 2001b). Thus, both genera are either paraphyletic or polyphyletic. The phylogenetic relationships of Dasyatis to other genera or of some species within the genus (see the following subsection) are uncertain. Traditionally, the genera were separated by the use of one character—the presence or absence of tail fin folds (folds occur in Dasyatis and are absent in Himantura, Bigelow & Schroeder 1953a). Resolution of the dasyatid clade would help elucidate evolutionary relationships across the myliobatids (de Carvalho et al. 2004).

Interspecific Relationships Using morphological characters in a parsimony analysis, Rosenberger (2001b) analyzed species relationships for

DASYATIDAE: WHIPTAIL STINGRAYS

143

Figure 4.2. Partial phylogeny of the family Dasyatidae containing 20 species, including the Atlantic Stingray, Dasyatis sabina (redrawn from Rosenberger 2001b).

14 of 35 species within Dasyatis and found several sisterspecies pairs (Fig. 4.2). Dasyatis sabina was basal to a large clade of eight other species of Dasyatis and the genera Himantura and Gymnura. From this work, the genus Dasyatis is not monophyletic (see also Nishidia 1990; Lovejoy 1996; McEachran et al. 1996).

FOSSIL RECORD The Myliobatiformes have a diverse and widespread fossil record that spans both freshwater and marine habitats, but most of the fossil specimens are incomplete and consist of only individual teeth, dermal denticles, and serrated caudal spines (de Carvalho et al. 2004). Fossil records span five geological epochs (55.5–2.6 mya): the Paleocene, Eocene, Oligocene, Miocene, and Pliocene (see compilation by de Carvalho et al. 2004). Presently, more complete fossils are known from only two localities, each representing different paleoenvironments: Monte Bolca Formation, northeastern Italy, and the Green River Formation, Wyoming. The Monte Bolca Formation was a tropical marine reef (de Carvalho et al. 2004) and the Green River Formation a collection of tropical to subtropical lakes. Both localities are Eocene (55.8–33.9 mya) in age (Grande 1984, 2001). The fossil record within Dasyatidae is complex and subject to debate about phylogenetic position of extant species or if the family should be assimilated into the

Potamotrygonidae. †Heliobatis radians, described from Fossil Butte (Eocene, 55.8–33.9 mya) of the Green River Formation, has been placed in Dasyatidae (Grande 1984) and Heliobatidae (de Carvalho et al. 2004). Synonyms of †Heliobatis include †Xiphotrygon acutidens and †Palaeodasybatis (de Carvalho et al. 2004). One other genus is based on incomplete fossil material and is subject to reinterpretation in the future. The genus †Coupatezia occurs from the Middle Eocene (48.6–37.2 mya) of Africa, Europe, and North America and was placed provisionally within Dasyatidae based on tooth structure. North American records include †C. woutersi from Chesapeake sediments (Eocene, 55.8–33.9 mya) (Ward & Weist 1990), †C. woutersi specimens from Lauderdale County, Mississippi (Early Eocene, 55.8–48.6 mya) (Case 1994), and †C. woutersi from Virginia (Early Eocene, 55.8–48.6 mya) (Kent 1999). No fossil records are specifically attributable to Dasyatis sabina.

MORPHOLOGY Stingrays are cartilaginous, dorsally flattened fishes, the flattened portion of the body being referred to as the disc (Figs. 4.3 and 4.4). The gill openings and mouth (with usually protrusible jaws) are on the ventral surface of the body. The pectoral fins are expanded laterally, becoming thin toward the outer edges to form wing-like structures that attach to the body in front of the gills on the sides of

144 FRESHWATER FISHES OF NORTH AMERICA

Figure 4.3. A single individual male of the Atlantic Stingray, Dasyatis sabina, from Jupiter, Florida (photo courtesy of, copyrighted by, and used with permission of David B. Snyder).

Figure 4.4. Ventral view of a female Atlantic Stingray, Dasyatis sabina, in the Mote Aquarium, Sarasota, Florida (photo courtesy of, copyrighted by, and used with permission of Larry Linton).

the head. The pelvic fins are expanded laterally with a convex lateral margin that is partially overlapped by the pectoral fins. The eyes and spiracles are located dorsally with the anterior portion of the head not elevating from the disc. The dorsal fin is absent, and the caudal fin tapers into a filament. The anterior vertebra is fused to form a

synarcual with the suprascapular of the pectoral girdles joined dorsally; the vertebral column is fused with the synarcual. All Whiptail Stingrays have a disc shape that is 380 mm DW with the largest males and females measuring 460 mm and 490 mm DW, respectively. The largest female weighed 5,433 g. Individuals from a Georgia sample had smaller maximum sizes (292 mm DW for males and 405 mm DW for females). Males and females along the Florida east coast and in the St. Johns River typically are smaller (≤300 mm and ≤400 mm DW, respectively) and show bimodal size distributions (Snelson et al. 1988; Johnson & Snelson 1996). The apparent bimodal size distributions of Florida and North Carolina populations, which may have been an artifact of sampling or based partly on misidentifications, likely were more attributable to differences in age, growth, and mortality characteristics of the respective populations (Snelson et al. 1988; Johnson & Snelson 1996). The larger size of the North Carolina samples led Snelson et al. (1988) to suggest the possibility that the North Carolina population is distinctly larger bodied or that size data are confused by misidentifications. A report of four specimens with disc widths (DW) ranging from 458 to 610 mm (Fowler 1926) was questioned as being too large and was almost certainly based on misidentified individuals (Lewis 1982; Snelson et al. 1988).

Spine Morphology and Replacement Stingray tail spines are composed of an inner core of vasodentine and a thin outer layer of enamel-like material (Halstead et al. 1955). Spines have retrorse serrations along the lateral margins, a sharp tip, longitudinal grooves on both the dorsal and ventral surfaces, and

146

FRESHWATER FISHES OF NORTH AMERICA

a raised longitudinal ridge along the ventral surface. The spines are covered by integumentary and glandular tissues. Venom is produced in the tissues covering the spine (Halstead et al. 1955; Haddad et al. 2004; Barbaro et al. 2007). The venom of Dasyatis spp., although apparently not as toxic as that of Neotropical freshwater Potamotrygon spp., nevertheless induces severe pain and depending on the dose may cause shortness of breath, agitation, convulsion, and swelling in victims (Barbaro et al. 2007). Spines are replaced periodically with many individuals in summer in both freshwater and marine populations having two spines. In late May or June an incipient secondary (replacement) spine becomes visible just posterior and ventral to the existing primary spine (Teaf & Lewis 1987; Amesbury & Snelson 1997). Replacement is rapid and synchronous with the mean spine count / individual increasing from one to two over a period of 3 weeks. Nearly all individuals exhibit two spines until early August when specimens with only a single spine become increasingly prevalent. At that time, individuals with a single spine often have a small white scar on the dorsal surface of the tail just anterior to the base of the existing spine, indicating the recent loss of the primary spine, rendering the secondary spine as the new primary spine.

Type of Locomotion Batoid fishes primarily propel themselves in two different ways: pectoral fin–based locomotion or axial-based locomotion (body and tail). Pectoral fin–based locomotion is of two types: undulation and oscillation. Undulation, or rajiform locomotion, occurs when >1 wave is present on the fins at a time; oscillation of the pectoral fins, or mobuliform locomotion, occurs when the fins move up and down with less than half a wave on each fin. The Atlantic Stingray, like most benthic Rays, moves through undulation (Rosenberger 2001a; Table 4.2). They also exhibit an augmented form of pelvic fin punting, or benthic walking, in which the pectoral fins are also undulated as the anterior lobe of each pelvic fin synchronously is protracted cranially, placed into the substrate, and retracted caudally, creating a forward thrust (Macesic & Kajiura 2010).

Morphology and Mechanosensory Function of the Lateral-Line System The lateral-line system in fishes is used to detect water movements relative to the organism’s body surface

Table 4.2. Summary of swimming kinematics for the Atlantic Stingray, Dasyatis sabina (data from Rosenberger 2001). Range Variable Fin-beat frequency (Hz) Mid-disc amplitude (proportion of disc width) Wavespeed (proportion of disc length) Wave number Stride length (cm) Phase velocity

Minimum

Maximum

Mean

Standard Deviation (SD)

1.91 0.09 1.65 1.08 53.33 0.46

3.41 0.2 3.46 1.56 221.59 0.97

2.51 0.15 2.59 1.31 104.9 0.76

0.36 0.03 0.46 0.12 33.71 0.13

Table 4.3. Summary of morphological features of the lateral-line system in the Atlantic Stingray, Dasyatis sabina (data from Maruska 2001). Surface

Dorsal

Ventral

SN location (number)

None

VS location (number)

Bilateral medial rows to end of tail (about 100/side) None

Pored canals Number of pores/tubule canals Non-pored canals

HYO, IO, SO, PLL 1 HYO, IO, SO

Bilateral rows along rostrum midline (6–10/row) HYO, MAN (SPL present) 1–20 IO, HYO

HYO = hydromandibular, IO = infraorbital, MAN = mandibular, SO = supraorbital, SPL = subpleural loop, SN = superficial neuromast, VS = vesicles of Savi.

DASYATIDAE: WHIPTAIL STINGRAYS

(Kalmijn 1989; Coombs 1994; Coombs et al. 1996); the functional units of the lateral-line system are neuromasts, which are mechanosensory hair cells and gelatinous support cells (Maruska & Tricas 1998; Maruska 2001). Elasmobranch fishes, including Stingrays, have several types of mechanosensory lateral-line organs that are morphologically and functionally distinct. Three types of neuromasts exist: the superficial neuromast, or pit organ, which is located on the skin surface with the cupula exposed to the water; the canal neuromasts, which are located in dermal or subdermal canals and make contact with water through pores on the surface of the skin; and the vesicles of Savi, which are mechanoreceptors found in subdermal pouches on the ventral surface (these are only found in some torpediniforms, Electric Rays, and dasyatids). The skin surface of the Atlantic Stingray contains superficial neuromasts, and canal neuromasts occur in the subdermal canals, the hyomandibular, infraorbital, supraorbital, posterior lateral line, and mandibular canals. The vesicles of Savi are located on the isolated subdermal pouches on the ventral surface of the rostrum (Table 4.3; Fig. 4.5) (Maruska & Tricas 1998; Maruska 2001). Little is actually known of the function of these systems, but superficial neuromasts appear best positioned to detect water movements along the transverse body axis (e.g., detecting movements produced by tidal currents, conspecifics, or predators) (Maruska & Tricas 1998; Maruska 2001). The pored, dorsal-canal system may detect water movements created by conspecifics, predators, or flow field distortions during swimming. Morphological examination of this system in the Atlantic Stingray led to the mechanotactile hypothesis, which contends that the ventral, nonpored canals and vesicles of Savi function as specialized tactile mechanoreceptors that help in detection and capture of small benthic invertebrate prey (Maruska & Tricas 1998).

Vision The Atlantic Stingray, like other studied elasmobranchs, exhibits a variety of advanced visual features such as mobile pupils, multiple visual pigments, visual streaks, and a moveable lens. The species possesses a ramp retina (Sivak 1975) that permits simultaneous focus of images at various distances (McComb & Kijiura 2008) and rods and cones (requisite for color vision). Cones decrease in density peripherally on the retina, but the species has a conerich band located along the horizontal axis of the retina,

147

SO IO

HYO SC

PLL

SO IO

VS MAN SPL

HYO

Figure 4.5. Distribution of the lateral-line canal system and vesicles of Savi on the dorsal (top) and ventral (bottom) surface of the Atlantic Stingray, Dasyatis sabina. Dorsal canals contain numerous lateral tubules that terminate in pores across the body surface. The infraorbital, supraorbital, and sections of the hyomandibular canal near the mouth and rostrum and along the ventral midline lack pores but do contain innervated neuromasts. Vesicles of Savi are located in bilateral rows on the ventral rostrum midline and are isolated from the surrounding water, but lumina of adjacent vesicles are connected via tubules. HYO = hyomandibular canal, IO = infraorbital canal, MAN = mandibular canal, PLL = posterior lateral-line canal, SC = scapular canal, SO = supraorbital canal, SPL = subpleural loop, VS = vesicles of Savi (redrawn from Maruska 2001).

148

FRESHWATER FISHES OF NORTH AMERICA

termed a horizontal visual streak, which likely enhances visual acuity within the horizontal monocular visual field (Logiudice & Laird 1994) (see physiology section). Atlantic Stingrays commonly partially bury themselves in the substrate, in part as a mechanism of predator avoidance. When the individual is buried, however, the eyes remain exposed. The visual streak might allow the individual to scan the horizon for predators (e.g., Sharks) without the revealing head and eye movements necessary in animals with a fovea (Logiudice & Laird 1994).

GE NE TICS The genetics of the Atlantic Stingray are little studied. The species is being used in molecular and functional characterization studies (e.g., urea transporters, Janech et al. 2006ab) and in comparative studies of genes and enzyme products such as those regulating important steroid hormones and those involved in acid-base and ion regulation.

Karyology In a sample of 10 Atlantic Stingrays from the Texas coast, the diploid (2n) chromosome count was 68; a congener, the Bluntnose Ray (Dasyatis say), also had a count of 68 (Donahue 1974). In the Atlantic Stingray, 28 chromosomes had median to submedian centromeres, and 40 had subterminal to terminal centromeres. Within the mediansubmedian group, 20 were large and 8 were relatively small. In the subterminal-terminal group, only the largest pair was clearly distinguishable as homologous. Male heterogamety was suggested by a pair in the mediansubmedian group, but the size distinction varied among the seven males examined.

Comparative Genomics The sequence of the cytochrome P450 aromatase (P450arom) of the Atlantic Stingray was characterized to provide insights into the evolution of this steroidogenic cytochrome (Ijiri et al. 2000). The P450arom enzyme mediates the conversion of androgens (i.e., testosterone and androstenedione) to estrogen and is a key enzyme in the steroidogenic pathway. Hence, in vertebrates it plays an important role in the onset of sexual maturity, development of gametes, expression of sexually dimorphic characters, and evoking of reproductive behaviors. With comparative data from mammals, birds, reptiles, and teleosts,

the phylogenetic relationship (neighbor-joining method, Kimurua protein distances) of the Atlantic Stingray P450arom suggested it as a unique evolutionary lineage having a common ancestral root with both the higher vertebrates and teleosts. Interestingly, especially given the relatively ancient origin of the elasmobranchs, including the Stingrays, the P450arom of the Atlantic Stingray was 56–59% identical to avian and mammalian forms of P450arom and 53–57% identical to that of bony fishes. In a physiologically oriented study using genetic techniques, the Atlantic Stingray yielded the first full-length H+ -K+ -ATPase (HKα1) transcript for any fish (Choe et al. 2004). The H+ -K+ -ATPases participate in systemic ion and acid-base regulation in animal stomachs, including that of the Atlantic Stingray (e.g., Smolka et al. 1994), but also are associated with similar functions in mammalian kidneys. Many of the functions of the mammalian kidney are accomplished by the gills in elasmobranchs. The reported similarities between mammalian type B intercalated cells in the cortical collecting duct of the mammalian kidney and elasmobranch epithelial cells (Piermarini et al. 2002) led to work revealing that HKα1 was expressed in the gills of the Atlantic Stingray (Choe et al. 2004). Within the gills, HKα1 expression did not increase under conditions of hypercapnia (increased carbon dioxide in the blood) but was increased in fresh water, suggesting an active role in osmoregulation via potassium transport. A phylogenetic analysis (rooted tree, ClustalV alignment) of the putative Atlantic Stingray HKα1 with available full-length HKα1 sequences (frog, rat, pig, rabbit) placed the Atlantic Stingray HKα1 as basal to the other taxa. This dates the origin of the H+ -K+ -ATPases to at least before the division of bony and cartilaginous fishes (Choe et al. 2004).

PHYSIOLOGY Several aspects of Atlantic Stingray physiology are documented. Studies examined urea biochemistry in muscle tissue (Treberg et al. 2006); interaction effects of osmoloytes on calcium binding (Heff ron & Moerland 2008); ion transport in gills (Piermarini & Evans 2001; Piermarini et al. 2002) and the alkaline gland (Grabowski et al. 1999); the role of the inter-renal gland in glucocorticoid and mineralocorticoid systems (Andrews et al. 2010); electro-olfactogram responses to amino acids (Silver 1979; Meredith & Kajiura 2010); reproductive histology and endocrinology (Maruska et al.

DASYATIDAE: WHIPTAIL STINGRAYS

1996; Büllesbach et al. 1997; Snelson et al. 1997; Volkoff et al. 1999; Forlano et al. 2000; Tricas et al. 2000; Piercy et al. 2003; see reproduction section); and biochemical, structural, and comparative neurology (Rosiles & Leonard 1980; Ritchie & Leonard 1983; Ritchie et al. 1984; Livingston & Leonard 1990; Bernau et al. 1991; Puzdrowski & Leonard 1993, 1994; Nunez & Trant 1999; Puzdrowski & Gruber 2009). The focus here is on metabolism, osmoregulation and salinity tolerance, sensitivity to conductivity, temperature tolerance, the visual field, and electroreception. Little else appears to be available on other environmental tolerances (e.g., pH, dissolved oxygen, turbidity).

Metabolism Metabolism of marine elasmobranch fishes differs slightly from that of bony fishes because of a low capacity for extrahepatic lipid oxidation, and thus an increased reliance on ketone bodies as the substrates for oxidation. This is most likely a consequence of the osmotic strategy (the retention of high levels of urea and methylamines in the tissues) because the urea disrupts the hydrophobic interactions that are necessary for the proper structure and function of proteins. The transport of fatty acids is dependent on binding to albumin; hydrophobic interactions mediate the binding of fatty acids to albumin, thus the increase in urea concentrations most likely led to the evolution of ketone body– and amino acid–based extra-hepatic metabolic organization instead of the fatty acid system seen in the bony fishes. In addition, within Himantura signifer (White-edged Freshwater Whip Ray), a negative relationship was seen in liver glutamate dehydrogenase activity and tissue and plasma urea levels, indicating that glutamate is deaminated in freshwater elasmobranchs because of the different levels of tissue and plasma urea seen in these species (Speers-Roesch et al. 2006).

Osmoregulation and Salinity Tolerance Chondrichthyan fishes are well adapted for living in a marine environment, a consequence of their unique ability to actively conserve urea and trimethylamine oxide (nitrogenous end products) in the blood and tissues. These organic solutes are freely filterable through the kidney and are actively reabsorbed by the renal tubules (Janech et al. 2003). Their subsequent accumulation in body fluids and tissues lowers the diff usion pressure of water such that it can be drawn into the animal from the surrounding environment

149

without the expenditure of free energy. Considering this unique adaptation, it follows that nearly all known species of chondrichthyans live exclusively in salt water. Organisms in fresh water must obtain ions from the environment through active branchial uptake to maintain hyperosmolarity of tissues relative to the surrounding environment; ions also are gained from the esophagus and intestines (Perry 1997). Within fresh water, euryhaline elasmobranchs maintain high concentrations of urea in plasma, around 100–250 mmol/l, and the Atlantic Stingray has substantially higher sodium, chloride, and urea concentration (almost double) compared with freshwater teleosts (Piermarini & Evans 1998). Early tests of acclimation to dilute seawater indicated Atlantic Stingrays are efficient regulators of osmolarity and may be better adapted to maintaining high plasma osmolarity than Sharks (De Vlaming & Sage 1973). When transferred to dilute sea water, body weight increased and hematocrit decreased but returned to normal within 6 days. Further, physiological studies measuring the ability of Atlantic Stingrays to regulate body fluids revealed a remarkable kidney (glomerular and tubular) functional reserve that allows the species to produce copious amounts of solute-free urine in fresh or low-salinity water (Janech & Piermarini 2002; Choe & Evans 2003; Janech et al. 2006a). Urine flow rate of Atlantic Stingrays in dilute sea water was nine times higher than that of individuals maintained in sea water. The glomerular filtration rate in dilute sea water is among the highest reported for any elasmobranch. This ability results in little increase in weight and little change in hematocrit values in fresh water. The Atlantic Stingray shows no difference in salinity tolerances between the freshwater and marine populations and even the freshwater population has not lost the ability to osmoregulate in salt water as have the Neotropical freshwater species in South America. A molecular and functional comparison of calcium-binding proteins (parvalbumins) in the presence of osmolytes, including urea, revealed no differences between Atlantic Stingrays from marine water and those from fresh water (Heffron & Moerland 2008). In the population of D. sabina in the upper St. Johns River, Florida, the gills are important for active ion uptake in fresh water, but the rectal gland is important for active sodium chloride excretion in sea water (Piermarini & Evans 1998, 2000). As such, the weight of the salt-secreting rectal gland in the freshwater population is 160 km from the Atlantic Ocean. The St. Johns River is unique among major North American rivers in that it originally was a marine embayment bordering the northeast Florida coast, which was partially isolated by a series of offshore barrier islands. Gradually these islands became coalesced to create the St. Johns River as it exists today. The river is fed by many freshwater streams, entering mostly from the west, together with numerous springs of varying chemical com-

position, ranging from fresh to slightly saline (Upchurch & Randazzo 1997). Invasion of marine species into lowsalinity waters in Florida was possible due to the presence of Pleistocene salt deposits that contribute to increased chloride levels in otherwise freshwater systems (Odum 1953). The complex water chemistry of the river, which is related to its unique origin and chemical composition of the underlying strata, means that pockets of brackish water are scattered throughout the drainage, which in turn has resulted in a fish fauna that is unique in its mixture of freshwater and marine species. Some of the marine species are rarely if ever encountered elsewhere in riverine environments. In addition to Dasyatis sabina, this includes marine teleost fishes such as Elops saurus (Ladyfish), Opisthonema oglinum (Atlantic Thread Herring), Mugil curema (White Mullet), Membras martinica (Rough Silverside), Syngnathus scovelli (Gulf Pipefish), Eucinostomus harengulus (Tidewater Mojarra), Micropogonias undulatus (Atlantic Croaker), Lutjanus griseus (Gray Snapper), Microgobius gulosus (Clown Goby), Gobiosoma bosc (Naked Goby), and Paralichthys lethostigma (Southern Flounder). Other freshwater records for Dasyatis sabina from localities well beyond tidal influence are from the Mississippi River in Louisiana and the Tombigbee River in Alabama. Gunter (1938) reported the Atlantic Stingray from the Mississippi River at New Orleans and Angola, Louisiana, the latter >322 km upstream of the river’s mouth. He also noted that Stingrays are occasionally caught during the summer months in the Atchafalaya River at Simmesport, Louisiana, >257 km upstream from the mouth. Boschung & Mayden (2004) reported the species from the Tombigbee River (Mobile Bay basin) upstream as far as Jackson, Clarke County, Alabama. Unlike the St. Johns River, freshwater reproduction for this species is not confirmed in either the Mississippi River or Mobile Bay basins, although this seems likely in the Mississippi River considering the distance this species has been found upstream. The source of a record for D. sabina from the Chattahoochee River (Apalachicola River drainage) in the extreme southeastern corner of Alabama is uncertain (see Ross & Burgess 1980; Boschung & Mayden 2004). No voucher specimens from this locality or elsewhere in the Apalachicola River drainage apparently exist, and none of the above authors can recall the basis for this record.

Sensitivity to Low Conductivity Indirect evidence suggests the freshwater population of Atlantic Stingrays in St. Johns River, Florida, can be

DASYATIDAE: WHIPTAIL STINGRAYS

affected negatively by low water conductivity (16–18°C. The species occurred, however, in waters as warm as 35°C, and individuals did not leave shallow water in mid-summer when afternoon temperatures were regularly >30°C (Snelson et al. 1988). Thermal preference in Atlantic Stingrays is affected by parturition and feeding. When acclimated at 29°C, males, pregnant females, and non-pregnant females all preferred temperatures lower than the acclimation temperature but well within the summer thermal milieu experienced in

151

the wild (Fangue & Bennett 2003). Males and pregnant females showed highest preferred temperatures of 26.1 and 25.9°C, respectively, which were not statistically distinguishable. Non-pregnant females preferred a slightly lower and statistically different average median temperature of 25.3°C. The preference of pregnant females for higher temperatures, even though seemingly slight, may increase embryonic growth rate and decrease the gestation period. A 1°C difference could reduce gestation time by ≥14 days. After feeding, females (pregnant and nonpregnant), but not males, selected lower temperatures. Although the mechanism is hypothetical and not entirely clear, migration to cooler waters after feeding might increase food absorption rates relative to evacuation rates and result in an energetic benefit (Wallman & Bennett 2006). Clearly, however, Atlantic Stingray movement and distribution are a function, at least in part, of the physiological effects of temperature.

Visual Field In an assessment of the visual field among four batoids, the Atlantic Stingray was characterized as having Type III vision (McComb & Kajiura 2008). The Type III visual field is characteristic of predators with large eyes that show a reduction in vigilance behavior. Visual fields were measured in three ways: the field of view of a single eye (monocular), the combined field of view of both eyes (cyclopean), and the overlap of the monocular fields (binocular) (Fig. 4.6). The point at which the monocular visual fields overlap is termed the “binocular convergence point,” and the distance from this point to the central point between Dasyatis sabina 72°

199°

20°

3 cm

115°

15 cm

(34°)

Figure 4.6. The static functional horizontal and vertical visual fields of the Atlantic Stingray, Dasyatis sabina. Values within the shaded area represent monocular visual fields (left) and the standardized convergence distance (right). Values shown outside the shaded areas represent binocular overlaps, and values in parentheses indicate blind areas (redrawn from McComb & Kajiura 2008).

152

FRESHWATER FISHES OF NORTH AMERICA

largest monocular visual field mea sured (McComb & Kajiura 2008).

SO M H

H

Figure 4.7. Distribution of the electrosensory canals of the ampullae of Lorenzini over the ventral (left) and dorsal (right) surfaces of the Atlantic Stingray, Dasyatis sabina. H = hyoid cluster group; M = mandibular cluster group; SO = superficial ophthalmic cluster group (redrawn from Sisneros & Tricas 2000 with permission of Timothy C. Tricas, University of Hawaii at Manoa).

the eyes (in the transverse plane) is called the “convergence distance.” A relatively short convergence distance provides depth perception beginning closer to the eyes, but a longer convergence distance conveys binocular vision farther from the eyes. Type III visual fields consist of broad binocular overlaps (~50 degrees) that are coupled with large posterior blind areas and are seen in fast-moving predators that may simultaneously use other sensory modalities (e.g., electroreception, Blonder & Alevizon 1988) just prior to prey capture. Because the Atlantic Stingray feeds primarily on benthic infauna, vision does not likely play an important role in prey detection. The Atlantic Stingray possesses anterior horizontal vision, conferring frontal vision, and dorsal binocular vision, conferring good dorsal vision, which is good for overhead predator detection, and a short convergence distance, allowing binocular vision at close distances (McComb & Kajiura 2008). Of the batoid species tested, the Atlantic Stingray had one of the broadest binocular fields, one comparable to species with nearly frontal-facing eyes (e.g., Rana pipiens, 90 degrees, Grobstein et al. 1980). The slight canting of the eye of D. sabina and retracted skin surrounding the anterior portion of the eye contributed to the large anterior binocular field. The large anterior binocular vision is beneficial as the Atlantic Stingray negotiates turbid shallow coastal lagoons with sea grass and sandy bottoms (Snelson et al. 1988) but is coincident with large blind areas behind the head. The Atlantic Stingray also had the

Electroreception Elasmobranch fishes possess a sensitive electrosensory system that is used to detect weak electric fields (11.0 g, the external gill filaments were reabsorbed.

157

The tail spine first began to appear in embryos about 60 mm DW, but did not become hardened until about 70 mm DW (20 June), when embryos were morphologically similar to adult Rays. The rate of growth from 4 June until parturition was less rapid than during the preceding interval. About a twofold increase in weight occurred between both of the semi-monthly collections from 4 June to 3 July. This rate of weight increase was smaller than that occurring earlier in development, but it represented a large increase in embryo mass. The mean weight gain during the 2-week period from 20 June and 3 July accounted for 42% of the mean weight of the embryo at parturition. Weight increased only 20% during the last 2 weeks of gestation. In the 4 June and subsequent collections, the gender of embryos was identifiable by the presence or absence of claspers, which could be seen at 40 mm DW. Of embryos that could be sexed, the sex ratio was 26F:19M, not significantly different from 1:1 (X2 = 1.089, df = 1, P < 0.297). Of embryos sampled on 20 June, males were slightly heavier and had larger disc widths than females. In the other three embryo collections, the average female embryo was significantly larger than the average male embryo. Female embryos were significantly heavier than the males (from 4 June to 17 July; 2-way ANOVA: F = 5.906, P = 0.020). The same results were obtained using embryo DW as the dependent variable. Embryo DW measurements by gender were (mean in mm±SD): 4 June, male 46.9±5.11 (n = 5), female 55.4±10.02 (n = 12); 20 June, male 73.8±4.73 (n = 5), female 70.4±7.05 (n = 8); 3 July, male 86.9±10.49 (n = 7), female 100.2±4.26 (n = 3); 17 July, male 91.4±5.09 (n = 2), female 99.9±8.37 (n = 3).

ECOL OGY

Habitat The Atlantic Stingray occurs in brackish bays and estuaries with a permanent freshwater population known to exist in the St. Johns River, Florida (McLane 1955; Tagatz 1968; Johnson & Snelson 1996). The marine populations occur in shallow, inshore regions over sand flats or sand-silt substrates in water 270 mm DW) were females (Schmid 1988; Snelson et al. 1988). Rings on vertebral centra ranged from 2–9 for males and 6–12 for females, but whether ring formation was annual or not was inconclusive (Schmid 1988).

Ontogenetic Shifts in Habitat Use Size frequency data are suggestive that in some populations habitat shifts occur with growth. Juveniles (160– 200 mm DW) were taken in low frequencies in studies in Indian River Lagoon, Florida (Schmid 1988; Snelson et al. 1988), and the Cape Fear River (Schwartz & Dahlberg 1978) despite the use of capture gear and techniques that were effective in taking other sizes of Atlantic Stingrays (e.g., trawls, gillnets, seines). Although the evidence is far from conclusive, juveniles may be using deeper or at least different habitats than adults. In any case the deficit of juveniles in samples is enigmatic (Snelson et al. 1988).

Effects on Prey Rays can alter the community structure through mechanical excavation of benthic substrates. In the Atlantic Stingray a patch of sediment could be excavated once every 70 days, creating a disturbance event that leads to infaunal recolonization. Nevertheless, harpacticoid copepod communities in disturbed sites were indistinguishable from undisturbed sites 29 h after the disturbance (Reidenauer & Thistle 1981). With this relatively high turnover rate, Atlantic Stingrays most likely alter largescale habitat and may smooth distributions of infaunal invertebrates (Hines et al. 1997), suggesting that digging Rays, such as the Atlantic Stingray, can structure benthic communities in many locations (VanBlaricom 1982).

Parasites Elasmobranch species offer a multitude of sites that can be parasitized by metazoan parasites. Six phyla are most common as parasites within elasmobranchs (i.e., Platyhelminthes, Arthropoda, Nematoda, Annelida, Acanthocephala, and Mollusca) with the phyla Platyhelminthes and Arthropoda representing the most diverse parasites. The Atlantic Stingray is no exception, containing many

DASYATIDAE: WHIPTAIL STINGRAYS

parasites spanning the phyla Platyhelminthes, Annelida, and Arthropoda. Common parasites found include the platyhelminths Entobdella corona and Prochristianella penaei and Thaumatocotyle (Hutton 1964; Aldrich 1965; Euzet & de Buron 2010); the annelids Branchellion ravenelii and Myliobaticola richardheardi (Hutton 1964; Bullard & Jensen 2008); and two parasitic copepods (Arthropoda), Brachiella concava and Caligus praetextus (Hutton 1964). A main parasite of the freshwater Atlantic Stingray is the genus Argulus, a fish louse that feeds on the skin mucus (Passarelli & Piercy 2009).

Predators The Atlantic Stingray, as well as other Stingrays living in shallow marine water, are a common food item for inshore Sharks, as attested by spines found embedded in the stomachs of such species as Tiger Sharks (Galeocerdo cuvier), Lemon Sharks (Negaprion brevirostris), Bull Sharks (Carcharinhus leucas), and Blacktip Sharks (Carcharhinus limbatus) (Bigelow & Schroeder 1953a). In addition, Great Blue Herons off the coast of Mississippi consumed Atlantic Stingrays residing in marine waters (Ajemian et al. 2011). In freshwater habitats American Alligators (Alligator mississippiensis) prey on Atlantic Stingrays (Passarelli & Piercy 2009).

CONSERVATION No fishery directly targets Dasyatis sabina, but Atlantic Stingrays are caught as bycatch in gillnets, Shark drift nets, and nearshore trawls that target commercially important species (although because most are released alive there is minimal impact on their population numbers) (Piercy et al. 2006b). Yet, the freshwater population of the Atlantic Stingray has declined in overall health and reproductive success as water quality has declined within the St. Johns River basin (Passarelli & Piercy 2009). Declines in health may be related at least in part to elevated organochlorine levels in tissues of Atlantic Stingrays that are associated with development in the St. Johns River basin, but reproductive impairment apparently is related to other as yet unidentified ecological factors (Gelsleichter et al. 2006).

COMMERCIAL IMPORTANCE The Atlantic Stingray is of little or no direct economic importance. It is often caught during commercial fishing op-

159

erations but is normally discarded as part of the bycatch. No fishery targets Atlantic Stingray and the bycatch mortality is considered low, resulting in an assessment label of Least Concern by the International Union for Conservation of Nature. Skates, and to a lesser extent Stingrays, are sometimes retained for the purpose of punching out circular sections of the wings, which are marketed as scallops. Because of its small size, Dasyatis sabina is rarely if ever used for this purpose. The species is of scientific interest because as a marine species it can live and sometimes spend its entire life in fresh water. The Atlantic Stingray also has scientific value in comparative studies of the evolution and function of genes and physiological systems (e.g., see physiology section).

LITERATURE GUIDE Bigelow and Schroeder’s (1953a) guide, Fishes of the Western North Atlantic: Sawfishes, Guitarfishes, Skates, Rays, and Chimaeroids, provides a still useful, albeit somewhat dated, overview of the Atlantic Stingray, including distribution and habitat. Boschung & Mayden (2004) also provide a good summary of the species and its biology. The works by F. F. Snelson Jr. and colleagues provides an excellent framework for both the ecology and reproduction of the freshwater populations (references in text). T. C. Tricas, J. S. Sisneros, S. M. Kajiura, and K. P. Maruska provide in-depth research of the mechanosensory and electrosensory systems within the Atlantic Stingray (references in text). In addition, T. C. Tricas, S. M. Kajiura, and K. P. Maruska cover, in great detail, the reproductive behavior and mating strategies for the Atlantic Stingray (references in text). For detailed information concerning the osmoregulation within the freshwater population, consult the work of P. M. Piermarini and colleagues (e.g., Piermarini & Evans 1998; Piermarini & Evans 2000; Piermarini & Evans 2001; Piermarini et al. 2002).

Acknowledgments We thank Amy Carson-Commens for help in obtaining literature and redrawing of several figures. We are also grateful to Gordon McWhirter, Anthony Rietl, Vicki Reithel, and Daniel Warren for assistance in proofing literature cited, figures, and tables. Larry Linton and Stephen Kajiura graciously allowed us to use their photographs.

Chapter 5

Acipenseridae: Sturgeons Bernard R. Kuhajda

The Acipenseridae consist of 25 extant species in 4 genera, including 17 species in Acipenser, 2 in Huso, 3 in Pseudoscaphirhynchus, and 3 in Scaphirhynchus (Birstein & Bemis 1997; Birstein et al. 1997a; Billard & Lecointre 2001; Ludwig 2008). The word “acipenser” is the Latin name for Sturgeon. Sturgeons occur on all continents in the Northern Hemisphere and are almost completely restricted to the northern temperate zone (Bemis & Kynard 1997; Choudhury & Dick 1998). The greatest diversity of Sturgeons is in western Europe (11 species) and central Asia, including the Mediterranean, Aegean, Black, Caspian, and Aral Seas, which is referred to as the Ponto-Caspian region (Bemis & Kynard 1997). Five species of Acipenser and three species of Scaphirhynchus occur in North America. Sturgeons are the largest freshwater fishes and are long-lived (e.g., >150 years old). Originally described as Sharks in the 1700s due to the cartilaginous skeleton, jaw structure, and shark-like tail, Sturgeons are actually ancient bony fishes with fossils that date to 175 mya. They have a toothless, protrusable mouth on the underside of the head with four barbels just before the mouth and five rows of bony plates along the body. Sturgeons cruise along

the bottom of rivers, lakes, and oceans, locating and eating invertebrates and fishes from the bottom with the aid of taste buds on the barbels and ampullary organs on the underside of the head, which use electroreception to detect the weak electrical fields emitted by prey. Although some Sturgeons spend most of their lives in the oceans or estuaries, they all spawn in freshwater rivers where they were hatched, some migrating ≤1,200 km (746 miles) to spawn. Newly hatched larvae passively drift in river currents ≤530 km (329 miles). Sturgeons are highly vulnerable to human activities, especially overfishing and dams that block migratory routes; therefore, almost all Sturgeons worldwide are considered imperiled. Their roe (eggs), which are processed into caviar, have been valued by humans for >1,000 years.

DIVERSITY AND DISTRIBUTION In North America, Sturgeons occur along both coasts and in inland fresh waters (Figs. 5.1 and 5.2). Two species of Acipenser (Fig. 5.1), the Green Sturgeon (Acipenser

Plate 5.1. White Sturgeon, Acipenser transmontanus

ACIPENSERIDAE: STURGEONS

Figure 5.1. Geographic range of Acipenser in North America.

Genus Acipenser

Figure 5.2. Geographic range of Scaphirhynchus.

Genus Scaphirhyncus

medirostris) and the White Sturgeon (Acipenser transmontanus), are restricted to the North Pacific region. Green Sturgeons occur along 3,000 km (1,864 miles) of coastal and estuarine areas from the Aleutian Islands, Alaska, to central California, and into Mexico (Lee et al. 1980; Page & Burr 1991; Moyle 2002; Wilson & McKinley 2004; Nelson et al. 2004). The White Sturgeon was once considered one of the few Sturgeon species found on two continents with an Asian distribution in China, northern Japan, Korea, and Russia north to the Amur River, but karyotypic and genetic data indicate that the Asian distribution represents a separate species, the Sakhalin Sturgeon (Acipenser mikadoi) (Birstein & Bemis 1997). White Sturgeons inhabit coastal areas from the Aleutian Islands, Alaska, to central California with a landlocked population naturally isolated by falls in the Kootenai River in the upper Columbia River drainage in Idaho, Montana, and British Columbia (Lee et al. 1980; Page & Burr 1991, 2011; Anders et al. 2002; Wilson & McKinley 2004). The Western Atlantic region is home to two species of Acipenser, the Shortnose Sturgeon (Acipenser brevirostrum)

161

and Atlantic Sturgeon (Acipenser oxyrinchus) (Fig. 5.1). Shortnose Sturgeons inhabit coastal waters along the eastern seaboard from New Brunswick, Canada, to northern Florida. Two subspecies of the Atlantic Sturgeon are recognized. The nominate form, A. oxyrinchus oxyrinchus, is distributed along the Atlantic Coast from northern Quebec and Newfoundland, Canada, to northern Florida. The Gulf Sturgeon (A. oxyrinchus desotoi) is restricted to coastal areas in the Gulf of Mexico from Tampa Bay, Florida, to Louisiana, and perhaps westward to the Rio Grande, Texas and Mexico (Lee et al. 1980; Page & Burr 1991; Wilson & McKinley 2004; Nelson et al. 2004). The Lake Sturgeon (Acipenser fulvescens) (Fig. 5.1) and all species of Scaphirhynchus (Fig. 5.2) occur in fresh water in middle North America. The Lake Sturgeon is found in lakes and rivers from Hudson Bay, Great Lakes, and St. Lawrence River drainages from Alberta to Quebec, Canada, south to the lower Mississippi and Coosa River drainages, Louisiana and Alabama. The Alabama Sturgeon (Scaphirhynchus suttkusi) is restricted to the Mobile Basin in large rivers in Alabama and formerly in Mississippi. The Pallid Sturgeon (Scaphirhynchus albus) is nearly restricted to the Missouri and Mississippi Rivers proper from Montana to Louisiana but does use the lower reaches of major tributaries. The Shovelnose Sturgeon (Scaphirhynchus platorynchus) is distributed throughout the Mississippi River basin, including the Missouri and Ohio River drainages, from Montana to Pennsylvania, south to Louisiana, and had a historical population in the Rio Grande, New Mexico (Lee et al. 1980; Page & Burr 1991; Wilson & McKinley 2004; Nelson et al. 2004).

Polyploidization and Diversity Polyploidization is one of the main genetic mechanisms of speciation within Acipenseriformes (Sturgeons) and has contributed to the diversification within Acipenser. The diploid ancestor of all Acipenseriformes had a karyotype of 60 chromosomes that produced a tetraploid (4n) ancestor with 120 chromosomes through a gene duplication event (Birstein et al. 1997b; Ludwig et al. 2001; see genetics

Plate 5.2. Shovelnose Sturgeon, Scaphirhynchus platorynchus

162

FRESHWATER FISHES OF NORTH AMERICA

section). A diploid condition may have been established in this ancestor before diversification into species of Polyodontidae (Paddlefishes) and Acipenseridae (Ludwig et al. 2001; Fontana 2002), but others consider extant species with 120 chromosomes as tetraploids (Birstein et al. 1997b; Kim et al. 2001). Only species within Acipenser have higher chromosome numbers with some species possessing 240–260 (8n) and others having ≤500 chromosomes (16n) (Birstein et al. 1997b; Fontana 2002). Relationships within Acipenser based on molecular data indicate that these high-ploidy species arose multiple times (Birstein et al. 1997b; Ludwig et al. 2001), and the likely method was reticular speciation (Vasil’ev 1999). A current model for reticulate speciation involves three steps. Interspecific hybridization between diploid bisexual species forms an all-female diploid species that produces diploid eggs. The all-female species then back-crosses with the parental species, resulting in a triploid unisexual species. Hybridization between the triploid species and a diploid bisexual species results in a tetraploid bisexual species. These intermediate unisexual diploid and triploid species occur in some other fishes, amphibians, and lizards but are not known from Sturgeons (Vasil’ev 1999, 2009). Polyploidization via reticulate speciation may be one of the primary sources of reduced genetic differences between Sturgeon taxa (Robles et al. 2005).

Anadromy and Diversity Anadromy is likely a secondary adaptation in Sturgeons with ancestors having evolved as freshwater fish. This is supported by several evolutionary life history constraints of Sturgeons (Sulak & Randall 2002). Eggs and larvae of all Sturgeons are intolerant of salt water (Kynard & Horgan 2002a; Kynard & Parker 2004). Juveniles and subadults of anadromous species return in the spring to a physiological refuge that fresh water affords them as they remain relatively dormant with little or no feeding during the warm summer months (Sulak & Randall 2002). Anadromy likely evolved as a means to exploit the marine and estuarine environment where the rich benthic invertebrate resources allow for dramatic increase in growth rates relative to those species foraging in freshwater environments. In Gulf Sturgeons feeding in salt water is confined to winter months when predation threats from sharks and competition from teleosts are lowest (Sulak & Randall 2002). The exploitation of the marine environment by Acipenser may be an additional factor linked to the diversity within this genus (Bemis & Kynard 1997).

Evolutionary Rate and Diversity Although some morphological, genetic, and ecological differences have evolved between Sturgeon genera, and to a lesser extent, among congeners, relatively little change has occurred in Acipenseridae over tens of millions of years. The basic body plan, osteology, and general habitats between extant and fossil Sturgeons from the Upper Cretaceous (about 99.6–65.5 mya) differ little (Bemis et al. 1997; Choudhury & Dick 1998; Hilton & Grande 2006; see fossil record section), and the rate of molecular evolution has been extremely slow for all Sturgeons, resulting in a lack of variation in genetic markers. This slow rate of change is likely due to Sturgeons having long generation times, low metabolic rates, and reduced rates of concerted evolution (Wirgin et al. 1997; de la Herrán et al. 2001; Krieger & Fuerst 2002b, 2009; Robles et al. 2004; Krieger et al. 2006; Dillman et al. 2007).

Relationships among Species Relationships among species of North American Sturgeons have received considerable attention. Within Scaphirhynchus, phylogenetic analysis of morphological data indicated Shovelnose and Alabama Sturgeons were sister-species (Mayden & Kuhajda 1996), but studies based on mitochondrial DNA sequence data were unable to address this hypothesis because of the slow rate of molecular evolution at these genes (Campton et al. 2000; Krieger et al. 2000; Simons et al. 2001). Within Acipenser, relationships based on morphological and karyological data show sister-species relationships between the two North American Pacific species, White and Green Sturgeon, and between Shortnose and Lake Sturgeon, with the Atlantic Sturgeon basal to these species (Artyukhin 1995; Choudhury & Dick 1998). Mitochondrial DNA analyses on only North American species showed these same relationships (J. R. Brown et al. 1996; Krieger et al. 2000), but when other species of Acipenser are included, these sister-species pairs are not recovered, but each species of the pair typically occur within the same clade (Birstein & DeSalle 1998; Ludwig et al. 2001; Birstein et al. 2002; Artyukhin 2006). Most species of Sturgeon are identified morphologically using adult characters, such as head and body measurements, including snout shape and eye diameter, as well as barbel placement and the number or type of scutes, squamation, and head spines (Scott & Crossman 1973; Mayden & Kuhajda 1996; Vecsei & Peterson 2004). But differential

ACIPENSERIDAE: STURGEONS

growth rates of distinguishing characteristics (allometry or heterochrony) can make species identification difficult for larvae, juveniles, and subadults, and other characters, such as spines and scutes, may be obscured or lost in older adults (Jordan & Evermann 1896; Bailey & Cross 1954; Vladykov & Greeley 1963; Scott & Crossman 1973; Mayden & Kuhajda 1996; Gisbert & Doroshov 2006).

Intraspecific Variation Identification of Sturgeons is further complicated by intraspecific variation in meristic and morphometric characters, especially between populations from separate drainages (Guénette et al. 1992; Keenlyne et al. 1994; Mayden & Kuhajda 1996; Walsh et al. 2001; North et al. 2002). This variation may be a result of strong homing capabilities by spawning adults to natal river systems or sites (Bemis & Kynard 1997). Even within a population, skeletal, rostral, and scute morphologies can show significant variation between specimens at the same life history stages, and skull features can vary between right and left sides within an individual (Hilton & Bemis 1999; Vecsei 1999; Vecsei & Peterson 2004). The large degree of intraspecific variation in Shortnose Sturgeon may be the result of high genetic diversity of founding populations during post-Pleistocene (≤10,000 years ago) colonization of Atlantic systems (i.e., minimal founder effect), resulting from either colonization by individuals from numerous adjacent systems or from glacial refugia that had maintained high haplotype diversity. Alternatively, dramatic declines in Atlantic Coast populations during the late Pleistocene (1.8–0.01 mya) may have caused a genetic bottleneck, making remnant populations susceptible to genetic drift that created morphological divergence (Vecsei & Peterson 2004).

Sturgeons as Non-Natives Only one species, the White Sturgeon, has been introduced outside its native range. White Sturgeons have been introduced via aquaculture to South America, the Middle East, and Europe. White Sturgeons were introduced in Chile sometime after 1981 and have been found in the Rio Maipo (Dyer 2000); they were found in the wild in Israel in 1997 but are not established (Roll et al. 2007). Introductions in Europe are the result of accidental releases or escapes from aquaculture facilities and perhaps deliberate releases by aquarists (Arndt et al. 2000, 2002). The numbers of exotic Sturgeons have in-

163

creased substantially since 1990 in Germany, coinciding with the rise of Sturgeon aquaculture and the first importation of White Sturgeons in 1992. They are not considered established because they have not reached maturity, but once mature they may hybridize with the native Sturgeons (Wolter & Röhr 2010); hybridization between native Sterlet (A. ruthenus) and introduced Siberian Sturgeon (A. baerii) has occurred in the upper Danube River on the Germany-Austria border even though these species differ in chromosome number (118 versus 248, respectively) (Ludwig et al. 2009). Other concerns in Europe include exotic Sturgeons thriving and interfering with restoration of the critically endangered native European Sturgeon (Acipenser sturio) and introduction of diseases and parasites. Disease transfer from White Sturgeons is of particular concern given the variety of viruses present in this species (Arndt et al. 2000, 2002), and the White Sturgeon herpes virus occurs in aquacultured White Sturgeons in Italy (Kurobe et al. 2008). From 1991 to 2000 White Sturgeons made up 1% of the total coastal and inland Sturgeon fishery in Poland, Germany, and the Netherlands, first appearing after 1993 in rivers and estuaries. Within the United States, White Sturgeons have been introduced in Arizona (Minckley 1971).

PHYLOGE NE TIC RELATIONSHIPS

Higher Relationships Researchers in the mid- to late 1700s considered Sturgeons (and the Paddlefish) closely related to Sharks given their similar jaw structure and cartilaginous skeletons (Linnaeus 1758). This relationship was rejected in the mid-1800s and Sturgeons (and the Paddlefishes) were considered bony fishes (osteichthyans) and placed in the basal grade Chondrostei (Grande & Bemis 1991, 1996; Bemis et al. 1997). Currently Sturgeons (Acipenseridae) and Paddlefishes (Polyodontidae) are placed in the suborder Acipenseroidei, and along with the fossil families †Chondrosteidae and †Peipiaosteidae, are placed in the order Acipenseriformes (Grande & Bemis 1991, 1996; Grande et al. 2002; Grande & Hilton 2006; Krieger et al. 2008; Hilton & Forey 2009; Hilton et al. 2011). Based on the distribution of the fossil families and the greatest current species diversity of the group in the Ponto-Caspian region, the order Acipenseriformes likely originated in Western Europe followed by diversification into central Asia with a later appearance in North America (Bemis & Kynard 1997). Acipenseriformes, Bichirs (order Polypteriformes),

164

FRESHWATER FISHES OF NORTH AMERICA

and other fossil orders were assigned to the subclass Chondrostei that was basal to all other neopterygian fishes (Nelson 1994). But subsequent studies found that Bichirs are extant basal actinopterygiian fishes (Rayfinned Fishes) (subclass Cladistia), and Acipenseriformes together with other fossil orders form the subclass Chondrostei that is sister to the subclass Neopterygii (Patterson 1982; Bemis et al. 1997; Nelson 2006).

Relationships within Acipenseridae Several hypotheses exist for the relationships of Sturgeon genera within Acipenseridae (genera Acipenser, Huso, Scaphirhynchus, Pseudoscaphirhynchus) (Fig. 5.3). The classic arrangement recognizes two subfamilies: Acipenserinae (genera Acipenser and Huso) and Scaphirhynchinae (genera Scaphirhynchus and Pseudoscaphirhynchus) (Berg 1940; Bailey & Cross 1954; Nelson 1994). This classification was supported by a phylogenetic analysis of morphological characters (Fig. 5.3a; Mayden & Kuhajda 1996; Artyukhin 2006). Several osteological studies using extant and fossil materials (Findeis 1993, 1997; Grande & Bemis 1996; Bemis et al. 1997) place Huso in the subfamily Husinae that is basal to the subfamily Acipenserinae; Acipenserinae contains the tribe Acipenserini (genus Acipenser) that is basal to the tribe Scaphirhynchini (genera Scaphirhynchus, †Protoscaphirhynchus, and Pseudoscaphirhynchus) (Fig. 5.3b). But subsequent studies discovered (1) a new well-preserved Sturgeon fossil, genus †Priscosturion, possesses numerous osteological characters previously not discernible in other fossil Sturgeons; (2) some characters from the fossil Sturgeon genus †Protoscaphirhynchus are inaccurate (see fossil record section); and (3) some characters thought unique to Sturgeons (e.g., pectoral spines) actually are shared with a primitive fossil Paddlefish (Grande et al. 2002; Grande & Hilton 2006, 2009). These new osteological data place the genus †Priscosturion in its own subfamily, †Priscosturioninae, which is basal to the subfamily Acipenserinae that now contains the genera Acipenser and Huso and the tribe Scaphirhynchini (genera Scaphirhynchus and Pseudoscaphirhynchus). Placement of †Protoscaphirhynchus is uncertain because of its lack of characters (Grande & Hilton 2006, 2009) (Fig. 5.3c). In contrast to morphological studies, phylogenetic relationships based on mitochondrial DNA sequence place the genus Scaphirhynchus (Birstein et al. 1997b, 1999; Birstein & DeSalle 1998; Krieger et al. 2000), the genus Scaphirhynchus plus Atlantic-European Sturgeons (Ludwig

Figure 5.3. Phylogenetic relationships of Sturgeon genera within the family Acipenseridae (genera Acipenser, Huso, Scaphirhynchus, Pseudoscaphirhynchus): (A) classic arrangement recognizing two subfamilies Acipenserinae and Scaphirhynchinae (redrawn from Mayden & Kuhajda 1996; Artyukhin 2006); (B) osteological studies using extant and fossil materials recognize the subfamily Husinae basal to the subfamily Acipenserinae, which contains the tribe Acipenserini basal to the tribe Scaphirhynchini (redrawn from Findeis 1993, 1997; Grande & Bemis 1996; Bemis et al. 1997); (C) revised osteological relationships recognize the subfamily †Priscosturioninae basal to the subfamily Acipenserinae, which contains the tribe Scaphirhynchini (redrawn from Grande & Hilton 2006, 2009); and (D) mitochondrial DNA sequence data place the genus Scaphirhynchus plus Atlantic-European Sturgeons basal to an unresolved clade of all other Sturgeons (redrawn from Birstein et al. 1997b, 1999, 2002; Birstein & DeSalle 1998; Fontana et al. 2001; Ludwig et al. 2001; Dillman et al. 2007; Krieger et al. 2008). Scaph. = Scaphirhynchus; Pseudoscaph. = Pseudoscaphirhynchus.

ACIPENSERIDAE: STURGEONS

et al. 2001; Fontana et al. 2001), or either group (unresolved, Krieger et al. 2008) as basal to all other extant Sturgeons. Additionally the subfamily Husinae is not recovered, the genus Huso is not monophyletic (Birstein & DeSalle 1998; Birstein et al. 1999, 2002; Ludwig et al. 2001; Krieger et al. 2008), and the genera Scaphirhynchus and Pseudoscaphirhynchus do not form a distinct subfamily or tribe (Birstein et al. 1997b, 2002; Dillman et al. 2007; Krieger et al. 2008) (Fig. 5.3d). Studies using sequence data from nuclear and nuclear satellite DNA support these relationships (de la Herrán et al. 2001; Robles et al. 2004, 2005; Krieger et al. 2006). Sturgeon relationships as revealed by molecular analyses appear to conform to transoceanic distributions for two major clades (Ludwig et al. 2000, 2001; Birstein et al. 2002; Dillman et al. 2007; Krieger et al. 2008), including the Pacific Sturgeon clade (White and Green Sturgeons, and Asian Acipenser and Huso) and the Atlantic Sturgeon clade (Shortnose and Lake Sturgeons, and European-Asian Acipenser, Huso, and Pseudoscaphirhynchus). Two other clades also are recognized, the basal sea Sturgeons (Atlantic and European Sturgeons) and the genus Scaphirhynchus (Robles et al. 2004, 2005; Krieger et al. 2008). A morphological phylogeny of 23 extant Sturgeon species (no fossils) supported several aspects of these molecular phylogenies, including sea Sturgeons as basal Acipenser, an Atlantic-Pacific split for some Acipenser species (but only those with ≥250 chromosomes), and lack of support for the subfamily Husinae (containing only Huso) but continued to support the monophyly of Huso and Scaphirhynchinae (Artyukhin 2006). Studies examining fewer extant Sturgeons but including extensive skeletal characters and fossil taxa were in agreement with molecular phylogenies on a close relationship between Pseudoscaphirhynchus spp. and the Stellate Sturgeon, Acipenser stellatus, rather than with Scaphirhynchus, and designated a new subfamily Pseudoscaphirhynchinae. The subfamily Husinae was also recognized, but only included one Huso species and the Sterlet, Acipenser ruthenus. The relationships of these clades and the genera †Priscosturion, Scaphirhynchus, and several other Acipenser species were unresolved within Acipenseridae (Hilton 2005; Hilton & Forey 2009; Hilton et al. 2011).

Evolutionary Considerations The greatest extant Sturgeon species diversity is in the Ponto-Caspian region (Bemis & Kynard 1997; Choudhury & Dick 1998) and biogeographical reconstruction based on a molecular phylogeny of extant Acipenseriformes implies

165

this region was the center of origin for Sturgeons (Peng et al. 2007). Estimates of the divergence time of major Sturgeon lineages using molecular data include the separation of the sea Sturgeons about 171 mya, Scaphirhynchus from the remaining Sturgeons nearly 151 mya, and the AtlanticPacific split around 121 mya (Peng et al. 2007).

FOSSIL RECORD †Priscosturion longipinnis is the only well-preserved, articulated Sturgeon fossil with a relatively complete skeleton (Fig. 5.4). The single specimen is 80 cm TL and is from the Upper Cretaceous Judith River Formation (about 78 mya), Montana (Grande & Hilton 2006, 2009). The species shares several features with living Sturgeons, including a thick pectoral spine, a cardiac shield formed by the shoulder girdle, a row of dorsal bony plates (scutes) from the back of the skull to the anterior edge of the dorsal fin, and a lateral-line canal system concealed under the lateral scutes. †Priscosturion longipinnis differs from living Sturgeons by having more dorsal scutes (20), a trunk almost completely covered with thick scales, and a highly falcate and extremely long dorsal fin (140 fin rays). A phylogenetic analysis of osteological features placed †P. longipinnis as basal to all living Sturgeons (Grande & Hilton 2006). The only other articulated Sturgeon fossil described is †Protoscaphirhynchus squamosus, but the specimen is only partially intact and is badly crushed and weathered (Wilimovsky 1956) (Fig. 5.5). The fossil is from the Upper Cretaceous Hell Creek beds (about 65 mya), Montana. Parts of the head, body, and tail are discernible, and the fossil resembles the genus Scaphirhynchus but differs in having its entire body covered by plates and scutes (Wilimovsky 1956). Findeis (1993) argued that only the posterior end of the fossil is completely armored, which is similar to Scaphirhynchus, and Bemis et al. (1997) suggested †P. squamosus may not warrant its own genus. Additionally, Hilton (2004) considered two caudal skeleton characters as shared between †Protoscaphirhynchus and Scaphirhynchus. In contrast, Gardiner (1984) did not consider †P. squamosus an Acipenseriformes based on scale type and skull features. This fossil was used in several phylogenetic studies of Sturgeons and of lower actinopterygians (Findeis 1993; Grande & Bemis 1996; Bemis et al. 1997; Jin 1999), but some characters detailed in the original description are questioned in a redescription of this fossil species (Hilton & Grande 2006) in which only the

166

FRESHWATER FISHES OF NORTH AMERICA

Figure 5.4. †Priscosturion longipinnis is the only well-preserved, articulated Sturgeon fossil with a relatively complete skeleton. This specimen (80 cm TL) is from the Upper Cretaceous Judith River Formation (about 78 mya), Montana (Grande & Hilton 2006, 2009; photograph by and used with permission of Eric Hilton).

Figure 5.5. Tail region of †Protoscaphirhynchus squamosus from the Upper Cretaceous Hell Creek beds (about 65 mya), Montana (Hilton & Grande 2006; photograph by and used with permission of Eric Hilton).

Scaphirhynchus-like caudal fin region offers any unambiguous details. The apparent close relationship between †P. squamosus and Scaphirhynchus indicates that North American riverine Sturgeons have existed largely unchanged for ≥65 million years. Both of these articulated fossil Sturgeon specimens, as well as a third as yet undescribed Sturgeon and a Paddlefish fossil, were discovered independently in abdominal areas of hadrosaurian (duck-billed) dinosaurs (Grande & Hilton 2006; Hilton & Grande 2006). Because these specimens are whole and hadrosaurs were plant eaters, the Sturgeons were not likely eaten. Instead, the dinosaur carcasses likely trapped the deceased Sturgeons and facilitated the rapid burial of these fishes by forming sedi-

ment traps in near-shore or river habitats (Grande & Hilton 2006). Fossils of Acipenser are only represented by scutes and pectoral fin spines from ≤10 described fossil species (Wilimovsky 1956; Gardiner 1984; Birstein & DeSalle 1998; Choudhury & Dick 1998; Grande & Hilton 2006). In North America, these are represented by †A. albertensis and †A. eruciferus from the Late Cretaceous (about 99.6– 65 mya), †A. ornatus from the Miocene (20–5 mya), and numerous other material only identified to genus from the Late Cretaceous to the Pliocene (5–2 mya). These fossils do not provide enough information to be useful in understanding relationships within Acipenseridae, and the validity of the described fossil species of Acipenser is questioned, given

ACIPENSERIDAE: STURGEONS

that the material cannot be readily distinguished from each other or from living Sturgeons (Findeis 1993, 1997; Bemis et al. 1997; Hilton & Bemis 1999; Hilton & Grande 2006). The fossils do indicate that Acipenser has existed for ≥65 million years. The oldest fossil Sturgeon is †Asiacipenser kotelnikovi from the Middle Jurassic (about 176–161 mya) in Asia (Nessov et al. 1990). This is the only Sturgeon fossil that predates the Cretaceous, but some consider that the undiagnostic nature of the material (pieces and fragments) makes its assignment to Acipenseridae unreliable (Grande & Hilton 2006). Most fossil Sturgeons, including the newly discovered †P. longipinnis, were recovered from deposits associated with freshwater and estuarine areas of coastal plains, the same environments used by present-day Sturgeons, indicating similar habitat for tens of millions of years (Choudhury & Dick 1998; Grande & Hilton 2006). These large riverine and near-shore habitats are highenergy environments that would facilitate the disarticulation of specimens before fossilization could occur, perhaps explaining the dearth of articulated Sturgeon fossil skeletons. The combination of fragmentary Sturgeon fossils and the overall conservative morphology of Sturgeons, including living species, make interpretation of the fossil record of Sturgeons difficult (Hilton & Grande 2006).

167

MORPHOLOGY Sturgeons have an elongate, robust body that is subcylindrical in cross-section with the dorsal fin well back on the body behind the pelvic fins. The tail is heterocercal. Four conspicuous barbels are suspended in front of the highly protrusible, subterminal mouth. A rather large pair of nostrils with 20–30 olfactory lamellae is anterior to the small eyes. The body is covered with five rows of bony plates (scutes, one dorsal, two lateral, and two ventral), and bony platelets occur between the rows of scutes. Rhomboid scales are on the caudal peduncle, caudal fin, and abdomen (Fig. 5.6) and round-based scales on the internal surface of the pectoral girdle (Fig. 5.7). The four gill arches have relatively few gill rakers. Vertebral centra are lacking and adults lack teeth. The shoulder girdle forms a cardiac shield. Internally, the stomach has numerous pyloric appendages and the intestine a spiral valve. The gas bladder is simple with a connection to the gut (physostomous). The gonads are thick and elongate, extending along each side of the air bladder near the dorsal surface of the body cavity (Berg 1940; Vladykov & Greeley 1963; Nelson 1994; Mayden & Kuhajda 1996; Bemis et al. 1997; Findeis 1997; Hilton et al. 2011). Not all of these characters are derived (synapomorphies) for the order Acipenseriformes or family Acipenseridae (characters Figure 5.6. Cleared and stained caudal region of a Shortnose Sturgeon (Acipenser brevirostrum) showing angled vertical rows of rhomboid scale and small, scattered bony platelets. Scale bar = 2 mm (Hilton et al. 2011; photograph by and used with permission of Eric Hilton).

168

FRESHWATER FISHES OF NORTH AMERICA

Figure 5.8. Dorsal caudal fulcra from a Shortnose Sturgeon (Acipenser brevirostrum) with anterior and more posterior fulcra on top and bottom, respectively (Hilton 2004; photograph by and used with permission of Eric Hilton). Figure 5.7. Scanning electron micrograph of small patch of skin on the internal surface of the pectoral girdle of a Shortnose Sturgeon (Acipenser brevirostrum) showing round-based scales. Scale bar = 0.25 mm (Hilton et al. 2011; photograph by and used with permission of Eric Hilton).

Greeley 1963; Grande & Bemis 1991; Bemis et al. 1997; Findeis 1997; Grande et al. 2002). The endoskeletal elements of the pectoral fins of Sturgeons (and Paddlefishes) have elements homologous to both the fin radials of teleosts and the limb bones of tetrapods (Davis et al. 2004).

unique to these taxa in Grande & Bemis 1991, 1996; Bemis et al. 1997; Findeis 1997; Hilton 2004; Grande & Hilton 2006; Hilton et al. 2011).

Benthic Cruisers

Ancient Body Plan The basic body plan of living Sturgeons generally reflects that found in fossil Acipenseriformes from the Lower Jurassic (201.6–176 mya) and closely resembles fossil Sturgeons from the Late Cretaceous (78–65 mya) (Bemis et al. 1997). These relic characters include a subcylindrical body, a heterocercal tail; the endocranium greatly extended into a rostrum; reduced ossification of the endoskeleton and a persistent notochord; the head covered with bony plates separated by prominent sutures; a subterminal, protrusable mouth in which the upper jaw does not articulate with the cranium and the jaws are suspended from a mobile hyoid arch; the loss of the maxillary and premaxillary bones; no opercle; the subopercle, supported by a reduced number of branchiostegal rays (one to three), acts as the gill cover; fin rays more numerous than their basal supporting elements; novel scales, including median predorsal, preanal, and precaudal scales (fulcra) (Fig. 5.8); and a stout spine along the leading edge of the pectoral fin made of rays encased in a dermal bone sheath (Vladykov &

Although North American Sturgeons have morphological specializations for close interactions with the bottoms of rivers, lakes, estuaries, and oceans (see next three paragraphs), they lack extreme modifications such as flattened bodies, extensive camouflage, and stationary lifestyles. Many aspects of Sturgeon morphology reflect an active benthic lifestyle termed “benthic cruising” (Findeis 1997), representing the ecological guild supra-benthos cruisers (Vecsei & Peterson 2004). Distinct adaptations for benthic cruising include the lack of articulation of the upper jaw with the cranium, the subterminal placement of the mouth, and a novel jaw protrusion mechanism that allows for jaw projection toward the benthos with an accompanying suction to facilitate the capture of prey items (Findeis 1997; Carroll & Wainwright 2003; Miller 2004). A novel palatal complex shears across a tongue pad (with biting ridges) that grips prey as the upper jaw is projected and retracted as ingested substrate is sieved through. In Sturgeons that prey on smaller items (e.g., aquatic insects), gill rakers are branched and flattened to assist in prey retention (Findeis 1997; Vasil’eva 1999). Normal gill ventilation in Sturgeons is accomplished with a buccal pump, where water is pumped into the mouth and out of

ACIPENSERIDAE: STURGEONS

the ventral opening of the subopercle. An accessory flow system is used when the mouth is blocked with prey and substrate during feeding; water is brought into the opercular chamber through the dorsal edge of the gill cover. Species of Acipenser possess a spiracle, a vestigial gill opening with a canal that leads to the opercular cavity. A small gill (pseudobranch) lines the internal opening of the spiracle canal, but these structures play no role in gill ventilation (Burggren 1978). All Sturgeons possess structures on the ventral surface of the rostrum called ampullary organs that likely use electroreception to detect the weak electrical fields emitted by prey items (Teeter et al. 1980; New & Bodznick 1985; Miller 2004). In Scaphirhynchus ampullary organs are found in clusters of 20 (on average) distributed on the dorsal, lateral, and ventral surfaces of the head and on the subopercle, with the largest number covering the entire ventral surface of the rostrum except for a small anterior midline oval area (Fig. 5.9). About 1,300 ampullary clusters were counted in an individual, indicating a total of >20,000 ampullary organs present (Northcutt 1986). Although the anterior lateral-line nerve innervates both ampullary organs and mechanoreceptive neuromasts of the cephalic lateral-line system, a clear division exists between these sensory systems in the hindbrain of Sturgeons (and the Paddlefish) (New & Bodznick 1985). Because the sensory epithelium and function of the ampullary organs of Sturgeons is similar to Lorenzinian ampullae found in elasmobranchs, the electrosensory system in Acipenseroidei likely is derived from a common ancestor shared with cartilaginous fishes (New & Bodznick 1985; Northcutt 1986). The overall flattening of the head of Sturgeons may be associated with the expansion of these electroreception ampullary fields and with the ventral placement of the mouth. The rostrum of Acipenser is subconical but Scaphirhynchus, the most benthic genus of North American Sturgeons, has an extremely flattened rostrum along with a flatter body and a flattened and completely armored caudal peduncle, all modifications reflecting a close association with river bottoms (Findeis 1997). Juveniles of Acipenser are more benthic than adults and correspondingly have more dermal armoring (Findeis 1997). Other specialized rostral structures include barbels and lip papillae that have chemoreceptors (taste buds) to detect benthic prey (Bemis et al. 1997; Kasumyan 1999, 2002). Some species of Scaphirhynchus have highly fringed barbels and fringed papillae on lobes on their lips to presumably increase detection of prey items (Miller 2004).

169

Figure 5.9. Drawing of the head of Scaphirhynchus stained with methylene blue showing ampullary organs in clusters on the dorsal, lateral, and ventral surfaces of the head and on the subopercle. Scale bar = 1 cm (Northcutt 1986; illustrated by and used with permission of Glenn Northcutt).

Sturgeons also have a lateral-line canal system on the underside of the rostrum, as well as on the lateral and dorso-lateral surface of the head. The lateral line extends concealed under the lateral row of scutes onto the upper lobe of the heterocercal caudal fin (Forbes & Richardson 1920; Norris 1924; Hilton 2004). Species of Scaphirhynchus possess a long caudal filament (circus) representing an extension of the notochord. The caudal filament likely has some sensory function because it has nerves and a lateral line along its length (Weisel 1978); this structure is often missing or incomplete in adults (Bailey & Cross 1954). In Scaphirhynchus the lower lobe of the deeply forked heterocercal caudal fin is reduced to allow for tail movement in close proximity to the bottom (Findeis 1997; Vecsei & Peterson 2004).

170

FRESHWATER FISHES OF NORTH AMERICA

Other modifications for a benthic lifestyle include the pectoral fin spines that some species of Scaphirhynchus use to walk along the substrate and an abdominal area protected with heavy scales (Findeis 1997). Species of Scaphirhynchus also possess varying degrees of head spines, including preorbital, parietal, post-temporal, and tabular spines with some species having recurved spines on the dorsal tip of the rostrum (Bailey & Cross 1954; Mayden & Kuhajda 1996; Vecsei & Peterson 2004).

Swimming Sturgeons are classified as benthic cruisers, swimming horizontally just above the substrate in search of food (Findeis 1997; Vecsei & Peterson 2004). The classic model explaining how fishes with heterocercal tails such as Sharks and Sturgeons maintain a horizontal cruising plane postulates that the lift and movement produced by

the heterocercal tail would pitch the head ventrally if not countered by lift produced by the pectoral fins. But a new model based on three-dimensional coordinate data from tank observations on juvenile White Sturgeons proposes that a horizontal cruising plane is maintained with lift from the dorsally angled ventral body surface both anterior and posterior to the body center that is countered with the negative buoyancy of the Sturgeon at the body center (center of mass); heterocercal tail oscillations only generate forward thrust and the pectoral fins do not generate lift. But the posterior portions of the pectoral fins are actively moved ventrally or dorsally to induce rising or sinking respectively by reorienting the head and body in the flow (Fig. 5.10). In current speed of 0.5 body length/s, Sturgeons have a positive body tilt of 20 degrees. In current speed of 1.0 body length/s Sturgeons have a positive body tilt of 8 degrees (Wilga & Lauder 1999).

Size

Figure 5.10. Diagram of vertical force balance on swimming Sturgeon with X as the center of mass and vertical arrows indicating force exerted by the fish as the posterior portions of the pectoral fins are actively moved ventrally or dorsally to induce rising or sinking, respectively, by reorienting the head and body in the flow. The tail arrow indicates forward force (Wilga & Lauder 1999; reproduced with permission of The Journal of Experimental Biology).

Sturgeons are the largest fishes found in fresh water (Tables 5.1 and 5.2). Acipenser contains the largest North American Sturgeons, ranging from the giant White Sturgeon that can reach 6.1 m TL (20 feet) and weigh 816 kg (1,800 pounds), to the Shortnose Sturgeon that attains 1.4 m TL (4.6 feet) and 23 kg (51 pounds). The Pallid Sturgeon is the largest species of Scaphirhynchus, reaching a maximum size of 1.7 m TL (5.6 feet) and 45 kg (99 pounds); the Alabama Sturgeon is the smallest at 78 cm TL (2.5 feet) and 3 kg (6.6 pounds) (Williams & Clemmer 1991; Cech & Doroshov 2004). Female Sturgeons attain larger sizes (and greater ages) compared with males, but otherwise show no obvious external sexual dimorphism except when they are ripe. Before spawning, females have swollen abdomens full of ripe eggs, and males are more elongate, but substantial overlap can occur between the sexes (Vladykov & Greeley 1963; Noakes et al. 1999; Kennedy et al. 2006; Van Eenennaam et al. 2006). Reliable minimally invasive techniques to determine the sex of Sturgeons and reproductive readiness include, in order of increasing accuracy, field ultrasound, endoscopy, and blood plasma assays (Colombo et al. 2004; Wildhaber et al. 2007; Craig et al. 2009; Divers et al. 2009; Heise et al. 2009).

Early Life Stages Sturgeons progress through large morphological changes in early life stages. Immediately upon hatching larvae have yolk sacs and are classified as proto-larvae that lack

Table 5.1. Life history attributes for five species of Sturgeons in the genus Acipenser in North America. Life History Attribute Strictly freshwater Maximum size recorded in length and weight Maximum age Age at first reproduction

Acipenser (five species, one with two subspecies) Anadromous (two), semi-anadromous (two), potamodromous (one) 1.43 m TL (4.7 feet) and 23 kg (51 pounds) to 6.1 TL m (20 feet) and 816 kg (1,800 pounds) 60–152 years Males 2–12 years and females 4–18 years to males 15–20 years and females 22–33 years

Iteroparous or semelparous Fecundity estimates (ovarian counts)

Iteroparous 27,000–208,000 to 400,000–2.6 million

Mature egg diameter

1.74–2.49 mm (0.07–0.10 inch), average 2.21 mm (0.09 inch), to 4.04–4.66 mm (0.16–0.18 inch), average 4.33 (0.17 inch)

Egg deposition sites

Hard or rocky substrate

Clutch size Range of spawning dates and temperatures

957–1,444 eggs/spawning bout (A. fulvescens) March–July; 8.8–23°C (47.8–73.4°F)

Habitat of spawning sites; average water depth

Hard or rocky substrate; 0.5–4.7 m (1.6–15.4 feet) to 2–27 m (6.6–88.6 feet)

Incubation period; larval type at hatching

2.3–2.5 days at 22.2–23.3°C (72–73.9°F) to 8–14 days at 10–14°C (50–57.2°F); all species protolarvae at hatching

Mean size at hatching

Average 7.1 mm (0.28 inch) TL to 12.6–15 mm (0.50–0.59 inch) TL

Parental care Major dietary items

None Nematodes, oligochaetes, amphipods, aquatic insects, mollusks, crayfishes, and fishes in fresh water; mysids, copepods, and fishes in brackish and marine waters Anadromous adults in near-shore marine waters and semi-anadromous adults in estuaries in brackish water, except when spawning; potamodromous and landlocked Sturgeon in large rivers or lakes with adults occasionally found in brackish water Migratory (one) and diadromous (four) All species except one (A. fulvescens) with a distinct population or all populations federally Endangered or Threatened

General year-round habitat

Migratory or diadromous Imperilment status

171

References Boreman 1997; Wilson & McKinley 2004 Cech & Doroshov 2004 Cech & Doroshov 2004 Dadswell 1979; Cochnauer et al. 1985; Smith 1985; Kynard 1997; McLeod et al. 1999; Van Eenennaam et al. 2006 Boreman 1997; Billard & Lecointre 2001 Dadswell 1979; DeVore et al. 1995; Boreman 1997; Van Eenennaam & Doroshov 1998; Van Eenennaam et al. 2001, 2006; Bruch & Binkowski 2002; Bruch et al. 2006 Scott & Crossman 1973; Dadswell 1979; Cherr & Clark 1985; Parauka et al. 1991; Van Eenennaam et al. 1996, 2006; Van Eenennaan & Doroshov 1998; Bruch et al. 2006 Parsley et al. 1993; Sulak & Clugston 1999; Bruch & Binkowski 2002 Bruch & Binkowski 2002 Dadswell 1979; Parsley et al. 1993; Sulak & Clugston 1999; Fox et al. 2000; Bruch & Binkowski 2002; Erickson et al. 2002; Perrin et al. 2003; Wilson & McKinley 2004; Van Eenennaam et al. 2005, 2006 Parsley et al. 1993; Fox et al. 2000; Bruch & Binkowski 2002; Perrin et al. 2003; Wilson & McKinley 2004 Smith et al. 1980; Buckley & Kynard 1981; Kempinger 1988; Parauka et al. 1991; Richmond & Kynard 1995; Kynard 1997; Van Eenennaam et al. 2001; Deng et al. 2002 Smith et al. 1980; Buckley & Kynard 1981; Richmond & Kynard 1995; Kynard 1997; Bardi et al. 1998; Van Eenennaam et al. 2001; Deng et al. 2002; Snyder 2002 Bruch & Binkowski 2002 Hatin et al. 2002; Jackson et al. 2002; Wilson & McKinley 2004

Wilson & McKinley 2004

Erickson et al. 2002; Wilson & McKinley 2004 Auer 2004; Wilson & McKinley 2004; NMFS 2006, 2010ab

172 FRESHWATER FISHES OF NORTH AMERICA

Table 5.2. Life history attributes for three species in the genus Scaphirhynchus (—, not applicable). Life History Attribute

Scaphirhynchus (three species)

References

Strictly freshwater Maximum size recorded in length and weight Maximum age Age at first reproduction

Potamodromous 0.78 m TL (2.5 feet) and 3 kg (6.6 pounds) to 1.68 m TL (5.5 feet) and 45 kg (99 pounds) 41–43 years Males, ≥5 years; females, 6–12 to 15–20 years

Iteroparous or semelparous Fecundity estimates (ovarian counts) Mature egg diameter

Iteroparous 7,069–65,490 to 170,000

Boreman 1997; Wilson & McKinley 2004 Williams & Clemmer 1991; Cech & Doroshov 2004 Keenlyne & Jenkins 1993; Everett et al. 2003 Keenlyne & Jenkins 1993; Keenlyne 1997; Kennedy et al. 2006 Boreman 1997; Billard & Lecointre 2001 Keenlyne et al. 1992; Kennedy et al. 2006

Egg deposition sites Clutch size Range of spawning dates and temperatures Habitat of spawning sites; average water depth Incubation period; larval type at hatching Mean size at hatching Parental care Major dietary items

General year-round habitat

Migratory or diadromous Imperilment status

2.0–2.5 mm (0.08–0.10 inch) to an average of 2.6 mm (0.10 inch) Hard or rocky substrate Unknown February–July; 16.9–21.2°C (62.4–70.2°F) Unknown 5 days at 20°C (68°F), 5–14 days at 14°C (57.2°F), 7–13 days at 13–16.5°C (55.4–61.7°F); all species proto-larvae at hatching 7–9 mm (0.28–0.35 inch) TL None Aquatic insects, snails, mussels, fish eggs, and occasionally fishes, switching to mostly piscivorous diet as juvenile and adult in S. albus Main channel, borders, or pools downstream of sand bars or wing dikes in large rivers in areas of high to moderate flows over stable substrate Migratory Two species federally Endangered and one species with a distinct population federally Threatened

obvious Sturgeon features. As proto-larvae grow they completely absorb their yolk sacs and develop a subterminal mouth and barbels shortly before active feeding (6–10 days post-hatch). Mesolarvae develop median fin elements, an elongated rostrum, teeth (lost as juveniles) (Fig. 5.11), and rows of scutes, and resemble miniature adults at the end of this stage (around 31–45 days post-hatch). Members of Acipenser metamorphose into juveniles at this point and are from 31 to 94 mm TL. In Scaphirhynchus, individuals pass into a third larval stage (metalarvae) at a size of 55–90 mm TL, where all fin rays are present but an anal fin fold persists, and this fold is present in specimens ≥200 mm TL (Bath et al. 1981; Richmond & Kynard 1995; Deng

Keenlyne et al. 1992; Kennedy et al. 2006; Bryan et al. 2007 Parsley et al. 1993; Sulak & Clugston 1999; Bruch & Binkowski 2002 — Keenlyne & Jenkins 1993; Mayden & Kuhajda 1997a; Wilson & McKinley 2004 — Snyder 2002; Colombo et al. 2007b

Snyder 2002 Bruch & Binkowski 2002 Carlson et al. 1985; Gerrity et al. 2006; Hoover et al. 2007; Keevin et al. 2007; Rapp et al. 2011 Keenlyne 1997; Mayden & Kuhajda 1997a; Wilson & McKinley 2004 Mayden & Kuhajda 1997a; Wilson & McKinley 2004 Auer 2004; Wilson & McKinley 2004; USFWS 2010a

et al. 2002; Snyder 2002; Colombo et al. 2007b). Species identification of proto-larvae within genera, although difficult, can be realized (Deng et al. 2002; Synder 2002).

Paedomorphosis and Peramorphosis Conventional views invoke paedomorphosis ( juvenile characters expressed in adults) to explain the secondary de-ossification of the largely cartilaginous skeleton and the loss of dermal elements in Sturgeons, and accept paedomorphosis as one of the driving forces in Sturgeon evolution (Traquair 1877; Goodrich 1909; Gregory 1933; Yakovlev 1977; Vecsei & Peterson 2004). Rare documented

ACIPENSERIDAE: STURGEONS

173

direct evidence for paedomorphosis includes delayed ossification of skeletal elements in adult Shortnose Sturgeons (Bemis et al. 1997). This, however, is not conclusive (Hilton & Bemis 1999), and no studies support the role of delayed ossification in evolution of Sturgeons (Bemis et al. 1997; Findeis 1997). In fact, phylogenies of acipenserids based on osteology demonstrate that additions and enlargement in skeletal elements and an increase in scalation occurred within Sturgeons at all phylogenetic nodes, which suggests that peramorphosis (addition of new or enlargement of existing structures compared with outgroups), rather than paedomorphosis, has played a central role in Sturgeon evolution (Findeis 1997).

GE NE TICS

Figure 5.11. White Sturgeon, Acipenser transmontanus, larva 12 days post-hatch (22.28 mm TL) showing teeth that are lost in juveniles and adults (photograph by and used with permission of Katie May Laumann).

Karyology Sturgeons possess a karyotype with large numbers of chromosomes. Species are separable into classes of about 120, 250, and perhaps 500 chromosomes (Birstein et al. 1997b; Fontana 2002). The latter group is tentative because it is based on DNA content and not on actual number of chromosomes (Blacklidge & Bidwell 1993). Up to half of these chromosomes are small microchromosomes with the remaining consisting of larger metacentric and submetacentric macrochromosomes. A diploid ancestor to all Acipenseriformes likely had 60 chromosomes, and a gene duplication event created a tetraploid (4n) ancestor (Birstein et al. 1997b; Ludwig et al. 2001). Some consider extant Sturgeons as tetraploid, octoploid (8n), and perhaps 16n-ploid species (Birstein et al. 1997b; Kim et al. 2001), which is the highest level of polyploidy in fishes (Vasil’ev 1999). Others treat Sturgeons with 120 chromosomes as functional diploids and those with 250 chromosomes as tetraploids (Ludwig et al. 2001; Fontana 2002; Fontana et al. 2004). The chromosome number for North American Sturgeons is known in 5 species with the Shovelnose Sturgeon and Atlantic Sturgeon in the 120 chromosome group, and the Lake Sturgeon, Green Sturgeon, and White Sturgeon in the 250 chromosome group. No direct data are available for the other species of Scaphirhynchus or Shortnose Sturgeon, but the genome size of the Shortnose Sturgeon suggests it may have 360 or 500 chromosomes (Birstein et al. 1997b; Fontana 2002; Fontana et al. 2004). Phylogenetic relationships and divergence times based on molecular data indicate that several genome duplication events created the diverse polyploidy in North American Sturgeons, ranging from 5°C for April–October, Lake Sturgeon populations at higher latitudes actually have a higher growth rate compared with more southern populations (Power & McKinley 1997).

BEHAVIOR

Diel Periodicity Sturgeons vary in their diel movements. Adult Scaphirhynchus do not display distinct diel movements, but Shovelnose Sturgeons tend to be more active at night compared with more daytime movement for Pallid Sturgeons. These behaviors are suggestive of temporal resource partitioning (Curtis et al. 1997; Bramblett & White 2001). Overall Shortnose Sturgeons do not demonstrate diel movement patterns, but individuals can occupy shoal habitat more often at night (Kynard et al. 2000). Some species show strong diel patterns with Gulf Sturgeons more active at night in all seasons except summer; spawning and downstream migration in autumn take place almost exclusively at night (Sulak & Clugston 1999; Wrege et al. 2011). Pronounced nocturnal behavior was observed in captive Green Sturgeons (Lankford et al. 2003), and Green Sturgeons were more active and occupied shallower depths at night in marine habitats (Erickson & Hightower 2007). In drift migration of early life history stages, activity differences occur within genera and species. Feeding larvae are diurnal in the Pallid Sturgeon but nocturnal in the Shovelnose Sturgeon, which may optimize larval migration and feeding for each species (Kynard et al. 2002a). In Atlantic Sturgeon larvae start as nocturnal migrators

180

FRESHWATER FISHES OF NORTH AMERICA

then switch to both nocturnal and diurnal migration (Kynard & Horgan 2002a), but larvae of other species of Acipenser show consistent nocturnal activity (Kempinger 1988; Richmond & Kynard 1995).

Movement and Non-spawning Migrations Generally Sturgeons migrate to optimize feeding and reproductive success. Downstream migrations are associated with feeding, especially for anadromous and semianadromous species; brackish and marine environments typically contain more food than freshwater habitats. Upstream migrations usually are associated with spawning (see reproductive section), although feeding may be involved (Auer 1996; Bemis & Kynard 1997). Migration also may be associated with avoidance of unfavorable environmental conditions (Auer 1996). Potamodromous and amphidromous Sturgeons typically remain in their natal river basins or estuaries throughout their lives, but anadromous and semianadromus Sturgeons can range widely along the coast (Kynard 1997) with movements of ≤968 and 1,000 km (602 and 621 miles) for the Green Sturgeon and White Sturgeon, respectively (Erickson & Hightower 2007; Welch et al. 2006). For semi-anadromous species movement into saltwater habitats may occur shortly after spawning and adults may remain in salt water for all or most of the year, but some populations upstream of dams remain in fresh water all year. Abundance of forage and suitable thermal regimes likely dictate movement for most non-spawning individuals (Kynard 1997), and this may be true for potamodromous and anadromous species as well. All ripe adult Sturgeons demonstrate upstream migration in autumn, winter, and spring followed by downstream movement in summer and autumn (see the following subsection and reproduction section). Gulf Sturgeon males and females average 6.4 and 16.0 km/day (4.0 and 9.9 miles/day), respectively, as they move downstream from rivers into marine environments in autumn (Parkyn et al. 2007). Spawning migrations can cover almost 500 km (311 miles) for both Acipenser and Scaphirhynchus species with rates of ≤22 km/day (13.7 miles/day) (Wilson & McKinley 2004; Parkyn et al. 2007). But movement by non-spawning adults and early life history stages can be extensive and complex. Non-migratory movement is typically limited to 6 m), slow water for the Lake Sturgeon (Rusak & Mosindy 1997) compared with shallow (1–2 m) channel crossovers with slow bottom flows for the Shovelnose Sturgeon (Quist et al. 1999). Semi-anadromous species overwinter in deep (10–27 m) saltwater estuaries and estuarine lakes (Wilson & McKinley 2004).

Other sensory systems besides olfaction and taste likely also play a role in feeding. Vision may play an important role in helping Acipenser larvae detect and capture prey items as they alternate between foraging on the bottom and in the water column during downstream migration. The swift pursuit and capture of zooplankton in the water column likely is guided by the ability of Sturgeons to visually detect moving objects in illuminated habitat (Kynard & Horgan 2002a). Ampullary electroreceptors concentrated on the underside of the snout of Sturgeons also are likely used in feeding, and these structures are well developed in feeding larvae of Acipenser (Northcutt 1986; Miller 2004). Early larval foraging behavior in the Shovelnose Sturgeon and Atlantic Sturgeon occurs predominantly in the water column with foraging at the water-air interface, including swimming inverted with the ventral surface up. As these larvae develop they switch to mostly benthic feeding. Larvae of the Lake Sturgeon use benthic feeding throughout their development (Ross & Bennett 1997; Kynard & Parker 2004).

Chemosensory Systems and Feeding Given the dorsolateral position of Sturgeons’ eyes and the deep, turbid environment they typically inhabit, vision does not play a dominant role in feeding (Sillman et al. 1999; Miller 2004). Instead well-developed chemosensory systems are used. Olfaction allows for location of food items; external taste buds on the barbels and lips (extraoral taste) trigger ingestion behavior, and internal oral taste buds determine whether food items are swallowed or rejected (Kasumyan 1999, 2002). In feeding larvae and juveniles of Acipenser and Scaphirhynchus, food odor stimulates feeding responses that include hovering close to the bottom with barbels trailing on the substrate as the fish moves in circular or S-shaped trajectories (scouring behavior). While swimming, the bodies of Sturgeons oscillate from side to side, allowing foraging over a wide area (Ross & Bennett 1997; Kasumyan 1999, 2002). Low concentrations of food odor (1 μM) elicit a feeding behavior response. The spectrum of amino acids that induce olfac-

Camouflage Coloration Larvae of some species of Acipenser and all species of Scaphirhynchus studied to date (Pallid Sturgeon and Shovelnose Sturgeon) have light bodies and black tails that may provide cryptic coloration in benthic and drifting (mid- to upper-water column) habitats, and the wigwag swimming motion of the black tail in feeding larvae may create a confusing strike zone for predators (Kynard et al. 2002b; Gisbert & Ruban 2003; Kynard & Parker 2005). Young Shortnose Sturgeons and Lake Sturgeons have black blotches that break up the body pattern and may act as camouflage; the blotched pattern disappears in subadults. Even so, bony plates and sharp scutes may provide all the protection from predation necessary for juvenile Sturgeons; hence, the advantage of blotched

ACIPENSERIDAE: STURGEONS

coloration is uncertain (Scott & Crossman 1973; Wallus 1990a).

Jumping and Sound Production Species of Acipenser frequently jump entirely out of the water. Jumping may occur to produce sounds used in communication in the Gulf Sturgeon (Sulak et al. 2002), but jumping observed in the Lake Sturgeon and Atlantic Sturgeon may serve to remove attached parasitic lampreys (Becker 1983; Scott & Scott 1988). Jumping also occurs in the Lake Sturgeon before and during spawning activities adjacent to spawning sites in association with a more common porpoising behavior (see reproduction section), but the purpose is unknown (Bruch & Binkowski 2002). Distinct aural signals are produced by Sturgeons (see also reproduction section). Gulf Sturgeons, as well as other Sturgeons, frequently jump out of the water, and the sound produced may be a form of communication to maintain group cohesion. Jumping is prevalent during the summer when Gulf Sturgeons congregate in deep holding areas in rivers; feeding does not occur during this time. The greatest number of jumps occurs in June with peak activity in all summer months near dawn and to a lesser degree near sunset. Sonograms revealed nine distinct sounds, including acoustic signatures produced by exit, reentry, and splash subsidence that differed from the sounds of jumping Striped Mullet (Mugil cephalus) and of dropped objects. Because of the significant energy cost of jumping during a fasting period, this behavior likely provides some benefit to the Gulf Sturgeon. This, coupled with the distinctive sounds produced, suggests that jumping is a form of communication (Sulak et al. 2002). Jumping also occurs in estuary and marine habitats for Gulf and Green Sturgeons (Edwards et al. 2007; Erickson & Hightower 2007).

REPRODUCTION

Seasonality Worldwide Sturgeons spawn in all seasons and in variable water flow and temperature conditions (Bemis & Kynard 1997). All North American Sturgeons typically spawn in the spring with some southern populations spawning as early as February and Sturgeons in northern areas extending spawning into late June (Tables 5.1 and 5.2; Wilson & McKinley 2004; Beamesderfer et al. 2007). Populations of the Atlantic Sturgeon in southern South Carolina can

183

make two spawning runs. Some migrate upriver from the Atlantic to spawn in autumn with males and females running ripe and spent females found later in the season; histological examination of gonad biopsies verified spawning. The two season spawning events are corroborated with strong bimodal length distribution of age-1 Atlantic Sturgeons in the Edisto River, South Carolina. Autumn migrants are typically smaller than those making spring spawning runs (Smith et al. 1984; Collins et al. 2000a; McCord et al. 2007). Small numbers of gravid and running Gulf Sturgeon females and males are captured in autumn in the Suwannee River, Florida, with late spring capture of small juveniles equivalent in size to 6- to 10-month-old fish (Sulak & Clugston 1998). Capture of Shovelnose Sturgeon milting males, females with ripe eggs, males and females with elevated plasma sex steroid concentrations, and larvae in the Mississippi River in autumn indicate an autumn spawn, but the contribution of autumn cohorts to the population is unknown (Divers et al. 2009; Stahl et al. 2009; Tripp et al. 2009b).

Age at Sexual Maturity Sturgeons have late sexual maturation (Tables 5.1 and 5.2) that is an important life history parameter to consider in their conservation and management (Birstein 1993; Bemis & Kynard 1997). Aging techniques, however, may underestimate the actual age at sexual maturity (see ecology section). Size actually may be more important than age in the initiation of maturation (Caron et al. 2002). Males mature sooner than females, but a wide range of maturation ages is present between and within species, particularly across sizes and latitudes. Small, more southern species or populations mature sooner than large, more northern species or populations (Dadswell 1979; Smith 1985; Billard & Lecointre 2001; Auer 1999). The Shovelnose Sturgeon is a smaller relative of the Pallid Sturgeon, and female Shovelnose Sturgeons mature at 6–12 years versus 15–20 years in the larger species; males of both species mature at ≥5 years old (Keenlyne & Jenkins 1993; Keenlyne 1997; Kennedy et al. 2006; Tripp et al. 2009b; Stahl et al. 2009). Shortnose Sturgeons, the smallest North American species of Acipenser, mature at 2–5 years (males) and 4–5 years (females) in southern populations, but northern populations mature at 11–12 years (males) and 12–18 years (females) (Dadswell 1979; Kynard 1997). This contrasts with the intermediate-sized Green Sturgeon, which matures at 14–16 years (males) and 16–20 years (females) (Van Eenennaam et al. 2006), and the

184

FRESHWATER FISHES OF NORTH AMERICA

larger Lake Sturgeon in which northern populations mature at older ages of 15–20 years (males) and 22–33 years (females) (McLeod et al. 1999). The largest species of Acipenser have a wide range of maturation ages, from 5 to 34 years for the Atlantic Sturgeon to 11 to 34 years for the White Sturgeon (Cochnauer et al. 1985; Smith 1985). Many Sturgeons mature earlier in captivity due to warmer water temperatures and high-quality commercial feed. Late-maturing, wild White Sturgeons typically attain early maturity in captivity at 3–4 years for males and 6– 14 years for females (Birstein 1993; Doroshov et al. 1997; Van Eenennaam et al. 2004).

Natal Fidelity Several species of Acipenser show high levels of fidelity to the rivers in which they were spawned, migrating in winter or spring into those rivers to spawn themselves. Genetic studies demonstrated natal fidelity in the Shortnose Sturgeon, Atlantic Sturgeon, and Gulf Sturgeon (e.g., King et al. 2001; Grunwald et al. 2002; Dugo et al. 2004). Likewise, molecular and tagging studies demonstrated fidelity to rivers and sites within rivers in adult Lake Sturgeons (Lyons & Kempinger 1992; Auer 1999; Knights et al. 2002; DeHaan et al. 2006). Site fidelity can arise from either particular characteristics of the site or from homing through olfactory imprinting in larvae as they begin to feed, but the specific means used by Sturgeons to return to their natal rivers is unknown and needs to be tested (Boiko et al. 1993; Bemis & Kynard 1997; Kasumyan 1999).

Spawning Migrations The distance covered during spawning migrations by Huso and Acipenser species is correlated positively with average adult size. The smallest North American species of Acipenser, the Shortnose Sturgeon, typically migrates ≤200 km (124 miles) compared with the largest species, the White Sturgeon, which can migrate ≤1,200 km (746 miles). Similarly, some large Asian Sturgeons have spawning migrations of 3,300 km (2,051 miles). But not all North American species follow this trend. Lake Sturgeons are mid-sized with short migrations of usually 200 km (124 miles) occur that may reflect greater energy resources due to warmer water and perhaps winter feeding (Kynard 1997). In contrast to the anadromous Shortnose Sturgeon, the potamodromous Lake Sturgeon has a relatively simple and consistent autumn pre-spawning migration from lake or downstream habitats to deep holes within tributary river systems. These winter staging areas are often adjacent to spawning sites (Kempinger 1988; McKinley et al. 1998; Bruch & Binkowski 2002).

ACIPENSERIDAE: STURGEONS

Spawning Territories Sturgeons do not exhibit any territoriality, but male Lake Sturgeons maintain position at a spawning site for the duration of spawning activities (2 weeks), and they will continue to stage near spawning grounds for ≤1 month if gravid females are present awaiting warmer water temperatures to spawn a second time (Bruch & Binkowski 2002).

Spawning Frequency Sturgeons typically do not spawn every year in the wild, and most have multiple years between spawning events. The reproductive hiatus is typically longer for females, northern populations within a species, and larger species. Male and female Shovelnose Sturgeons and male Pallid Sturgeons spawn at 2- to 3-year intervals, but female Pallid Sturgeons range from 3 to 10 years between spawning events (Mayden & Kuhajda 1997b; Kennedy et al. 2006; Tripp et al. 2009b). Large species of Acipenser have extensive ranges for spawning intervals, including 3–11 years for both sexes in White Sturgeons and 2–7 years for male and 3–12 years for female Lake Sturgeons (Cochnauer et al. 1985; Auer 1999; Peterson et al. 2003). The smaller Shortnose Sturgeon has relatively short spawning intervals of 2 years for males and 3–5 years for females in northern populations compared with every year for males and perhaps every 1–2 years for females in more southern populations (Dadswell 1979; Kieffer & Kynard 1993; Kynard 1997). Green Sturgeons also have a relatively short spawning frequency of 2–4 years (Beamesderfer et al. 2007; Erickson & Webb 2007). The Atlantic Sturgeon is one of the largest North American species, but males spawn every year and females every 3 years with both sexes having intervals ≤5 years (Smith 1985; Collins et al. 2000a; Fox et al. 2000). With less variable temperatures and more food availability in captivity, some species show a shortened time between spawning events relative to wild populations. In captive Siberian Sturgeon populations (Acipenser baerii), e.g., males spawn every year and females spawn every 1.5–2 years (Birstein 1993).

Spawning Modes and Location Sturgeons typically are considered broadcast spawners, releasing large numbers of eggs over extensive areas of river bottom (Tables 5.1 and 5.2; Cherr & Clark 1985; Parsley et al. 1993, 2002), more specifically referred to as lithophilic riverine spawners (explicitly the genus

185

Acipenser), meaning spawning in close proximity or directly on hard or rocky substrates (Fig. 5.12; Sulak & Clugston 1999; Bruch & Binkowski 2002). Upstream spawning migrations occur in all species with potamodromous Sturgeons entering tributaries to lakes or large rivers or using borders of main channels. Semianadromous and anadromous species leave marine, brackish, or estuarine habitats to spawn in the river main stem. Spawning occurs at temperatures from 17 to 21°C (63– 70°F) in Scaphirhynchus and 8.8 to 21.5°C (47.8–70.7°F) for most Acipenser; Gulf Sturgeons spawn in slightly warmer waters (18.3–23°C, 64.9–73.4°F) (Bruch & Binkowski 2002; Wilson & McKinley 2004; Beamesderfer et al. 2007; Erickson & Webb 2007). All North American Sturgeons require strong flows over hard substrates at spawning sites, often in the swiftest water available. Spawning substrates are typically hard bottoms consisting of hard clay, gravel, rubble, boulders, bedrock, and rocky ledges in high-velocity areas ≤2.4 m/s (7.9 feet/s) near the substrate (Fig. 5.12; Parsley et al. 1993; Wilson & McKinley 2004). Some spawning sites for Lake Sturgeons have flow of 1 individual). This includes polyandry with each female spawning with numerous males both within a single spawning bout and over the 8- to 12-h spawning period, and polygyny, in which males remain at the spawning site for the duration of spawning activity (1–4 days) and spawn with numerous females. This polygamous breeding system maximizes the genetic diversity of offspring (Bruch & Binkowski 2002). In contrast, Shortnose Sturgeons might form pair bonds as suggested by the capture of the same individuals side by side after 1- to 3-year intervals (Dadswell 1979). Sex ratios in Sturgeons can vary between and within species, but general patterns are evident when comparing the general population versus populations at spawning sites and comparing different size classes. When populations are sampled in all or several seasons or data are gathered from sport or commercial harvest, sex ratios are typically nearly equal for the Lake Sturgeon, White Sturgeon, Green Sturgeon, and some populations of the Atlantic Sturgeon and Shovelnose Sturgeon (Threader & Brousseau 1986; DeVore et al. 1995; Van Eenennaam & Doroshov 1998; Auer 1999; Colombo et al. 2007c; Webb & Erickson 2007) but favor females in ratios ≥2.8F:1M for the Shortnose Sturgeon and Pallid Sturgeon, and some populations of the Atlantic Sturgeon and Shovelnose Sturgeon (Dadswell 1979; Smith et al. 1984; Carlson et al. 1985). Higher numbers of females likely are the result of their longer lifespan, a supposition supported by dominance of females in larger size classes (Probst & Cooper 1955; Dadswell 1979; Beamesderfer et al. 1995). In contrast, males outnumber females during spawning runs and at spawning sites for all species of Sturgeons examined with ratios ≥9.6M:1F (Lyons & Kempinger 1992; Carr et al 1996; Van Eenennaam et al. 1996, 2006; Auer

Sturgeons are large fishes that release large numbers of eggs over extensive areas of river bottom (Tables 5.1 and 5.2; Cherr & Clark 1985; Parsley et al. 1993, 2002), e.g., ≤2.6 million eggs in the Atlantic Sturgeon (Van Eenennaam & Doroshov 1998). But most Sturgeon species have substantially lower fecundity compared with other fishes with similar spawning modes; that fact, along with their late maturation, are among the main factors that make Sturgeons susceptible to overfishing (Boreman 1997). As with most fishes, egg production increases with female Sturgeon size, both among and within species, and more fecund species have smaller mature (stage-5) eggs. The two largest North American Sturgeons, Atlantic and White Sturgeons, produce ≤2.6 and 1.5 million mature eggs/female (absolute fecundity) for a 231 cm FL (91 inch) and a 309 cm TL (122 inch) female, respectively, but smaller females (183 and 135 cm FL or 72 and 53 inches) have absolute fecundities of 400,000 and only 98,200 eggs. Relative fecundity (number of eggs/kg or pound of body weight) range from 7,000 to 22,000 eggs/kg (3,175– 9,980/pound) for Atlantic Sturgeons. Mature egg diameter is 2.38–2.93 mm (0.09–0.12 inch) for Atlantic Sturgeons and 3.5–4.0 mm (0.14–0.16 inch) for White Sturgeons (Cherr & Clark 1985; Dettlaff et al. 1993; DeVore et al. 1995; Van Eenennaam et al. 1996; Boreman 1997; Van Eenennaam & Doroshov 1998). The Lake Sturgeon, a mid-sized Acipenser species, averages 383,529– 425,000 eggs/female and has an average relative fecundity of about 11,000 eggs/kg (5,000/pound) at a female weight of 21.3–42 kg (47–93 pounds). Mature egg diameters are 2.6–3.5 mm (0.10–0.14 inch) (Scott & Crossman 1973; Bruch & Binkowski 2002; Bruch et al. 2006). Absolute fecundity for the smaller Shortnose Sturgeon ranges from only 27,000 to 208,000 eggs, but has a similar average relative fecundity of 11,568 eggs/kg (5,247/pound) at a female weight of 3–19 kg (7–42 pounds). Mean egg diameter is 3.1 mm (0.12 inch) (Dadswell 1979). Although inter-

ACIPENSERIDAE: STURGEONS

mediately sized, the Green Sturgeon is relatively less fecund than other species, producing only 52,000– 242,000 eggs/female or a relative fecundity of 1,900– 4,200 eggs/kg (862–1,905/pound) at a female size of 153– 203 cm FL (60–80 inches). But they have the largest egg diameter of any North American Sturgeon at 4.04– 4.66 mm (0.16–0.18 inch) (Van Eenennaam et al. 2001, 2006). Within Scaphirhynchus, a large female Pallid Sturgeon (41 years old and 140 cm FL, 55 inches) produced 170,000 eggs or a relative fecundity of 9,936 eggs/kg (4,506/pound). Mean egg diameter was 2.0–2.5 mm (0.08– 0.10 inch) (Keenlyne et al. 1992). Smaller Shovelnose Sturgeons have a similar average egg diameter (2.6 mm, 0.10 inch), but lower absolute fecundity and higher relative fecundity. Fecundity can vary between Shovelnose Sturgeon populations with 7,069–16,708 eggs/female in the Missouri River compared with 13,241– 65,490 eggs/female in the Wabash River, Indiana (62– 86 cm FL or 24–34 inches for Wabash River), but relative fecundity can be similar with 13,090–25,080 eggs/kg (5,938–11,376/pound) in the Missouri River and 11,220– 23,956 eggs/kg (5,088–10,866/pound) in the Wabash River (Kennedy et al. 2006; Bryan et al. 2007). In the middle Mississippi River absolute and relative fecundity varies greatly between individuals, from 5,733–81,842 eggs/female (55.9–76.7 cm FL or 22.0–30.2 inches) and 6,220–46,230 eggs/kg (2,821–20,970/pound) (Tripp et al. 2009b).

Gametes Sturgeons (and the Paddlefish) are unique in that sperm possess an acrosome even though eggs have a micropyle, and Sturgeon eggs are unique in having numerous micropyles (3–15). An acrosome is a cap over the sperm head that contains enzymes involved in sperm penetration through the extracellular envelope (chorion) of the egg. Most Ray-finned Fishes (Actinopterygii) have eggs with a single funnel-shaped micropyle at the animal pole that allows a single sperm to make direct contact with the inner plasma membrane of the egg for fertilization, eliminating the necessity for an acrosome on sperm. Sturgeon sperm, however, retain an acrosome even though their eggs have numerous micropyles. Within vertebrates, only Lampreys (Petromyzontidae), Hagfishes (Myxinidae), Sturgeons, and the Paddlefish possess an acrosomal process. When the Sturgeon’s acrosome is activated, a fertilization filament (acrosomal process) forms that aids in fertilization, but the exact function of the filament is unknown. Nu-

189

merous micropyles and other gamete morphologies and functions are potential reproductive adaptations for Sturgeons to successfully broadcast-spawn in a riverine environment (see physiology section; Cherr & Clark 1985; Psenicka et al. 2011).

Embryo Characteristics and Development Fertilized eggs (embryos) are adhesive and therefore attach to the surface of hard substrates; they are found in interstitial spaces 10–20 cm (4–8 inches) below the substrate surface (Bruch & Binkowski 2002). The timing of development is dependent on incubation temperatures with faster development at higher temperatures (4.6–5.8 days at 16–20°C, 60.8–68°F) for the Atlantic Sturgeon, Shortnose Sturgeon, and Shovelnose Sturgeon and only 2.3–2.5 days at 22.2–23.3°C (72–73.9°F) for Gulf Sturgeons (Smith et al. 1980; Parauka et. al 1991; Kynard 1997; Colombo et al. 2007b). Typically, however, Sturgeon embryos hatch in 7–13 days at 12–16.5°C (53.6–61.7°F) (Richmond & Kynard 1995; Kynard 1997; Deng et al. 2002; Snyder 2002). Viable temperatures for embryonic development range from 8.3°C (46.9°F) for northern Lake Sturgeon populations to ≤20°C (68°F) for Gulf Sturgeons (Bolker 2004). Thirty-six developmental stages, from the unfertilized egg through hatching, are identified for Sturgeons (Dettlaff et al. 1993; Colombo et al. 2007b) (Table 5.3). Major stages include fertilization, cleavage, blastula formation, gastrulation, neurulation, and organogenesis. Sturgeons undergo holoblastic cleavage in which each cleavage divides the entire egg cytoplasm, including the yolk. Because Sturgeon eggs are large and have significant amounts of yolk concentrated at the vegetal pole, the process of cleavage is slow except at the animal pole, so that when the first cleavage is complete at the vegetal pole several more already have formed at the animal pole (Bolker 2004; Colombo et al. 2007b). Sturgeon embryos have hatching glands (at the ventral base of the head in the Shovelnose Sturgeon) that secrete a hatching enzyme that makes the chorion soft and then fragile to assist embryos in hatching (Deng et al. 2002; Colombo et al. 2007b). The size of Sturgeons at hatching is related to the size of the egg and its yolk. Hatchlings of the Atlantic Sturgeon and Gulf Sturgeon are the smallest of all North American Acipenser, with larvae averaging 7.1 and 8.3 mm TL (0.28 and 0.33 inch) and having the smallest mature eggs (average 2.62 and 2.21 mm, 0.10 and 0.09 inch diameter) (Smith et al. 1980; Parauka et al. 1991; Bardi et al. 1998;

Table 5.3. Stages of embryonic development of Shovelnose Sturgeons, Scaphirhynchus platorynchus, reared at 20°C (68°F) (Colombo et al. 2007b). Stage

Time Post-Fertilization

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22

0h 0.75 h 1h 2h 3h 4h 5h 6h 7h 8h 9h 11 h 17 h 18 h 23 h 25 h 27 h 29 h 29 h 31 h 35 h 35 h

23 24 25

36 h 39 h 45 h

26

52 h

27

54 h

28 29

60 h 63 h

30 31

81 h 85 h

32 33 34 35

87 h * * 93 h

36

102 h

Description of Embryo Fertilization, formation of light spot at apex of animal pole Disappearance of polar spot after 1 mitotic interval Formation of eccentric crescent band in animal pole Formation of first cleavage furrow after 2.5 mitotic intervals Formation of second cleavage furrow after 4 mitotic intervals Formation of third cleavage furrows after 5 mitotic intervals, furrow into vegetal pole Formation of fourth cleavage furrow after 6.5 mitotic intervals, full cleavage in vegetal pole Formation of fifth cleavage Continued division in vegetal pole Formation of cleavage cavity in animal pole Early blastula, small blastomeres still visible in animal pole Late blastula, animal pole surface smooth Onset of gastrulation, band between animal and vegetal poles Formation of dorsal blastopore lip, blastocoel seen through thin cell layer at animal pole Two-thirds of embryo covered by blastoderm, primitive gut present Large yolk plug evident, blastocoel ring-like on ventral surface Small yolk plug, blastocoel dark spot on ventral surface, blastoderm covering most of embryo Gastrulation complete, formation of slit-like blastopore Onset of neurulation, neural plate and neural groove present Neural groove wider, neural fold appears in head region Excretory rudiments evident, neural folds thicken Excretory rudiments elongate, neural folds close anteriorly, neural tube begins to close in caudal region Neural tube closed, different brain regions begin to form Excretory rudiments thicken anteriorly, eye vesicles form, first pair of visceral arches appear Eye rudiments clearly visible, second pair of visceral arches appear, lateral plates evident to anterior of head, tail thickens, pronephros present as tubes, first somites appear Heart rudiment forming with fusion of lateral plates and prosencephalon, tail rudiment separating from yolk sac, third pair of visceral arches appears Heart rudiment as short tube, eye rudiments as slits, head thickens and separates from yolk sac, somites covering body Heart rudiment elongates, tail continues to elongate, fin fold appears Onset of heartbeat, heart S-shaped, embryo bent, hatching glands pronounced, olfactory sacs present Tail begins to straighten, fin folds evident, eye cups evident, hatching glands thickened Tail approaches heart, head deep and most cranial part separated from yolk-sac membrane, pronephros difficult to distinguish Tail end reaches head, fin fold formed, head continues to separate from yolk sac Tail straightens fully if embryo removed Embryo capable of movement Hatching of advanced embryos, embryo clear with no pigment in eye cups, yolk sac yellow, pigment plug not fully formed Mass hatching; well-developed yolk plug, proto-larvae swim up in water column

* Stage not seen.

190

ACIPENSERIDAE: STURGEONS

Van Eenennaam & Doroshov 1998). Newly hatched larvae of the White Sturgeon and Green Sturgeon are the largest at 10–11 and 12.6–15 TL mm (0.39–0.43 and 0.50–0.59 inch) and have the largest mature eggs (3.5–4.0 and 4.04–4.66 mm, 0.14–0.16 and 0.16–0.18 inch diameter), respectively (Cherr & Clark 1985; Van Eenennaam et al. 2001, 2006; Deng et al. 2002). The Shortnose Sturgeon is intermediate in hatchling and mature egg size, averaging 9.5 mm TL (0.37 inch) and 3.1 mm (0.12 inch) diameter, respectively (Dadswell 1979; Buckley & Kynard 1981). Investment in a greater amount of egg yolk produces hatchlings that are longer and heavier (Van Eenennaam et al. 2001; Deng et al. 2002). Lake Sturgeons have a wide range of hatchling sizes from 7–12 mm TL (0.27–0.47 inch). Within Scaphirhynchus both Pallid Sturgeon and Shovelnose Sturgeon hatchlings are 7–9 mm TL (0.27–0.35 inch) (Snyder 2002). Larval development and behavior are discussed in the morphology and behavior sections, respectively.

ECOL OGY

Habitat Sturgeons inhabit fresh, brackish, and marine waters in North America. All three species of Scaphirhynchus spend their entire lives in large rivers, and Lake Sturgeons occur in large rivers or lakes (potamodromous) with adults occasionally entering brackish water. Adult Shortnose Sturgeons and White Sturgeons use brackish water (semi-anadromous or amphidromous), although a few individuals of Shortnose Sturgeon use marine habitat, and some populations of White Sturgeon are landlocked. Green Sturgeons and Atlantic Sturgeons move into the sea as adults (anadromous) using near-shore areas. All Sturgeons must spawn in fresh water (Bemis & Kynard 1997; Boreman 1997; Wilson & McKinley 2004). In general each phase of an adult’s life history (spawning, postspawning, non-spawning, overwintering, and feeding) and each life stage (embryo, larva, juvenile, and subadult) requires a different habitat (Auer 1996; Parsley et al. 2002). Potamodromous, semi-anadromous, and anadromous species differ greatly in habitat use for many of these life history phases or stages (Rochard et al. 1990). Adult and subadult Scaphirhynchus inhabit the main channel, channel borders, or pools downstream of sand bars or wing dikes in large rivers in areas of moderate to high flows. They typically occur over sand but also are found over silt, gravel, rubble, and bedrock at depths from

191

0.9 to 10 m (3–32.8 feet) (usually 2–6 m, 6.6–19.7 feet) (Keenlyne 1997; Wilson & McKinley 2004; Gerrity et al. 2008). Pallid Sturgeons use sandy substrates, greater depths, wider river channels, and mid-channel bars more often than co-occurring Shovelnose Sturgeons (Bramblett & White 2001). Pallid Sturgeons prefer downstream island tips and areas between wing dams in the Mississippi River (Hurley et al. 2004b) compared with main channel habitat without islands in the Missouri River (Gerrity et al. 2008). Bottom current velocities and substrate appear to play a more important role in habitat selection than does depth (Quist et al. 1999). The Lake Sturgeon inhabits large lakes and rivers usually at 4–9 m (13.1–29.5 feet) depth (≤43 m, 141 feet), over a variety of substrates including clay, mud, sand, gravel, and rock (Wilson & McKinley 2004). Greatest availability of food within a water body predicts habitat preferences (Rusak & Mosindy 1997). Semi-anadromous Sturgeons (White Sturgeons and Shortnose Sturgeons) can occupy large pools or deep channels in the main stem of rivers, the mouth of rivers in either freshwater or estuarine conditions, or near-shore marine environments from 3 to 30 m (9.8–98.4 feet) in depth, typically occurring over sand, gravel, and cobble (Kynard et al. 2000; Wilson & McKinley 2004). Anadromous species (Green Sturgeon and Atlantic Sturgeon) enter fresh water to spawn and generally remain in rivers after spawning (summer and early autumn) in deep holding areas (≤14 m), then return to salt water in the winter to feed (Erickson et al. 2002; Wilson & McKinley 2004; Heise et al. 2005), using shallow estuary and near-shore habitats typically over sandy substrate (Fox et al. 2002). Summer holding areas are inhabited by both recently spawned and non-spawned adult and subadult anadromous species (Green Sturgeon and Atlantic Sturgeon) and are located downstream of spawning areas in the lower reaches of rivers that are still fresh water but may experience tidal influences. Sturgeons occupy deep areas over sand, clay, and gravel substrates; holding areas can average ≤28°C (82.4°F) in the South. In these habitats, individuals show little movement until migration to a marine environment occurs in autumn (Moser & Ross 1995; Sulak & Clugston 1999; Erickson et al. 2002; Heise et al. 2005). The Gulf Sturgeon does not feed while occupying summer holding areas (Mason & Clugston 1993), but nematodes, oligochaetes, and amphipods are found in the stomachs of adult and subadult Atlantic Sturgeons within these aggregation areas (Hatin et al. 2002). Coastal rivers are typically cooler than near-shore Gulf of Mexico waters

192

FRESHWATER FISHES OF NORTH AMERICA

ceans, Lancelets, annelids, sand dollars, polychaetes, and small bivalves (Fox et al. 2002; Edwards et al. 2007; Ross et al. 2009; Parauka et al. 2011). Adult Green Sturgeons inhabit deeper near-shore Pacific waters typically at depths of 40–70 m (131–230 feet) (Erickson & Hightower 2007). Anadromous Sturgeons from mixed stocks can be found in the same areas of concentration, possibly due to prey concentrations (Edwards et al. 2007; Erickson & Hightower 2007; Laney et al. 2007; Lindley et al. 2008; Ross et al. 2009).

Ontogenetic Shifts in Habitat Use Figure 5.13. A Gulf Sturgeon, Acipenser oxyrinchus desotol, in Fanning Springs, Suwannee River drainage, Levy County, Florida, 26 May 1989 (photograph by and used with permission of Noel M. Burkhead).

in the summer and may act as thermal refugia, but specific holding areas are not cooler than surrounding river water and are likely selected for depth as refuge from high-velocity currents (Fig. 5.13) (Sulak & Clugston 1999; Hightower et al. 2002; Sulak et al. 2007). Some nonspawning adult and subadult Atlantic Sturgeons may remain exclusively in a marine environment for several years, including the summer (Bain 1997), and a few Gulf Sturgeons have remained in bays over the summer (Duncan et al. 2011). Green Sturgeons from all known spawning populations were detected in Willapa Bay estuary, Washington, during the summer when water temperatures exceed coastal water temperatures by ≥2°C (3.6°F), exhibiting rapid and extensive intra- and inter-estuary movement. Green Sturgeons are likely using this habitat for foraging; they are not present in the winter when water temperatures are 10°C (50°F) (Moser & Lindley 2007). Overwintering habitat for adult potamodromous Sturgeons includes deep (>6 m, 19.7 feet), slower water for Lake Sturgeons (Rusak & Mosindy 1997) compared with shallow (1–2 m, 3.3–6.6 feet) channel crossovers with slow bottom flows for Shovelnose Sturgeons (Quist et al. 1999). Semi-anadromous species inhabit deep (10–27 m, 32.8–88.6 feet), lower saltwater estuaries and estuarine lakes (Wilson & McKinley 2004), but ripening females can overwinter in deep freshwater sites adjacent to spawning grounds (Dadswell 1979). Anadromous Sturgeons feed over the winter and this is reflected in their habitat use. For Gulf Sturgeons, males usually inhabit bays and estuaries, but females use near-shore marine habitat; both are found typically 2–6 m (6.6–19.7 feet) deep over sandy or shell hash substrate containing crusta-

Sturgeon embryos are attached to hard substrates, including gravel, cobble, boulders, hard clay, and bedrock, in fast-flowing water at the spawning site. Depths vary from 0.6 to 13 m (2 to 42.7 feet) (Fox et al. 2000; Wilson & McKinley 2004). Upon hatching larvae of some species may remain at the spawning site, while others may relocate downstream (Kynard et al. 2002a; Kynard & Horgan 2002a; Kynard & Parker 2005). Green Sturgeon larvae initiate downstream migration shortly after exogenous feeding begins, and laboratory studies indicate that habitat with slate-rock increases foraging effectiveness and growth rate and reduces mortality compared with sand and especially cobble substrate (Nguyen & Crocker 2007). Juvenile Green Sturgeons spend 1–4 years in freshwater and estuary habitats before entering the marine environment (Beamesderfer et al. 2007). Subadults use shallow regions of the San Francisco Bay estuary 160 countries to control international trade (Raymakers & Hoover 2002; Léonard et al. 2004; Pikitch et al. 2005). Under CITES, all Acipenseriformes are listed as Appendix I (threatened with extinction) or Appendix II species (uncontrolled trade might threaten their existence) that require permits and certifications for international trade involving Sturgeons and the Paddlefishes along with additional initiatives for their protection, especially against illegal trade (Raymakers & Hoover 2002; Léonard et al. 2004).

Artificial Propagation and Stocking With coastal Sturgeon fisheries in North America currently regulated, the lack of recruitment in extant populations and reestablishment of extirpated populations are the main obstacles facing Sturgeon recovery (Secor et al. 2002; Coutant 2004). Lack of recruitment stems from a historical reduction of spawning adults, reduction or loss of cues (temperature and flow) to initiate proper egg production and spawning migrations, limited access to or lack of quality spawning habitat, and improper conditions and habitats for embryos and larvae to develop and grow (Parsley et al. 2002). As such, a main tool used to recover Sturgeon populations and species is the stocking of hatchery-reared specimens that can bypass these recruitment constraints and produce stock for reintroduction into historical habitats. But consideration must be given to intraspecific genetic variation in natural populations due to fidelity for spawning sites, maintaining genetic diversity within wild populations, and the release of progeny that are free of disease and can adapt from a hatchery to a natural setting. Ultimately habitat restoration must occur to address recruitment failure (Andreasen 1999; Auer 2004; Pikitch et al. 2005). Reintroduction efforts are on-

going for all species of North American Sturgeons except the Alabama Sturgeon (Rider & Hartfield 2007; Koch & Quist 2010). Successful reintroductions include fast growth rates for hatchery-reared Lake Sturgeons reintroduced into a New York lake with potential spawning habitat available to allow for a self-sustaining population (Jackson et al. 2002), the reestablishment of an extirpated population of Lake Sturgeons in a Wisconsin river via stocking and adult translocation (Runstrom et al. 2002), hatchery-reared juveniles of the endangered population of Kootenai White Sturgeons becoming established (Ireland et al. 2002), and the successful reintroduction of hatchery-reared endangered Pallid Sturgeons in a riverine section of the upper Missouri River isolated by dams and reservoirs (Jordan et al. 2006; Shuman et al. 2011). Other populations of extant Pallid Sturgeons in the Missouri River with little or no natural recruitment for >20 years due to lack of proper habitat are being successfully augmented with hatchery-reared progeny to avoid extirpation (Gerrity et al. 2008; Braaten et al. 2009b; Steffensen et al. 2010; Shuman et al. 2011). Hatchery-reared Shortnose Sturgeons (both fertile and sterile) have the same seasonal movements and microhabitat selection as wild individuals, indicating that hatchery Sturgeons can be integrated into wild populations and can be used as surrogates for behavioral studies (Trested et al. 2011). Most of these reintroductions are too recent to determine if these populations will become self-sustaining. Unfortunately stocking has also created potential problems for imperiled Sturgeons. Numerous hatchery-reared Shortnose Sturgeons released in the Savannah River, North Carolina and Georgia, have been found in other river systems that could disrupt genetic structure in distinct populations (Smith et al. 2002b). Shortnose Sturgeons within the Edisto River system, South Carolina, have depressed genetic diversity that may be due to former stocking programs not capturing genetic diversity in released progeny (Quattro et al. 2002). The source population is critical even in reintroductions into recovered habitat. Hatchery-reared Lake Sturgeons from Lake Winnebago stock moved rapidly downstream and out of the target area in the Menominee River, but native translocated riverine Lake Sturgeons stayed in the intended area (Thuemler 1988). Overstocking can potentially exceed contemporary carrying capacity. A Pallid Sturgeon augmentation program in the upper Missouri River targeted 1,700 adults and predicted adult number may reach 3,900 by 2038, but historical abundance estimates indicate a maximum of 968 adults (Braaten et al. 2009b). These problems can be minimized

ACIPENSERIDAE: STURGEONS

by developing peer-reviewed stocking plans that address imprinting and genetic diversity issues inherent to any hatchery program (Smith et al. 2002a; Ludwig 2006; George et al. 2009). For species that lack female brood stock for an artificial propagation program, fully viable androgenic nuleocytoplasmic hybrids were produced in Eurasian endangered Sturgeons. This was accomplished by inactivation of the egg nuclei from a donor species, dispermic fertilization using sperm from the imperiled species, and heat shock to the embryo to enhance fusion of male pronuclei to form a diploid individual. Because sturgeon eggs have numerous micropyles, simultaneous penetration of two spermatozoa is possible. The progeny have the nuclear DNA of the paternal species and the mtDNA of the maternal species. If sperm are from two different males the level of heterozygosity is similar to that of typical hybrids. Expression of morphological characters varies with development, with maternal effects expressed at 6 months and the final development of paternal characters at 1 and 3 years of age. Because the mechanism of sex determination in Sturgeons is unclear, it is unknown what sex ratios would be realized. The use of cryopreserved sperm is successful, but with a decreased rate of fertilization and survival. Viable progeny have only been produced between species with the same ploidy level (Grunina & Recoubratsky 2005; Grunina et al. 2006, 2009).

Habitat Restoration The creation of artificial spawning grounds can produce successful hatching provided the structures are placed at the proper depth and current velocity, have adequate surface area, consist of the appropriate substrate size, and have sediment-free interstitial spaces that are maintained. Artificial spawning habitat was successful for Lake Sturgeons in the St. Lawrence River, Michigan, and Des Prairies River, Ontario (Johnson et al. 2006b; Roseman et al. 2011; Dumont et al. 2011). Discrete choice modeling of gravid female Shovelnose Sturgeon habitat selection in the highly altered lower Missouri River indicated that restoration efforts need to concentrate on channel complexity because of the importance of variability in surrounding depths to gravid females (Bonnot et al. 2011).

Fisheries Overfishing contributed greatly to the decline of North American Sturgeons in the late 1800s and during various

199

periods in the 1900s, nearly causing the extinction of some species (Williamson 2003; Auer 2004; Wilson & McKinley 2004; see conservation section). Legal Sturgeon harvest persists today as commercial, sport, and subsistence fisheries for the Atlantic Sturgeon, Lake Sturgeon, White Sturgeon, Green Sturgeon, and Shovelnose Sturgeon (see commercial importance section). For all species of Acipenser, strict regulations are typically in place for sport and subsistence fisheries, including minimum size lengths, restricted open seasons, and quotas. Two Canadian provinces allow sportfishing for the Atlantic Sturgeon, but any harvest is illegal in the United States. Three provinces and three states have a sport harvest for the Lake Sturgeon, which includes spearing in some states. In 2000, >5,000 fish were harvested with hook and line in Ontario and >2,500 taken with spears in Wisconsin. An annual commercial harvest of 80 mt (88.2 tons) of the Lake Sturgeon over the last 10 years in the Quebec portion of the lower St. Lawrence River has been maintained due to restrictive regulations, close monitoring of the commercial catch, and periodic assessment of the population (Mailhot et al. 2011). Washington, Oregon, and California allow sportfishing for White and Green Sturgeons. Sportfishing for the White Sturgeon is popular in western Canada and the northwestern United States (Figs. 5.14 and 5.15). In the Columbia River and its tributaries, harvest peaked at 62,400 Sturgeons in 1987. With regulatory changes, the annual harvest was lowered and has ranged between 33,500 and 45,100 since 1992. Sportfishing for the less desirable Green Sturgeon is minimal with ≤100 harvested annually in the lower Columbia River since

Figure 5.14. Because of their size, White Sturgeons, Acipenser transmontanus, are the focus of a popular, albeit somewhat specialized, sport fishery in portions of their range. This individual was caught, recorded, and released in the upper Fraser River, British Columbia (photograph by and used with permission of Kevin Estrada, Sturgeon Slayers, www .sturgeonslayers.com, “Catch-Record-Release”).

200

FRESHWATER FISHES OF NORTH AMERICA

Figure 5.15. A White Sturgeon, Acipenser transmontanus, taken from the Columbia River, Washington. Note the protrusibility of the mouth (photograph by G. Burton).

1988, and total harvest declined from 6,494 annually in 1985–1989 to only 512 in 2003. Thirteen states permit sportfishing for the Shovelnose Sturgeon, but there is little oversight and restrictions are few and variable. Only one state enforces a season (Arkansas); five other states (Minnesota, Missouri, Montana, Nebraska, Wyoming) have creel or possession limits statewide, and no states maintain data on annual take. Regulations are inadequate (minimum length or harvest slot limit) in the commercial harvest of Shovelnose Sturgeons in all eight states where it is permitted, causing concern for the species due to the recent increase in demand and prices for North American caviar (Morrow et al. 1998a; Mosher 1999; Williamson 2003; Pikitch et al. 2005; Adams et al. 2007; Koch & Quist 2010). For example, declines occur in adult abundance, length, weight, and age of Shovelnose Sturgeons with increased harvest in the upper Mississippi River and with insufficient regulations in many states to maintain sustainable stocks (Colombo et al. 2007a; Koch et al. 2009c). Commercial and sport fisheries targeting other fishes can negatively impact Sturgeons through incidental catch and bycatch. Fisheries for the American Shad (Alosa sapidissima), Goosefish (Lophius americanus), Dogfish Sharks (Squalidae), shrimp, and other species along the Atlantic Coast use gillnets, trawls, and pound nets that also are effective at capturing the endangered Shortnose Sturgeon and Atlantic Sturgeon, especially juveniles and subadults. Mortality estimates from bycatch range from 10 to 22% with another 20% injured (Collins et al. 1996, 2000b; Stein et al. 2004; Spear 2007); this is in addition to natural annual mortality rates of 7–12% for most adult

Acipenser (DeVore et al. 1995; Boreman 1997). Annual bycatch mortality for Atlantic Sturgeons from Maine to North Carolina is estimated as 1,500 individuals/year, and 4.2 mt (4.6 tons) of bycatch of this species were reported in Canadian waters in 1998, although these statistics likely underestimate total losses (Williamson 2003; Stein et al. 2004; Kahnle et al. 2007). Bycatch is also a concern for the Green Sturgeon, including the threatened southern DPS, which is highly migratory along the Pacific Continental Shelf and overwinters in Canadian waters, where it is subject to intensive bottom trawl fisheries (Lindley et al. 2008). Even if mortality or injury is avoided, repeated capture and excessive handling by commercial fishers disrupt spawning migrations in Shortnose Sturgeons (Moser & Ross 1995). High mortality rates (16–17%) for the threatened Gulf Sturgeon in the Suwannee River, Florida, likely reflect death due to bycatch in the Gulf of Mexico (Sulak & Clugston 1999; Pine et al. 2001). In the Mississippi River, the endangered Pallid Sturgeon is similar in appearance to the sympatric and commercially harvested Shovelnose Sturgeon (Murphy et al. 2007a), and there are calls to ban all commercial fisheries for Sturgeons in the basin (Graham & Rasmussen 1999). Pallid Sturgeons have annual mortality rates of 11 and 30% downstream and upstream of the mouth of the Ohio River, respectively. The higher rate is attributed to Pallid Sturgeons being commercially harvested with Shovelnose Sturgeons, whose mortality rate in the upper Mississippi River upstream of the mouth of the Ohio River is 37% (Colombo et al. 2007a; Killgore et al. 2007). Direct evidence of bycatch of Pallid Sturgeons includes 2 of 113 total Scaphirhynchus spp. taken by commercial roe fishers even with observers on board, and an additional Pallid Sturgeon found dead in a retrieved ghost net (Bettoli et al. 2009a). Because of the commercial value of Sturgeon products, especially caviar, illegal harvest of Sturgeons has and continues to be a threat to all species (Ludwig 2006). Sturgeons pursued for caviar are easy targets as they migrate in river systems and concentrate on spawning grounds, and protection can be difficult and costly because of the large distances Sturgeons travel (Auer 2004). North American species most affected by illegal harvest include Atlantic Sturgeon, Lake Sturgeon, White Sturgeon, and Shovelnose Sturgeon, and two endangered species, Shortnose Sturgeon and Pallid Sturgeon (Williamson 2003; Colombo et al. 2007a; Waldman et al. 2008b). Illegal take can be large. For example, a single poaching ring killed 2,000 adult White Sturgeons from the Columbia River to obtain 1,352 kg (2,981 pounds) of caviar over a 5-year period

ACIPENSERIDAE: STURGEONS

(Cohen 1997). Recent advances in genetics allow differentiation of Sturgeon species from their caviar and flesh, which have given law enforcement agencies genetic forensic tools to identify protected species in the trade of Sturgeon commercial products (DeSalle & Birstein 1996; Birstein et al. 1999; Wolf et al. 1999; Congiu et al. 2002). But some species complexes are not readily identifiable with these forensic tools and data are lacking on identifying population-level variation, which is complicated in many instances by stocking programs that mix genotypes. Alternative molecular techniques such as single nucleotide polymorphism (SNP) markers may address some of these issues (Ludwig 2008; Waldman et al. 2008b). Aquaculture and wild-origin caviar can potentially be distinguished by examining fatty acid composition, but this would require the use of specific additives to formulated diets by aquaculturists (Gessner et al. 2008).

Dams Unaltered free-flowing large river habitat was all but eliminated in the Northern Hemisphere in the 1900s, transforming rivers into a discontinuous series of large pools and deep runs (Sulak & Randall 2002). For Sturgeons, dams can negatively affect spawning migrations, quality and quantity of spawning habitat, embryo and larval development, genetic diversity, and growth. All Sturgeons have upstream migrations associated with spawning, and dams obstruct these movements. Semi-anadromous and anadromous species typically have longer migrations and fewer spawning sites than potamodromous species and are therefore more susceptible (Cooke et al. 2002; Jager et al. 2007; Mora et al. 2009). Barriers to spawning areas can cause females to resorb eggs or not spawn. (Auer 1996). Disruption of natural flows downstream of dams can alter migratory cues for potential spawners (Mayden & Kuhajda 1997a). The presence of navigation locks may allow for limited upstream movement in Lake Sturgeons (Knights et al. 2002), but Shortnose Sturgeons do not use these as upstream passages (Cooke et al. 2002). Fish ladders can pass Sturgeons, even large White Sturgeons, but successful passage is limited and influenced by ladder construction (Parsley et al. 2007). A current velocity of 0.33m/s (1.1 feet/s) is sufficient to guide White Sturgeons to a horizontal ramp structure for dam bypass in a laboratory setting. An experimental flume at a 4% bed slope with baffles and average water velocities of 1.7–2.1 m/s (5.6–6.9 feet/s) representing the midsection of a fishway was capable of passing

201

adult White Sturgeons, but field assessment of such structures, stress responses of Sturgeons, and their spawning success after ascent need to be examined (Cheong et al. 2006; Cocherell et al. 2011). Even if fishways are available for Sturgeons to bypass dams, upstream spawning sites may be inundated (Knights et al. 2002). If Sturgeons spawn above hydroelectric dams, post-spawning adults and any juveniles from successful spawns are at risk of entrainment and death on turbines during downstream migrations (Kynard & Horgan 2001; Jager et al. 2007). Artificial lowering of spring and summer discharges from hydropower operations and for flood control reduces the availability and quality of spawning habitat for many Sturgeon species. For example, these actions buried cobble and gravel spawning sites for the Kootenai River White Sturgeon under fine sediments due to reduced velocities and shear stress. These effects are compounded by changes to channel morphology below dams (Rochard et al. 1990; Parsley & Beckman 1994; Paragamian et al. 2009). Controlled dam releases also can produce artificially high flows, which create high bottom velocities that preclude or reduce spawning (Buckley & Kynard 1985). Even if spawning is successful below dams, eggs may be laid in masses that have reduced survival from poor hatch success and increased predation and disease (Kempinger 1988; Auer 1996). Reduced flows can also lead to warmer water temperatures that negatively affect embryo development and hatching success (Van Eenennaam et al. 2005). Fluctuating water levels due to dam releases can dislodge embryos during high flows and expose and desiccate embryos during low flows (Kempinger 1988). Dams trap sediments and decrease turbidity, which may cause Sturgeons to use deeper and more restricted habitat for spawning and development of early life stages (Perrin et al. 2003) and likely increases predation on eggs and larvae (Gadomski & Parsley 2005a). Larval Sturgeons can drift great distances below spawning sites as they grow and develop (e.g., Pallid Sturgeon drift for ≤530 km, 329 miles). Dams produce reservoirs with greatly reduced flow that may not provide habitat necessary for larval Sturgeons to survive, and multiple dams on a river often restrict riverine reaches to lengths too short for larval development (Auer & Baker 2002; Smith & King 2005a; Kynard et al. 2007; Braaten et al. 2008). Dams have dramatic negative effects on Scaphirhynchus with no documented recruitment for Pallid Sturgeons in the highly fragmented upper Missouri River basin for the past 35 years or for Alabama Sturgeons in the highly impounded Alabama River for decades (Mayden &

202 FRESHWATER FISHES OF NORTH AMERICA

Kuhajda 1997a; USFWS 2000; Webb et al. 2005). For some river systems maintaining downstream reservoirs below maximum pool can increase the amount of riverine habitat (Gerrity et al. 2008). Dams and altered habitat downstream of dams and within impoundments also affect genetics and growth of Sturgeons. Metapopulation modeling shows that multiple dams also restrict gene flow and hence genetic diversity within artificially segmented populations with upstream populations at risk of extirpation due to higher downstream and lower upstream migrations (Jager et al. 2001, 2007). Channelization and dredging of waterways in the riverine reaches above and below dams to allow for navigation also can destroy both river and estuary Sturgeon habitat (Burke & Ramsey 1995; Kynard 1997). Altered habitat below dams and within impounded areas can lead to slower growth rates for Sturgeons due to decreased productivity (McKinley et al. 1993; Beamesderfer et al. 1995; Everett et al. 2003).

Agriculture Construction of levees to protect farmlands from winter and spring high water and the filling of isolated oxbows and backwaters to expand floodplain agriculture has led to the virtual elimination of riparian and backwater habitat in large-river ecosystems. Many rivers now are confined to narrow channels and cut off from the major source of organic nutrients. Sturgeons require this floodplain habitat for many important aspects of their life history, including feeding areas for all life stages and habitat for larvae and juveniles (USFWS 1993; Coutant 2004); year-class strength in Sturgeons is primarily dependent on survival in the first 2–3 months (Secor et al. 2002; Parsley et al. 2002). In addition, runoff from agriculture introduces numerous pollutants, including herbicides, pesticides, and fine sediments. For example, DDT occurred in tissues of Pallid Sturgeons 14 years after its ban in 1975 (Ruelle & Keenlyne 1993), and high silt loads on spawning substrates can lead to larval mortality in the Lake Sturgeon (Nichols et al. 2003).

Pollution Sturgeons are affected by numerous forms of water pollution. They are susceptible to the accumulation of contaminants in their flesh and eggs because they are long-lived; ingest benthic organisms and organic material from the bottom of rivers, lakes, estuaries, and near-shore marine

habitats; and have high lipid content (Auer 2004). Bioaccumulation of contaminants is exacerbated by the low capacity of Sturgeons to detoxify various compounds relative to other bony fishes (Singer & Ballantyne 2004). Instream cage studies on juvenile Shortnose Sturgeons to assess suitability of water quality in the Roanoke River led to a 91% mortality compared to only 0.6% mortality for the Fathead Minnow (Pimephales promelas), indicating that this standard toxicity test organism is an inappropriate surrogate for Shortnose Sturgeons in water quality or toxicity tests (Cope et al. 2011). Contaminants are introduced into aquatic systems from numerous sources such as mining, industry, urban discharge, and agricultural and urban runoff (Ruelle & Keenlyne 1993; Hinton 1998). Given the imperiled status of Sturgeons, surprisingly few studies have examined the effects of pollutants, but the fact that they are imperiled restricts the collection of tissues for such studies (Auer 2004; Singer & Ballantyne 2004). Limited studies do show that Sturgeons accumulate heavy metals and organic chemicals (e.g., PCBs, DDT, and chlordane) that may lead to developmental and behavioral abnormalities and interfere with reproduction; concentrations within Sturgeons increase with age (Ruelle & Keenlyne 1993; MacDonald et al. 1997; Doyon et al. 1999; Alam et al. 2000; Auer 2004). Heavy metals and detergents impair olfactory sensory abilities (Kasumyan 2002), and sediment pollutants may place stress on the overall viable reproduction of Sturgeons (Kruse & Scarnecchia 2002). Elevated concentrations of organochlorines in intersexual and male Shovelnose Sturgeons with limited reproductivity (low gonad somatic index, GSI) were restricted to the brain-hypothalamus-pituitary complex. This indicates these pollutants alter hormone production and reception necessary for proper gonadal development (Koch et al. 2006). Intersexuality rates in Shovelnose Sturgeons are 2 to 7.5% based on examination of gonads and the use of sexually dimorphic gene expression as a biomarker (Colombo et al. 2007c; Divers et al. 2009; Amberg et al. 2010). Sturgeons are less tolerant of hypoxia than other fishes, and heavy nutrient loading can lead to reduced oxygen levels, resulting in impaired respiratory metabolism, reduction of foraging activity and growth, or death (Secor & Gunderson 1998; Campbell & Goodman 2004; Cech & Doroshov 2004; Niklitschek & Secor 2005). The improper disposal of trash leads to deformities in the Pallid Sturgeon and Shovelnose Sturgeon because rubber bands, gaskets, and other ring-shaped objects are swam into and ultimately cause notching of the rostrum, loss of

ACIPENSERIDAE: STURGEONS

barbels, broken scutes, and deformed pectoral fins (Murphy et al. 2007b).

Industrial Use of Waterways All life stages of Sturgeons are impacted by commercial vessel passage. Shovelnose Sturgeon larvae are highly vulnerable to stranding from commercial vessel drawdown and subsequent dewatering of littoral zones when simulated in a laboratory setting. Larvae are positively rheotactic (swim toward current) and are more likely to swim toward the shoreline as water recedes (Adams et al. 1999b). Larval Sturgeons are also susceptible to propeller-induced shear stress because internal organs are underdeveloped and integument is fragile. In laboratory experiments replicating shear stress from vessels, small Lake Sturgeon larvae (mean 11 mm TL, 0.43 inch) had higher mortality (≤86%) relative to larger larvae (≤58%) (mean 14 mm TL, 0.55 inch) (Killgore et al. 2001). Juvenile and adult Shovelnose Sturgeons can be entrained by towboat propellers and have a high probability of being struck by a propeller due to their large size. An estimated 0.5 Shovelnose Sturgeons / km (0.8/mile) of towboat travel are entrained and killed in the Mississippi River navigation pool. The threat of entrainment is increased in narrow, shallow, and sluggish reaches of rivers, and higher propeller speeds increase the risk of a propeller strike (Gutreuter et al. 2003; Killgore et al. 2011). Unlike many large-river fishes, Shovelnose Sturgeons are concentrated within the navigation channel on the Mississippi River regardless of flow, temperature, or towboat traffic disturbance, putting them at high risk for entrainment mortality (Gutreuter et al. 2006, 2010). Juvenile Sturgeons are also susceptible to entrainment during channel dredging and water diversions. Laboratory experiments show that dredge entrainment of juvenile White Sturgeons (80–100 mm TL, 3.2–3.9 inches), Pallid Sturgeons (122–168 mm FL, 4.8–6.6 inches), and Lake Sturgeons (120–173 mm FL, 4.7–6.8 inches) is likely within a radius of 1.25 m (4.1 feet) of the cutterhead of hydraulic dredges used to maintain navigation channels. Escape speeds differed between populations and sizes. Pallid Sturgeons from the Yellowstone River, North Dakota (versus Atchafalaya River, Louisiana), Lake Sturgeons from Lake Winnebago, Wisconsin (versus Wisconsin River), and smaller Sturgeons in all species demonstrated weaker swimming and thus were more susceptible to entrainment (Boysen & Hoover 2009; Hoover et al. 2011a). Water diversions in the lower Mississippi River to control flood-

203

ing and restore wetlands entrain Pallid Sturgeons and Shovelnose Sturgeons into non-riverine habitat (Hoover et al. 2011b). Monitored diversions exporting water from the Sacramento–San Joaquin River, California, took on average an estimated 1,621 juvenile Green Sturgeons / year before 1986, which has decreased to an average of 79 juveniles/year from 1986 forward due to decreased abundance (Adams et al. 2007). Disposal of dredged sediments during navigation channel maintenance can have negative impacts on Sturgeon habitat. A significant reduction in Atlantic Sturgeon relative abundance occurred in and downstream of an area where dredged sediment was disposed in the St. Lawrence River estuary, sand dunes created by dredge disposal had lower densities and biomass of prey items for juvenile and subadult Atlantic Sturgeons relative to unaffected areas, and modeling for sediment transport over 10 years predicted sediments will impact the core area used by early juvenile Atlantic Sturgeons (Hatin et al. 2007a; Nellis et al. 2007ab).

Invasive Species Invasive species negatively affect early life stages of Sturgeons. The introduced Common Carp (Cyprinus carpio) feeds on the eggs of the White Sturgeon in the Columbia River (Miller & Beckman 1996), and the exotic Round Goby (Neogobius melanostomus) preys on eggs and perhaps larvae of the Lake Sturgeon in tributaries of the Great Lakes (Nichols et al. 2003). Invasive Zebra Mussels (Dreissena polymorpha) can densely colonize and essentially coat the sandy and silty habitats used for foraging by juvenile Lake Sturgeons. In habitat-choice experiments, Lake Sturgeon juveniles avoided areas with Zebra Mussels and showed reduced foraging, which may be detrimental to growth (McCabe et al. 2006). The invasive Overbite Clam (Potamocorbula amurensis), a known bioaccumulator of selenium, has replaced native food items of Green Sturgeons in the Sacramento–San Joaquin River, California (Adams et al. 2007). Programs to control the introduced Sea Lamprey (Petromyzon marinus) in the Great Lakes indirectly affect Lake Sturgeons by restricting upstream movement to spawning grounds with low-head barriers and killing of Sturgeon early life stages with lampricides (Auer 2004; Léonard et al. 2004).

Global Climate Change Sturgeon populations will be negatively affected by the predicted impacts of global climate change on river

204 FRESHWATER FISHES OF NORTH AMERICA

ecosystems worldwide. Reduced precipitation and flow, increased temperature and frequency of droughts, and more extreme floods will likely exacerbate many of the existing threats to Sturgeons and their habitats. Given that many species and populations of Sturgeons are already near the upper limit of temperature for proper development and growth, they may face extinction and extirpation as global temperatures increase (Brander 2007; Ficke et al. 2007; Kingsford 2011).

COMMERCIAL IMPORTANCE Before Europeans settled in North America, Native Americans used the flesh, roe, and oil of Sturgeons as a food source. Sturgeon skins were used to store the preserved flesh, and glue was made from a gelatin prepared from the swim bladder (isinglass) that later was used by European settlers as a clarifying agent for beer and wine. Sturgeons were a major fisheries resource for Native Americans living along the Atlantic Coast, in the Great Lakes, and in the Pacific Northwest (Ferguson & Duckworth 1997; Holzkamm & Waisberg 2004). The caviar trade has a long history in North America. Although prized by royalty in Europe and Asia, where caviar was considered a delicacy, European settlers initially found Sturgeons distasteful and considered them a pest that could destroy fishing nets and create a hazard for small boats in some New England rivers (Bogue 2000; Saffron 2002; Spear 2007). For example, in the early days of the Lake Sturgeon fishery (ca. 1850s) in Lake Erie only a portion of the fish caught were used. At the time the huge and then seemingly worthless, net-wrecking Lake Sturgeons were mostly thrown on the beach to rot, fed to hogs, or placed in large piles and burned (Trautman 1981; Becker 1983). The species was even stacked like cordwood on the dock along the Detroit River, Amherstburg, Ontario, and used to fire the boilers of steamboats traveling along the river (Scott & Crossman 1973). The flesh eventually was used to feed servants and slaves. The first caviar business in North America in the 1840s shipped all of its products overseas. In the 1860s–1870s local markets finally developed along the Great Lakes and Atlantic Coast, and a caviar rush ensued. But overfishing and pollution led to a collapse of the industry in 1900, and it completely ceased in 1925. Maximum catch of Lake Sturgeons reached 2,770 mt (3,053 tons) in 1885. That year, Lake Erie alone yielded 2,270 mt (2,502 tons), but the catch dropped to only 91 mt (100 tons) in 1895. Like-

wise catch of Atlantic Sturgeons peaked in the 1880s and early 1890s with ≥3,348 mt/year (3,691 tons) (Ferguson & Duckworth 1997; Saffron 2002; Williamson 2003). During the late 1800s a large fishery also was underway on the Pacific Coast for White Sturgeons with peak catches of 2,500 mt (2,756 tons) in the Columbia River in 1892, but unregulated overfishing led to a collapse of the fishery by the early 1900s (Williamson 2003; Van Eenennaam et al. 2004). Even the smaller so-called Shovelnose Sturgeons (actually Alabama Sturgeons) were used near the end of the fisheries with 19 mt (21 tons) harvested in the Mobile Basin in Alabama alone in 1894 (Smith 1898). Although once supplying >75% of the world’s caviar, after the North American fisheries collapsed, the caviar trade shifted to the Caspian Sea region (Van Eenennaam et al. 2004). But with the fall of the Soviet Union in 1991, unrestricted harvesting of Caspian region Sturgeons drove caviar prices down and increased the demand for this delicacy. Increased demand allowed the North American market to expand with the United States becoming the largest importer of caviar in 1999 (Raymakers & Hoover 2002; Saffron 2002). Recently production of caviar has decreased worldwide from 326 to 184 mt (359 to 203 tons) from 1995 to 2000 with prices increasing from $184 to $332/kg ($83 to $151/pound) (Raymakers & Hoover 2002). If Asian caviar exports continue to decrease from unrestricted exploitation, the demand for North American Sturgeon and Paddlefish roe likely will increase both domestically and internationally (Graham & Rasmussen 1999; Saffron 2002; Léonard et al. 2004). In addition to caviar demand, the popularity of Sturgeons for meat and as an ornamental fish is rising internationally. Worldwide exports increased from 44.7 to 178 mt (49.3 to 196.2 tons) of flesh from 1998 to 1999 with Lake Sturgeon meat exported from Canada accounting for ≥40% of the total. Demand for live specimens, including fertilized eggs and fry for aquaculture and fingerlings or juveniles for the ornamental fish industry, is also on the rise, increasing from 0.5 to 7 million individuals over the same time frame (Raymakers & Hoover 2002). White Sturgeons were first exported to Europe in 1981 to a fish farm in Italy and in 1996 made up 51% of all Sturgeon aquaculture production in western and central Europe with all White Sturgeons (450 mt, 496 tons) produced in Italy (Bronzi et al. 1999). Commercial harvest in North America is ongoing in four species of Acipenser and one species of Scaphirhynchus (for sport harvest, see conservation section). Canada allows harvest of Atlantic and Lake Sturgeons in only a few

ACIPENSERIDAE: STURGEONS

provinces; White Sturgeons are exploited in Alaska, Washington, and Oregon; and Green Sturgeons are harvested in Alaska and allowed as bycatch in Oregon and Washington. Species of Acipenser are highly regulated to avoid overharvesting. Modern-day commercial harvest of Atlantic Sturgeons peaked in 1988 at 44 mt (49 tons) and has declined in recent years, the decline attributable to a decrease in licensed fishers. In contrast, the harvest of Lake Sturgeons remained consistent, averaging 227 mt/ year (250 tons/year) through the 1990s (Williamson 2003). On the Pacific Coast, commercial catch of White Sturgeons remained fairly stable from 1990 to 1995, ranging from 89.2 to 155.4 mt/year (98.3 to 171.3 tons/year). Although a coastal marine fishery exists, most harvest is

A

B

205

from the Columbia River with an average of 10,500 specimens/year taken from 1998 to 2002 (Todd 1999; Williamson 2003). Commercial harvest of Green Sturgeons was 81.2 mt (89.5 tons) in 1991 but dropped to 10.2 and 13.8 mt (11.2 and 15.2 tons) in 1994 and 1995, respectively (Todd 1999). Many consider the flesh and roe of this species of lesser quality than that of the White Sturgeon, and this reduction in harvest may reflect a reduction in effort (Williamson 2003). Shovelnose Sturgeons are exploited commercially in eight states within the Mississippi River basin. Commercial harvest of Shovelnose Sturgeons is poorly regulated with only half of the states allowing commercial harvest reporting harvest estimates (Todd 1999). The lack of

Figure 5.16. (A) Ventral view of the head of a White Sturgeons, Acipenser transmontanus, cruising with other White Sturgeon in a hatchery pond at the Yakama Nation Fish Hatchery, Benton County, Washington (photograph by and used with permission of Dave Herasimtschuk of Freshwaters Illustrated). (B) Hatcheryreared juvenile Lake Sturgeon, Acipenser fulvescens, from Wolf River, Wisconsin (photograph by B. M. Burr).

206

FRESHWATER FISHES OF NORTH AMERICA

regulation stems from the perception that Shovelnose Sturgeons are common and that this species is not desirable because individuals yield low amounts of roe relative to other Sturgeons. Nonetheless harvest has increased dramatically through the 2000s. Harvest increased 51 and 87% in 2000 and 2001, respectively, in the Mississippi River in Missouri relative to the mean harvest from 1988 to 1998, and 86.5 and 5 mt (95.3 and 5.5 tons) of flesh and caviar, respectively, were taken in 2001 overall. In Iowa, harvest of Shovelnose Sturgeons more than doubled from 1997 to 2003, and the 2005 harvest was 1.6 mt (1.8 tons) valued at $158,000. Retail prices for Shovelnose Sturgeon roe climbed to $381–$1,340/kg ($173–$608/pound) in 2004 and fishing pressure is expected to increase. As such, concern is mounting for this species and there is a pressing need for regulations (Morrow et al. 1998a; Quist et al. 2002; Williamson 2003; Pikitch et al. 2005; Kennedy & Sutton 2007; Koch et al. 2009c). Between 2002–3 and 2005–6 the exploited Shovelnose Sturgeon population in the middle Mississippi River has experienced a change in sex ratio (from 1M:1F to 1.14M:1.00F), an increase in annual mortality rate (from 37% to 44%), and a decrease in recruitment through time (29%/year), a trend that will lead to a decline in population density by an order of magnitude in one decade (Tripp et al. 2009a). With the collapse of the North American caviar industry in the late 1800s, artificial reproduction of Sturgeons was attempted with limited success in the early 1900s. Advances made in Soviet Union hatcheries over the next several decades provided the necessary information for Sturgeon facilities to begin to appear in North America in the 1960s. Aquaculture of White Sturgeons became fully established in the mid-1990s (Fig. 5.16a). Today markets

exist for yolk-sac larvae for other aquaculture facilities, 2-year-old juveniles as live fish, 3- to 4-year-old subadults as food fish, and 7- to 8-year-old females for caviar and meat. In 2003, California facilities produced >6 mt (6.6 tons) of caviar with wholesale prices at $300/kg ($136/pound) and retail prices ≤$1,000/kg ($454/pound) (Van Eenennaam et al. 2004), and recent production from just 2 California farms has increased to 15 mt (16.5 tons) annually (Zhang et al. 2011). The Atlantic Sturgeon, Shortnose Sturgeon, and Lake Sturgeon are also being raised at commercial aquaculture facilities (Fig. 5.16b) (Waldman et al. 2008b).

LITERATURE GUIDE Several sources provide excellent overviews on the Sturgeons in North America. These include edited volumes and books that cover many aspects of Sturgeon biology, management, propagation, and conservation (Willot 1991; Birstein et al. 1997c; Hochleithner & Gessner 1999; Rosenthal et al. 1999, 2002b; Van Winkle et al. 2002; LeBreton et al. 2004; Munro et al. 2007) and those that focus on harvest and conservation (Williamson et al. 1999; Williamson 2003).

Acknowledgments The Department of Biological Sciences at the University of Alabama provided support for this project. Thanks to all of those who provided figures or gave permission for figure use and to the editors for their comments that greatly improved the original manuscript.

Chapter 6

Polyodontidae: Paddlefishes Bernard R. Kuhajda

The family Polyodontidae, the Paddlefishes, has only two living species, the Chinese Paddlefish, Psephurus gladius, and the North American Paddlefish, Polyodon spathula (Fig. 6.1), although numerous fossil Paddlefishes date back to >100 mya. The Paddlefish was originally described as a Shark (Chondrichthyes, Cartilaginous Fishes) in the late 1700s due to its cartilaginous skeleton, jaw structure, and shark-like tail, but Paddlefishes are actually ancient bony fishes. Both species get large, with the North American Paddlefish reaching 2.15 m (7.1 feet) and 74 kg (163 pounds). Paddlefishes get their common names and the specific epithet spathula (spatula) of the North American species from the long spatula- or paddle-shaped snout that overhangs extremely small eyes and a large mouth. The Paddlefish also goes by the name spoonbill catfish in reference to the paddle and the lack of obvious scales on the body, but Paddlefishes are not closely related to Catfishes (Siluriformes). The paddle is absent in small larvae but is one-half of the body length in juveniles and one-fourth to one-third of body length in adults. The North American Paddlefish is a riverine species that feeds on zooplankton using tens of thousands of electrosensory organs covering its paddle to detect

weak electric fields emitted by its prey. Zooplankton is captured by filter feeding in adults; Paddlefish pass large volumes of water into their huge mouths and filter it across extremely numerous long gill rakers, capturing the zooplankton. The name of the family and genus is derived from “poly-” (many) and “-don” (tooth), referring to the many gill rakers. The spawning habitat for the Paddlefish was unknown for >100 years until the spring of 1960 when spawning was observed on a flooded gravel bar after a rapid river rise of 2.7 m (9 feet). Upstream spawning migrations can cover >322 km (200 miles). Paddlefish populations are far below historical levels due to commercial harvest and large-river alterations (dams and channelization) that block migratory routes and destroy habitat. Paddlefish eggs are processed into caviar and are used as a substitute for Sturgeon (Acipenseridae) caviar.

DIVERSITY AND DISTRIBUTION The Polyodontidae are restricted to the Northern Hemisphere and consist of only two extant species. The Chinese

Plate 6.1. Paddlefish, Polyodon spathula

208

FRESHWATER FISHES OF NORTH AMERICA

Figure 6.1. The unique Paddlefish, Polyodon spathula, the largest planktivorous fish in North American fresh waters, cruises in the Osage River, Missouri, one of few known spawning streams for the species (photograph by and used with permission of W. N. Roston).

Figure 6.2. Geographic range of the Paddlefish, Polyodon spathula, the only extant representative of the Polyodontidae in North America. Genus Polyodon

Paddlefish (Psephurus gladius) occurs in the Yangtze River drainage, China, with adults occasionally migrating into the East China and Yellow Seas (Liu & Zeng 1988). The Paddlefish (Polyodon spathula) is a freshwater species that historically occurred in the Mississippi River basin of North America from New York to Montana and south to Louisiana and in adjacent Gulf Coast drainages from the Mobile Basin, Alabama, to Galveston Bay, Texas, as well as the Great Lakes in the United States and Canada (Table 6.1; Fig. 6.2). Although still present in most of its historical distribution in 22 states, the Paddlefish has disappeared from part of its peripheral range within the Mississippi River basin, from several western Gulf Coast drainages, and throughout the Great Lakes (Hubbs & Lagler 1964; Burr 1980; Gengerke 1986; Parker 1988; Page & Burr 1991; K. Graham 1997; Reid et al. 2007).

Inter- and Intraspecific Variation Other than a footnote alluding to an undescribed species of Polyodontidae in the Mississippi River basin (Myers 1949), no other species of Paddlefish or any subspecies of Polyodon spathula are known from North America. Lower Mississippi River Paddlefish reportedly attain a greater size than those from the Ohio and upper Mississippi Rivers (Stockard 1907), but the largest and heaviest specimens recorded are from the upper Mississippi and Ohio Rivers (Nichols 1916; Forbes & Richardson 1920). Variation occurs in riverine versus oxbow lake adults from the lower Mississippi River with riverine Paddlefish possessing more slender bodies and shorter and broader paddles (Fig. 6.3) (Stockard 1907). Morphological variation in juvenile Paddlefish (61.9–403.7 mm TL, 2.4–15.9 inches) exists between populations in the southeastern United States. Hatchery-reared Paddlefish from the Mermentau River, Louisiana, possess shorter, narrower leaf-shaped paddles and asymmetrical caudal-fin lobes; hatchery-reared specimens from the Tombigbee River, Alabama, have longer, broader spoon-shaped paddles and more symmetrical caudal lobes; and field-collected juveniles from the Mississippi River, Mississippi, have the longest and broadest (paddleshaped) paddles and symmetrical caudal lobes. Overall larger basins with higher gradients and discharges have juvenile Paddlefish with longer and broader paddles and

POLYODONTIDAE: PADDLEFISHES

209

Table 6.1. Life history traits of the Paddlefish, Polyodon spathula (Polyodontidae)(—, not applicable). Trait

Description

References

Number of extant North American species Strictly freshwater

One

Burr 1980

Yes, few records from estuaries

Maximum size recorded in length and weight Maximum age Age at first reproduction

2.15 m (7.1 feet); 74 kg (163 pounds)

Iteroparous or semelparous Fecundity estimates (ovarian counts) Mature egg diameter

Iteroparous 82,397–>1,000,000

Gunter 1942; Vladykov & Greeley 1963; Graham 1997; Paukert & Fisher 2000; Wojtenek et al. 2001 Nichols 1916; Forbes & Richardson 1920; Hochleithner & Gessner 1999 Scarnecchia et al. 2007 Lein & DeVries 1998; Scarnecchia et al. 1996b, 2007, 2011 Russell 1986 Needham 1965; Robinson 1966; Russell 1986

Egg deposition sites

56 years Males, 5–12 years; females, 6–19 years

2.0–3.2 mm (0.08–0.13 inch)

Clutch size

Over gravel, gravel and sand, and bedrock substrates No nests

Range of spawning dates and temperatures Habitat of spawning sites; average water depth Incubation period; larval type at hatching Mean size at hatching

Early March to mid-June; peak spawning occurs from 12 to 20°C (53.6 to 68°F) In 1.2–7.7 m of water (3.9–25.3 feet) in velocities ranging from 0.39 to 1.06 m/s (1.3–3.5 feet/s) 6.5 to 10 days, 14.0–18.8°C (57.2–65.8°F); yolk-sac larvae 8–9 mm (0.32–0.35 inch) TL

Parental care Major dietary items

None Crustacean zooplankton, consisting mainly of copepods and cladocerans Riverine and reservoir habitats that typically have low current velocities and high concentrations of zooplankton Potamodromous, migrating to spawn or forage wholly within freshwater river systems Paddlefish are considered Vulnerable or of Special Concern in the southeastern United States and throughout their range and are protected by 11 states

General year-round habitat

Migratory or diadromous Imperilment status

symmetrical caudal-fin lobes (Hoover et al. 2009a). These geographic differences are not documented for adults, but high variation of adult paddle morphology noted in large samples may obscure the existence of any distinct paddle morphotypes (Hoover et al. 2000). Similar-sized Paddlefish also have a high degree of variation in osteological features (Grande & Bemis 1991). Genetic data indicate distinctiveness of the population in Mobile Basin and perhaps other populations (see genetics section).

Larimore 1950; Reed et al. 1992; Scholten & Bettoli 2005 O’Keefe et al. 2007 Larimore 1950; Reed et al. 1992; Scholten & Bettoli 2005 Wallus 1986a; Hoxmeier & DeVries 1997; Firehammer et al. 2006 O’Keefe et al. 2007 Ballard & Needham 1964; Yeager & Wallus 1982; Bemis & Grande 1992 Ballard & Needham 1964; Yeager & Wallus 1982; Bemis & Grande 1992 — Eddy & Simer 1929; Rosen & Hales 1981; Hageman et al. 1986 Rosen et al. 1982; Southall & Hubert 1984; Moen et al. 1992; Zigler et al. 2003 Bemis & Kynard 1997 Williams et al. 1989; Graham 1997; Warren et al. 2000; IUCN 2011; Jelks et al. 2008

Paddlefish as Non-Natives Paddlefish were first introduced outside of the United States in 1974 into Russia and are now part of polyculture systems for meat and caviar in several European, Middle Eastern, and Asian countries (Hoover 1999; Billard & Lecointre 2001; Vedrasco et al. 2001; Lobchenko et al. 2002; Raymakers 2002; Hubenova et al. 2007; Lenhardt et al. 2011). From 1993 to 1997, >1 million live Paddlefish and

210

FRESHWATER FISHES OF NORTH AMERICA

strated infertility between the Paddlefish and Shovelnose Sturgeon, Scaphirhynchus platorynchus, of North America (Mims et al. 1997, 2009; Mims & Shelton 1998). In North America before 2001 hatcheries shared eggs and fry for stocking across the range of the Paddlefish, and Missouri River progeny were stocked in the upper Ohio and Allegheny Rivers through 2001 (Argent et al. 2009; Grady & Elkington 2009). Currently efforts are underway to ensure the genetic integrity of Paddlefish stocks are maintained by stocking with relatively local fish (Carlson et al. 1982; Epifanio et al. 1996; Heist & Mustapha 2008; Grady & Elkington 2009; Sloss et al. 2009), but some proponents of propagation and culture of Paddlefish consider genetic diversity among populations small and concern over gene-pool contamination as essentially a nonissue relative to stocking (Mims et al. 2009).

PHYLOGE NE TIC RELATIONSHIPS

Higher Relationships Figure 6.3. Large adult Paddlefish, Polyodon spathula, from the lower Mississippi River drainage demonstrating morphological variation between riverine versus oxbow lake populations, with riverine Paddlefish (middle specimen) possessing more slender bodies and shorter and broader paddles (from Stockard 1907).

fertilized eggs were exported from the United States with China importing 37% of these and establishing a culture program, raising concerns that specimens escaping into the wild could hybridize with the Chinese Paddlefish, Psephurus gladius (Raymakers 2002; Mims et al. 2009). Adult and juvenile Paddlefish have been captured in the lower Danube River in Serbia and Bulgaria, indicating that a population has become established, likely originating from aquaculture escapees in Bulgaria or Romania during floods (Simonović et al. 2006; Lenhardt et al. 2011). Paddlefish have also been captured in the Danube River in Austria and Slovakia (Holčik 2006). Some researchers believe a strong potential for hybridization exists between Sturgeons of the Danube River (two Acipenser and one Huso species) and introduced Paddlefish since they likely share spawning sites (Simonović et al. 2006), but others believe the threat of hybridization between these acipenserids is limited or nonexistent due to lack of appropriate habitat for the establishment of a Paddlefish population (Holčik 2006) and the demon-

Early researchers considered the Paddlefish (and Sturgeons) to be freshwater sharks based on their cartilaginous skeletons and jaw structures. The Paddlefish was originally described as Squalus spathula, the genus referring to Dogfish Sharks (Squalidae) (Walbaum 1792). Another researcher later described a new genus of shark, Proceros, supposedly based on a Paddlefish (Rafinesque 1820a), but the information used may have originated as a hoax (Markle 1997). The genus Polyodon was created when yet another Paddlefish specimen was described as P. feuille (Latinized to folium) (Lacepède 1797; McKinley 1984); the original specific name was later placed in this genus. The relationship to sharks was rejected in the mid1800s and Paddlefishes (and Sturgeons) were considered bony fishes (Osteichthyes) and were placed in the basal grade Chondrostei (Grande & Bemis 1991, 1996; Bemis et al. 1997). Currently Paddlefishes (Polyodontidae) and Sturgeons (Acipenseridae) are placed in the suborder Acipenseroidei, and along with the fossil families †Chondrosteidae and †Peipiaosteidae, are placed in the order Acipenseriformes (Grande & Bemis 1991, 1996; Grande et al. 2002; Grande & Hilton 2006; Krieger et al. 2008; Hilton & Forey 2009; Hilton et al. 2011). Acipenseriformes, along with Bichirs (Polypteriformes) and other fossil orders, were assigned to the subclass Chondrostei, which was basal to all other Neopterygii (Nelson 1994). But several other studies revealed that Bichirs are

POLYODONTIDAE: PADDLEFISHES

the extant basal actinopterygian (subclass Cladistia), and Acipenseriformes together with other fossil orders, form the subclass Chondrostei, which is sister to Neopterygii (Patterson 1982; Bemis et al. 1997; Nelson 2006).

Relationships within Polyodontidae The first phylogenetic hypothesis for relationships between living and fossil polyodontids did not consider the fossil †Paleopsephurus wilsoni a member of this family but placed it as a close relative to Sturgeons (Gardiner 1984). This relationship was based largely on the erroneous reported absence of stellate bones in the paddle in the original description of the species (MacAlpin 1947) and on considering characters of Polyodon as primitive within the family, which was also incorrect (Grande & Bemis 1991). Current hypotheses regard the Asian fossil †Protopsephurus liui as the basal polyodontid, followed by the North American fossil †Paleopsephurus wilsoni. The living Chinese Paddlefish (Psephurus gladius) is basal to a monophyletic group of North American taxa, including the fossil †Crossopholis magnicaudatus and the genus Polyodon that includes the fossil †Polyodon tuberculata and the living Paddlefish (Fig. 6.4) (Grande & Bemis 1996; Grande et al. 2002).

Evolutionary Considerations Fossil and recent Paddlefishes demonstrate a trans-Pacific pattern of relations dating to at least the Late Cretaceous (99.6–66.5 mya) based on minimum ages of fossils (Grande & Bemis 1991). This is corroborated with estimates of the divergence time of living Paddlefishes using molecular data that date to about 68 mya (Peng et al. 2007). Ram ventilation is used by both species of living Paddlefishes; therefore, it preceded the evolution of filter feeding in Polyodon (Burggren & Bemis 1992).

Figure 6.4. Phylogenetic hypothesis for relationships between living and fossil Polyodontidae (from Grande & Bemis 1996; Grande et al. 2002; used with permission of Lance Grande).

211

FOSSIL RECORD Four species of Paddlefish fossils are currently described, three found in North America. They all possess a paddle, an extreme elongation of the snout supported by a series of long median dorsal and ventral rostral bones. They also have stellate bones present in the paddle that form a dense interlocking network and support the lateral aspects of the paddle in more advanced species. An additional polyodontid character is the presence of tiny non-interlocking scales or denticles bearing one to three anterior knobs and a posterior fringe of spines (microctenoid scales, Fig. 6.5) that cover the trunk; this character was lost in more derived Paddlefish species (Grande & Bemis 1991; Grande et al. 2002). The fossil most closely related to the living Paddlefish is †Polyodon tuberculata from the Lower Paleocene Tullock Formation, Montana, dating to about 60 mya. This species is represented by only one specimen consisting of a nearly complete, but somewhat crushed, skull and part of the caudal fin. This fossil is included in the genus Polyodon because it shares numerous characters with P. spathula not found in other polyodontid fossils, including numerous long gill rakers highly modified for filter feeding, a close attachment of the upper jaw to the braincase, and an elongated and thin lower jaw. The major differences between †P. tuberculata and P. spathula may be partially developmental because †P. tuberculata is estimated to have reached about 2 m TL (6.6 feet), much larger than any P. spathula available for comparison, and Paddlefishes continue ontogenetic development their entire lifespans. Differences are basically the extent of ossification with †P. tuberculata possessing a heavily ossified skull with deep crests and tubercles on the dorsal surface (Fig. 6.6). Because this fossil occurs within the range of P. spathula, the genus Polyodon

Figure 6.5. Microctenoid scales present on a specimen of the fossil polyodontid †Protopsephurus liui; these scales are lost in living Paddlefishes (from Grande et al. 2002; used with permission of Lance Grande).

212 FRESHWATER FISHES OF NORTH AMERICA

Figure 6.6. Fossil polyodontid †Polyodon tuberculata from the Lower Paleocene Tullock Formation, Montana, dating to about 60 mya (from Grande & Bemis 1991; used with permission of Lance Grande).

Figure 6.7. Fossil polyodontid †Crossopholis magnicaudatus from the Lower Eocene (50 mya) of the Green River Formation, Wyoming (from Grande & Bemis 1991; used with permission of Lance Grande).

Figure 6.8. The oldest fossil polyodontid †Protopsephurus liui, found in formations from the Lower Cretaceous (>100 mya) in China (from Grande et al. 2002; used with permission of Lance Grande).

likely has been in the Montana area for 60 million years (Grande & Bemis 1991). †Crossopholis magnicaudatus is represented by numerous specimens, some complete, from the Lower Eocene (50 mya) of the Green River Formation, Wyoming (Fig. 6.7). This fossil possesses a dense covering of microctenoid scales over most of its body not found in Polyodon, and its paddle tapers from a wide base to a narrow anterior end compared with the straight or anteriorly expanded paddle in Polyodon. Specimens range in size from 26 cm to almost 1.5 m TL (0.9–4.9 feet). Based on fossilized stomach contents, its diet was fishes, which is the same as the diet of the living Chinese Paddlefish (Psephurus gladius) (Grande & Bemis 1991). †Paleopsephurus wilsoni is represented by a single skull, shoulder girdle, partial caudal peduncle region, and fin; these may be from 1 to 4 specimens about 56 cm TL (22 inches) that were found in Upper Cretaceous deposits (65 mya) of the Hell Creek Formation, Montana (Grande & Bemis 1991). This Paddlefish specimen, as well as three articulated Sturgeon fossils, was independently discovered in abdominal areas of hadrosaurian (duck-billed) dinosaurs (Grande & Hilton 2006; Hilton & Grande 2006).

Because these fossil specimens are whole and hadrosaurs were plant eaters, the fishes likely were not eaten. Instead, the dinosaur carcasses trapped the deceased acipenseriforms and facilitated their rapid burial by forming sediment traps in near-shore or river habitats (Grande & Hilton 2006). †Paleopsephurus wilsoni possesses unique skull features and a stouter and more heavily ossified shoulder girdle than other North American polyodontids. As in †Crossopholis, †P. wilsoni possesses microctenoid scales on the few body parts available for examination (Grande & Bemis 1991). Past researchers did not consider †Paleopsephurus a Paddlefish but rather the sister-group to Sturgeons (Acipenseridae) (Gardiner 1984). This view was based partially on the reported absence of stellate bones in the paddle in the original description (MacAlpin 1947), but stellate bones were found upon additional preparation of the skull (Grande & Bemis 1991). The oldest fossil Paddlefish by about 50 million years is †Protopsephurus liui, found in formations from the Lower Cretaceous (>100 mya) in China (Fig. 6.8). Although †P. liui possesses several polyodontid characters (e.g., elongate snout supported by a series of long median dorsal and ventral rostral bones and the presence of stellate bones and

POLYODONTIDAE: PADDLEFISHES

microctenoid scales), †P. liui lacks several derived characters found in other species and possesses a pectoral spine similar to that found in Sturgeons, indicating that this is the most basal Paddlefish. All fossil Paddlefishes, including †P. liui, are from freshwater deposits, indicating that their basic habitat has not changed in tens of millions of years (Grande et al. 2002).

213

A

MORPHOLOGY

Ancient Body Plan Although the North American Paddlefish has many derived characters relative to other extant and fossil Paddlefishes, all members of this family share a basic morphology common to other acipenseriforms, including fossil taxa from the Lower Jurassic (200 mya). These relic characters include a subcylindrical body, heterocercal tail, reduced ossification of the endoskeleton and a persistent notochord, the mouth on the lower surface of the head, loss of the maxillary and premaxillary bones, fin rays more numerous than their basal skeletal supports, body scaling reduced to tiny isolated elements, and novel median V-shaped scales (fulcra) at the dorsal and ventral base of the caudal fin (Fig. 6.9). In the suborder Acipenseroidei (Paddlefishes and Sturgeons) relic characters include the endocranium greatly extended into a rostrum, dorsal fin behind the pelvic fins, pectoral fins extending laterally from a ventral insertion on the pectoral girdle, electrosensory (ampullary) organs concentrated on the underside of the rostrum, an intestine with spiral valve, a simple gas bladder with a connection to the esophagus, and no opercle with the subopercle acting as the gill cover supported by a reduced number of branchiostegal rays (one to three) (Vladykov & Greeley 1963; Grande & Bemis 1991; Bemis et al. 1997; Grande et al. 2002; Mabee & Noordsy 2004). The endoskeletal elements of the pectoral fins of Paddlefishes and Sturgeons have elements homologous to both the fin radials of teleosts and the limb bones of tetrapods (Davis et al. 2004).

Unique Characters All Paddlefishes are easily recognized by their extremely elongated snout (rostrum), referred to as a paddle. The paddle is supported by long medial rostral bones and an interdigitating network of lateral rostral bones called stellate bones that get their name from the star-like points radiating from their center (Fig. 6.10). Other characters

B

Figure 6.9. Novel median V-shaped scales (fulcra) along the dorsal edge of the caudal fin (A) and a single ventral fulcrum (B) at the caudal fin base in the caudal skeleton of a Paddlefish, Polyodon spathula (from Grande & Bemis 1991; used with permission of Lance Grande).

unique to all Polyodontidae include a reduction in the relative size of teeth as individuals grow, the subopercle with radiating pattern of splint-like rods dorsally and posteriorly, a lateral-line canal enclosed in an ossified tube for the entire length of the caudal fin, a series of heavy dorsal caudal fulcra, a single ventral caudal fulcrum (Fig. 6.9), interlocking rhomboid scales on the upper caudal lobe, round-based scales on the shoulder and isthmus regions, and fringed (microtenoid) scales covering the trunk in fossil species (Fig. 6.5) that are replaced by vestigial denticular scales in living species. Except for the oldest and most primitive polyodontid fossil (genus †Protopsephurus), Paddlefishes also possess branchiostegal rays modified into a single bone with a branched posterior edge and lack a pectoral spine present in other Acipenseriformes. Living Paddlefishes have a pair of nostrils at the base of the paddle anterior and dorsal to extremely small eyes, a pair of minute barbels on the underside of the paddle anterior to

214 FRESHWATER FISHES OF NORTH AMERICA

Figure 6.10. Skull of a Paddlefish, Polyodon spathula, with paddle supported by dorsal and ventral sheets of numerous and densely packed interdigitating stellate bones, which are attached to each other and to the median rostral bones (from Grande & Bemis 1991; used with permission of Lance Grande).

A

Figure 6.11. The huge gape of the Paddlefish, Polyodon spathula, allows a large volume of water to enter its mouth and facilitates ram ventilation and filter feeding on zooplankton (photographs by and used with permission of (A) ©Engbretson Underwater Photography, taken in Table Rock Lake, Branson, Missouri, October 2005, and (B) Todd Stailey, Tennessee Aquarium).

B

the mouth, a small spiracle, a fleshy tapering posterior expansion of the gill cover, and pyloric caeca present as a broad and branching organ (Forbes & Richardson 1920; Weisel 1975; Grande & Bemis 1991; Bemis et al. 1997; Grande et al. 2002; Hilton 2004).

American Paddlefish reaches a maximum of 2.15 m (7.1 feet) and 74 kg (163 pounds) (Nichols 1916; Forbes & Richardson 1920; Hochleithner & Gessner 1999).

Size

Unlike the Chinese or most fossil Paddlefishes that use protrusable jaws (similar to Sturgeons) to feed on a variety of prey, including fishes, the North American species is a filter feeder and possesses numerous morphological specializa-

The Chinese Paddlefish is the largest extant species reaching 3.6 m (11.8 feet) and 300 kg (660 pounds). The North

Filter Feeding and Electrosensory Organs

POLYODONTIDAE: PADDLEFISHES

tions that accommodate this mode of feeding. A large volume of water can enter the mouth of a Paddlefish because the upper jaw is firmly attached to the braincase by a short ligament and both jaws are elongate, giving the Paddlefish a huge gape as it drops its lower jaw (Fig. 6.11). Collection of zooplankton is accomplished by filtering water across extremely numerous long and flattened gill rakers (Fig. 6.12) found in a double series on gill arches compressed to waferlike plates (Fig. 6.13). The paddle is supported by dorsal and ventral sheets of numerous and densely packed stellate bones that are attached to each other and the median rostral bones (Fig. 6.10) (Forbes & Richardson 1920; Grande & Bemis 1991; Bemis et al. 1997). The Paddlefish can detect concentrations of zooplankton with electrosensory (ampullary) organs covering the relatively broad paddle. Both dorsal and ventral sides of the paddle are covered by dark pores that lead to ampullary organs that can detect weak electric fields emitted by zooplankton. Pores are partially filled with a jelly-like substance and occur in clusters (commonly 10–20/cluster) between stellate bones; ≤4 ampullae may share a single pore (Fig. 6.14). Ampullary organs also occur on the head and subopercle. Paddlefish have more ampullary organs (50,000–75,000) than any other fish (Collinge 1894; Kistler 1906; Nachtrieb 1910; Jørgensen et al. 1972; New & Bodznick 1985; Wilkens et al. 2002). One ampullary organ has sensory epithelium with ≤400 electrosensitive hair cells at the end of a short duct. The duct length (100–250 μm) is much shorter in Paddlefish and other freshwater fishes relative to the Lorenzinian ampullae of elasmobranchs, perhaps the result of differences in conductivity between fresh and salt water (Jørgensen et al. 1972; Neiman et al. 2000). Although the anterior lateralline nerve innervates both ampullary organs and mechanoreceptive neuromasts of the cephalic lateral-line system, a clear division exists between these sensory systems in the hindbrain of Paddlefish (and Sturgeons) (New & Bodznick 1985). Paddlefish are unique among bony fishes in having three electrosensory pathways to deliver information from the hindbrain to the midbrain, where information for orienting movements and prey capture is processed (Hofmann et al. 2002; Wilkens et al. 2002; see behavior section). Because the sensory epithelium and the function of the ampullary organs of Paddlefish are similar to those of Lorenzinian ampullae found in elasmobranchs, the electrosensory system in Acipenseroidei likely is derived from a common ancestor shared with cartilaginous fishes (Jørgensen et al. 1972; New & Bodznick 1985; Northcutt 1986; Wilkens & Hofmann 2007).

215

Figure 6.12. Extremely numerous long and flattened gill rakers in an adult Paddlefish, Polyodon spathula, allow for collection of zooplankton as water is filtered through these structures. Note the trapped zooplankton under the gill rakers (from Bemis et al. 1997; used with permission of Willy Bemis).

Figure 6.13. Gill arch from a Paddlefish Polyodon spathula laterally compressed to wafer-like plates that allows for ram ventilation. Anterior (top) and dorsal (bottom) view (from Grande & Bemis 1991; used with permission of Lance Grande).

Vision and Chemosensory System The extraordinary development of electroreception for filter feeding is countered by a reduction of visual and chemosensory structures. Paddlefish eyes are extremely small, and the electrosensory system appears to be partially replacing the visual system in the midbrain of Paddlefish (Hofmann et al. 2002). Paddlefish have only two minute barbels anterior to the mouth on the underside of the

216

FRESHWATER FISHES OF NORTH AMERICA

Figure 6.14. Enlarged view of dorsal surface of a Paddlefish, Polyodon spathula, paddle showing clusters of dark pores between stellate bones (visible through the skin of the paddle) that lead to electrosensory (ampullary) organs (from Grande & Bemis 1991; used with permission of Lance Grande).

Therefore they use ram ventilation, which involves steady and continuous swimming (1.25 body lengths/s for juveniles) with the mouth open slightly to allow a continuous flow of water through the oral cavity, over the gills, and out the gill cover. The long, tapering gill cover allows water to flow out of the gill chamber dorsally, caudally, or ventrally. Gill arches are held in a position that exposes the gill filaments to the direct flow of water. No other fishes as small or as slow as juvenile Paddlefish have ever been documented as ram ventilators. Ram ventilation in Paddlefish is only possible because of the extremely wide mouth gape and the lateral flattening of the gill arches. If Paddlefish are forced to stop swimming or to swim at slow speeds ($200/kg ($69.85–$90.72/pound) wholesale and $381–$1,340/kg ($172.82–$607.82/pound) retail (Pikitch et al. 2005; Scholten & Bettoli 2005, 2007; Bettoli et al. 2009b). Nets are the most efficient method to commercially harvest Paddlefish. When a Paddlefish makes contact with a net, it makes feeble efforts to free itself and can be removed by hand with little struggle (Stockard 1907; Alexander 1914; Coker 1923). In the early 1900s, 3.2 km (2 miles) seines, 9.1 m (30 feet) deep, were used to capture Paddlefish in backwaters and floodplain lakes in the lower Mississippi River basin by encircling an area >1.6 km (1 mile) in circumference and hauling in the seine on a huge reel by the crew walking up the spokes of the wheel (Fig. 6.20) (Stockard 1907; Hussakof 1911). Another method involved dragging a 183 m (600 feet) seine between 2 boats near the surface up and down a lake all day with a row boat moving along the seine every 0.5 h to remove Paddlefish from the net (Alexander 1914). Drifting trammel nets were used in rivers to capture Paddlefish concentrated below dams (Coker 1923). Currently fishers typically use large-meshed gill and trammel nets (127–152 mm bar mesh, 5–6 inches) to harvest Paddlefish, which are effective (Quinn 2009). Two fishers harvested 5,443 kg (12,000 pounds) in 2 nights in a subimpoundment of a reservoir (Semmens & Shelton 1986). Gillnets cannot be easily deployed in riverine sections of reservoirs when discharge is high. The number of Paddlefish harvested in Kentucky Lake is positively related to the number of fishable days in a season, with greater harvest in drought years (Scholten & Bettoli 2005).

241

Figure 6.20. Huge reel mounted in a flat-bottomed boat winding up a seine to catch Paddlefish, Polyodon spathula, by the crew walking up the spokes of the wheel (from Hussakof 1911).

Because Paddlefish are zooplanktivores they do not take bait, so snagging is the only common method used in recreational harvest. Snagging involves jerking a large treble hook and a lead weight through the water on a heavy line with a stout spinning rod and reel (Purkett 1963a; Scarnecchia et al. 1996a; Hansen & Paukert 2009). Although Paddlefish incur injury when snagged, hooking mortality may be low and a catch-and-release sport fishery is feasible (Scarnecchia et al. 1996a; Scarnecchia & Stewart 1997). A small archery fishery also exists (Rosen & Hales 1980). As of 2006, sportfishing for Paddlefish is allowed in 14 of 22 states with extant Paddlefish populations (Bettoli et al. 2009b; Hansen & Paukert 2009). From 2000 to 2006 annual sport harvest of Paddlefish ranged between 15,000 and 20,000, with 7 fisheries harvesting >1,000 Paddlefish annually (Quinn 2009).

Sport Fisheries Most sport fisheries for Paddlefish developed after construction of dams on rivers. Paddlefish concentrating in tailwaters below dams provide a substantial fishery. Those populations that support a self-sustaining fishery are in the upper and central Mississippi River basin. Paddlefish harvested are typically immature, but a few sport fisheries concentrate on adult Paddlefish during spawning migrations. Large harvest coincides with high spring flows (Carlson & Bonislawsky 1981; Combs 1982; Rosen et al. 1982). Tailwater fisheries can be affected by power generation negatively since low flows decrease Paddlefish abundance, especially over weekends when energy demands are low (Unkenholz 1986).

Aquaculture Artificial propagation of Paddlefish is used for stocking to enhance sport and commercial fisheries, reestablishing extirpated populations, and raising Paddlefish outside of North America for harvest. The preferred method of rearing young Paddlefish is to stock larvae in heavily fertilized ponds, which promotes dense growth of zooplankton. Polyculture in ponds with Channel Catfish and other fishes also is used, with excess food replacing fertilizer to promote zooplankton growth (Graham et al. 1986; Semmens & Shelton 1986; Mims et al. 2009). Paddlefish also can be raised in tanks and trained to eat an artificial diet of sinking pellets, but they switch to filter feeding on

242 FRESHWATER FISHES OF NORTH AMERICA

zooplankton when placed in ponds or reservoirs (Graham et al. 1986; Kroll et al. 1992, 1994; Mims 2001). In China, Paddlefish are raised in floating cages with an artificial diet for production of flesh (Mims et al. 2009). Paddlefish were successfully established in an Ozark reservoir by stocking fingerlings, and a sport fishery was realized in 10 years (Graham 1986). A commercial fishery for Paddlefish can be created by stocking fingerlings in small reservoirs and harvesting in $650) (Mims 2001; Bettoli & Scholten 2006; Bettoli et al. 2007). To increase caviar yields attempts are being made to produce all-female (monosex) populations of Paddlefish for culture, which involves gynogenesis (producing embryos containing only maternal chromosomes), steroid-induced sex reversal of some individuals to produce neomales (produce only X-determinant sperm), and the use of these neomales to produce all-female progeny (Mims & Shelton 1999). Paddlefish were introduced into the former Soviet Union and Europe beginning in 1974 for aquaculture in ponds, and demand for live Paddlefish

and eggs has continued overseas. Paddlefish are now part of polyculture systems for meat and caviar in several European, Middle Eastern, and Asian countries (Hoover 1999; Billard & Lecointre 2001; Vedrasco et al. 2001; Lobchenko et al. 2002; Hubenova et al. 2007; Mims et al. 2009).

LITERATURE GUIDE Several sources provide excellent overviews on the Paddlefish in North America. These include edited volumes and books that cover many aspects of Paddlefish biology, management, propagation, and conservation (Dillard et al. 1986; Hochleithner & Gessner 1999; LeBreton et al. 2004; Paukert & Scholten 2009) and those that focus on harvest and conservation (Williamson et al. 1999; Williamson 2003).

Acknowledgments The Department of Biological Sciences at the University of Alabama provided support for this project. Thanks to all of those who provided figures or gave permission for figure use, and the editors for their comments that improved the original manuscript.

Chapter 7

Lepisosteidae: Gars Anthony A. Echelle and Lance Grande

Living Gars are easily identified by their torpedo-shaped bodies encased in hard, rhombohedral-shaped scales, their posteriorly set median fins, and their elongate bills lined with sharply pointed teeth. Today’s Gars (order Lepisosteiformes) include seven species, five in eastern North America and one each in Cuba and the tropics of Central America. They are living fossils in the sense of Hubbs & Lagler (1958:30) and Wiley & Schultze (1984). That is, the family is ancient (Gars are >100 million years old), the living species show numerous primitive neopterygian traits, and the closest relatives are extinct. The living-fossil status and the air-breathing habit of Gars have generated a great deal of interest among fish systematists, developmental and evolutionary biologists, and physiologists. Gars are sometimes considered undesirable, a view fueled by their reputation for competing with and consuming more desirable gamefishes, their potential for fouling commercial fishing nets, and sometimes even the perception that the larger species pose a threat to humans (Scarnecchia 1992; Spitzer 2010). Most of the perceived problems posed by Gars are minor relative to the negative reactions they often receive from fishers and, until rather recently, most state and federal fishery management programs (Scarnecchia 1992; Spitzer 2010). Gars are now beginning to be appreciated for their own sake and as integral components of healthy aquatic ecosystems. As top predators they perform ecosystem functions that may be important in stabilizing populations of gamefish and their prey (Cross 1967; Scarnecchia 1992). At the same time, their unusual, prehistoric appearance makes Gars valued attractions at public zoos and aquaria. This is particularly true for the larger species, one of which, the Alligator Gar, Atractosteus spatula, reaches about 3 m TL.

DIVERSITY AND DISTRIBUTION The Gars inhabit waters of eastern and central North America, Meso-America, and Cuba (Figs. 7.1 and 7.2). The fossil record extends the time range of the Lepisosteidae to >100 mya and extends the geographic range to western North America, Europe, Africa, India, and South America. Gars, particularly Alligator Gars, occasionally occur in coastal brackish or marine habitats, but the extant Gars are effectively freshwater fishes (Table 7.1). The same appears true for fossil Gars, which are almost entirely known from freshwater deposits (Grande 2010). The living Gars are placed in two genera: Lepisosteus with four species (Figs. 7.3 and 7.4) and Atractosteus with three species (Fig. 7.5). The extant members of Lepisosteus are restricted to eastern North America from Montana and Texas eastward. The distribution of Lepisosteus osseus (Longnose Gar, Aguja in Mexico) includes southern Quebec, Canada, south to Florida and northern Mexico, and westward from the Great Lakes region to Montana. The Longnose Gar is also referred to vernacularly as the common Gar, billfish, and needlenose Gar. Lepisosteus oculatus (Spotted Gar, Catán Pinto in Mexico) occurs from the Great Lakes south to the Gulf Coast of Texas, northern Mexico, and east to northwestern Florida. Lepisosteus platyrhincus (Florida Gar, Florida spotted Gar in earlier literature) is restricted to Florida and lowlands of southern Georgia. Lepisosteus platostomus (Shortnose Gar, vernacularly known as the duckbill garfish) occurs in lowgradient regions of the Mississippi River basin, from northeastern Texas north to Montana, east to Ohio, and south to Mississippi.

Figure 7.1. Geographic range of Lepisosteus.

Figure 7.2. Geographic range of Atractosteus.

Genus Lepisosteus Genus Atractosteus

Table 7.1. Life history characteristics of Gars, Lepisosteidae. Characters

Characteristics

Comments

Number of extant species

Seven in two genera

Salinity preferences Maximum size recorded in m TL

Fresh and brackish waters Alligator Gar, 3 Cuban Gar, 2 Longnose Gar, 1.6 Others, 1.1–1.3 Alligator Gar, 50 years Longnose Gar, 32 years Others, 8 ppt. Significant mortality and reduced growth occurred at 8 ppt, and no survival occurred at 10, 12, and 14 ppt. All larvae survived at 0, 2, 4, and 6 ppt with no differences in growth rate. Early juveniles (20 days post-hatching) survived at 12 ppt with stepwise acclimation to the end of the study 31 days later, but growth was slower and mortality was higher than at 4 and 8 ppt. Juveniles at 50 days post-hatching showed no difference in mortality and little difference in growth at salinities of 0, 6, 12, and 18 ppt. Juvenile Alligator Gars and Spotted Gars are tolerant of higher salinities than are juvenile Paddlefish and Lake Sturgeons (Suchy 2009). Non-acclimated Paddlefish and Lake Sturgeons showed 100% mortality at 16 ppt, but Spotted Gars and Alligator Gars did not reach 100% mortality until 20 ppt and 36 ppt, respectively. The Paddlefish and Lake Sturgeon showed no improvement with acclima-

259

tion, but with acclimation, juvenile Spotted Gars (average TL 122 mm) and Alligator Gars (199 mm TL) tolerated salinities ≤30 and 37 ppt, respectively. Florida Gars apparently resemble Spotted Gars in salinity tolerance; juveniles tolerated 75% sea water (about 26 ppt) during a 2-week acclimation period before osmoregulation experiments, but they were unable to tolerate full-strength sea water (about 35 ppt; Zawodny 1975).

pH Tolerance Gars appear to be relatively tolerant of acidic waters compared with other fishes. Survival time was not affected in juvenile Florida Gars exposed to pH as low as 3.8, but at 3.6 there was a significant effect with an average survival time of about 40 h (Krout & Dunson 1985). Death at low pH was associated with a large increase in net sodium loss, suggesting that the cause of death involves ionic imbalance. Tolerance to low pH was correlated with a high capacity for resisting the net sodium loss usually associated with pH toxicity in fishes.

Pollution Tolerance Air breathing apparently makes Gars more tolerant of pollutants affecting uptake of oxygen at the gills. For example, in 1988, an estimated 37,000 fishes died in a series of fish-kills in the Fox River, Wisconsin, from carbon monoxide pollution produced by motorboat engines. Gars were present in the river but were not included in the kills (Kempinger et al. 1998). On the other hand, Gars are undoubtedly adversely affected by many aquatic pollutants. As long-lived top predators, they are susceptible to bioaccumulating compounds such as organophosphate pesticides and heavy metals (Cruz et al. 2010). Spotted Gars from a petroleumcontaminated site in Louisiana showed gonadal cysts in both sexes (4 of 30 males; 1 of 42 females). The sample from the control site was relatively small (17 males, 15 females), but no cysts were detected (Thiyagarajah et al. 2000). Gars seem to have relatively high tolerance to dissolved ammonium concentrations. This is particularly true of Alligator Gars, which in a recent review had the highest tolerance known for fishes (Boudreax et al. 2007b). The Spotted Gar had greater tolerance than all except 2 of 12 non-Gar species included in the review. The exceptions were relatively hardy species (Common Carp, Cyprinus carpio, and Red Shiner, Cyprinella lutrensis). The 96-h

260

FRESHWATER FISHES OF NORTH AMERICA

median-lethal concentrations for total ammonia and unionized ammonia were 135 and 4.30 ppm, respectively, for Alligator Gars. The corresponding values for Spotted Gars were 35 and 1.82 ppm. High ammonia tolerance in Gars might be explained as an adaptation to their high-protein, therefore high-nitrogen diets (Boudreax et al. 2007b). Gars tend to concentrate aquatic nitrite internally (Boudreax et al. 2007a). This is because they retain the ancestral actinopterygian mechanism for chloride uptake (see osmoregulation and exchange of carbon dioxide and ammonia subsection) and nitrite competes with chloride for this mechanism of transport across the gill membranes. Because of this competitive relationship, tolerance to nitrite pollution is higher in waters containing higher chloride content. Gars exposed to 1 ppm nitrite and 20 ppm chloride survived better than those exposed to 1 ppm nitrite and 8,600 cells) (Popper & Northcutt 1983). The otolith chambers are innervated by nerves connecting to lateral-line sensory pores that convey information on orientation, but other nerves from the lateral line connect to the central ner vous system (McCormick 1981, 1982). The Bowfin, Gars, and Hagfishes (Myxinidae) are the only non-teleosts that lack the dorsal nucleus, a nexus of nerves responsible for electroreception, and thus species in these families cannot detect electric fields (Bullock et al. 1983).

Olfaction and Chemosensation In Bowfin, the nasal opening is at the end of an extended tube, similar to that in Bichirs, and the olfactory organ itself is relatively near the anterior portion of the head as in most actinopterygians (e.g., Bichirs, lepisosteids). The olfactory rosette, the main site of olfaction, is made up of many folds of tissue, or lamellae, which can number >100 in Bowfin—much greater than in acipenserids (20–32), polyodontids (13–18), and lepisosteids (8–10). Secondary lamellae are absent (Chen & Arratia 1994). Like most fishes, Bowfin taste buds contain two main cell types, electron-light and electron-dark, which form the epithelium of the taste bud and likely serve as secondary sensory cells. Within each of these cell types, Bowfin further exhibit two subtypes of taste bud microvilli (apical areas of taste bud cells that house the receptors). Although fishes from several diverse families exhibit variation in microvilli subtype in light cells, morphological variation in dark cell microvilli is much less frequent, and one of the subtypes appears unique to Bowfin. The significance of variation in taste bud microvilli subtype is unclear, but these appear to be speciesspecific and may have evolved as each species came to occupy a specific niche (Reutter et al. 2000; Reutter & Hansen 2005).

Digestion Digestion rate in young-of-the-year Bowfin was slower than that of the Longnose Gar, Lepisosteus osseus, requiring 28–32 h to complete digestion of a food ration equaling 4.9% of body weight when held at a constant 21°C (69.8°F) (Herting & Witt 1968). An adult Bowfin had par-

289

tially digested a large tadpole after 2.5 h, though the holding temperature was not given (Hopkins 1890). A Bowfin kept in captivity survived for 20 months without food (Smallwood 1916).

BEHAVIOR The daily activities and behaviors of Bowfin are inadequately studied (but see reproduction section). Adults are mostly solitary and sedentary predators that spend considerable time in an almost motionless state. In winter, in Oconomowoc Lake, Wisconsin, Bowfin were “in schools closely huddled together in the bottom of pockets or shallow depressions of the gravelly bed of the lake, among the water-weeds . . . They lie so close together that occasionally two individuals are impaled on the fish-spear by one throw. When thus disturbed they scatter from their resting places. Moving out a short distance to return quickly after the first few disturbances” (Ayers in Reighard 1903:65). Observers have considered Bowfin in natural settings to be ambush predators. In swamps they can be heard hitting the surface of the water to presumably gulp air, and predation is often a crepuscular activity (Reighard 1903).

Movement and Dispersal Little information is available on dispersal and movement in the Bowfin. Differences were not detected in diel motor activity (Reynolds et al. 1978), but no studies have examined longer-distance movements from day to night. Bowfin may move en masse into inundated floodplain habitat or tributaries during high-water periods, but this is likely associated with spawning activities (Eddy & Underhill 1974; Simon 1990). In the upper Mississippi River, Bowfin, along with other fishes like Northern Pike, Esox lucius, and crappies (Pomoxis spp.), actively moved out of backwater habitats toward the main channel with lowering water levels and current flow toward the main channel, unlike Bluegill and Largemouth Bass that tended to be trapped in isolated pools. In a Wisconsin study of fish movement under winter ice conditions between a backwater lake and the Mississippi River, >600 Bowfin were recorded in trap nets leaving the backwater and 206 individuals were detected entering the backwater; although all were marked, few individuals were recaptured (Greenbank 1956). Total fish movement, including Bowfin (January–February), was positively correlated with snow

290

FRESHWATER FISHES OF NORTH AMERICA

cover depth on the ice, and many more Bowfin entered the lake during this time than emigrated (Greenbank 1956). In the Savannah River, North Carolina, Bowfin migrate in spring from the river to lower portions of Steele Creek, returning to the river in autumn when the creek temperatures fall below that of the river (Marcy et al. 2005). Fish traps set over a 162-day period spanning 2 years revealed the Bowfin as one of the most abundant fishes (101 individuals, averaging 477 mm TL, 18.8 inches) moving from the Okefenokee Swamp through a spillway to the Suwannee River (Holder 1970). Via extrapolation of trap area to spillway area the author estimated about 6,400 Bowfin moved from the swamp to the river. Nearly all the movement occurred during December to February when water temperatures were at annual lows. Of 266 tagged Bowfin in a year-long study in a North Carolina swamp stream, only 35 were recaptured, and none of the recaptures had moved >0.2 km (0.12 mile) from the original tagging site (Whitehurst 1981).

Schooling Yolk-sac larvae have a prominent adhesive organ on the snout that is used for attachment to plant materials. Once the yolk is absorbed the young school in a large black pod or swarm and are accompanied by the former nestguarding male (Fig. 8.12). Most of the young in the pod usually swim at the same rate of speed, face the same direction, and move in concert (i.e., the school is polarized). Young remain with the adult male until about 100 mm TL (3.9 inches), but schools become smaller, more loosely aggregated, less polarized, and not well guarded after the young reach about 35 mm TL (1.4 inches). Brood protec-

tion then ends, and the young scatter to sheltered areas and begin to live and feed entirely on their own. In aquaria, young Bowfin can be quite active in swimming and in pursuit of live food. Within a few months they usually become too large for most aquarium keepers and are less active during the day.

REPRODUCTION One aspect of Bowfin natural history that is generally ignored in the phylogenetics of so-called ancient fishes is the fact that Amia calva builds a nest and only the male guards its young. This highly specialized reproductive mode is unlike any of the living ancient fish lineages in North America (e.g., Sturgeons, Paddlefishes, Gars), which are made up of broadcast spawners that do not build nests or show parental care-giving behaviors. Only in Ictaluridae (North American Catfishes) and Centrarchidae (Sunfishes), lineages distantly related to Bowfins, do we encounter nest building, parental guarding, and care-giving behaviors.

Sexual Maturity Male Bowfin mature at ages 2–4 (0.25 mm. During the peak spawning period, oocytes can hydrate and develop from 0.35 to 0.90 mm (spawning size) in 24 h. Hydration consists of oocytes rapidly absorbing fluids (4–14 h) of lower specific gravity than sea water and yolk granules fusing into yolk plates (Luo & Musick 1991; Bassista & Hartman 2005). Spawning occurs nightly at about the same time (Kuntz 1915) with high fertilization rates (94–100%) (Peebles 2002). Spawning typically starts between 1800 and 2100 h (e.g., 1800 h, Beaufort, North Carolina, Kuntz 1915; 2000 h, Manatee River Estuary, Florida, Peebles 2002; 2000 h in Hudson River Estuary, New York, Bassista & Hartman 2005; 2100 h in Barnegat Bay, New Jersey, Vouglitois et al. 1987). Nightly timing also can be dynamic in the same estuary as summer progresses. Bay Anchovy in Chesapeake Bay started spawning at 2000 h in June, 2100 h in July, and 2300 h in August (Luo & Musick 1991). Nightly spawning duration is typically 1–4 h (Luo & Musick 1991; Zastrow et al. 1991; Peebles 2002) but can last ≤10 h (Bassista & Hartman 2005). Bay Anchovy can spawn >50 times in a season. Almost all Bay Anchovy females in Chesapeake Bay apparently spawned nightly for about 50 nights during the peak period in 1986–1987 (Zastrow et al. 1991).

Fecundity, Egg Densities, and Reproductive Allocation Bay Anchovy are highly fecund and can be the most abundant taxon of fish eggs and larvae collected in estuaries (Dovel 1981; Olney 1983; Leak & Houde 1987). Mean batch fecundity (eggs/spawn) ranges from 429 to 1,186 in Lower York River, Chesapeake Bay (Luo & Musick 1991), and from 514 to 2,026 in Patuxent River, Chesapeake Bay

(Zastrow et al. 1991). Relative batch fecundity (eggs spawn−1 g−1 ovary-free female body weight) ranged from 334 to 803 (Luo & Musick 1991) and 441 to 959 (Zastrow et al. 1991) in the same studies. In the Hudson River Estuary, the mean relative batch fecundity was 506 with mean batch fecundities of 1,233 in 1996 and 1,508 in 1997 (Bassista & Hartman 2005). Length and weight of Bay Anchovy females are correlated positively with batch fecundity (Luo & Musick 1991; Wang & Houde 1994; Bassista & Hartman 2005). A linear regression model (r 2 = 0.70) predicted batch fecundity (hydrated oocytes per batch) of about 750 for a 1 g female (ovary-free body weight), 1,500 for a 2 g female, and 2,250 for a 3 g female (Luo & Musick 1991). Adult prey availability may affect egg production and fecundity (Peebles et al. 1996; Peebles 2002). Egg densities >200 eggs/m3 occur during spawning, peak at 1,098 for Great South Bay, New York (Monteleone 1992), range from 0.64 to 30.77 seasonally in Biscayne Bay, Florida (Houde & Lovdal 1984), range from 32 to 140 with a peak of 800 in Chesapeake Bay (Olney 1983), and range from 5 to 132 during peak spawning in Barnegat Bay, New Jersey (Vouglitois et al. 1987). Overall estimates of 45,110 eggs female−1 season−1 (Luo & Musick 1991) and 4.25 × 1012 (June 1993) to 8.43 × 1012 (July 1993) eggs/day baywide (Chesapeake Bay, Rilling & Houde 1999) provide insight into how fecund Bay Anchovy really are. Gonad somatic indexes (GSI, gonad weight / gonad-free body weight × 100) are highest in the spring and summer during the Bay Anchovy spawning season and rapidly decline afterward. In Chesapeake Bay males have higher GSI than females (Zastrow et al. 1991); this may be true for all populations because males have larger gonads than females of the same body weight (Wang & Houde 1994). From February to November in Chesapeake Bay, GSI ranged from 0.47 to 7.40% for females and from 0.13 to 10.95% for males. During peak spawning season, females averaged 4.44% and males 7.18% (Zastrow et al. 1991). Female GSI ranged from 5.74 to 17.13% and male GSI from 6.06 to 15.09% during the spawning season (June– September) in the Hudson River Estuary (Bassista & Hartman 2005). The differences for these two estuaries may be because Zastrow et al. (1991) used Bay Anchovy >40 mm FL and Bassista and Hartman (2005) used individuals >60 mm TL for their analyses.

Embryonic Development Bay Anchovy eggs occur throughout the water column at salinities of 8–15 ppt, tend to concentrate near the surface

ENGRAULIDAE: ANCHOVIES

at salinities of 30–35 ppt, and remain buoyant until hatching (Jones et al. 1978; Morton 1989). This is a function of egg buoyancy since buoyancy decreases at low salinities and eggs remain buoyant at salinities >14 psu (= 14 ppt) (Peebles 2002). Fertilized eggs of Bay Anchovy are relatively small, transparent, and planktonic; lack oil globules; and contain little yolk material (Kuntz 1915; Jones et al. 1978). Yolk that is present is separated into large masses appearing as big broken cells (Fig. 11.8). The eggs are

347

slightly elongate; the major axis ranges from 0.65 to 1.33 mm long; and the minor axis is 0.1– 0.3 mm shorter (Kuntz 1915). Egg size decreases with increasing water salinity and as the spawning season progresses (Jones et al. 1978). Eggs weigh 15 μg (Tucker 1989). Development of Bay Anchovy eggs (detailed by Kuntz 1915) is not different from development of typical pelagic teleostean eggs. Development and hatching times are relatively rapid. The egg is in the late morula stage (≥16 cells) 5 h after fertilization. At 10 h the blastopore closes, the embryo is >0.5 of the egg circumference, and it lies parallel with the major axis of the egg. Up to hatching, the embryo keeps elongating and can extend around the entire circumference of the yolk (Fig. 11.8) (Kuntz 1915). Hatching time varies slightly with temperature: 24 h at 27°C (Kuntz 1915; Jones et al. 1978), 23–24 h at 25°C (Peebles 2002), and 24–27 h at 24–25°C (Fives et al. 1986). At hatching, the larvae are long and slender, weigh 14 μg, are 1.8– 3.75 mm TL, and have a posterior anus; the large, segmented yolk tapers to a point posteriorly (Jones et al. 1978; Tucker 1989). The head is deflected downward over the yolk, the body is flattened and slender, the fin fold is continuous, and no pigmentation is evident (Fig. 11.8) (Kuntz 1915).

Larval Development

Figure 11.8. Embryonic and larval development of the Bay Anchovy, Anchoa mitchilli. (A) Egg with two-celled blastoderm and yolk (X60). (B) Egg with embryo curved around the yolk (X60). (C) Newly hatched larvae, 1.9 mm TL. (D) 12-h post-hatch larvae, 2.7 mm TL. (E) 3.4 mm TL. (F) 5 mm TL. (G) 7.5 mm TL. (H) 10.0 mm TL. (I) Adult fish, 7 cm TL (modified from Kuntz 1915).

Bay Anchovy larvae are abundant in estuaries and adjoining fresh water, and larvae may experience upriver or tidal transport upstream (Dovel 1981; Kimura et al. 2000; Schultz et al. 2003). They also are found at all sizes offshore (MacGregor & Houde 1996). Larvae might demonstrate diel behavior, being more abundant near the surface at night (Schultz et al. 2003). Larvae must develop rapidly due to the small amount of yolk available (Table 11.4; Fig. 11.8). The yolk is absorbed completely in the first 27–52 h after hatching, occurring faster at higher temperatures (Houde 1974; Tucker 1989). Larvae are vulnerable to food shortages during the second and third days after hatching (Kuntz 1915; Tucker 1989). Larvae differ from adults in that they are slender; have a terminal mouth; have a short, round maxillary; and are mostly transparent because they lack pigment (Hildebrand & Schroeder 1928). Bay Anchovy do not have positive growth until after the yolk sac is absorbed completely (Tucker 1989) and quickly develop into adults within 2.5–3 months (Table 11.4).

348

FRESHWATER FISHES OF NORTH AMERICA

Table 11.4. Bay Anchovy, Anchoa mitchilli, larval and juvenile development by length (TL, unless otherwise stated) and age post-hatch (h, hours). Larval Length/Age 27–32 h 27–52 h 2.6–2.8 mm 2.7 mm, 36 h 2.9 mm 44 h 102–122 h 3.7 mm 5 mm 7.5 mm 7–8 mm 11–12 mm 15.5 mm SL 19.5 mm

20–25 mm 22.5 mm SL 34–40 mm 43 mm 60 mm

Development

References

Eye pigmentation Yolk completely absorbed Head no longer deflected down; yolk elongated along body Mouth terminal and apparently functional; anlage of pectoral present; first daily otolith ring created Fin fold somewhat constricted in caudal region First feeding Starvation (with no food provided) Incipient rays in caudal fin; few chromatophores between anal and caudal fin and along midline below gut Incipient rays in caudal and anal fins; muscular rings developed along hindgut; intestine convoluted Urostyle oblique Some with full dorsal and anal fin counts; chromatophores along ventral aspect of thoracic region and at base of caudal fin Full dorsal and anal fin counts Metamorphosis into juveniles begins Chromatophores between operculum and pelvic fins, anal to caudal fins, midlateral row on dorsolateral surface, dark blotch between eyes on top of head, and caudal fin heavily pigmented Projecting snout; body depth increases from 9 times in body depth to 5.5 times Metamorphosis into juveniles essentially complete Minimum length at maturity (2.5 months) Row of chromatophores along anal base and onto caudal fin, a few on the head Adult characteristics present

Leak & Houde 1987; Tucker 1989 Leak & Houde 1987; Tucker 1989 Kuntz 1915 Kuntz 1915; Jones et al. 1978; Leak & Houde 1987 Jones et al. 1978 Leak & Houde 1987 Leak & Houde 1987; Tucker 1989 Jones et al. 1978

Mating System Bay Anchovy are iteroparous broadcast spawners with high fecundities. Spawning energy is derived from daily feeding, not energy stores (Luo & Musick 1991; Wang & Houde 1994). Peebles et al. (1996) described Bay Anchovy in Tampa Bay, Florida, as an income breeder in which seasonal and spatial patterns in egg production can be explained by adult metabolic rate and ration. This fish produces daily cohorts of eggs and larvae in great abundances that in turn experience rapid growth along with high mortality (MacGregor & Houde 1996). Bay Anchovy are therefore strongly r-selected and well adapted to persist in the dynamic estuarine habitats in which they reside (Lapolla 2001a). They also are described as opportunistic life history strategists because of their early maturation, batch spawning, rapid larval growth, and rapid population turnover (Rose et al. 1999).

Kuntz 1915; Jones et al. 1978 Jones et al. 1978 Kuntz 1915; Jones et al. 1978 Jones et al. 1978 Jones et al. 1978 Jones et al. 1978

Jones et al. 1978; Morton 1989 Jones et al. 1978 Jones et al. 1978 Jones et al. 1978 Houde 1974

ECOL OGY

Habitat Bay Anchovy occupy diverse coastal habitats of the United States and Mexico primarily because of their high tolerances of varying salinity and temperature. They typically live in shallow water (90% of total annual production for an entire population. One estimate of annual production of upper and middle Chesapeake Bay was >211 million kg (Wang & Houde 1995).

Importance as Predators Bay Anchovy feed on a wide range of prey items (Fig. 11.9), but eat primarily zooplankton. Because Bay Anchovy are usually abundant, they can impact their prey populations. Short-term reductions in zooplankton are correlated with high densities of this species (Johnson et al. 1990), and young-of-the-year alone can consume ≤50% of the zooplankton standing stock during peak biomass (Cowan et al. 1999). During the summer, juvenile Bay Anchovy may consume 60% of their body weight/day (Luo & Brandt 1993), and adult fish consume ≤25% (Hartman et al. 2004). During autumn and winter when prey availability decreases, consumption also sharply decreases for both young-of-the-year and adult fish. Bay Anchovy also can af-

fect the life history patterns of crabs by feeding on their larvae (Morgan 1990). In response, crab larvae can develop defensive morphologies to avoid being eaten or disperse to predator-free areas.

Importance as Prey This fish serves as a critical trophic link in many food webs due to its ability to transfer so much planktonic energy into available energy for other predators (Johnson et al. 1990; Luo & Brandt 1993; Newberger & Houde 1995; Dorsey et al. 1996; Scharf et al. 2002). Many economically important predatory fishes use Bay Anchovy as their primary forage, especially during times of peak abundances (Morton 1989; Juanes et al. 1993; Hartman & Brandt 1995; Buckel et al. 1999; Scharf et al. 2002). In Chesapeake Bay, Bay Anchovy provided 70% (summer), 90% (autumn), and 60% (spring) of the total energy intake of the truly carnivorous fishes (e.g., Pomatomus saltatrix, Bluefish; Cynoscion regalis, Weakfish; Paralichthys dentatus, Summer Flounder; Morone saxatilis, Striped Bass). These predatory fishes consumed 66% (summer), 80% (autumn), 5% (winter), and 54% (spring) of Bay Anchovy secondary production (Baird & Ulanowicz 1989). Bay Anchovy also composed >40% (by weight) of the diet of 100,000 (9)

1,400–2,800 (1)

829–3,602 (12)

144–1,200 (13)

Egg deposition sites

Aquatic vegetation (6)

Large gravel (4)

Not observed

Adhesive eggs deposited on vegetation and debris (9)

Small depressions in sand or gravel (4)

Adhesive demersal eggs on gravel, vegetation (12)

Adhesive eggs in cavities under flattened cobble (13, 14)

Clutch size

Unknown

N/A

Unknown

N/A

200–1,200 (8)

Unknown

9–260 (13, 14)

Range of nesting and spawning dates and temperatures

April–July (6); 11– 15°C (11)

Late April–early June; 19.4°C (4)

June–August; 18– 25°C (13)

February–July; 14– 19°C (9)

June–early July (10); 11– 23°C (4)

May–August; 14.5–18°C (10)

Late March–early June; 16–21°C (13). September (23°C) also reported (14)

Habitat of spawning sites; average water depth

Vegetated areas of still or slow water (6)

Stream runs; 0.3–1 m deep (4)

Unknown

Flooded vegetation (9)

Shallow areas of fast, flowing water (4)

Riffles; 0.1 m (10)

Flowing water; 6–21 cm deep (13)

Incubation period

Unknown

Unknown

Unknown

3–7 days (9)

7–10 days at 15.6°C (8)

3 days at 21°C, 15 days at 12°C (12)

6 days at 21°C (13)

Mean size at hatching

Unknown

Unknown

Unknown

Unknown

4.5–5.9 mm TL (2)

5.3 mm TL (12)

Unknown

Parental care

None

None

None

None

Eggs defended by at least one parent (10)

None

Males defend eggs from predators (14)

Major dietary items

Small invertebrates (6)

Benthic insect larvae and other benthic inverts (4)

Insects (10), small fishes, and the occasional young rodent (8)

Detritus, benthic invertebrates (9)

Aquatic insects (10)

Aquatic invertebrates (10), fish eggs, and small fishes (12)

Benthic aquatic insects (13)

General year-round habitat

Still to slowflowing water (6)

Clear, moderategradient streams (4)

Muddy, turbid, large rivers (10)

Sloughs, lakes, rivers (9)

Fast-flowing streams, lakeshores (10)

Streams and lakes (7, 12)

High- to moderategradient large streams (13)

Migratory or diadromous

None

None

Small streams to spawn (10)

Upstream to spawn (9)

None

Migrate to streams to spawn (12)

None

Imperilment status; number of species

Endangered; one Vulnerable; two (3)

None

None

Vulnerable; one (3)

Vulnerable; two Various populations and subspecies of R. cataractae and R. osculus considered Endangered, Threatened, and Vulnerable (3)

None

Threatened; one (3)

Table 12.4A. Life history data for type species of Agosia, Alburnops, Algansea, Aztecula, Codoma, Cyprinella, Dionda, and Ericymba. Parenthetical numbers refer to 1Barbour & Miller 1978; 2 Becker 1983; 3Cross 1967; 4Etnier & Starnes 1993; 5Fuiman et al. 1983; 6Gale 1986; 7Gibson & Fries 2005; 8Jelks et al. 2008; 9Jenkins & Burkhead 1994; 10Miller et al. 2005; 11Minckley & Barber 1971; 12Minckley & Vives 1990; 13Phillips et al. 2009; 14Sublette et al. 1990; 15Trautman 1981; 16Vives 1993; 17 Wayne 1979; 18Wayne & Whiteside 1985. Life History Traits

Agosia chrysogaster

Alburnops blennius

Algansea tincella

Aztecula sallaei

Codoma ornata

Cyprinella lutrensis

Dionda episcopa

Ericymba buccata

Clade

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Number of extant species

Two

Twenty

Seven

Two

One

Thirty

Six

Two

1 or 2 degree freshwater

1

1

1

1

1

1

1

1

Maximum size recorded in length

65 mm SL (14)

132 mm TL (2)

175 mm SL (1)

88 mm SL (10)

58 mm SL (10)

90 mm TL (4)

64 mm SL (14)

97 mm TL (15)

Maximum age

3 years (14)

4 years (2)

Unknown

Unknown

Unknown

3 years (4)

Unknown

4 (9)

Age and size at first reproduction

1 year; size not given (14)

Males, 1 year, 46–64 mm TL; females, 2 years, 70–84 mm TL (2)

Unknown

Unknown

Unknown

2 years (2); 30 mm SL (4)

Presumably age 0–1; 25 mm SL (7)

1 (9)

Iteroparous or semelparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Fecundity estimates (ovarian counts)

Unknown

1,895–2,840 (2)

Unknown

Unknown

Unknown

485–684 (2)

2–584, based on D. nigrotaeniata (17, 18)

150–1,350 (9)

Egg deposition sites

Saucer-shaped depressions in sand (11)

Over sand and gravel bars (15)

Unknown

Unknown

Adhesive eggs deposited in crevices, under, or between rocks (10, 12)

Adhesive eggs deposited along margins of Sunfish nests, over gravel, or in crevices (2, 3, 16)

Non-adhesive demersal eggs on gravel, based on D. diaboli (13, 17, 18)

Sand and gravel (9)

Clutch size

Unknown

N/A

Unknown

Unknown

85–115 (12)

585 (6)

N/A

N/A

Range of nesting and spawning dates and temperatures

Unknown

June–August (2)

May–July (10)

February–May (10)

March–October

May–August, depending on location (2); 15.6–29.4°C (3)

Spring months; >17–18°C (14)

March–May 10–20°C (9)

Habitat of spawning sites; average water depth

Sandy substrate and slight current; 5–20 cm deep (11)

Large rivers (2)

Unknown

Unknown

Clear, flowing pools; between 2.5 and 15 cm deep (12)

Highly variable

Springs (14)

Unknown

Incubation period

4 days at >24°C (11)

Unknown

Unknown

Unknown

Unknown

Unknown

Unknown

Unknown

Mean size at hatching

Unknown

Unknown

Unknown

Unknown

Unknown

4–5 mm TL (5)

7 mm TL, based on D. diaboli (7)

Unknown

Parental care

None (11)

None

Unknown

Unknown

Males defend eggs (12)

None

None

Unknown

Major dietary items

Detritus, zooplankton, aquatic insects, filamentous algae (14)

Aquatic insects (2)

Unknown

Unknown

Unknown

Plant material, aquatic insect larvae, microinvertebrates (2)

Algae, diatoms (14)

Small, benthic invertebrates (9)

General year-round habitat

Small, clear streams (11)

Large rivers in water 1–3 m deep (2)

Small streams to large lakes (1)

Spring-fed ponds, lakes, streams (10)

Pools and riffles of clear streams and rivers (10)

Turbid and silty streams, rivers, and lakes (2)

Vegetated lowgradient rivers and streams (14)

Creeks and rivers (9)

Migratory or diadromous

None

None

None

Unknown

None

None

None

None

Imperilment status; number of species

Vulnerable; one (8)

Vulnerable; two Threatened; one (8)

Endangered; four Vulnerable; two (8)

Vulnerable; two (8)

Vulnerable; one (8)

Endangered; seven Threatened; two Vulnerable; three (8)

Endangered; four (8)

None

Table 12.4B. Life history data for type species of Erimonax, Graodus, Hudsonius, Hybognathus, Hybopsis, Luxilus, Lythrurus, and Miniellus. Parenthetical numbers refer to 1Becker 1983; 2 Contreras-MacBeath & Rivas 2007; 3Cross 1967; 4Etnier & Starnes 1993; 5Hunter & Hasler 1965; 6Jelks et al. 2008; 7Jenkins & Burkhead 1994; 8Miller et al. 2005; 9Scott & Crossman 1973; 10Trautman 1981. Life History Traits

Erimonax monachus

Graodus boucardi

Hudsonius hudsonius

Hybognathus nuchalis

Hybopsis amblops

Luxilus chrysocephalus

Lythrurus umbratilis

Miniellus procne

Clade

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Number of extant species

One

Three

Three

Seven

Seven

Nine

Eleven

Four

1 or 2 degree freshwater

1

1

1

1

1

1

1

1

Maximum size recorded in length

92 mm SL (7)

66 mm SL (8)

147 mm TL (10)

152 mm TL (10)

99 mm TL (10)

240 mm TL (10)

67.4 mm SL (1)

65 mm SL (7)

Maximum age

3 years (7)

Unknown

5 years (7)

3 years (1)

Unknown

6 years (7)

3 years (1)

3 years (7)

Age and size at first reproduction

2 years; 53 mm SL (7)

Unknown

1–4 years (1); 55 mm SL (7)

2 years (1); 65 mm TL (1)

1 year (7); 58 mm TL (10)

2 years; 60 mm SL (7)

1 year; 28–32 mm SL (4)

2 years (7); 39 mm SL (7)

Iteroparous or semelparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Fecundity estimates (ovarian counts)

157–791 (7)

Unknown

100–8,898 (7)

2,054–3,105 (1)

Unknown

900–1,150 (7)

219–887 (1, 4)

Unknown

Egg deposition sites

Adhesive eggs in rock crevices or between rocks (7)

Rocky substrate (2)

Sand, gravel, algae (1)

Non-adhesive eggs deposited on vegetation, H. regius (4)

Unknown

Nocomis or Campostoma nests; males may excavate shallow pits (7, 9)

Gravel and sand in stream or green Sunfish nests (5, 9)

Sand, Sunfish and cyprinid nests (7)

Clutch size

Unknown

N/A

N/A

N/A

Unknown

50 (9)

Unknown

Unknown

Range of nesting and spawning dates and temperatures

Mid-May–midAugust; 26.1–27.2°C (7)

January–April (8)

April–early September (7)

April–July (1)

Late spring and early summer (4)

May–June; 16–26.7°C (7)

June–August (1, 9); 21°C or higher (3)

May–July (7); 25.6°C (7)

Habitat of spawning sites; average water depth

Moderate current in shallow stream runs (7)

Shallow water during low-flow conditions (2)

Over sand and algae in shallow riffles; lakes up to 4.6 m (1, 7, 9)

Unknown

Unknown

Streams over gravel substrate (7, 9)

Green Sunfish nests; 4–35 cm (1)

In moderate current (7)

Incubation period

Unknown

Unknown

Unknown

Unknown

Unknown

Unknown

Unknown

Unknown

Mean size at hatching

Unknown

Unknown

Unknown

Unknown

Unknown

6.9 mm TL (1)

Unknown

Unknown

Parental care

Unknown

None (2)

None

None

Unknown

None

None

None

Major dietary items

Benthic aquatic insect larvae, 99% immature midge and blackfly larvae (7)

Small, benthic, aquatic insects (2)

Larval insects, algae, microcrustaceans, mollusks, small fishes (7)

Algae, blue-green algae, diatoms (1)

Microcrustaceans and midge larvae (7)

Aquatic and terrestrial insects

Aquatic and terrestrial insects, filamentous algae (1)

Aquatic invertebrates, filamentous algae (7)

General year-round habitat

Warm, clear, medium-size streams of moderate gradient (7)

Shallow streams with rocky substrate (2)

Large lakes and rivers (1)

Silty creeks (4) and medium to large rivers in clear waters (1)

Clear creeks, streams, and rivers (7)

Creeks, streams, and rivers

Weedy or turbid pools of lowgradient streams (1, 9)

Pools of large creeks and rivers (7)

Migratory or diadromous

None

Unknown

None

None

None

None

None

Unknown

Imperilment status; number of species

Threatened; one (6)

Threatened; two (6)

None

Endangered; one Vulnerable; two (6)

Vulnerable; two (6)

None

Vulnerable; one (6)

Endangered; one (6)

Table 12.4C. Life history data for type species of Notropis, Opsopoeodus, Pimephales, Pteronotropis, Tampichthys, and Yuriria. Parenthetical numbers refer to 1Ankley et al. 2001; 2 Becker 1983; 3Boschung & Mayden 2004; 4Etnier & Starnes 1993; 5Jelks et al. 2008; 6Jenkins & Burkhead 1994; 7Markus 1934; 8Mayden & Simons 2002; 9Miller et al. 2005; 10 Page & Johnston 1990a; 11Roberge et al. 2002; 12Scott & Crossman 1973. Life History Traits

Notropis atherinoides

Opsopoeodus emiliae

Pimephales promelas

Pteronotropis hypselopterus

Tampichthys rasconis

Yuriria alta

Clade

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Shiner (OPM)

Number of extant species

Fifty-six

One

Four

Ten

Six

One

1 or 2 degree freshwater

1

1

1

1

1

1

Maximum size recorded in length

97 mm TL (2)

66 mm TL (4)

102 mm TL (6)

58 mm SL (3)

53 mm SL (9)

200 mm SL (9)

Maximum age

4 years (6)

2 years (2)

3 years (11)

Unknown

Unknown

Unknown

Age and size at first reproduction

2 years; 46 mm SL (6)

1 year; 48 mm TL (2)

50 mm TL (1)

Unknown

2 years; size not given (13)

2 years; 45 mm SL (8)

Probably 2–4+ years; 120 mm SL (8)

Iteroparous or semelparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Fecundity estimates (ovarian counts)

6,200 (15)

230–862 (1)

Unknown

600–45,125 (13)

250–2,000 (7)

Unknown

Egg deposition sites

Demersal, adhesive eggs on gravel, cobble, boulders (7, 15)

Gravel substrate, Nocomis and Semotilus nests (1, 4)

Unknown

Adhesive eggs on gravel or rocks (13)

Adhesive eggs deposited in rock crevices (8)

Adhesive eggs deposited on shallow gravel and rock (8)

Clutch size

N/A

N/A

Unknown

N/A

N/A

N/A

Range of nesting and spawning dates and temperatures

Late May–July; 13–18°C (9, 15)

March–July, depending on location (1, 3)

Unknown

June–July when temperatures reach 18.3°C (12)

March–July when temperatures reach 16°C (8)

May–August; 15–18°C (8)

Habitat of spawning sites; average water depth

Streams and lakes (7)

Streams with gravel or pebbles (1)

Unknown

Streams in water; < 9 m deep (12)

Shallow areas of streams (8)

Lakes and rivers; 0.3–0.5 m (8)

Incubation period

16 days at 12°C, 6 days at 18°C (7)

Unknown

Unknown

Unknown

2–3 days (8)

Unknown

Mean size at hatching

8 mm TL (7)

Unknown

Unknown

Unknown

Unknown

Unknown

Parental care

None

None

Unknown

None

None

None

Major dietary items

Diatoms, algae (11)

Algae, diatoms (1)

Algae and invertebrates (14)

Aquatic insects, macroinvertebrates, and fishes (2)

Filamentous algae, aquatic insects, and crustaceans (8)

Aquatic and terrestrial insects (8)

General year-round habitat

Large streams, lakes (7)

Small, clear streams (1)

Hot springs, between 30.6 and 33.9°C (6)

Pools and rapids of medium to large rivers (13)

Small streams and rivers (8)

Small streams, rivers, and lakes (8)

Migratory or diadromous

None

None

None

None

None

None

Imperiled status; number of species

None

Endangered; one Threatened; one Vulnerable; one (5)

Threatened; one (5)

Endangered; seven Threatened; four Vulnerable; four (5)

Vulnerable; one (5)

None

Table 12.5B. Life history data for type species of Lavinia, Mylopharodon, Orthodon, Ptychocheilus, Relictus, and Siphateles. Parenthetical numbers refer to 1Hankins 1995; 2Hubbs et al. 1974; 3Jelks et al. 2008; 4McPhail 2007; 5Moyle 2002; 6Roberge et al. 2002; 7Scoppettone 1988; 8Scott & Crossman 1973. Life History Traits

Lavinia exilicauda

Mylopharodon conocephalus

Orthodon macrolepidotus

Ptychocheilus oregonensis

Relictus solitarius

Siphateles bicolor

Clade

Western

Western

Western

Western

Western

Western

Number of extant species

One

One

One

Four

One

Three

1 or 2 degree freshwater

1

1

1

1

1

1

Maximum size recorded in length

350 mm SL (5)

600 mm SL (5)

500 mm FL (5)

>400 mm FL (4); >1.8 m in P. lucius (5)

99 mm SL (2)

420 mm SL (5)

Maximum age

6 years (5)

10 years (5)

>5 years (5)

20 years (6)

Unknown

32 years (7)

Age and size at first reproduction

1 year (5); 49 mm SL (5)

Unknown

2 years (5); 170 mm (5)

4 years; 30 cm (6)

Unknown

2–7 years (5)

Iteroparous or semelparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Iteroparous

Fecundity estimates (ovarian counts)

3,000–63,000 (5)

7,000–24,000 (5)

14,700–346,500 (5)

5,000–95,000 (4)

Unknown

11,200–25,000 (5)

Egg deposition sites

Non-adhesive, demersal eggs deposited over gravel (5)

Not observed

Adhesive eggs deposited on vegetation (5)

Adhesive, demersal eggs on gravel and cobble (4, 8)

Aquatic vegetation (2)

Vegetation, algae-covered substrate, sand (5)

Clutch size

N/A

Unknown

N/A

N/A

N/A

N/A

Range of nesting and spawning dates and temperatures

14–26°C (5)

Unknown

March–July; 12–14°C (5)

May–July; 12–18°C (6)

June– September (2)

13–17°C (5)

Habitat of spawning sites; average water depth

Riffles of streams with gravel substrate (5)

Presumed spawning individuals observed in stream riffles (5)

Shallow areas with aquatic plants (5)

Streams, lakeshore (6)

Springs

Vegetation, algae-covered substrate (5); < 1.5 m (5)

Incubation period

3–7 days at 15–22°C (5)

Unknown

Unknown

6 days at 18°C (4)

Unknown

3–6 days (5)

Mean size at hatching

Unknown

Unknown

Unknown

8 mm TL (4)

Unknown

Unknown

Parental care

None

Unknown

None

None

Unknown

None

Major dietary items

Algae, zooplankton, aquatic and terrestrial insects (5)

Benthic invertebrates, aquatic plants, zooplankton (5)

Phytoplankton and zooplankton (5)

Crayfish, fish (6)

Insects (1)

Detritus, plant fragments, zooplankton, macroinvertebrates (5)

General year-round habitat

Low-gradient creeks, slow-moving rivers, lakes (5)

Large streams and rivers (5)

Slow-moving, turbid rivers and lakes (5)

Near shore areas of lakes and large rivers (6)

Clear springs (2)

Springs, slow-moving streams, rivers, and lakes (5)

Migratory or diadromous

To streams to spawn (5)

To streams to spawn (5)

None

Upstream to spawn in tributary rivers (6)

None

None

Imperiled status; number of species

Vulnerable; one (3)

None

None

Endangered; one (3)

Vulnerable; one (3)

Endangered; one Threatened; one Various subspecies and populations considered Endangered, Threatened, and Vulnerable (3)

Plate 12.1. Emerald Shiner, Notropis atherinoides

Plate 12.2. Alabama Shiner, Cyprinella callistia

Plate 12.3. Weed Shiner, Alburnops texanus

Plate 12.4. Humpback Chub, Gila cypha 378

CYPRINIDAE: CARPS AND MINNOWS

Notropis is a taxonomic repository for small, silvery fishes of unknown relationship and thus is polyphyletic. This genus may ultimately be reduced to many fewer species once phylogenetic relationships of the shiners are refined. Although absent from Pacific Ocean drainages and the northern Canadian Shield, species of Notropis occur nearly everywhere else on the continent, including the endorheic Ríos Nazas-Aguanaval basin, Mexico, and Gulf Coast drainages as far south as the Rio Pánuco drainage, Mexico (Miller et al. 2005). One of the most widely distributed members of the genus and the type species, the Emerald Shiner, Notropis atherinoides, is found from the Mackenzie River in northern Canada, which flows into the Arctic Ocean, in the St. Lawrence River, the Hudson River, and the Mississippi River drainage south to the Gulf of Mexico, and in Gulf Slope drainages from Mobile Basin, Alabama, to Galveston Bay, Texas (Page & Burr 2011). The Emerald Shiner is the northernmost occuring species and the Pygmy Shiner, “Notropis” tropicus, is the southernmost species, occuring in the Rio Pánuco system, Mexico.

Genus Cyprinella The genus Cyprinella contains 30 species (Schönhuth & Mayden 2010). These striking and often colorful fishes display fascinating courtship and mating behaviors (see reproduction section). The southeastern United States is the center of diversity for Cyprinella. Species diversity is well understood in this genus, although four widespread taxa (Red Shiner, Cyprinella lutrensis; Spotfin Shiner, Cyprinella spiloptera; Blacktail Shiner, Cyprinella venusta; and Steelcolor Shiner, Cyprinella whipplei) exhibit morphological or genetic variation that suggests cryptic species may be subsumed in these taxa (Schönhuth & Mayden 2010). Species of Cyprinella occur throughout the Mississippi River basin and surrounding rivers ranging from the Red River of the North, a Hudson Bay drainage in southern Manitoba, Canada; in Atlantic Slope drainages from the St. Lawrence River to the Altamaha River, Georgia; and in Gulf Coast drainages from the Suwannee River, Florida, to the Rio Pánuco, Mexico (Miller et al. 2005; Page & Burr 2011; Fig. 12.2). Species of Cyprinella occur in the endorheic Ríos Nazas-Aguanaval basin and the Rio Yaqui, which drains into the Pacific Ocean (Miller et al. 2005). The Spotfin Shiner is the northernmost occurring species and one of the most widespread, inhabiting the Red River of the North, the St. Lawrence River, and drainages of the Great Lakes (except Lake Superior)

379

(Page & Burr 2011). It is also widespread in eastern tributaries of the Mississippi River and present to the west in isolated populations in the Ozark and Ouachita Mountains. The southernmost member, the Red Shiner, also a widely distributed species, ranges from Missouri River tributaries, South Dakota, south to eastern Mexico where it occurs in the Ríos Tamesí and Pánuco (Miller et al. 2005; Page & Burr 2011).

Genus Alburnops The genus Alburnops contains 20 species that were formerly included in the genus Notropis (Mayden et al. 2006). Much of the species diversity of Alburnops is located in the southeastern United States, particularly in the Mobile Basin (Fig. 12.4). Alburnops contains at least one undescribed species, and there may be additional undescribed species in the genus. Some species, such as the Rainbow Shiner, Alburnops chrosomus, are brilliantly colored when spawning. Species of Alburnops occur in Atlantic and Gulf Slope drainages from the Hudson River, New York, almost to the southern tip of Florida and west to the Nueces River, Texas; in the Mississippi River north to Minnesota, and in the Hudson Bay drainages, Red River of the North, Minnesota and Manitoba, and Assiniboine River, Alberta, Saskatchewan, and Mantitoba (Page & Burr 2011). Most species in this group have relatively restricted ranges, although three have large ranges. The River Shiner, Alburnops blennius, is widespread, occuring in the Hudson Bay drainages, Red River of the North, the Assiniboine River system, and the Mississippi River system south to the Gulf of Mexico. The Ironcolor Shiner, Alburnops chalybaeus, occurs in lowlands of the Atlantic Slope and east Gulf Slope from the Hudson River, New York, to southern Florida and west to the Sabine River, Louisiana and Texas, western drainages of the Mississippi embayment, with isolated populations in the San Marcos River, Texas, the Illinois River system, Illinois and Indiana, the Wisconsin River and Lake Winnebago system, Wisconsin, and Lake Michigan drainages, Michigan and Indiana. The Weed Shiner, Alburnops texanus, occurs in Gulf Coast drainages from the Suwannee River, Florida and Georgia, to the Nueces River, Texas; in western tributaries of the Mississippi River in the Mississippi Embayment; in the upper Mississippi River system in Minnesota, Iowa, Wisconsin, Illinois, Indiana, and Michigan; and in the Hudson Bay drainage in the Red River of the North, Minnesota (Lee et al. 1980; Page & Burr 2011).

380

FRESHWATER FISHES OF NORTH AMERICA

Genus Gila The genus Gila contains 18 species and has much of its diversity in Pacific Slope drainages of western North America, although a few taxa inhabit Gulf of Mexico drainages (Fig. 12.3). The Colorado River system contains the highest number of Gila species, including the Humpback Chub, Gila cypha, and Bonytail Chub, Gila elegans (Fig. 12.56e). Colorado River species get quite large (>300 mm SL) (Lee et al. 1980) and can live >40 years (Minckley 1991a; Table 12.5a), clearly an advantage in desert rivers that exhibit unpredictable annual flow patterns. Species of Gila occur from the upper Missouri River basin, Montana, where the Utah Chub, Gila atraria, has been introduced; the upper Snake River; the Sacramento–San Joaquin River system, California; the Lake Bonneville system, Utah; the Colorado River

system, and a number of smaller Pacific Coast drainages of southern California and northern Mexico (Fig. 12.3). Species of Gila also occur in the upper Rio Grande, New Mexico, and Rio Conchos, Mexico. The Utah Chub is the northernmost species native to the upper Snake River, Idaho and Wyoming; the Lake Bonneville Basin, Utah; and the Sevier River system, Idaho and Utah; and has been introduced to the upper Missouri system, Montana (Lee et al. 1980; Page & Burr 2011). The southernmost species is the Nazas Chub, Gila conspersa, which is restricted to the endorheic Ríos Nazas-Aguanaval basin, Mexico (Miller et al. 2005).

Genus Lythrurus The genus Lythrurus contains 11 species, although additional cryptic taxa may be present in this group (Pramuk

Figure 12.56. (A) Spotfin Chub, Erimonax monachus, male (top) and female (bottom) (photograph by and used with permission of N. M. Burkhead). (B) Bluehead Chub, Nocomis leptocephalus, constructing nest (photograph by and used with permission of W. N. Roston). (C) Central Stoneroller, Campostoma anomalum, school of tuberculate males (photograph by and used with permission of W. N. Roston). (D) Rosyside Dace, Clinostomus funduloides, school of adults in spawning color (photograph by and used with permission of W. N. Roston). (E) Bonytail Chub, Gila elegans, Dexter Fish Hatchery, New Mexico (photograph by and used with permission of B. M. Burr). (F) Bluehead Shiner, Pteronotropis hubbsi, breeding secondary male (photograph by and used with permission of W. N. Roston). (G) Blackside Dace, Chrosomus cumberlandensis, spawning over nest of Creek Chub, Semotilus atromaculatus (photograph by and used with permission of R. R. Cicerello). (H) Bleeding Shiner, Luxilus zonatus, school of adults in spawning color (photograph by and used with permission of W. N. Roston).

CYPRINIDAE: CARPS AND MINNOWS

381

Plate 12.5. Scarlet Shiner, Lythrurus fasciolaris

Plate 12.6. Speckled Chub, Macrhybopsis aestivalis

et al. 2007). Species of Lythrurus are widely distributed throughout much of the Mississippi River basin and in Gulf Coast drainages from the Navidad River, Texas, east to the Chattahoochee River, Alabama and Georgia (Lee et al. 1980; Pramuk et al. 2007; Fig. 12.5). The Redfin Shiner, Lythrurus umbratilis, is the most widespread species in the genus, occurring in the Mississippi and Ohio River basins, southern Great Lakes drainages, and Gulf Coast drainages west to the San Jacinto River, Texas (Page & Burr 2011). Other members occupy much smaller ranges, such as the Ouachita Mountain Shiner, Lythrurus snelsoni, which is restricted to tributaries of the Little River, Arkansas and Oklahoma, above the Fall Line (Taylor & Lienesch 1996; Page & Burr 2011).

Genus Macrhybopsis The genus Macrhybopsis contains eight described and three undescribed species. These fishes inhabit large, of-

ten turbid rivers across the eastern United States. Often found over sandy substrates, these fishes have relatively small eyes, elongate maxillary barbels, and many tastebuds across the surface of the body. Species of Macrhybospsis occur in Hudson Bay drainages from the Assiniboine and Red Rivers of the North in Manitoba south to Minnesota; in several rivers draining into Lake Erie; in suitable habitat throughout the Mississippi River system from Minnesota south to the Gulf of Mexico, and from the upper Ohio River, New York and Pennsylvania, west to the Republican River of western Nebraska; and in Gulf Coast drainages from the Mobile Basin west to the Rio Grande (Lee et al. 1980; Eisenhour 2004; Fig. 12.6). The Silver Chub, Macrhybopsis storeriana, is the most widespread species in the genus, ranging from Hudson Bay drainages in the North to the lower Mississippi River basin and Mobile Basin in the south and from the Kansas River drainage in southern Nebraska to the upper Ohio River drainage, New York. Three species, in the Mobile Basin and

382 FRESHWATER FISHES OF NORTH AMERICA

other Gulf Coast drainages, Alabama and Florida, are undescribed and additional cryptic species may exist in this genus (Eisenhour 2004).

Genus Pteronotropis The genus Pteronotropis contains 10 species (Suttkus & Mettee 2001; Mayden et al. 2006; Page & Burr 2011). With the exception of the Redeye Chub, Pteronotopis harperi, male Pteronotropis are colorful with enlarged dorsal and anal fins (Fig. 12.56f). Species of Pteronotropis occur in swamps and streams of the Atlantic and Gulf Coasts and in the Mississippi Embayment from the Red and Ouachita River drainages of Arkansas and Texas in the west, along the lower reaches of Gulf Coast drainages, much of Florida and north along the Atlantic Slope to the

Pee Dee River, South Carolina (Lee et al. 1980; Boschung & Mayden 2004; Page & Burr 2011; Fig 12.7). An isolated population of the Bluehead Shiner, Pteronotropis hubbsi, was present in Wolf Lake, southwest Illinois, but is now extirpated (Burr & Warren 1986a).

Genus Luxilus The genus Luxilus contains nine species distributed across much of the eastern United States (Fig. 12.8). Two widespread species occupy most of the range of this genus. The Common Shiner, Luxilus cornutus, occurs from southern Canada south to central Nebraska in the west and southern Virginia in the east, and the Striped Shiner, Luxilus chrysocephalus, occurs in southern Great Lakes drainages, the Mississippi River basin, south to the Gulf Coast and in

Plate 12.7. Aplachee Shiner, Pteronotropis grandipinnis

Plate 12.8. Striped Shiner, Luxilus chrysocephalus

CYPRINIDAE: CARPS AND MINNOWS

Gulf Coast drainages from the Mobile Basin west to Lake Ponchartrain (Lee et al. 1980; Page & Burr 2011). The ranges of these two species overlap in some places and hybridization is reported (see genetics section).

Cyprinids as Non-Natives Cyprinids are popular in the aquarium trade, and a number are important fishes in aquaculture. As a result, a number of non-indigenous species have been released in North America, although few have become established. Unfortunately, the few that are established have either caused or have the potential to cause substantial damage to aquatic ecosystems. In addition to exotic species, a large number of North American species have been introduced (transplanted) outside their native range in North America. At least 104 exotic and native cyprinid species may have been introduced outside their native range, although it often is unclear if the actual introductions occurred and in other cases little or no environmental impact is identified (Fuller et al. 1999). Here, we focus on cyprinid species that are widely established outside their native range or that have the potential to negatively impact native species and the systems they inhabit (see genetics section). The spread of non-native cyprinids occurs via a number of mechanisms. Several species were stocked intentionally. Stocking may be used to develop a fishery (e.g., Common Carp, Cyprinus carpio) or to provide forage for gamefishes (e.g., Red Shiner, Cyprinella lutrensis, and Fathead Minnow, Pimephales promelas) (Fuller et al. 1999). Other species are stocked in restricted areas such as aquaculture facilities and subsequently escape; examples include the recent escapes of species of three genera of large Asian carps: Hypopthalmichthys, Ctenopharyngodon, and Mylopharyngodon (Schofield et al. 2005). Baitbucket introductions, fishes used as bait and deliberately or accidentally released, account for the spread of many native fishes outside of their native range, particularly the Golden Shiner, Notemigonus crysoleucas (Fig. 12.40), and the Fathead Minnow, both favored as baitfishes (Litvak & Mandrak 1993). There is a baseless perception among some anglers that releasing unused baitfish benefits the environment (Litvak & Mandrak 1993). Cyprinids may also be spread as stock contaminants with fish that are stocked intentionally (e.g., Fathead Minnow) (Woodling 1985). Once released, non-native cyprinids may disperse throughout a river system and in some cases become so common that they are considered native. For example, the Common Carp, a native of Eurasia, is sometimes referred

383

to as the native carp in parts of the Midwest. Competition or predation may restrict the spread of cyprinids. For example, multiple predators (Sculpins, Cottidae; Sacramento Pikeminnow, Ptychocheilus grandis, also introduced) restrict the range and distribution of introduced Speckled Dace, Rhinichthys osculus, in the Eel River, California (Harvey et al. 2004). Possibly the first exotic fish species to be released in North America was the ornamental Goldfish, Carrassius auratus, originally native to Asia. These apparently were first introduced in the late 1600s (Dekay 1842), and repeated introductions have occurred since then. Although Goldfish are captured widely in North America, locations for established reproducing populations are scattered and localized (Schofield et al. 2005). They appear to establish in highly degraded habitats (Fuller et al. 1999), but continuing introductions complicate the picture. The Common Carp was introduced into North America in the late 1800s. This species was stocked intentionally and widely and is established throughout much of southern Canada (Scott & Crossman 1973), the United States (Schofield et al. 2005), and Mexico (Miller et al. 2005). Common Carp cause a tremendous amount of damage to the environment, especially when feeding. Their rooting feeding mode disrupts the sediment and uproots aquatic plants, often significantly impacting water quality (Roberts et al. 1995; Titus et al. 2004; Miller & Crowl 2006). Two species of Hypopthalmichthys have been introduced into North America: Silver Carp, Hypopthalmichthys molitrix, and Bighead Carp, Hypopthalmichthys nobilis. These large Asian carps were introduced by the aquaculture industry to improve water quality in Catfish ponds and to provide an additional aquaculture crop (Fuller et al. 1999; Schofield et al. 2005). Both species feed on plankton and have the potential to compete with large native planktivores (e.g., Paddlefish, Polyodon spathula), and they may have the ability to alter plankton communities with negative impacts on invertebrates and other fishes (Sampson et al. 2009). Both are established in the Mississippi River and its larger tributaries such as the Missouri and Ohio Rivers and are steadily increasing their distribution. Significantly, they may invade other river systems, including the St. Lawrence River system and the Mobile Basin via navigation canals, causing substantial ecological disruption in the process. Silver Carp are also well known for their habit of leaping clear of the water when startled and have injured boaters. Black Carp, Mylopharnyngodon piceus, another native of Asia, were introduced by the aquaculture industry to

384 FRESHWATER FISHES OF NORTH AMERICA

control snails infested with parasites in Catfish ponds. Black Carp feed on invertebrates, particularly mollusks, and thus can reduce the numbers of snails that serve as the intermediate host for a trematode that parasitizes Catfish (Schofield et al. 2005). Black Carp may have established reproductive populations in the lower Mississippi River and are considered a major threat not only to endangered freshwater mussels and snails but to the entire mollusk fauna in the Mississippi River system (Nico et al. 2005). Grass Carp, Ctenopharyngodon idella, yet another Asian native, were introduced to North America by the aquaculture industry to control macrophytes. These large carp can consume up to 45 kg/day of plant material (Fuller et al. 1999) and have been eagerly cultivated and stocked by state agencies as well as aquaculturists because of their ability to control macrophytes (Mitzner 1978). Grass Carp are likely the source of introduction of the Asian tapeworm, Bothriocephalus opsarichthydis, to North America (Hoffman & Schubert 1984; Ganzhorn et al. 1992). Grass Carp have been stocked widely within the United States but are established only in the Mississippi River basin and in the Trinity River, Texas (Schofield et al. 2005). The Tench, Tinca tinca, a Eurasian native, was introduced to North America by the U.S. Fisheries Commission in 1877. Several subsequent introductions have occurred, and although Tench were widely stocked, they did not become widely established (Schofield et al. 2005). Reproductive populations occur in the Connecticut River system, Massachusetts and Connecticut; parts of the upper Columbia River system, British Columbia, Idaho, and Washington; the upper Colorado River system, Colorado; the Salinas River system, California; and localities around Puget Sound, Washington. The Rudd, Scardinius erythropthalmus, a Eurasian species, has been widely introduced and transported as a baitfish (Litvak & Mandrak 1993). Numerous baitbucket introductions have occurred, particularly in New England and the south-central United States (Schofield et al. 2005). The Rudd is difficult to distinguish from the Golden Shiner, which is native to North America but also widely introduced outside its native range (Fuller et al. 1999), and thus the overall extent of Rudd introduction and establishment is not clear. Established reproductive populations are present in Arkansas, Maine, Massachussetts, Michigan, Missouri, Nebraska, New York, Ontario, Texas, Virginia, and Washington (Litvak & Mandrak 1993; Schofield et al. 2005). The Bitterling, Rhodeus sericeus, a Eurasian species, was introduced into the Sawmill and Bronx Rivers, New

York, sometime before 1923 (Lee et al. 1980). This species, which deposits its eggs in freshwater mussels, persists in the Bronx River. Apparently its numbers are decreasing coincident with decreasing water quality, which is impacting the freshwater mussel population (Schoefield et al. 2005). The Red Shiner, Cyprinella lutrensis, a North American cyprinid native to the central United States and eastern Mexico, has been widely introduced outside its range in the Mobile Basin, Colorado River system, San Joaquin River, California, as well as the Yadkin and Roanoke Rivers (Schoefield et al. 2005). This species preys on larval cyprinids, including the endangered Colorado Pikeminnow, Ptychocheilus lucius (Bestgen 1996). Red Shiners are also thought to compete with other endangered native fishes (Rinne 1991) and act as vectors of the Asian tapeworm that is a threat to endangered cyprinids (Brouder 1999; Choudhury et al. 2004; Pullen et al. 2009). The Fathead Minnow, Pimephales promelas, another North American native, is widely produced for use as a baitfish (see commercial importance section) and has also been stocked outside its native range as a forage fish. The native range is difficult to determine but likely included much of North America east of the Rocky Mountains (Fuller et al. 1999). It has been introduced and is now widespread in drainages along the Pacific Coast from southern British Columbia to California (Lee et al. 1980; Fuller et al. 1999). The Golden Shiner, another North American native, is also widely used as a baitfish (see commercial importance section) and is widely distributed in southeastern Canada and the eastern United States (Lee et al. 1980; Fig. 12.40). This may be its native range, although this is unclear. The Golden Shiner was introduced intentionally into the Colorado and Sacramento River systems as a forage fish for introduced gamefishes (Lee et al. 1980; Sigler & Sigler 1987).

PHYLOGE NE TIC RELATIONSHIPS The phylogenetic relationships of North American cyprinids have been and continue to be confused. The confusion is a function of the large number of species and their apparent morphological conservatism. An historical proliferation of generic and specific names has resulted in complex synonymies. Lack of a coherent philosophy governing the field of systematics and the inability of investigators to deal with large numbers of characters, particularly in such a species-rich group, contributed further to

CYPRINIDAE: CARPS AND MINNOWS

385

Plate 12.9. Golden Shiner, Notemigonus crysoleucas

the confusion for much of the 20th century. Phylogenetic systematics (Hennig 1966) dealt with the first problem; the development of computer algorithms and rapid increases in computing power have helped with the second. These changes, coupled with the recent ability to access large numbers of molecular characters, have dramatically clarified cyprinid relationships. We anticipate that future work, particularly multi-locus, species-tree analyses (see Hollingsworth & Hulsey 2011), will dramatically transform our understanding of North American cyprinid relationships. Two major clades of cyprinids represented in the North American fauna are generally recognized by cyprinid systematists: the leucisins, containing the single species Golden Shiner, Notemigonus crysoleucas, and the phoxinins, containing all remaining species (Cavender & Coburn 1992; Coburn & Cavender 1992). Relationships among phoxinins are being resolved gradually, but several major questions still remain unanswered regarding phylogenetic relationships of North American cyprinids. Notably, what are the relationships among species, and are the currently recognized genera monophyletic? What are the relationships among major clades of North American cyprinids? Are the North American phoxinins a monophyletic group? In the mid-1900s, the North American cyprinid fauna was considered to contain two distinct groups, a western fauna and an eastern fauna. The western fauna contained taxa native to the Pacific Slope, and the eastern fauna contained taxa native to the Mississippi River basin, other Gulf of Mexico drainages, and Atlantic Slope drainages. This view persisted even though several taxa did not conform to this pattern. The genera Rhinichthys and Clinostomus (Fig. 12.56d) were both considered members of the western fauna even though they are distributed widely

east of the Pacific Divide (Figs. 12.51 and 12.15, respectively), and the genus Oregonichthys was considered part of the eastern fauna even though it is native to Pacific Slope drainages (Fig. 12.42). The classification of North American cyprinids reflected the tensions in systematic biology. Classifications represented a compromise between evolutionary history and convenience (Mayr 1953), and unsurprisingly, this led to nomenclatural instability and confusion. The tensions in systematic biology were reflected in the different approaches taken by Carl Hubbs and Reeve Bailey, both major figures in North American cyprinid systematics. Hubbs’s work focused on morphological differences between taxa, but Bailey concentrated on similarities. In the 1950s, Bailey dramatically revised the classification of several groups of North American cyprinids (Bailey 1951, 1956). He synonymized several genera based on a few morphological characters that included the presence and position of a maxillary barbel, scalation, and gut morphology. Several workers responded negatively to these taxonomic changes (Lachner & Jenkins 1967; McPhail & Lindsey 1970; Jenkins & Lachner 1971). Hubbs & Miller (1977:275) stated: “It is abundantly obvious that much of the generic placement in American cyprinids is in a chaotic state, and that the prime significance attributed to intestinal coiling vs. the single compressed-S configuration, and to the presence vs. absence of a maxillary barbel, in the taxonomy of the group has been very considerably discredited.” Nevertheless, Bailey’s changes had a profound impact on classification. Taxonomic changes that he made in two footnotes in identification keys (Bailey 1951, 1956) now require large amounts of data to overturn. The first explicitly phylogenetic analysis of relationships among North American cyprinids was Mayden’s

386

FRESHWATER FISHES OF NORTH AMERICA

(1989) examination of relationships in Cyprinella based largely on morphological characters. In order to clarify the appropriate outgroups to polarize characters within Cyprinella, he examined a large number of cyprinid taxa. Ironically, the phylogeny produced by his search for outgroup taxa has had a greater impact on cyprinid systematics than his phylogeny of Cyprinella. Mayden recognized a large monophyletic group of primarily eastern cyprinids (but also including the genera Oregonichthys and Richardsonius, both found in Pacific Slope drainages) (Figs. 12.42 and 12.50) that he referred to as the open posterior myodome (OPM) clade. This clade was supported by a single synapomorphy (shared, derived character), an opening in the floor of the posterior myodome bounded by the paras-

Gila Ptychocheilus lucius Klamathella Acrocheilus Relictus Siphateles Eremichthys Ptychocheilus oregonensis Lavinia Orthodon Chrosomus A) Simons et al. 2003 Acrocheilus Klamathella Gila Ptychocheilus lucius Ptychocheilus grandis Eremichthys Relictus Hesperoleucas Lavinia Mylopharodon Siphateles Ptychocheilus oregonensis Orthodon Chrosomus B) Smith et al. 2002 Acrocheilus Gila Relictus Siphateles Eremichthys Ptychocheilus oregonensis Ptychocheilus lucius Lavinia Orthodon Chrosomus C) Bufalino & Mayden 2010b

Figure 12.57. Hypotheses of relationships in the western clade. (A) Bayesian analysis of mitochondrial 12S and 16S rRNA sequences (redrawn from Simons et al. 2003). (B) Parsimony analysis of mitochondrial cytochrome b sequences (redrawn from Smith et al. 2002). (C) Bayesian analysis of nuclear Rag1 and S7 sequences (redrawn from Bufalino & Mayden 2010b).

phenoid and the basioccipital (Coburn 1982). Further, he made several taxonomic recommendations to maintain consistency with his hypothesized phylogenetic relationships. He removed the genera Cyprinella, Luxilus, Lythrurus, and Pteronotropis from synonymy with Notropis and dismembered the genus Hybopsis, recognizing the genera Erimystax, Extrarius, Macrhybopsis, and Platygobio. Coburn & Cavender (1992) produced the most comprehensive phylogenetic hypothesis based on morphology to date. They included representatives of nearly all North American cyprinids and recognized three major clades: the western clade, the chub clade, and the shiner clade. Members of Mayden’s OPM clade were distributed among these clades. This was due to the use of different character systems and also disagreement between Coburn & Cavender (1992) and Mayden (1989) over interpretation of some characters, including the OPM (Simons & Mayden 1999). Their analysis also included a number of Asian phoxinins, including the genera Tribolodon and Rhynchocypris; these were part of their chub clade (Coburn & Cavender 1992). The relationships of North American phoxinins have been investigated using DNA sequences of the mitochondrial 12S and 16S ribosomal RNA genes (Simons & Mayden 1997, 1998, 1999; Simons et al. 2003) and the nuclear Rag1 and S7 genes (Schönhuth et al. 2008; Bufalino & Mayden 2010ab; Schönhuth & Mayden 2010). These analyses consistently identified three major clades: western clade, Creek Chub + plagopterin clade, and OPM clade. Simons et al. (2003) considered the OPM clade sister to the Creek Chub + plagopterin clade with these sister to the western clade; however, Bufalino & Mayden (2010) could not resolve sister-group relationships among these three groups. These clades differed in various ways from the clades identified by Coburn & Cavender (1992) and Mayden (1989). A phylogeny of western cyprinids based on mitochondrial cytochrome b sequences (Smith et al. 2002) was largely consistent with Simons et al. (2003) and Bufalino & Mayden (2010ab). The western clade contains the genera Acrocheilus, Chrosomus, Eremichthys, Gila, Hesperoleucas, Klamathella, Lavinia, Mylopharodon, Orthodon, Ptychocheilus, Relictus, and Siphateles (Fig. 12.57). The genus Chrosomus is the sister to all other taxa in the western clade (Bufalino & Mayden 2010b; Simons & Mayden 1998; Smith et al. 2002) and is the only member of this group present in eastern North America (Fig. 12.14). The genus Orthodon is the sister to all other members of the western clade, but little consensus of relationships of the remaining genera is

CYPRINIDAE: CARPS AND MINNOWS

present among the various analyses (Fig. 12.57). The Moapa Dace, Gila coriacea, was classified formerly in the monotypic genus Moapa because of its distinctive appearance (Hubbs & Miller 1948), but Smith et al. (2002) considered it a member of Gila based on their analysis (Fig. 12.57b). Smith et al. (2002) resurrected the genus Klamathella from synonymy with Gila (Fig. 12.57b), and Simons & Mayden (1999) and Simons et al. (2003) considered the genus Klamathella to be sister to the rest of Gila (Fig. 12.57a); Smith et al. (2002) considered Klamathella sister to Acrocheilus (Fig. 12.57b). We concur with the resurrection of Klamathella. The pikeminnows, genus Ptychocheilus, may not be a monophyletic group. Mitochondrial data place the Colorado Pikeminnow, P. lucius, as either closely related to Gila (Fig. 12.57a) (Simons & Mayden 1998; Simons et al. 2003) or sister to the Klamathella + Acrocheilus + Gila clade (Fig. 12.57ab) (Smith et al. 2002). Nuclear gene data provide little resolution to this issue (Fig. 12.57c) (Bufalino & Mayden 2010b). The Creek Chub + plagopterin clade contains the genera Couesius, Lepidomeda, Meda, Plagopterus, Margariscus, Hemitremia, and Semotilus (Dowling et al. 2002; Simons et al. 2003: Fig 12.58). This is a small clade that presents a complex problem. The genera Couesius, Margariscus, Hemitremia, and Semotilus are similar morphologically; however, no unequivocal morphological features unite this group. Couesius, Margarsiscus, and Semotilus have a short, triangular, preterminal maxillary barbel, although the barbel is missing in some individuals and populations (McPhail & Lindsey 1970; Scott & Crossman 1973; Jenkins & Burkhead 1994). Semotilus and Hemitremia are recovered consistently as sister-taxa in analyses of morphological (Coburn & Cavender 1992) and molecular (Simons & Mayden 1997; Simons et al.

387

2003; Bufalino & Mayden 2010ab) data (Fig. 12.58bc). These four genera form a monophyletic group in maximum likelihood and Bayesian analyses of mitochondrial 12S and 16S ribosomal RNA genes (Simons et al. 2003) and nuclear Rag1 and S7 genes (Bufalino & Mayden 2010b) (Fig. 12.58bc). Parsimony analysis of the same data set revealed they were paraphyletic with respect to the plagopterins. Resolution of relationship of Semotilus + Hemitremia, Couesius, and Margariscus varied depending on weights applied to various character transformations (Simons & Mayden 1997). Unweighted morphological (Coburn & Cavender 1992) and molecular analyses of the mitochondrial cytochrome b gene and the Rag1 and S7 nuclear genes (Dowling et al. 2002; Bufalino & Mayden 2010ab) also resulted in paraphyly of these taxa (Fig. 12.58a). The plagopterins, or spine-fins, contain the genera Plagopterus, Lepidomeda, and Meda (Miller & Hubbs 1960). They are characterized by spine-like rays in the dorsal and pelvic fins; tiny scales; and bright, silvery body coloration. Molecular evidence supported inclusion of the Leatherside Chub, Lepidomeda copei, formerly placed in Gila and Snyderichthys, in the plagopterin group (Simons & Mayden 1997), a relationship alluded to by Miller & Hubbs (1960:6): “Resemblance is particularly close between Lepidomeda and several species referred to the genus Gila, and even more strikingly with a species of the Bonneville system, copei that has been referred to a monotypic genus (Miller 1945). It is not now apparent whether such relationship extending even to details of coloration is indicative of intimate relationship.” A study of cytochrome b variation across the range of L. copei revealed evidence for two clades, a northern clade in the Snake and Bear Rivers of Idaho and Wyoming, respectively, and

Plate 12.10. Chiselmouth, Acrocheilus alutaceus

Plate 12.11. Northern Redbelly Dace, Chrosomus eos

Plate 12.12. Desert Dace, Eremichthys acros

Plate 12.13. California Roach, Hersperoleucas symmetricus

Plate 12.14. Hardhead, Mylopharodon conocephalus 388

Plate 12.15. Sacramento Blackfish, Orthodon microlepidotus

Plate 12.16. Colorado Pikeminnow, Ptychocheilus lucius

Plate 12.17. Relict Dace, Relictus solitarius

Plate 12.18. Mohave Tui Chub, Siphateles bicolor mojavensis 389

390

FRESHWATER FISHES OF NORTH AMERICA Semotilus Couesius Margariscus Lepidomeda Meda Plagopterus A) Dowling et al. 2002 Semotilus Hemitremia Margariscus Couesius Lepidomeda Meda B) Simons et al. 2003 Semotilus Hemitremia Lepidomeda Meda Margariscus Couesius C) Bufalino & Mayden 2010b

Figure 12.58. Hypotheses of relationships in the creek chub + plagopterin clade. (A) Parsimony analysis of mitochondrial cytochrome b sequences (redrawn from Dowling et al. 2002). (B) Bayesian analysis of mitochondrial 12S and 16S rRNA sequences (redrawn from Simons et al. 2003). (C) Bayesian analysis of nuclear Rag1 and S7 sequences (redrawn from Bufalino & Mayden 2010b).

a southern clade in the Utah Lake drainage and Sevier River of Utah (Johnson & Jordan 2000; Dowling et al. 2002). In an analysis of plagopterin relationships including representatives from both clades, L. copei was polyphyletic; Lepidomeda mollispinis (Virgin Spinedace), Lepidomeda albivallis (White River Spinedace), and L. copei from the Snake and Bear River drainages formed an unresolved trichotomy sister to Lepidomeda vittata (Little Colorado Spinedace). This group was sister to L. copei from the Sevier River and Utah Lake drainages (Dowling et al. 2002). Mitochondrial and nuclear DNA sequences, morphology, and ecology present compelling evidence that there are two species of the Leatherside Chub, the Southern Leatherside Chub, Lepidomeda aliciae, and the Northern Leatherside Chub, Lepidomeda copei (Johnson et al. 2004). The genus Lepidomeda is the sister-taxon to Meda plus Plagopterus (Fig. 12.58a) (Dowling et al. 2002). For ease of discussion, the OPM clade is divided into four groups, including a paraphyletic basal grade of taxa and three monophyletic clades: Platygobio clade, Phenacobius clade, and shiner clade. Simons & Mayden (1999) discussed

Plate 12.19. Lake Chub, Couesius plumbeus

Plate 12.20. White River Spinedace, Lepidomeda albivallis

Plate 12.21. Spikedace, Meda fulgida

Plate 12.22. Pearl Dace, Margariscus margarita

Plate 12.23. Flame Chub, Hemitremia flammea

Plate 12.24. Creek Chub, Semotilus atromaculatus

391

392

FRESHWATER FISHES OF NORTH AMERICA

the homology of the OPM and the inclusion of taxa not identified by Mayden (1989) as members of this group. The OPM clade is the largest group of North American cyprinids, and much work remains to adequately resolve the relationships among the included taxa, thus the classification and relationships described here are necessarily preliminary. The basal grade contains the genera Campostoma, Clinostomus, Exoglossum, Iotichthys, Mylocheilus, Nocomis, Oregonichthys, Pogonichthys, Rhinichthys, Richardsonius, and Tiaroga (Smith et al. 2002; Simons et al. 2003; Bufalino & Mayden 2010ab). Simons et al. (2003) divided these taxa into a series of clades, but not all these are supported by nuclear data (Bufalino & Mayden 2010ab); however, some groups are always recovered (Fig. 12.59). The Phenacobius clade contains the genera Erimystax and Phenacobius (Dimmick 1993; Simons et al. 2003; Hollingsworth & Hulsey 2011). Mayden (1989) described a hypothesis of relationship among included species based on morphological data. Dimmick & Burr (1999) examined species relationships in Phenacobius using data from morphology, allozymes, and DNA sequences. Simons (2004) described relationships among species in Erimystax based on analysis of mitochondrial cyto-

Shiner Clade Macrhybopsis Platygobio Erimystax Phenacobius Oregonichthys Tiaroga Exoglossum Rhinichthys Campostoma Nocomis Clinostomus Richardsonius Mylocheilus Pogonichthys A) Simons et al. 2003 Shiner Clade Erimystax Phenacobius Macrhybopsis Platygobio Campostoma Nocomis Rhinichthys Tiaroga Oregonichthys Exoglossum Mylocheilus Pogonichthys Clinostomus Richardsonius B) Bufalino & Mayden 2010

Figure 12.59. Hypotheses of relationships in the open posterior myodome clade. (A) Bayesian analysis of mitochondrial 12S and 16S rRNA sequences (redrawn from Simons et al. 2003). (B) Bayesian analysis of nuclear Rag1 and S7 sequences (redrawn from Bufalino & Mayden 2010b).

chrome b sequences. Hollingsworth & Hulsey (2011) used a multi-locus, coalescent-based approach, corroborating a sister-group relationship between Phenacobius and Erimystax but recovered a different species-level phylogeny in Phenacobius than that of Mayden (1989) and Dimmick & Burr (1999). The Phenacobius clade is either sister to the Platygobio clade plus the shiner clade (Fig. 12.59a) (Simons et al. 2003) or to the shiner clade (Fig. 12.59b) (Bufalino & Mayden 2010b). Schönhuth & Mayden (2010) placed Erimystax in the shiner clade (see shiner clade below). The Platygobio clade contains two genera, Macrhybopsis and Platygobio. Placed in Hybopsis by Bailey (1951), these taxa were classified subsequently in three genera: Extrarius, Macrhybopsis, and Platygobio (Mayden 1989). Subsequent authors placed the genus Extrarius in Macrhybopsis (Coburn & Cavender 1992; Dimmick 1993; Simons & Mayden 1999). Relationships within Macrhybopsis are not well resolved. Dimmick (1993) placed the Sturgeon Chub, Macrhybopsis gelida, plus the Sicklefin Chub, Macrhybopsis meeki, as sister to the M. aestivalis complex. Eisenhour (1999) reviewed the systematics of the Peppered Chub, Macrhybopsis tetranema, and Eisenhour (2004) described variation in the M. aestivalis complex, elevating several species from synonymy with the Speckled Chub, Macrhybopsis aestivalis. He also presented a hypothesis of relationships based on 17 morphological characters for these species. No comprehensive phylogenetic hypothesis is available for all species in this clade, and several species await description. Substantial support exists for a sister-group between Platygobio and Macrhybopsis (Fig. 12.59ab) (Simons et al. 2003; Bufalino & Mayden 2010b). The Campostoma clade contains the genera Campostoma and Nocomis; the phylogenetic position of this group depends on the data and analysis. Bayesian analysis of mitochondrial 12S and 16S sequences places the Campostoma clade as sister to a group containing the shiner clade and the genera Platygobio, Macrhybopsis, Erimystax, Phenacobius, Oregonichthys, Tiaroga, Exoglossum, and Rhinichthys (Fig. 12.59a) (Simons et al. 2003). Bayesian analysis of combined nuclear Rag1 and S7, however, places this clade in a monophyletic group with Rhinichthys and Tiaroga (Fig. 12.59b) (Bufalino & Mayden 2010b). Similarly, the relationships of Exoglossum and Oregonichthys vary depending on the data and analysis (Fig. 12.59ab). Resolution of these relationships will require more intensive taxon sampling and additional data.

Plate 12.25. Gravel Chub, Erimystax x-punctatus

Plate 12.26. Stargazing Minnow, Phenacobius uranops

Plate 12.27. Flathead Chub, Platygobio gracilis

Plate 12.28. Mexican Stoneroller, Campostoma ornatum 393

Plate 12.29. Redspot Chub, Nocomis asper

Plate 12.30. Oregon Chub, Oregonichthys crameri

Plate 12.31. Longnose Dace, Rhinichthys cataractae

Plate 12.32. Redside Dace, Clinostomus funduloides 394

CYPRINIDAE: CARPS AND MINNOWS

The genera Clinostomus (Fig. 12.56d), Richardsonius, Mylocheilus, and Pogonichthys are in a basal position with respect to the rest of the OPM clade. Clinostomus and Richardsonius are sister-taxa (Simons & Mayden 1999; Simons et al. 2003; Bufalino & Mayden 2010b), have long been recognized as close relatives (Uyeno 1961b), and have an interesting biogeographical distribution with Clinostomus found in eastern North America (Fig. 12.15) and Richardsonius found west of the western Continental Divide (Fig. 12.50). Mylocheilus and Pogonichthys were only recently recognized as close relatives (Simons & Mayden 1999) and occur in western North America where the genus Mylocheilus is native to the Columbia River, Oregon, north to the Nass River, British Columbia (Fig. 12.37), and the genus Pogonichthys is native to the Sacramento River system, California (Fig. 12.48). Clinostomus plus Richardsonius are either the sister to all other members of the OPM clade (Fig. 12.59b) (Bufalino & Mayden 2010b) or sister to Mylocheilus plus Pogonichthys (Fig. 12.59a) (Simons & Mayden 1999; Simons et al. 2003).

395

The shiner clade contains ≥170 species. This clade, as currently defined, includes the genera Agosia, Alburnops, Algansea, Aztecula, Codoma, Cyprinella, Dionda, Ericymba, Erimonax, Graodus, Hudsonius, Hybognathus, Hybopsis, Luxilus, Lythrurus, Miniellus, Notropis, Pimephales, Pteronotropis, Tampichthys, and Yuriria (Simons et al. 2003; Mayden et al. 2006; Schönhuth et al. 2008). In addition, the genus Erimystax (and presumably Phenacobius) may also be part of this group (Schönhuth & Mayden 2010). Relationships among these genera are largely unresolved because no single study exists with sufficient taxon and character sampling to produce a well-supported, comprehensive hypothesis of relationships. Several genera were recently elevated from synonymy with Notropis and a number of species, although clearly not members of Notropis, cannot be assigned to a par ticular genus; herein these are referred to as “Notropis” (following Mayden et al. 2006). Thus, our discussion of relationships must be viewed as provisional, illustrating areas for future research. The shiner clade is well supported based on analysis of mitochondrial 12S and

Plate 12.33. Redside Shiner, Richardsonius balteatus

Plate 12.34. Peamouth, Mylocheilus caurinus

396

FRESHWATER FISHES OF NORTH AMERICA

Plate 12.35. Splittail, Pogonichthys macrolepidotus

Plate 12.36. Longfin Dace, Agosia chrysogaster

Plate 12.37. Roundnose Minnow, Dionda episcopa

16S ribosomal RNA (Simons et al. 2003); Mayden et al. (2006) included many more taxa in an analysis of mitochondrial cytochrome b sequences. This analysis resolved relationships within species groups and illustrated that the genera Notropis, Luxilus, and Pteronotropis are not monophyletic as currently defined. Mayden et al. (2006) resurrected several genera to partially resolve these taxonomic problems. Schönhuth et al. (2008) examined relationships of Mexican shiners in an analysis of one mitochondrial (cytochrome b) and three nuclear genes (S7, Rhodopsin, and Rag1) and illustrated that the

genus Dionda was not monophyletic as currently defined. A new genus, Tampichthys, was described to contain species formerly included in Dionda, and genera were resurrected to contain taxa formerly classified in Notropis. Schönhuth & Mayden (2010) examined the relationships of Cyprinella based on separate analyses of mitochondrial cytochrome b and nuclear Rag1 genes and included several outgroup taxa that also shed light on relationships in the shiner clade. Papers by Mayden et al. (2006), Schönhuth et al. (2008), and Schönhuth & Mayden (2010) form the basis of the discussion of

CYPRINIDAE: CARPS AND MINNOWS

phylogenetic relationships in the shiner clade (Figs. 12.60–12.62). The genus Dionda is the sister-group to all other shiner taxa included in the Schönhuth et al. (2008) and Schönhuth & Mayden (2010) analyses (Figs. 12.60–12.62). The Tampichthys Codoma Cyprinella Hybognathus Dionda sp. c.f. ipni. (Rio Axtla) Aztecula Graodus Yuriria Algansea Agosia Dionda Schönhuth et al. 2008

Figure 12.60. Hypothesis of relationships among Mexican cyprinids in the shiner clade. Bayesian analysis of mitochondrial (cytochrome b) and nuclear (Rag1, Rhodopsin, Rag1) sequences (redrawn from Schönhuth et al. 2008).

genus Dionda was not included in Mayden et al. (2006), but a sister-group relationship of Dionda to all other shiners is not inconsistent with their results. The genus Agosia is sister to Algansea (Schönhuth et al. 2008). These comprise the sister-group to all other shiners exclusive of Dionda; Mayden et al. (2006) did not include Algansea but did include Agosia. Schönhuth et al. (2008) described interspecific relationships within Dionda and Algansea. The genera Aztecula, Graodus, and Yuriria form a monophyletic group (Mayden et al. 2006; Schönhuth et al. 2008; Schönhuth & Mayden 2010), but their relationship to the rest of the shiners is unclear (Figs. 12.60–12.62). Schönhuth et al. (2008) and Mayden et al. (2006) described interspecific relationships within these genera. The genus Codoma is sister to Tampichthys (Schönhuth et al. 2008; Schönhuth & Mayden 2010), and these are closely related to Cyprinella, Miniellus, and four species: Nazas Shiner, “Notropis” nazas; Bigmouth Shiner, “Notropis” dorsalis; and Sandbar Shiner, “Notropis” scepticus

Tampichthys Codoma Cyprinella (in part) Cyprinella callistia Luxilus zonatus group Luxilus chrysocephalus group Luxilus coccogenis Notropis buchanani Lythrurus Notropis braytoni Aztecula Yuriria Graodus Ericymba Hybognathus Notropis nubilus Erimonax Pimephales + Opsopoeodus Erimystax Dionda A) Schönhuth & Mayden 2010, Cytochrome b Tampichthys Codoma Cyprinella (in part) Cyprinella (in part) Pimephales +Opsopoeodus Erimonax Hybognathus Notropis braytoni Notropis buchanani Notropis nubilus Graodus Aztecula Yuriria Ericymba Luxilus chrysocephalus Lythrurus Erimystax Dionda B) Schönhuth & Mayden 2010, Rag1

397

Figure 12.61. Hypotheses of relationships of members of the shiner clade. (A) Maximum likelihood and Bayesian analyses of mitochondrial cytochrome b sequences. (B) Maximum likelihood and Bayesian analyses of nuclear Rag1 sequences (redrawn from Schönhuth & Mayden 2010).

398

FRESHWATER FISHES OF NORTH AMERICA Cyprinella Miniellus “Notropis” scepticus “Notropis” nazas “Notropis” dorsalis “Notropis” photogenis “Notropis” telescopus Ericymba “Notropis” bifrenatus Hybognathus Hybopsis Luxilus “Luxilus” zonatus group “Luxilus” cerasinus “Luxilus” coccogenis group Lythrurus “Notropis” melanostomus “Notropis” scepticus “Notropis” semperasper Notropis “Notropis” longirostris Alburnops Aztecula Graodus Yuriria “Pteronotropis” “Notropis” heterolepis “Notropis” rupestris “Notropis” greenei Pimephales + Opsopoeodus “Notropis” scabriceps “Notropis” simus Hudsonius Pteronotropis Agosia

Mayden et al. 2006

Figure 12.62. Hypotheses of relationships of members of the shiner clade. Bayesian analysis of mitochondrial cytochrome b sequences (redrawn from Mayden et al. 2006).

(Mayden et al. 2006). Codoma has been considered a close relative of Pimephales (Page & Johnston 1990a), Cyprinella (Mayden 1989, 2002), and Tampichthys (formerly Dionda) (Simons et al. 2003). Apparent eggclustering behavior in Codoma (Minckley & Vives 1990) was used as evidence for relationship with Pimephales, also an egg clusterer (Page & Johnston 1990; Table 12.4C). Subsequently, S. J. Vives indicated Codoma is actually a crevice spawner (pers. comm. cited by Mayden & Simons 2002; Table 12.4A). Mayden (2002) considered Codoma a close relative of Cyprinella based on morphology (see Mayden 1989), spawning behavior, and molecular data, but Simons et al. (2003) considered Codoma a close relative of Tampichthys (formerly Dionda) based on molecular data. Reproductive behavior also supports this hypothesis as some members of Tampichthys are also crevice spawners (Mayden & Simons 2002; Table 12.4C). Phylogenetic issues associated with Cyprinella relationships include relationships among species, membership of

the genus, and relationship of Cyprinella to other North American cyprinid genera. Most species included in Cyprinella are unequivocally members of this group (but see Schönhuth & Mayden 2010; Fig. 12.61) given their distinctive physiognomy and spawning behavior. Other species that have been referred to Cyprinella include the Ornate Shiner, Codoma ornata; Spotfin Chub, Erimonax monachu; Thicklip Chub, Cyprinella labrosa; and Santee Chub, Cyprinella zanema. Mayden (1989) published a phylogenetic hypothesis of species-level relationships based on morphology and also detailed morphological variation within the clade. He placed Cyprinella in a monophyletic group with Luxilus and Lythrurus. He included the genus Codoma within Cyprinella, sister to the C. lutrensis clade. He referred the Spotfin Chub to Erimystax based on barbel morphology and included the Thicklip Chub and Santee Chub in Hybopsis based on nine morphological characters. In contrast, Coburn & Cavender (1992) included the Spotfin Chub, Thicklip Chub, and Santee Chub in Cyprinella. They placed the genus Cyprinella in a monophyletic group with Opsopoeodus plus Pimephales based on three morphological characters. Their work was focused on generic relationships and thus did not address relationships within Cyprinella. Dimmick (1993) used variation in allozymes to examine relationships of barbelled cyprinids, formerly classified in Hybopsis. Dimmick identified a sister-group relationship between the Thicklip Chub and Whitetail Shiner, Cyprinella galactura. In addition, the Thicklip Chub plus the Santee Chub formed the sister-group to the Bigeye Chub, Hybopsis amblops, and the Rosyface Chub, Hybopsis rubrifrons. Dimmick’s (1993) analysis may suffer from low taxon sampling and choice of outgroup (genus Campostoma). Broughton & Gold (2000) examined relationships of species of Cyprinella based on mitochondrial ND2 and ND4L genes and found significantly different relationships than did Mayden (1989). No evidence supported inclusion of the Spotfin Chub in Cyprinella; however, inclusion of the Thicklip Chub and Santee Chub in Cyprinella was supported as in Schönhuth & Mayden (2010). In Simons et al. (2003), no evidence supported a relationship of Cyprinella with Pimephales, Opsopoeodus, or the Spotfin Chub. Parsimony analyses placed Cyprinella sister to some species of Notropis, but likelihood analyses placed Cyprinella as the sister to the rest of the shiner clade. Schönhuth et al. (2008) identified Cyprinella as sister to Codoma and Tampichthys based on their analyses of one mitochondrial and three nuclear genes and this relationship is supported by similarity in spawning behavior

Plate 12.38. Spotfin Chub, Erimonax monachus

Plate 12.39. Spottail Shiner, Hudsonius hudsonius

Plate 12.40. Rio Grande Silvery Minnow, Hybognathus amarus

Plate 12.41. Bigeye Chub, Hybopsis amblops 399

400

FRESHWATER FISHES OF NORTH AMERICA

Plate 12.42. Topeka Shiner, Miniellus topeka

Plate 12.43. Bluntnose Minnow, Pimephales notatus

(see previous paragraph). A study using both mitochondrial (cytochrome b) and nuclear (Rag1) genes that included all species of Cyprinella (Schönhuth & Mayden 2010) painted a complex picture. Although recovered topologies differed somewhat for the two genes, both data sets demonstrated that the Thicklip Chub and Santee Chub should be included in Cyprinella. Interestingly, the genera Codoma and Tampichthys rendered Cyprinella paraphyletic in these analyses (Fig. 12.61). The genus Alburnops contains taxa formerly included in the subgenus Alburnops, subgenus Hydrophlox, and texanus species group (sensu Swift 1970). Mayden et al. (2006) described the monophyly of and relationships in Alburnops. Cashner et al. (2011) redefined the subgenus Hydrophlox to include only five species. The genus Ericymba is morphologically distinctive with large cephalic canals (see morphology section) and has been placed in Notropis and Hybopsis as well as Ericymba. Ericymba was part of an unresolved polychotomy

at the base of the Hybopsis clade in Mayden’s (1989) analysis. Coburn & Cavender (1992) included Ericymba in Notropis but commented on its morphological similarity to the Hybopsis group (also included in Notropis by Coburn & Cavender 1992). Raley & Wood (2001) examined cytochrome b sequences and included the Silverjaw Minnow, Ericymba buccata, in Notropis based on a close relationship with the “Notropis” dorsalis species group. In Simons et al. (2003), the position of Ericymba was unstable and analysis dependent. Using cytochrome b, Schönhuth & Mayden (2010) found a sister-group relationship with the clade Aztecula + Yuriria + Graodus (Fig. 12.61a), but analysis using Rag1 placed Ericymba sister to the Striped Shiner, Luxilus chrysocephalus (Fig. 12.61b). Mayden et al. (2006) recovered a relationship with the Silver Shiner, “Notropis” photogenis, and the Telescope Shiner, “Notropis” telescopus (Fig. 12.62). Here, we recognize the genus Ericymba, awaiting a more in-depth analysis of relationships in the shiner clade.

CYPRINIDAE: CARPS AND MINNOWS

The genus Erimonax (Fig. 12.56a) has a complex taxonomic history, having been included in Hybopsis, Erimystax, and Cyprinella. Classified in Erimystax based on the stellate morphology of the barbel as well as other morphological characters (Mayden 1989), it was subsequently included in Cyprinella based on osteology and scale morphology (Coburn & Cavender 1992), biochemical data (Dimmick 1993), and spawning behavior (Jenkins & Burkhead 1994). In Simons et al. (2003) and Schönhuth & Mayden (2010), the phylogenetic position was unstable and dependent on the data set and analysis, but Erimonax was never recovered as a close relative of Cyprinella (Fig. 12.61ab). We concur with Mayden et al.’s (1992) resurrection of Erimonax (Jordan 1924); this is supported by several analyses (i.e., Simons et al. 2003; Bufalino & Mayden 2010ab; Schönhuth & Mayden 2010). The genus Hudsonius contains three species: Spottail Shiner, Hudsonius hudsonius; Highfin Shiner, Hudsonius altipinnis; and Dusky Shiner, Hudsonius cummingsae (Mayden et al. 2006). Mayden et al. (2006) discussed relationships in this group and noted that the apparent paraphyly of the Highfin Shiner may indicate cryptic diversity within this species, a point alluded to by Hubbs & Raney (1948), who identified six subspecies within the species. The genus Hybognathus, considered a member of the chub clade by Mayden (1989), was placed in a polychotomy with Exoglossum and the Campostoma clade (genera Nocomis, Campostoma + Dionda). Coburn & Cavender (1992) included Hybognathus in their chub clade, sister to Dionda + Campostoma. Mayden et al. (1992) considered Hybognathus sister to Dionda + Campostoma. Schmidt (1994) examined phylogenetic relationships of species in this genus. Simons et al. (2003) included Hybognathus in the shiner clade but could not identify the sister-taxon of this clade. Schönhuth et al. (2008) identified Hybognathus plus an undescribed species (Dionda sp. cf. ipni from Rio Axtla, San Luis Potosi, Mexico) as the sister to Cyprinella + Codoma and Tampichthys supporting membership of Hybognathus in the shiner clade (Fig. 12.60). In Schönhuth & Mayden (2010), phylogenetic positions differed for Hybognathus depending on the gene used in the analysis (Fig. 12.62ab). The taxon sampling and data used in these studies are insufficient to determine the phylogenetic position of Hybognathus with confidence. Bailey (1951) dramatically expanded the genus Hybopsis to include all North American cyprinids with a terminal maxillary barbel. The taxa Nocomis and Couesius were re-

401

elevated subsequently to generic status (Lachner & Jenkins 1967; McPhail & Lindsey 1970; Jenkins & Lachner 1971). Mayden (1989) restricted the genus Hybopsis to the subgenus Hybopsis; the dorsalis species group; the Thicklip Chub, Cyprinella labrosa; Santee Chub, C. zanema; Balsas Shiner, Graodus boucardi; Yellow Shiner, Aztecula calientis; Whitemouth Shiner, “Notropis” alborus; and Bridle Shiner, “Notropis” bifrenatus. Several of these taxa have since been removed from Hybopsis. Shaw et al. (1995) and Grose & Wiley (2002) examined relationships of the H. amblops species group. We restrict Hybopsis to five species (as in Mayden et al. 2006): Bigeye Chub, H. amblops; Highback Chub, Hybopsis hypsinotus; Lined Chub, Hybopsis lineapunctatus; Rosyface Chub, H. rubrifrons; and Clear Chub, Hybopsis winchelli. The genus Luxilus was removed from synonymy with Notropis by Mayden (1989). Mayden considered Luxilus as the sister to Cyprinella. Coburn & Cavender (1992) considered Luxilus sister to a clade containing the genera Lythrurus, Cyprinella, Pimephales, and Opsopoeodus. Mayden (1989) identified three characters supporting monophyly of Luxilus, including retrorse preorbital tubercles, cleithral region with intense pigmentation, and epibranchial 3 with an elongate and curled uncinate process. Interspecific relationships were examined using morphology (Gilbert 1964), biochemical data (Buth 1979a), and DNA sequences (Dowling & Naylor 1997). Gilbert (1964) recognized three species groups: the L. coccogenis group, the L. zonatus group (Fig.12.56h), and the L. cornutus group. He noted that significant morphological breaks occurred between each group and that they did not appear closely related to one another. Buth (1979) also identified three species groups but argued that the Crescent Shiner, Luxilus cerasinus, was not part of the L. cornutus group and should be considered a separate lineage. Dowling & Naylor (1997) also recovered three species groups, but bootstrap support was low for relationships among these groups and the Crescent Shiner. Mayden et al. (2006) identified four species groups but found no evidence that Luxilus is monophyletic (Fig. 12.62). Schönhuth & Mayden (2010) found support for a sister-group relationship between the L. zonatus and L. chrysocephalus groups, but the Warpaint Shiner, Luxilus coccogenis, was recovered as sister to the Ghost Shiner, Notropis buchanani, rather than other Luxilus (Fig. 12.61a). Additional work is needed to determine the extent, membership, and relationships of Luxilus. Mayden (1989) considered the genus Lythrurus sister to Luxilus + Cyprinella. Coburn & Cavender (1992) placed

402 FRESHWATER FISHES OF NORTH AMERICA

Lythrurus as sister to a clade containing the genera Cyprinella, Pimephales, and Opsopoeodus. Molecular analyses do not provide a clear picture of the phylogenetic position of this genus (Figs. 12.61 and 12.62). Schmidt et al. (1998) examined interspecific relationships in Lythrurus based on DNA sequences of the mitochondrial cytochrome b gene. The genus Miniellus is a monophyletic group containing the Blackchin Shiner, Miniellus heterodon; Sand Shiner, Miniellus stramineus; Topeka Shiner, Miniellus topeka; and Swallowtail Shiner, Miniellus procne. Mayden (1989) and Schmidt & Gold (1995) recognized a group containing the Sand Shiner, Topeka Shiner, and Swallowtail Shiner using morphology and cytochrome b sequences, respectively. Mayden et al. (2006) added the Blackchin Shiner to this group and recommended that the group be removed from synonymy with Notropis. Bailey (1951) placed nearly all non-barbelled shiners (except Hybognathus and Pimephales) in the genus Notropis. Here, we restrict the genus Notropis to the former subgenus Notropis (see Bielawski & Gold 2001) and a number of species of uncertain relationship (“Notropis” of Mayden et al. 2006). Bielawski & Gold (2001) resolved interspecific relationships in the subgenus Notropis. Wood et al. (2002) and Berendzen et al. (2008) examined relationships and species boundaries in the N. rubellus species group, presenting evidence for undescribed diversity. The genus Pimephales has long been considered a close relative of the Pugnose Shiner, Opsopoeodus emiliae, based on morphology (Coburn & Cavender 1992), reproductive behavior (Page & Johnston 1990a), and molecular data (Simons et al. 2003). Schmidt et al. (1994) and Bielawski et al. (2002) examined phylogenetic relationships of Pimephales; however, their analyses may be compromised by outgroup choice. Both studies used the Pugnose Shiner as the outgroup, making the assumption that Pimephales was monophyletic. Larger-scale molecular phylogenetic analyses that have included species of Pimephales and the Pugnose Shiner suggest that Pimephales is paraphyletic with respect to the Pugnose Shiner (Mayden 2002; Simons et al. 2003; Mayden et al. 2006; Schönhuth & Mayden 2008); however, support values for these relationships are weak. Pugnose Shiners exhibit a suite of morphological characters that distinguish them from Pimephales; thus, we continue to recognize the genus Opsopoeodus until these relationships are resolved. The genus Pteronotropis usually is divided into two groups based on overall morphological similarity. The Broadstripe Shiner, Pteronotropis euryzonus, Sailfin

Shiner, Pteronotropis hypselopterus, Orangetail Shiner, Pteronotropis merlini, and Flagfin Shiner, Pteronotropis signipinnis, for example, are relatively deep-bodied fishes with a prominent, dark, wide, lateral band; nuptial males have bright orange, red, and blue colors on the body and fins and develop enlarged dorsal and anals fins. The Bluehead Shiner, Pteronotropis hubbsi, and Bluenose Shiner, Pteronotropis welaka, are more terete than other members of the genus, but the nuptial males develop dramatically enlarged dorsal, anal, and pelvic fins as well as an intense blue color on the head or snout (Fig. 12.56f). Phylogenetic analyses of this group have produced conflicting results with little evidence for monophyly of the group. Dimmick (1987) found evidence for a monophyletic group containing the Flagfin Shiner, Sailfin Shiner, and Bluenose Shiner, based on allozyme electrophoresis of 21 gene loci. His analysis did not provide evidence of a close relationship of the Bluehead Shiner with these taxa. Mayden (1989) recognized four species in Pteronotropis, but placed the Bluehead Shiner in Notropis. Simons et al. (2000) examined relationships among species of Pteronotropis using DNA sequences of the mitochondrial cytochrome b gene. Outgroup taxa were included to represent taxa previously proposed as close relatives of Pteronotropis. They recovered two monophyletic groups, the first contained the Bluehead Shiner, Flagfin Shiner, and Bluenose Shiner, and the second contained the Broadstripe Shiner and Sailfin Shiner. Simons et al. (2000) found no evidence that these two clades formed a monophyletic group. Interestingly, in a larger-scale analysis of representatives of nearly all North American cyprinid genera, Simons et al. (2003) consistently recovered the Bluehead Shiner and Broadstripe Shiner as sister-taxa. Unfortunately, these were the only putative Pteronotropis included in their study. Mayden et al. (2006) included six species of Pteronotropis in their analysis and, similar to previous studies, did not recover a monophyletic Pteronotropis. The Broadstripe Shiner, Sailfin Shiner, and Flagfin Shiner formed a monophyletic group, but no evidence existed for a sister-taxon relationship with the other group of Pteronotropis consisting of the Redeye Chub, Pteronotropis harperi (not formerly considered part of the Pteronotropis group), Bluehead Shiner, and Bluenose Shiner. The phylogenetic positions of Evarra and Stypodon are unclear. Both genera are extinct, and little is known of their morphology. The North American cyprinid fauna is not monophyletic. Less clear is the number of monophyletic North American clades of cyprinids. As noted, North American

CYPRINIDAE: CARPS AND MINNOWS

cyprinids contain representatives of two major cyprinid clades: phoxinins and leuciscins (Cavender & Coburn 1992). Phoxinins are Laurasian in distribution and may have representatives southward into China. Leuciscins, in contrast, are confined to North America and Europe. The lone leuciscin native to North America is the Golden Shiner, Notemigonus crysoleucas, which is related to European cyprinids, specifically the genera Abramis, Blicca, Chondrostoma, Leuciscus, and Scardinius (Briolay et al. 1998; Cunha et al. 2002). Bailey (1951) placed the North American genus Pfrille into synonymy with Chrosomus. Banarescu (1964) then placed Chrosomus into synonymy with Phoxinus, thus creating the only cyprinid genus containing both Eurasian and North American species. This change was based on the remarkable similarity of Pfrille and Chrosomus to the Eurasian Minnow, Phoxinus phoxinus. Chen (1994) hypothesized that the genus Phoxinus was monophyletic and sister to a clade of Asian phoxinins including the genera Eupallasella, Lagowskiella, and Rhynchocypris. Chen suggested that Phoxinus and the Eupalasella clade were sister to a clade containing Margariscus, Couesius, and Semotilus. Simons & Mayden (1998) presented evidence that North American species of Phoxinus were sister to the western clade. They suggested that if Phoxinus was truly the sister to the western clade, the creek chub + plagopterin clade and the OPM clade would have Asian representatives or would compose the sister group to an Asian taxon. This prediction was partially supported by Strange & Mayden (2009), who identified the Asian genus Rhynchocypris as sister to a clade containing Hemitremia, Semotilus, Couesius, and Margariscus (plagopterins were not included in their analysis). Strange & Mayden (2009) also clarified the paraphyly of Phoxinus (sensu Banarescu) and recommended that the North American taxa be removed from synonymy with Phoxinus and assigned to Chrosomus. Howes (1984) proposed another transcontinental relationship suggesting a sistergroup relationship between Pogonichthys and Tribolodon. The genus Pogonichthys is native to the Sacramento River system, California, and the genus Tribolodon is native to Japan, China, and Korea. Coburn & Cavender (1992) disagreed with Howes, placing Pogonichthys in their western clade. Coburn & Cavender (1992) did consider Tribolodon closely related to the North American fauna; Tribolodon and Rhynchocypris were resolved as part of their Chub clade. Simons & Mayden (1999) described evidence for considering Pogonichthys sister to Mylocheilus, native to Pacific drainages from the Columbia River north to the Nass River, British Columbia. They cited genetic data as

403

well as a shared tolerance of brackish water, unusual in cyprinids. Simons & Mayden (1999) speculated that Pogonichthys and Mylocheilus were related closely to Tribolodon based on tolerance of brackish water and similar male nuptial coloration. Resolution of the relationships of the North American cyprinid fauna, both identifying relationships of the native genera and species and determining the relationship of the fauna with that of other continents, will require substantial taxon sampling and large-scale analyses with the inclusion of multiple loci.

FOSSIL RECORD The history of North American cyprinid paleoichthyology extends from Cope’s work in the late 1800s to the present day. Early work was primarily descriptive, defining new species, specimens, or the composition of faunas from a geological formation. More contemporary integrative approaches used the cyprinid fossil record to examine faunal changes and rates of diversification (G. R. Smith 1981; Smith et al. 2002). The North American cyprinid fossil record is limited to the Middle Cenozoic with most fossils described from Pliocene (5.3–1.8 mya) and Pleistocene (1.8 mya to 10,000 years ago) localities (Fig. 12.63; reviewed by G. R. Smith 1981). The oldest fossils, all of which represent extinct taxa, are from the Oligocene (34–32 mya) of northwestern North America. Other fossils indicate that most cyprinid trophic guilds were already in place by the end of the Oligocene (Cavender 1991). Sadly, many of these Oligocene cyprinids are not formally described. Fossil cyprinids from the Miocene (23–5.3 mya) are much better known than those of the Oligocene and include representatives from seven extant genera (Acrocheilus, Gila, Mylocheilus, Mylopharodon, Notropis, Orthodon, and Ptychocheilus) and one extinct genus (†Idadon; G. R. Smith 1981). Fossil localities are widespread and include western Montana (Cavender 1991); Sentinel Butte, North Dakota (Cavender 1991); Kansas (G. R. Smith 1981); and Oregon (Kimmel 1975). These taxa with representatives from the western clade and the OPM clade (Simons & Mayden 1999) indicate that much of the diversity of the North American cyprinid fauna had evolved by the Miocene (about 23–5.3 mya). A wide range of trophic morphologies occurs in Miocene cyprinids. Inferences based on tooth morphology suggest this fauna included carnivorous, herbivorous, and molluskivorous cyprinids (Kimmel 1975).

404 FRESHWATER FISHES OF NORTH AMERICA

Taxon

Miocene

Pliocene

Early Late Pleistocene Pleistocene

Holocene

Acrocheilus †Aphelichthys Campostoma †Evomus Gila Hybognathus †Idadon Lavinia Luxilus Mylocheilus Mylopharodon Nocomis Notemigonus Notropis Orthodon Phoxinus Pimephales Platygobio Pogonichthys Ptychocheilus Rhinichthys Richardsonius Semotilus Siphateles

Figure 12.63. Fossil record of North American cyprinid genera. Solid line indicates the presence of fossils and dashed line indicates inferred presence based on previous fossil data.

The best known locality for North American fossil cyprinids is Lake Idaho. This locality was studied sporadically since Cope (1870) first described the fauna and contains both Miocene (23–5.3 mya) and Pliocene (5.3–1.8 mya) deposits (Kimmel 1975; Smith 1975). Uyeno (1961a) detailed the fauna of Pliocene Lake Idaho, describing a new species of cyprinid and revising the generic assignments of Cenozoic North American fishes. Smith (1975) reviewed and described the fishes of the Pliocene Glenns Ferry Formation, part of fossil Lake Idaho. This lake was large and stable for 4–10 million years before its capture by the Columbia River system. The Glenns Ferry Formation contained about 30 species, including ≥10 cyprinids. The cyprinid fauna was trophically diverse, as inferred by pharyngeal tooth morphology with several specialized forms, including †Ptycho-

cheilus arciferus, a large piscivore; †Acrocheilus latus and †Orthodon hadrognathus, presumably algivores; and †Mylocheilus robustus and †M. inflexus, presumed molluskivores. Extant species from genera that were members of the Lake Idaho fossil fauna are not as specialized as the extinct forms (Smith 1975). Thus the fossil fauna of Lake Idaho is similar to that of other lacustrine systems in which closely related sympatric taxa develop specialized trophic morphologies (Greenwood 1984; Smith & Todd 1984; Nagelkerke et al. 1994). The North American cyprinid fossil record increases in the Pliocene and Pleistocene (1.8 mya to 10,000 years ago) but is still relatively poor (Fig. 12.63). Most fossils are from lacustrine environments, and this could bias the represented diversity because most contemporary cyprinid diversity exists in streams and rivers, habitats less amenable to fossil preservation or discovery. The fossil record of pre-Pleistocene cyprinids is dominated by western North American genera, including Gila, †Idadon, Mylocheilus, Mylopharodon, Orthodon, Ptychocheilus, and Siphateles (G. R. Smith 1981). This also points to bias in preservation as most of the extant diversity of North American cyprinids is present in the Mississippi River basin. The cyprinid fossil record also may be biased because the bones of these fishes are small and delicate and may be missed in excavations or misidentified in museums. Most cyprinid fossils from Pliocene and Pleistocene localities represent extant genera, if not species (Uyeno & Miller 1963; G. R. Smith 1981). The fossil evidence of a paucity in large-scale taxonomic changes in cyprinids over the past 5 million years is corroborated by molecular evidence (Dowling et al. 2002; Simons 2004; Berendzen et al. 2008b). Interestingly, cyprinid fossils show similar diversity throughout the past 2 million years, which could indicate that the Pleistocene mammalian megafaunal extinctions were not mirrored in cyprinids (Gobalet & Fenenga 1993). Diversification of North American cyprinids is likely much older than generally recognized. The fossil record of Great Basin cyprinids, together with examination of molecular phylogenies (see phylogenetic relationships section) indicates that the divergence between major cyprinid lineages had occurred by the Miocene. Genetic divergence between allopatric taxa was often older than the most recent geographic connection between the drainages the taxa inhabit (Smith et al. 2002). Thus, the most recent biogeographic event may not be the event responsible for the initial vicariance, and historical inferences based only on extant taxa may be misleading

CYPRINIDAE: CARPS AND MINNOWS

because of a tendency to consider the most recent events as responsible for diversity (G. R. Smith 1981). Clearly, an accurate assessment of biogeographic history requires well-supported phylogenies and molecular clocks well calibrated against the fossil record.

MORPHOLOGY Cyprinids, despite their huge diversity, maintain a relatively conserved overall body plan. A single dorsal fin, whose anterior-most ray may be enlarged and stiff, usually sits midway along the body. The pelvic fins are also typically midway along the body, and an anal fin slightly smaller than the dorsal sits between the pelvic fins and the caudal fin. The caudal fin of most cyprinids is widelobed with rounded edges, erupting from a caudal peduncle larger than typical of most other groups of fishes. Mouth position ranges from inferior to terminal. All species have cycloid scales, although these are sometimes minute and embedded in the skin, and many have a silvery appearance. A lateral-line canal can be seen on the flanks of most species, and sometimes the sensory canals are expanded on the head. Because of so many conserved features, cyprinids often are considered to be morphologically monotonous, a misconception that has impeded the understanding of the functional and evolutionary biology of these organisms. Most North American cyprinids are small (14°C. This variation in CTM was >60% of the seasonal ambient temperature range (21.5°C). In winter, the CTM of 17.5°C was lower than ambient summer water temps of 19–21.5°C. The Red Shiner, Cyprinella lutrensis, is abundant across the central United States and thrives in seemingly harsh habitats. Matthews (1986d) conducted a large-scale geographic study of the CTM over an 1,100 km north-south span of the range of the Red Shiner. No statistically significant geographic trends emerged in CTM, and thus, he hypothesized that CTM is conserved within the species. This is interesting in light of the findings on the Red Shiner that high thermal instability in some areas of the Brazos River, Texas (as a result of Sheppard Dam), corresponded to high genetic heterozygosity at the Mdh-B locus of Red Shiners and in significantly different proportions than HardyWeinberg equilibrium would predict (Zimmerman & Richmond 1981; see genetics section). This was interpreted as an adaptive response to the unstable temperatures of that area. Similarly, the preferred temperature of Red Shiners also is plastic and subject to selection (Calhoun et al. 1982). Though the Cyprinidae are not the most thermally tolerant group of fishes in North America, the range of CTM observed in the family is impressive, and the most tolerant cyprinids are more tolerant of high temperatures than most other North American freshwater fishes (Beitinger et al. 2000).

Salinity Tolerance Cyprinids have long been classified as primary freshwater fishes (i.e., restricted to fresh water; Myers 1938). Salinity tolerances are fairly conserved within the family Cyprinidae; few species can tolerate any appreciable

salinity. Most of these stenohaline fishes are not found in areas where salinity is >3 ppt (Young & Cech 1996). The Mojave Tui Chub, Siphateles bicolor mohavensis, shows increased tolerance of just over 10 ppt (McClanahan et al. 1986). Nevertheless, the Smalleye Shiner, Alburnops buccula, Sharpnose Shiner, “Notropis” oxyrhynchus, and Plains Minnow, Hybognathus placitus, have LC50 levels >15 ppt (Ostrand & Wilde 2001). This is higher than normal for members of this family. These heightened tolerances may help explain distributions and persistence in areas such as the Brazos River, Texas, where evaporation in pools can create temporarily increased salinities (Ostrand & Wilde 2001). Even higher salinity tolerance is observed in two other North American cyprinids, the Splittail, Pogonichthys macrolepidotus, and the Peamouth Chub, Mylocheilus caurinus. Since these are sister-species (Simons & Mayden 1999), increased salinity tolerance likely was inherited from a common ancestor. Splittails can tolerate salinities of >28 ppt (Young & Cech 1996) and occur naturally in environments with salinities ≤18 ppt (Meng & Moyle 1995). Higher apparent sensitivity of young fish than adults to salinity may be a function of body surface area:volume ratio or incomplete development of the osmoregulatory system (Young & Cech 1996). The curious distribution of the Peamouth Chub is the likely result of migration through brackish water (Clark & McInerney 1974). The species is native to rivers and lakes of the Pacific Northwest from the Columbia River, Oregon and Washington, north to the Nass River, British Columbia, including fjords that do not contain other primary freshwater fishes (Fig. 12.37). It is the only cyprinid species present on Vancouver Island off the coast of British Columbia. The species can move through brackish water and occurs in the sea off Spanish Banks, Vancouver, British Columbia (Carl et al. 1967). Their presence in areas that have no other cyprinids or primary freshwater fishes likely involved dispersal through brackish water during floods or other periods of high runoff. This dispersal may have occurred during glacial retreat at the end of the Pleistocene glaciations. The large amount of meltwater produced by glaciers may have created a brackish water corridor for migration. Heightened salinity in fresh waters can cause stress to fish, altering other physiological parameters. Toepfer & Barton (1992) investigated the effect of salinity on oxygen consumption in the Southern Redbelly Dace, Chrosomus erythrogaster, compared with the Northern Studfish (Fundulus catenatus). The authors chose these taxa because the

CYPRINIDAE: CARPS AND MINNOWS

Southern Redbelly Dace is not closely related to brackish or marine fishes, but the Northern Studfish is related to euryhaline species. At higher salinities (4 and 10 ppt) both species increased the rate of oxygen consumption, suggesting a higher metabolic rate. Nevertheless, the effects of salinity on Southern Redbelly Dace were significantly greater than those on Northern Studfish, suggesting that phylogenetic inertia plays a role in the osmoregulation and heightened metabolism of these fish.

Acidification Tolerance Acidification of water bodies has become an important issue in recent years, especially as evidence mounts that acidification can be anthropogenic. Agriculture, air pollution, and other anthropogenic causes can acidify lakes and rivers, resulting in increased stress to fishes and other inhabitants. As lakes begin to acidify, the fraction of cyprinid species present is one of the first aspects of the ecosystem to change, which can make them good bioindicators of anthropogenic change to an ecosystem (Matuszek et al. 1990). Declining populations of Fathead Minnows, Pimephales promelas, Common Shiners, Luxilus cornutus, Bluntnose Minnows, Pimephales notatus, and Blacknose Shiners, “Notropis” heterolepis, are good indicators of declining ecosystem health as a result of anthropogenic acidification of lakes and rivers (Mills & Schindler 1986; Matuszek et al. 1990). The lower tolerance of minnows for acidity is typically pH 4–5 (Matthews & Hill 1977), which may be one explanation for their use as an early warning sign of acidification. Upper pH levels of nearly 11 can be tolerated by the Klamath Tui Chub, Siphateles bicolor bicolor (Falter & Cech 1991).

Turbidity Tolerance Turbidity (suspended sediment particles in the water column) is another factor that can negatively affect minnows (e.g., foraging, reproduction), likely via respiratory or visual impairment (Vinyard & Yuan 1996; Sutherland & Meyer 2007; Hazelton & Grossman 2009). Minnows are among the most tolerant of fish groups to turbidity (Trebitz et al. 2007) but can still show signs of stress at even miniscule (25 mg/l) concentrations of suspended sediments (Sutherland et al. 2008). In a laboratory study of turbidity and competition effects, Rosyside Dace, Clinostomus funduloides, a species of clear, flowing streams, exhibited reduced reactive distance to prey (3.5 cm/10 nephelometric turbidity unit, NTU, increase,

427

range 10–30 NTU) in turbidity, but turbidity interacted complexly with prey capture when trials included intraand interspecific competitors (i.e., Yellowfin Shiners, Alburnops lutipinnis). Interestingly, the trials suggested increased turbidity decreased prey capture success forward of the fish’s location, possibly representing an overall increase in energy costs associated with foraging (Hazelton & Grossman 2009), which could affect growth and overall fitness. Similarly, feeding trials with the Lahontan Redside, Richardsonius egregius, revealed a strong linear decrease in predation rates on zooplankton prey with even slight increases in turbidity (about 95% prey capture at 3.5 NTU to 20% at 25 NTU with 1.7 mm prey) (Vinyard & Yuan 1996). Turbidity decreases the effort that the Whitetail Shiner, Cyprinella galactura, invests in reproduction as well as the number of propagules spawned (Sutherland 2007). As is typical of other physiological tolerances mentioned here, minnows show diversity in their ability to tolerate turbidity. Comparative data for the Whitetail Shiner and Spotfin Chub (Erimonax monachus) show that Spotfin Chubs are less resilient to turbidity than Whitetail Shiners. Some authors suggest low turbidity tolerance of the Spotfin Chub is responsible for its declining numbers, leading to its Threatened status (Sutherland & Meyer 2007; Sutherland et al. 2008) (see also ecology and conservation sections).

Oxygen Tolerance Low amounts of dissolved oxygen (hypoxia) also can negatively affect minnow species. As with other tolerance metrics discussed above, we know that a diversity of physiological requirements exists among species in this large clade, but we do not yet fully understand how physiologically diverse the clade is. The Red Shiner, Cyprinella lutrensis, can survive (though stressed) at dissolved oxygen levels between 0.9 and 1.5 ppm. Above that level, this species is not stressed. Red Shiners can survive (though stressed) for hours at 1.0 ppm. In general, these oxygen requirements are not impressively high or low. For example, the Black Bullhead (Ameiurus melas) and Pumpkinseed (Lepomis gibbosus) can survive at dissolved oxygen levels much 113,000 eggs in a 21-month laboratory test under 24-h light conditions (Gale 1986). Most species of minnows lay 800–1,500 eggs/season (Etnier & Starnes 1993). Variation in fecun-

CYPRINIDAE: CARPS AND MINNOWS

dity is not only observed in distantly related species. Bluntnose Minnows, Pimephales notatus, lay 200–500 eggs/season (Hubbs & Cooper 1936), and a close relative, the Fathead Minnow, P. promelas, lays 4,000–5,000 eggs/ season (Markus 1934). A female Fallfish, Semotilus corporalis, contained >12,000 eggs immediately before spawning (Reed 1971).

Broadcast and Crevice Spawning Spawning modes in cyprinids are diverse and interconnected. Some cyprinids prepare the substrate for spawning, but others do not. Those species with no substrate preparation have two spawning modes: broadcast spawning and crevice spawning. Broadcast spawning is considered the ancestral state for breeding in cyprinids (Page & Johnston 1990ab; Mayden & Simons 2002). This mode involves scattering of gametes over the substrate and occurs in most species of North American minnows (Johnston & Page 1992; Tables 12.2–2.5). Six species in the Rio Grande (Hybognathus placitus, Hybognathus amarus, Rio Grande Silvery Minnow, Macrhybopsis aestivalis, Notropis girardi, Notropis jemezanus, Rio Grande Shiner, and “Notropis” simus, Bluntnose Shiner) compose a reproductive guild of pelagic broadcast spawners. During increases in stream flow, these pelagic spawners broadcast semi-boyant eggs that remain suspended in current during development (Dudley & Platania 1999). Several species of cyprinids are nest associates, spawning over the nests of other fishes (see nest associates subsection). In general, broadcast-spawning males do not occupy territories (Raney 1939a). In the Northern Pikeminnow, Ptychocheilus oregonensis, groups of ≤8,000 individuals (males outnumbering females 50M:1F to 200M:1F) congregate and swarm over coarse gravel where spawning occurs (Patten & Rodman 1969). Most groups of broadcast spawners are not nearly as large. Generally, a female and one to several males will break away from the spawning swarm, and one male clasps the female. Sperm and eggs are released by the pair and the surrounding males. In the Spikedace, Meda fulgida, the spawning group (one female and one to several males) makes abrupt vertical dashes during the spawning act, led by the female, touch the surface of the water, and then make a steep dive toward the substrate. Some members of the group actually touch the substrate (Barber et al. 1970). In some broadcasting cyprinids, such as the Eastern Blacknose Dace, Rhinichthys atratulus, the pair vibrates during spawning while resting on the sub-

433

strate, which can result in burying of the eggs (Johnston & Page 1992). Crevice spawning occurs in four genera: Cyprinella, Codoma, Erimonax, and Tampichthys. Here we describe reproduction in Cyprinella (Fig. 12.76cd), although our description is consistent with accounts of Codoma, Erimonax, and Tampichthys. Male Cyprinella defend territories over submerged logs with loose bark, fissures in large rocks, spaces between adjacent rocks, or other crevices for egg deposition (Pflieger 1965; Gale 1986; Mayden & Simons 2002). The males are tenacious defenders of their territories. Territorial male Spotfin Shiners, C. spiloptera, defend their respective crevices using their mouths to grab threatening males by the pelvic or anal fins and drag them away from the crevice area (Gale & Gale 1977). Males display by making passes across the crevices and leading females to the territory (detailed by Stephens & Mayden 1998; Mayden & Simons 2002). Male Spotfin Shiners usually extend a pectoral fin into the crevice while making these display passes (Gale 1986). A female enters the territory, and the pair swims together through the crevice with their vents oriented toward the crevice, vibrating during the spawning act itself as the eggs are deposited into the crevice (Fig. 12.76d). This may be repeated several times by the same pair before the female leaves. Males may mate with multiple females, and spawning crevices may be used sequentially by multiple males.

Substrate Preparers, Clumpers, and Clusterers Many species of cyprinids practice careful preparation of the substrate before breeding. About 8% of North American minnows build nests for spawning not including crevice, clump, or cluster spawners (Johnston 1989). Species that prepare the substrate can build saucershaped, pit, and ridge nests with pits traversing the ridges. Others build gravel mounds, and some species clump or cluster their eggs in specific pre-cleaned cavities. In North American cyprinids, only the male prepares spawning sites. Saucer building occurs only in the Longfin Dace, Agosia chrysogaster (Minckley & Barber 1971). Males construct and maintain large, saucer-shaped depressions on sandy substrates. These depressions can occur in relatively high densities, 25/m2, and can reach 25 cm in diameter and 6 cm in depth. A rim around the saucer extends about 1 cm above the substrate. Males do not seem to defend territories, although they will display while over a nest.

434 FRESHWATER FISHES OF NORTH AMERICA

In spawning, one to four males align next to a receptive female, but only one or rarely two actually mate with her. The pair or trio vibrates close to the substrate during gamete release and raise a cloud of silt that buries the eggs. After spawning is complete, neither parent guards the nest or provides any parental care. Members of Campostoma and some members of Luxilus spawn in pits that males dig in gravel. Males move gravel aside with their snouts and mouths (in Campostoma). In some cases, Luxilus males (Figs. 12.76ab and 12.81) will breed as nest associates with species of Nocomis (see nest associate subsection). The pits are generally shallower than the saucer depressions of male Agosia. Though male Campostoma defend territories (Fig. 12.77c), they often switch between pits. The dominant male positions himself on the upstream edge of the pit; groups of females stay outside of the pits, entering individually when ready to spawn. Once a female enters the pit, one or more males align with her, they all vibrate, and gametes are subsequently released (Miller 1962). Species of Campostoma also spawn over nests of other cyprinids as nest associates (see nest associates subsection) (Fig. 12.77b). Within Luxilus, at least three members (Common Shiner, L. cornutus, Duskystripe Shiner, L. pilsbryi, and Bleeding Shiner, L. zonatus) nest in pits. Several other species (Crescent Shiner, Luxilus cerasinus, Warpaint Shiner, L. coccogenis, White Shiner, L. albeolus, and Bandfin Shiner, L. zonistius) spawn as nest associates. Pit-digging males defend territories and display over the pits for females. When a female enters the nest, the male tilts to one side; the female lays on the substrate. The male then clasps her with his curved body and pushes her into the substrate. Upon vibration, the pair releases gametes and departs (Gleason & Berra 1993; Maurakis & Woolcott 1993). Perhaps the most peculiar reproductive mode in North American minnows is the pit-ridge building by male Semotilus. In these fishes, males construct nests in about 0.5 m of water (Ross & Reed 1978). A male excavates a pit in the gravel substrate and piles all of the excavated stones directly upstream of the pit. Males can move individual stones weighing ≤168 g in the excavation process (anecdotal accounts suggest weights of 500 g). An entire nest can weigh ≤80 kg (Reed 1971); Semotilus males rarely reach 1.6 kg in body weight. A receptive female can apparently only enter the nest area when the nest-building male is distracted by chasing another fish or building the nest itself, or he will chase her out as well (Ross 1976). Once over the nest pit, the male clasps the female by placing his pectoral fin under her anterior ventral surface and his

caudal peduncle over her dorsal surface, twisting his body into a U shape. The female is forced into a nearly vertical position with her tail pointed toward the substrate immediately before gamete release. After the spawning bout is finished, the female typically floats belly up with the current for a short distance. She then proceeds to mate again with either the same or another male. Male Fallfish, Semotilus corporalis, do not use a spawning clasp. As a result, access to the female is not monopolized by a single male, and sneaker males, non-nest-building mature males who attempt copulation on other males’ nests, are common (Ross & Reed 1978). Immediately after spawning, the male covers the pit (and eggs) with rocks from the downstream edge of the pit. As he continues to spawn, the pit is moved farther and farther downstream with the gravel from the downstream edge of the pit moved to the upstream edge. In this way, a ridge forms as successive spawning pits are filled. This ridge can be >2 m in length and 0.6 m in height (Reed 1971), an impressive feat for a species whose maximum recorded length is 0.5 m (Page & Burr 2011). The male guards the nest aggressively against other males and potential predators until hatching (Reed 1971; Ross 1977a; Johnston & Page 1992). On occasion, a large male may challenge the nest-building male, and the two begin a rapid upstream parallel swim. During this 1- to several meter swim, the males periodically push against each other. Finally, the winning male swims back to the nest, and the losing male swims away. This parallel swim as well as lateral displays and head butting are used to establish a hierarchy among the breeding males in a population (Ross 1977ab). About 10% of mature males build nests in any given year; the remaining males either do not spawn or act as sneaker males, stealing chances to mate with a female on the nest of another male. Interestingly, despite their aggressive territoriality, guarding males tolerate Blacknose Dace, Striped Shiners, Luxilus chrysocephalus, Common Shiners, L. cornutus, Blackside Dace, Chrosomus cumberlandensis, and White Suckers (Catostomus commersoni), all of which are known nest associates of Semotilus (Ross 1977a; see nest associates subsection; Fig. 12.56g). Males may use the same spawning site within and among spawning seasons (Ross & Reed 1978). Species in the genera Nocomis and Exoglossum are mound builders in which males build large mounds of gravel with their mouths in slow- to moderately flowing water (Lachner 1952; Raney 1939b; Van Duzer 1939). A single male builds a nest about 0.5 m in diameter out of pebbles brought from ≤25 m from the nest (Fig. 12.78). If uninterrupted, a male can build the nest in a single day.

CYPRINIDAE: CARPS AND MINNOWS

Multiple males can work simultaneously on the same nest (Johnston 1991). Mound construction consists of four stages: a cavity is excavated; a platform 0.5 m in diameter is built over the excavation; a mound is constructed over the platform; and finally, small pits or troughs are dug by the male on top of the mound (Maurakis et al. 1991a; Johnston & Page 1992). Hornyhead Chub, Nocomis biguttatus, mounds can consist of a collection ≤300 pebbles, with a combined weight of 11 kg (Wisenden et al. 2009b). Males clasp females by placing the pectoral fin on the ventral side of the female and the caudal peduncle on the dorsal side of the female. On occasion two males clasp the same female simultaneously (Maurakis et al. 1991a); satellite or sneaker males may fertilize eggs before the nest-building male (Sabaj et al. 2000). Male Central Stonerollers, Campostoma anomalum, may accompany a female Bluehead Chub, Nocomis leptocephalus, and get caught in the spawning clasp of a male Bluehead Chub (Fig. 12.78ab; Sabaj et al. 2000). Aside from these anomalous occurrences, generally a single male will guard his nest against other males. Males in Nocomis generally cover the eggs with gravel after spawning and continue to guard the nest until hatching (Johnston 1991). Egg clustering coupled with intense parental care occurs only in Codoma, Pimephales, and Opsopoeodus (Gale 1983; Page & Ceas 1989; Minckley & Vives 1990; Page & Johnston 1990a; Mayden & Simons 2002; Table 12.4ac). Spawning in these fishes is initiated when a male establishes a territory under a horizontally flattened rock (or perhaps another solid surface). In Pugnose Minnows, O. emiliae, males rapidly raise and lower the dorsal fin immediately before spawning occurs. The guardian male lines up broadside and head-to-head with the female, and the pair makes several passes under the rock, laying and fertilizing several eggs during each spawning pass (Page & Johnson 1990a). Egg deposition in Pimephales involves females attaching eggs to the nest-stone with a rapid lateral undulation; the male uses his body to press the female’s side against the roof of his territory, where she deposits the eggs (detailed by Page & Ceas 1989). The male remains in his territory, and several females may sequentially mate with him, leaving the eggs in his care. Males will continue to guard eggs until they hatch.

Spawning Mode and Male-Male Competition Given such a high diversity of spawning behaviors, one would expect some spawning modes to show greater male-male competition than other modes. Pyron (2000)

435

hypothesized that due to greater sperm competition between males of spawner groups (i.e., broadcast, saucer building) compared with pair-wise spawners (i.e., egg clustering, pit building, mound building), the former would have a greater testes:body mass ratio. Nevertheless, no significant differences were detected in regressions of testes mass against body mass for species with different spawning modes. The influence of spawning mode on male-male competition is not yet fully resolved.

Nest Associates Many species of minnows, especially broadcast spawners, are nest associates, using the nests of other fishes, usually species of Nocomis (Figs. 12.77b and 12.78) or Sunfishes (Centrarchidae). Nest association is known in only three non-cyprinid taxa: Gars (Lepisosteidae) (Goff 1984), Lake Chubsuckers (Erimyzon sucetta), and Creek Chubsuckers (Erimyzon oblongus) (Page & Johnston 1990b). Interestingly, the Creek Chubsucker will breed over nests constructed by Central Stonerollers, Campostoma anomalum, and Creek Chubs, Semotilus atromaculatus. Most species of nest-associating minnows spawn on nests constructed by other cyprinids, but some minnow taxa spawn in non-cyprinid nests. Golden Shiners, Notemigonus crysoleucas, spawn over active nests of Largemouth Bass (Micropterus salmoides) and may account for ≥78% of all eggs in the nest (Kramer & Smith 1960). Bluehead Shiners, Pteronotropis hubbsi, Bluenose Shiners, P. welaka, Golden Shiners, Notemigonus crysoleucas, Dusky Shiners, Hudsonius cummingsi, and three species of Miniellus (Topeka Shiners, Swallowtail Shiners, and Dusky Shiners) spawn over centrarchid nests (Carr 1946; Chew 1974; Noltie & Smith 1988; Fletcher & Burr 1992; Johnston & Page 1992; Fletcher 1993; Shao 1997ab; Johnston & Knight 1999; Boschung & Mayden 2004). Redfin Shiners, Lythurus umbratilus, spawn over Green Sunfish (Lepomis cyanellus) nests (Hunter & Wisby 1961). Odors from milt and ovarian fluids are the primary attractant of Redfin Shiners to the nests of Green Sunfish, rather than the nest itself (Hunter & Hasler 1965). In all these cases, the centrarchids only rarely chase off the minnow species, although the nests are guarded vigorously by the male Sunfish from other species of fishes, especially other centrarchids. In several species of Luxilus, the nests of Bluehead Chubs, Nocomis leptocephalus, are used as a spawning substrate without modification. The White Shiner, L. albeolus, however, does alter the nest slightly; males dig furrows in

436

FRESHWATER FISHES OF NORTH AMERICA

the nest (about the length of the male) using their snouts. The male holds a territory around the furrow and waits for a receptive female. Male Crescent Shiners, Luxilus cerasinus, in contrast, do not dig furrows but still maintain territories on the mound (Maurakis & Woolcott 1993). The Rainbow Shiner, Alburnops chrosomus, and the Rough Shiner, A. baileyi, also spawn in depressions in the nests of the Bluehead Chub; however, they use pre-existing depressions in the nest rather than creating them (Johnston & Kleiner 1994). These behaviors are typical of many nestassociating minnows, such as the Yellowfin Shiner, A. lutipinnis (McAuliffe & Bennett 1981; Wallin 1989), and Bandfin Shiner, L. zonistius (Johnston & Birkhead 1988).

Advantages and Disadvantages of Nest Association Since nest association is common in North American minnows, it clearly confers some advantages to the associate. Surprisingly, even potential predators of the nest associate can occasionally serve as host to nest-associating minnows. Bowfin, Amia calva, which typically prey on minnows and other small fishes, can act as host to Golden Shiners, Notemigonus crysoleucas, and it is unknown why the host fish do not consume the spawning minnows or their offspring (Katula & Page 1998). The association of the Dusky Shiner, Hudsonius cummingsae, is verifiably detrimental to its host, the Redbreast Sunfish (Lepomis auritus), because the shiner actually consumes the host’s eggs (Fletcher 1993). Egg-eating leeches can occur in spawning nests that would presumably consume host and associates alike, thus encouraging hosts to allow associate spawning (Light et al. 2005). This mutualistic dilution has been argued as an explanation for protection in multiple host species (Johnston 1994; Shao 1997). Nest association can also cause direct hybridization in some minnow species (Scribner et al. 2000). Nest association in North American minnows has evolved several times with several hosts (Johnston & Page 1992; Mayden & Simons 2002). Invasive, non-native nest associates are competitively excluded from prime spawning sites by native nest associates (Herrington & Popp 2004), which is indicative of the tight coupling of host and associate. The precursor to nest association likely is broadcast spawning, although some species, such as members of Campostoma, are known nest associates but generally construct pits for spawning, suggesting plasticity in spawning mode. Several theories attempt to explain the selective pressure driving a shift to nest association as

well as the selective pressure on the host species to allow for nest associates to mate over the nest (Johnston 1991). One obvious advantage to the associate is parental care without parental investment. The associate can deposit eggs on clean, well-aerated substrate, and the nest receives protection against predators from the host (Shao 1997). The associate parents do not invest any effort in either preparation of the substrate or protection of the eggs or fry after spawning. According to the selfish-herd hypothesis (Hamilton 1971), nest association could increase survivorship because eggs of the associate are less likely to be eaten by egg predators because other eggs are also available. McKaye (1981) discussed mutual benefits of group spawning aggregations in his adaptive or mutualistic hypothesis. He described how, unlike brood parasites such as birds (e.g., cuckoos, Cuculidae), the alien or associate young do not remove or displace the host young, attacks by predators lead to only partial loss of brood (arguably less than without the associate brood), and the extra energetic input by the host species needed is low or nonexistent. Nests could also act to synchronize spawning events. For example, the Redfin Shiner, Lythurus umbratilus, is attracted to the milt and ovarian fluid of the Green Sunfish (Lepomis cyanellus) with which it is a nest associate. This olfactory attraction, rather than the physical nest itself, should attract minnows from greater distances, resulting in a widespread signal of breeding time and location (Hunter & Hasler 1965; Johnston & Page 1992). The use of host species’ milt to induce spawning in the associate Blackside Dace, Chrosomus cumberlandensis, has allowed for captive breeding that is otherwise impossible in this species (Rakes et al. 1999). Another interesting issue of nest associate recognition is that, although hosts with nest associates voraciously guard their nests from potential predators, nest associates are not chased away (e.g., Kramer & Smith 1960; Hunter & Wisby 1961; Ross 1977a). One possible explanation is that males of the host species may be energetically incapable of driving away nest associates (Johnston 1991). The defense of the nest may simply come to a point of diminishing returns; protection afforded the nest by chasing away potential spawning associates is negligible when compared with the protection with that energy focused on chasing away potential egg predators.

Evolution of Spawning Modes As noted, broadcast spawning generally is accepted as the plesiomorphic mode in North American minnows (John-

CYPRINIDAE: CARPS AND MINNOWS

ston & Page 1992) as well as in the shiner clade (Mayden & Simons 2002). Evidence indicates crevice spawning evolved directly from broadcast spawning (Johnston & Page 1992; Mayden & Simons 2002). This evolutionary shift occurred independently at least three times in the shiner clade alone (Mayden & Simons 2002). Pit building is most likely derived from broadcast spawning as well, but the ancestral state of saucer building in Agosia is equivocal (Mayden & Simons 2002). Pit-ridge building in Semotilus may be derived from pit building (Johnston & Page 1992). Differences in behavior suggest two separate origins of mound building for Nocomis and Exoglossum; Exoglossum does not perform the initial excavation stage seen in Nocomis (Johnston & Page 1992). Egg clustering, observed in Codoma, Opsopoeodus, and Pimephales, is hypothesized to be derived from crevice spawning due to the relative similarity of those spawning modes (egg attachment, male territorial behavior, and male parental care) as opposed to broadcast spawning (Johnston & Page 1992). Nevertheless, phylogenetic support for this hypothesis is weak, and the ancestral state for the derivation of egg clustering is, at best, equivocal (Mayden & Simons 2002). Without greater taxon sampling, phylogenetic resolution, and intense study of reproductive behaviors, theories on the evolution of reproductive modes will remain speculative, especially due to the homoplastic nature of many of these behaviors. In addition, categories such as “broadcast spawning” likely need further subdivision (e.g., broadcast spawning with and without male territoriality), all of which have not been sufficiently investigated. One major difficulty with several studies of the evolution of spawning mode is that they were not conducted with reference to a well-resolved, explicit phylogenetic hypothesis because none were available. Use of reproductive characters alone to understand their evolution will underestimate the number of independent origins of a particular behavior. An understanding of the evolution of spawning mode will only be reached when investigated in the context of other phylogenetically informative characters.

Egg Characteristics Many minnow species spawn multiple clutches (Heins & Rabito 1986; Johnston & Kleiner 1994). Heins & Rabito (1986) defined a clutch as comprising propagules that undergo synchronized development and observed multiple events of synchronous development in a single season in two species of minnows from the genus Alburnops. Clutch

437

size often is correlated with female length (Heins & Baker 1992; see also Haag et al. 2007). Stream runoff significantly is correlated with egg size (after correcting for body size) in the Blacktail Shiner, Cyprinella venusta, such that high runoff appeared to select for larger egg size in these minnows (Machado et al. 2002). In this case, egg size and clutch size were unrelated. In contrast, an apparent tradeoff occurred between clutch size and egg size in the Weed Shiner, Alburnops texanus, with larger eggs (and decreased clutch sizes) being associated with higher mean annual runoff in the Gulf Coastal Plain (Heins & Rabito 1988). Even within the Gulf Coastal Plain, large (150 km), migrations to use prime spawning and feeding habitats. These fish evolved during periods of episodic fluctuations in river levels and are quite capable of surviving under these conditions (Tyus 1986). Most authors agree that construction of dams in the 1960s was the proximate cause of decline of this species. The main detrimental effects of these dams include flowing water habitat loss due to reservoirs, cold flows downstream of dams decreasing suitable habitat and reproductive success, and loss of backwater habitat (due to lack of spring flows and fluctuations) that is used extensively by young individuals (Holden & Wick 1982). A low growth rate coupled with

445

lowered temperature (as a result of dam construction) are likely other key reasons for the decline of these fish (Kaeding & Osmundson 1988). Interestingly, floodplain isolation seems to have served to protect the surviving populations of the Oregon Chub, Oregonichthys crameri (Endangered), which is endemic to the Willamette Valley, Oregon. This minnow was once distributed throughout that valley; however, the Oregon Chub now inhabits 150 km) migrations and was once widespread and common in large rivers and streams of the Colorado River basin. Historically, large specimens were commonly caught in the basin as exemplified by these historical photographs dating to the early 20th century (Quartarone 1995). Today, large individuals are rarely seen, population sizes are dramatically reduced, and the species is limited in distribution by the negative effects of dams and attendant reservoirs as well as by non-native fishes introduced in the basin. The species is protected under the U.S. Endangered Species Act as an endangered species and extensive efforts are underway to prevent extinction of this unique fish. (A) Two large Colorado Pikeminnow hang off a burro. (B) Colorado Pikeminnow caught near Jensen, Utah (Green River drainage), in the 1930s. (C) A farmer carries two Colorado Pikeminnows caught as he irrigated a hayfield in about 1934 (photographs courtesy of the Upper Colorado River Endangered Fish Recovery Program, U.S. Fish and Wildlife Ser vice, Denver, Colorado).

Slender Chub, Erimystax cahni (Endangered), historically occurred in the Holston, Clinch, and Powell Rivers in Tennessee. Populations in both the Holston and Powell Rivers now are extirpated, probably as a result of industrial pollution (Jenkins & Burkhead 1994). The Turquoise Shiner, Erimonax monachus (Threatened; Fig. 12.56a), is another endemic of the Tennessee River drainage. This

CYPRINIDAE: CARPS AND MINNOWS

species disappeared from most of its range; however, the population in the North Fork Holston River is rebounding, presumably as a result of pollution abatement in Saltville, Virginia, where previous pollution had lowered its abundance (Jenkins & Burkhead 1994; Fig. 12.21). The Blue Shiner, Cyprinella caerulea (Endangered), was historically endemic to the Cahaba and Coosa River systems within the Mobile Basin (Boschung & Mayden 2004). Currently, six extant isolated populations remain in the upper Coosa River in northeast Alabama, northwest Georgia, and southeast Tennessee. The Cahaba River populations are no longer extant. Extirpation from the Cahaba River occurred as nitrification increased from high sewage loads (Stephens & Mayden 1999). Sediments also can harbor chemical pollutants that may impact cyprinids. Lake Pinchi, a large lake (>5,300 ha) in northern British Columbia, was mined for mercury from 1940 to 1944, and the fishes of that region continue to exhibit effects of contamination. Northern Pikeminnows, Ptychocheilus oregonensis, have methylmercury concentrations about 2–4 times higher than Northern Pikeminnows from similar, but unpolluted, nearby lakes and higher methyl-mercury concentrations than similar size, sympatric Rainbow Trout (Oncorhynchus mykiss). Body size and methyl-mercury concentration are correlated in the Northern Pikeminnow but not in the Rainbow Trout. This is attributed to the higher trophic position and slower growth rate of the Northern Pikeminnow, resulting in biomagnification of the mercury (Weech et al. 2004). Rapid biomass accumulation in the Rainbow Trout results in tissue mercury concentrations similar to that of prey items, but Northern Pikeminnows do not convert prey biomass as efficiently, resulting in heightened levels of methyl-mercury (Weech et al. 2004).

Non-Native Species The introduction of non-native species has profound, usually negative, effects on native species. Introduced predators can have severe and long-lasting impacts on native fauna. Cyprinids are no exception. This can be particularly acute if the native taxon is naive to an introduced predator. The Little Colorado Spinedace, Lepidomeda vittata (Threatened), occurs in disjunct populations in northern Arizona and is extant in only three counties where it occupies a wide variety of habitats. The abundance and range of these fish has diminished since the introduction of Rainbow Trout in the mid-1800s (Blinn et al. 1993). Experiments demonstrated high Rainbow Trout

447

predation on the Little Colorado Spinedace, which exhibited almost no predator avoidance (Blinn et al. 1993). The introductions of Rainbow Trout may be partly responsible for the scattered, relatively small distribution of the Little Colorado Spinedace. Centrarchid invasives also have detrimental effects on native fish populations. In the Umpqua River system in the Pacific Northwest, the Smallmouth Bass (Micropterus dolomieui) is implicated in the decline of the Umpqua Chub, Oregonichthys kalawatseti (Vulnerable). Smallmouth Bass were stocked in Oregon in 1924 and 1925 (Simon & Markle 1999) but did not gain access to the Umpqua River system until 1964 during a flooding event, after which they became distributed across the Umpqua River basin by the late 1970s. This was followed by a substantial decline in the Umpqua Chub, which by 1998 disappeared from 50% of the sites it previously occupied, while Smallmouth Bass abundance increased at all spots from which Umpqua Chubs were extirpated. Other species in the system, such as the Umpqua Pikeminnow, Ptychocheilus umpquae, disappeared from 13% of formerly occupied sites (Simon & Markle 1999). Introduced predators such as the Northern Pike (Esox lucius), Largemouth Bass (Micropterus salmoides), and Smallmouth Bass can have a dramatic influence on species richness, reducing the species richness of native cyprinids by as much as two-thirds, compared with similar, predator-free lakes (Findlay et al. 2000). Introduced sightfeeding predators have also been implicated in the disappearance or reduction in numbers of native cyprinids in rivers (see previous two paragraphs). Introduced North American species may also compete with native cyprinids. Red Shiners, Cyprinella lutrensis, have been introduced into the Colorado River system (see diversity and distribution section) as forage for gamefishes and has become widespread. In the Gila River, a tributary to the Colorado River, Red Shiners overlap in range with Spikedace, Meda fulgida (Endangered). The two cyprinids share similar spawning habitat and have general (though not exact) habitat overlap. Red Shiners are more tolerant of turbidity, high temperatures, and low dissolved oxygen. These factors make Red Shiners an effective competitor, especially against Spikedace, and may be influential in restricting its range (Rinne 1991). Introduced parasites pose a threat to native cyprinids. The western cyprinids Roundtail Chub, Gila robusta (Vulnerable), and Woundfin, Plagopterus argentissimus (Endangered), e.g., are greatly affected by Bothriocephalus acheilognathi, an Asian tapeworm. Red Shiners, Cyprinella

448 FRESHWATER FISHES OF NORTH AMERICA

lutrensis (an invasive species in that area), may have carried B. acheilognathi from other parts of the country, where the tapeworm was originally introduced from Asia via Grass Carp, Ctenopharyngodon idella (Heckmann et al. 1986). Bothriocephalus acheilognathi generally has two hosts, a copepod and a cyprinid; the tapeworm can block the gastrointestinal tract, perforate the intestine, and destroy the internal mucosa. They generally concentrate in the anterior portion of the gastrointestinal tract. Dense infestations can result in reduced growth, deformities, suppressed swimming, and lowered reproduction. The number of tapeworms infesting an individual is negatively correlated with fish length and weight (Brouder 1999). This parasite is generally not alone when infesting fish species; in addition to the Asian fish tapeworm, six other parasite species occurred in specimens of Woundfin (Heckmann et al. 1986). Bothriocephalus acheilognathi occurred in every cyprinid sampled in Arizona and Nevada reaches of the Virgin River (Heckmann et al. 1993). Interestingly, despite the presence of infested host fishes in tributaries of the Colorado River, B. acheilognathi is unknown from hosts in the main stem of the river. This may be due to lower temperatures in the Colorado River compared with those of its tributaries because B. acheilognathi requires water temperatures >20°C (Brouder & Hoff nagle 1997). Invasive species can also impact bioaccumulation of toxic substances as exemplified by the association of the non-native Overbite Clam, Potamocorbula amurensis, and the native Splittail, Pogonichthys macrolepidotus (Endangered). The Overbite Clam is an invasive species in the Sacramento River estuary, California. This species, introduced from Asia, has a substantially different physiology from that of native invertebrates. The Overbite Clam was first captured in the estuary in 1986 and currently accounts for more than 95% of the benthic invertebrate biomass in some areas (Carlton et al. 1990). This bivalve possesses a selenium loss rate constant an order of magnitude lower than that of other co-occurring invertebrates and is much higher in selenium concentration than other invertebrate food sources (Stewart et al. 2004; Linville et al. 2002). Although selenium is an essential nutrient, the difference between the nutritional and toxic concentrations is small and at high levels can be toxic to fishes (Hamilton 2004; see physiology section). Splittails feed extensively on these bivalves, particularly when numbers of their natural prey, mysid shrimp, decline, and high selenium concentrations could further endanger Splittail populations (Feyrer et al. 2003).

Naturally Small Populations The Moapa Dace, Gila coriacea (Endangered), endemic to the upstream-most 4 km of the 40 km long Muddy River, Nevada (formerly the Moapa River), is a spring-dependent species and is protected as Endangered under the U.S. Endangered Species Act. This fish, although locally abundant, has an exceedingly limited range, making it vulnerable to local perturbations. These include a flash fire in 1994 that is believed to have reduced the population from about 500 to 34 individuals (Scoppettone et al. 1998); invasive species, including the Shortfin Molly (Poecilia mexicana) and Blue Tilapia (Oreochromis aurea) (Scoppettone 1993; Scoppettone et al. 1998); and groundwater pumping (Mayer and Congdon 2008). The Desert Dace, Eremichthys acros (Threatened; Fig. 12.19), is also a spring-dwelling endemic, residing in several, isolated, hot springs in Soldier Meadow, Nevada (Vinyard 1996, 1997). These fish occur in waters much warmer than typical of cyprinids, ranging from 13 to 38°C. Alterations to habitat such as water withdrawals for agriculture significantly reduced abundance of this fish (Vinyard 1997). In addition, threats from introduced predators and competitors, such as the Largemouth Bass and Goldfish, Carrasius auratus, are of special concern to the conservation of the Desert Dace (Vinyard 1997). The small extent of habitat (and lowered total population numbers as a result) makes for an especially precarious position of these fish. Perturbations normally absorbable by higher population levels may have a significant impact on population processes in these situations. The Borax Chub, Siphateles boraxobius (Endangered), another example of a range-restricted cyprinid, is endemic to the Alvord Basin, southeast Oregon and northwest Nevada, where it is restricted to Borax Lake and some tributaries. The small range of this fish is a primary concern, but other factors also contribute to its endangered status. The downstream reservoir of Borax Lake, termed Lower Borax Lake, previously sustained the Borax Chub, though reproduction was not confirmed (Williams & Bond 1983). In 1979, the outflow to Lower Borax Lake was diverted, and Borax Chubs are no longer found in the reservoir (Williams & Bond 1983), thus further restricting the habitat available to the species. The Bluehead Shiner, Pteronotropis hubbsi (Vulnerable; Fig. 12.56f), inhabits lowland streams and swamps in Illinois, Arkansas, Texas, Oklahoma, and Louisiana. Though widely distributed, these fish are not abundant (except in spawning aggregations), and the few populations are iso-

CYPRINIDAE: CARPS AND MINNOWS

lated. The only population in Illinois occurred in Wolf Lake, about 443 km from the nearest known population in Arkansas. The Wolf Lake population was extirpated by chemical spills in 1974 and 1979 as a result of train derailments (Burr & Warren 1986a). This disjunct distribution suggests recent habitat alteration (e.g., channelization, dredging, sedimentation), bolstered by the fact that much of the Missouri and Arkansas lowlands were once covered by cypress swamps before settlement of these areas (Fletcher & Burr 1992).

COMMERCIAL IMPORTANCE Commercial exploitation of North American cyprinids is largely limited to the bait industry. Though cyprinids are widely bought and sold in the aquarium industry, nearly all species in that trade are not native to North America. Though acknowledging the beauty, diversity, and lucrative nature of cyprinids in aquaria, we restrict discussion here to North American cyprinids used as baitfish or food. Although several species of cyprinids are sold as bait, the Fathead Minnow, Pimephales promelas, and Golden Shiner, Notemigonus crysoleucas, are among the most widely sold and support a relatively extensive pond culture bait industry. In Arkansas alone, an estimated six billion bait minnows (mostly Golden Shiners and Fathead Minnows) are raised each year and shipped around the country (Stone et al. 2009). The relative ease of commercial pond culture, as well as the popularity and hardiness of these two minnow species, have led to their introduction across the United States well outside of their native range. The Fathead Minnow has been introduced in 24 states and the Golden Shiner in 21 states outside of its natural range (Boydstun et al. 1995). Other species commonly sold as bait include the Spottail Shiner, Hudsonius hudsonius, Sand Shiner, Miniellus stramineus, Hornyhead Chub, Nocomis biguttatus, Creek Chub, Semotilus atromaculatus, Finescale Dace, Chrosomus neogaeus, and Pearl Dace, Margariscus margarita (Meronek et al. 1997b). Estimates of the total value of baitfish sold in six states (Illinois, Michigan, Minnesota, Ohio, South Dakota, and Wisconsin) was >$145 million in 1992 (Meronek et al. 1997a). In 1998, the farm-gate value of baitfish production in Arkansas was estimated at $23 million with an economic impact of 6 to 7 times that amount (Stone et al. 2009). In Ontario, an annual baitfish business reached $12.4 million, which consisted of roughly 11 million dozen fish sold (OMNR 1986). This economic value is likely >$1 billion annually in Canada alone (Litvak &

449

Mandrak 1993). Baitfishes, which are typically cyprinids, are used across North America, so the total commercial importance of North American cyprinids as baitfishes is likely much higher. Large-scale collection for bait has likely played a role in the decline of some species; of the species that are legally sold as baitfishes in Canada, 12 are listed as Vulnerable and 3 as Threatened (Campbell 1992). The only North American cyprinid that is commercially important as a food fish is the Sacramento Blackfish, Orthodon microlepidotus. This species was described originally from specimens obtained from the San Francisco fish market (Ayres 1854). The Sacramento Blackfish is harvested from Clear Lake, San Luis reservoir, and Lahontan reservoir for sale in Asian markets in several California cities (Moyle 2002). In 1988, more than 363,800 kg of Sacramento Blackfish were harvested from the Lahontan reservoir (Sevon 1988). This species is considered a possible candidate for aquaculture because of its high production potential (Cech et al. 1979; Capagna & Cech 1981). Mylocheilus caurinus was served historically as whitefish to customers of hotels in the Columbia River Basin, British Columbia (Jordan & Evermann 1902; McPhail & Lindsey 1970), and Eastern Silvery Minnows are reputed to be tasty when deep fried (Schwartz 1963). Tastes have changed, and sadly, the days when one could enjoy a hearty plate of minnows are long past.

LITERATURE GUIDE Cyprinid Fishes by Winfield & Nelson (1991) provides an excellent overview and introduction to the biology and distribution of the family Cyprinidae. Several regional faunal guides contain excellent information on the biology, distribution, and identification of North American cyprinid fishes. Six of these stand out for the depth of the information they provide and the quality of their illustrations. Freshwater Fishes of Canada (Scott & Crossman 1973) provides detailed information, including biology, distribution, and parasitism of cyprinids in that country. Fishes of Alabama (Boschung & Mayden 2004) is notable for the outstanding illustrations by Joe Tomelleri. The Fishes of Tennessee (Etnier & Starnes 1993) contains excellent color photographs and a well-researched identification key. Freshwater Fishes of Virginia (Jenkins & Burkhead 1994) also contains excellent color photographs of some species, and much interesting information on natural history. Together, the latter three texts describe the biology of the majority of North America’s cyprinid fauna. The

450

FRESHWATER FISHES OF NORTH AMERICA

biology and conservation status of many western cyprinids are reviewed by Moyle (2002) in Inland Fishes of California and McPhail (2007) in The Freshwater Fishes of British Columbia. Freshwater Fishes of Mexico (Miller et al. 2005) provides information on the distribution and identification of the Mexican cyprinid fauna. G. R. Smith (1981) detailed the fossil record of North American fishes and contains a valuable list of references for any reader interested in the distribution and age of cyprinid fossils. Reviews of ecology, systematics, and behavior can be found in Matthews & Heins (1987), Matthews (1998), and Mayden (1992).

Acknowledgments We thank Elizabeth Johnson and Brett Nagle for their help in compiling literature; Mel Warren for his support and countless hours of editing; and Tom Chart and Debra Felker for their time and effort in making available historical photographs of Colorado Pikeminnows in the archives of the Upper Colorado River Endangered Fish Recovery Program. Financial support was provided by the Department of Fisheries, Wildlife, and Conservation Biology and the Bell Museum of Natural History at the University of Minnesota.

Chapter 13

Catostomidae: Suckers Phillip M. Harris, Gregory Hubbard, and Michael Sandel

Catostomids (order Cypriniformes) are commonly called Suckers because these fishes use their downward-directed mouths like vacuum cleaners to suck up small organisms, organic matter, and some detritus. The family name, Catostomidae, is derived from the Latinized Greek roots “kato,” meaning “downward,” and “stoma,” meaning “mouth” (Boschung & Mayden 2004). Mullet is another name commonly applied to catostomids throughout North America, although that name refers to fishes in the family Mugilidae. Suckers inhabit a variety of freshwater ecosystems from large rivers and lakes to small headwater streams, making species of the family important functional components of aquatic habitats (e.g., Li et al. 1987; Schlosser 1987; Gido & Propst 1999; Clarkson & Childs 2000).

DIVERSITY AND DISTRIBUTION Catostomidae consists of 12 genera and at least 76 recent species; 75 species of Suckers occur in North America (Table 13.1; Nelson et al. 2004). These 75 species constitute about 8% of the North American ichthyofauna with only minnows (Cyprinidae, Carps and Minnows) and darters (Percidae, Perches) having more species (Warren et al. 2000). The 2 most speciose genera are Catostomus (finescale Suckers, 26 species) and Moxostoma (redhorse and jumprock Suckers, 22 species), followed by Ictiobus (buffalofishes, 5 species), Chasmistes (lake Suckers, 4 species), Carpiodes (carpsuckers), Erimyzon (chubsuckers), Hypentelium (hog Suckers) and Thoburnia (torrent Suckers), each with 3 species, Cycleptus (blue Suckers, 2 species), and the monotypic genera Deltistes (D. luxatus, Lost River Sucker), Minytrema (M. melanops, Spotted Sucker), Myxo-

cyprinus (M. asiaticus, Chinese Sucker), and Xyrauchen (X. texanus, Razorback Sucker). At least three genera, however, contain recognized undescribed species (e.g., Catostomus, Wall Canyon Sucker, Little Colorado River Sucker; Cycleptus, Rio Grande Blue Sucker; Moxostoma, Sicklefin Redhorse, Apalachicola Redhorse). In addition, several genera (e.g., Catostomus, Carpiodes, Cycleptus, Ictiobus, Moxostoma) have species showing evidence of polytypy. The Summer Sucker (Catostomus utawana), a former subspecies of White Sucker (Catostomus commersonii), is the most recently described species (Morse & Daniels 2009). Kettratad & Markle (2010) recognized the Tyee Sucker (Catostomus tsiltcoosensis) as a distinct species, resurrecting it from the synonomy of Largescale Sucker (Catostomus macrocheilus). Scharpf (2006) reviewed the described and undescribed taxa in Catostomidae. One member of the family, the Harelip Sucker (Moxostoma lacerum), is particularly noteworthy because of its morphological distinctiveness and extinction in historical time. This species was described originally as Lagochila lacera on the basis of the unusual upper and lower lip morphology (Jordan & Brayton 1877; see morphology section). Although Jenkins (1970) regarded Lagochila as a terminal, derived offshoot within Moxostoma, G. R. Smith’s (1992) phylogenetic hypothesis resolved Lagochila embedded within Moxostoma. This species was last collected in 1893, and only 33 specimens are deposited in ichthyological collections (Jenkins & Burkhead 1994). A phylogenetic hypothesis based on gene sequences might provide additional insights into the evolutionary relationships of this unique species, but unfortunately, current molecular methods cannot produce high-quality DNA from formalin-fixed specimens.

Table 13.1. Classification of extant Catostomidae (Suckers) (Harris & Mayden 2001; Harris et al. 2002). Number of species based on Nelson et al. (2004), in part. Information on intraspecific variation (number of species, number of subspecies) including references from Lee et al. (1981) and Scharpf (2006) unless noted. ? = no published study. Superscripted numbers refer to 1Endemic to Yangtze and Minjiang Rivers in China (Gao et al. 2008); 2Sun et al. (2004); 3Kirsch (1889); 4Jordan (1917); 5Berendzen et al. (2003); 6 Jenkins (1970).

Number of Species

Subfamily

Genus

Type Species

Myxocyprininae

Myxocyprinus Gill 18781

Myxocyprinus asiaticus

1

Ictiobinae

Carpiodes Rafinesque 1820b Ictiobus Rafinesque 1820b

Carpiodes cyprinus Ictiobus bubalus

3 5

Cycleptinae

Cycleptus Rafinesque 1819

Cycleptus elongatus

2

Catostomus Lesueur 1817b Chasmistes Jordan 1878 Deltistes Seale 1896 Xyrauchen Eigenmann & Kirsch 18893 Erimyzon Jordan 1876 Minytrema Jordan 1878 Thoburnia Jordan and Snyder 1917 4 Hypentelium Rafinesque 1818 Moxostoma Rafinesque 1820b

Catostomus catostomus Chasmistes liorus Deltistes luxatus Xyrauchen texanus

Catostominae Tribe Catostomini

Tribe Erimyzonini Tribe Thoburnini

Tribe Moxostomatini

Number of Subspecies

Other Geographic or Phylogeographic Structure? Yes2

2

Yes Yes Yes

26 4 1 1

14 2

Yes No No No

Erimyzon oblongus Minytrema melanops Thoburnia rhothoeca

3 1 3

4

? ? No

Hypentelium nigricans

3

Moxostoma anisurum

22

Yes5 2

Plate 13.1. Longnose Sucker, Catostomus catostomus

Plate 13.2. Silver Redhorse, Moxostoma anisurum 452

Yes6

Plate 13.3. Smallmouth Buffalo, Ictiobus bubalus

Plate 13.4. June Sucker, Chasmistes liorus

Plate 13.5. Quillback, Carpiodes cyprinus 453

454 FRESHWATER FISHES OF NORTH AMERICA

Plate 13.6. Creek Chubsucker, Erimyzon oblongus

Plate 13.7. Northern Hog Sucker, Hypentelium nigricans

Plate 13.8. Torrent Sucker, Thoburnia rhothoeca (photograph by N. Burkhead and R. Jenkins, courtesy Virginia Department of Game and Inland Fisheries; used with permission from Noel Burkhead and the American Fisheries Society, copyright 1994)

Native Range Suckers are Holarctic in distribution, but the family has a disjunct distribution between eastern Asia and North America. In eastern Asia, the Chinese Sucker (Myxocyp-

rinus asiaticus) occurs in the Yangtse and Minjiang River drainages of eastern China (Gao et al. 2008); the Longnose Sucker (Catostomus catostomus) invaded eastern Siberia from Alaska during the last Pleistocene interglacial period (Berra 2001). Catostomids are widespread in North America with the northern limit being Arctic Ocean drainages (C. catostomus; Fig. 13.1) with the southern limit being the Rio Usumacinta of southeastern Mexico and northern Guatemala (Ictiobus bubalus, Smallmouth Buffalo; Lee et al. 1981; Berra 2001; Miller et al. 2005). The genera Catostomus, Chasmistes, Deltistes, and Xyrauchen occur west of the Continental Divide (Figs. 13.2–13.5), although two species, C. catostomus and C. commersonii, are distributed across much of northern North America (Lee et al. 1981). Most species diversity within Catostomus is associated with the complex, and often endorheic, basins of the western United States and northwestern Mexico. Several species of Catostomus, as well as Chasmistes and Deltistes, exhibit

Plate 13.9. Blue Sucker, Cycleptus elongatus

Plate 13.10. Lost River Sucker, Deltistes luxatus

Plate 13.11. Spotted Sucker, Minytrema melanops

455

456

FRESHWATER FISHES OF NORTH AMERICA

Plate 13.12. Chinese Sucker, Myxocyprinus asiaticus, juvenile (illustration used with permission of and copyrighted by Emily S. Damstra)

Plate 13.13. Razorback Sucker, Xyrauchen texanus

highly localized distributions and often are confined to a single basin. The Wall Canyon Sucker (Catostomus sp.) probably has the most limited distribution in the genus, occurring in a single stream system in northwestern Nevada. The monotypic genus Xyrauchen is endemic to the Colorado River basin (Fig. 13.5). Suckers found primarily east of the western Continental Divide associated with large river systems (genera Carpiodes, Fig. 13.6; Cycleptus, Fig. 13.7; Erimyzon, Fig. 13.8; Hypentelium, Fig. 13.9; Ictiobus, Fig. 13.10; Minytrema, Fig. 13.11; and Moxostoma, Fig. 13.12) are more widespread in their distributions. Exceptions to this pattern are species of

headwaters, small streams, or rivers (e.g., torrent Suckers, Thoburnia, Fig. 13.13; Roanoke Hog Sucker, Hypentelium roanokense; and jumprock Suckers, Moxostoma), which can exhibit highly localized distributions. Two species of Moxostoma, Mexican Redhorse (Moxostoma austrinus) and Mascota Jumprock (Moxostoma mascotae), occur on the Pacific Slope of Mexico.

Suckers as Non-Natives Unlike some other families of North American freshwater fishes (e.g., Centrarchidae, Sunfishes; Salmonidae, Trouts

CATOSTOMIDAE: SUCKERS

457

Figure 13.1. The Longnose Sucker, Catostomus catostomus, shown here cruising in a tributary to Great Slave Lake (Northwest Territories, Canada) in May 2012, is the most widespread species of Sucker (Catostomidae) in North America (Page & Burr 2011). The species ranges across Arctic, Pacific, and Atlantic Ocean drainages throughout Canada, Alaska, and much of the northern United States (photograph by and used with permission of ©Paul Vecsei / Engbretson Underwater Photography).

Figure 13.2. Geographic range of the genus Catostomus.

Figure 13.3. Geographic range of the genus Chasmistes.

Genus Catostomus Genus Chasmistes

and Salmons), catostomids have not been extensively introduced outside of North America. Most introductions were unsuccessful, in terms of establishing wild, reproductive populations, but a few exceptions exist. Only one

introduction is thought to be associated with the ornamental fish industry. A single C. commersonii was collected in England not far from ornamental fish ponds containing Goldfish (Carassius auratus) imported from North Amer-

Figure 13.4. Geographic range of the genus Deltistes.

Figure 13.8. Geographic range of the genus Erimyzon.

Genus Erimyzon

Genus Deltistes

Figure 13.5. Geographic range of the genus Xyrauchen.

Figure 13.9. Geographic range of the genus Hypentelium.

Genus Hypentelium

Genus Xyrauchen

Figure 13.6. Geographic range of the genus Carpiodes.

Figure 13.10. Geographic range of the genus Ictiobus.

Genus Ictiobus

Genus Carpiodes

Figure 13.7. Geographic range of the genus Cycleptus.

Figure 13.11. Geographic range of the genus Minytrema.

Genus Minytrema Genus Cycleptus

458

CATOSTOMIDAE: SUCKERS

Figure 13.12. Geographic range of the genus Moxostoma.

Genus Moxostoma

Figure 13.13. Geographic range of the genus Thoburnia.

Genus Thoburnia

ica (Copp et al. 1993); this specimen died shortly after collection. The Quillback (Carpiodes cyprinus) and buffalofishes (Ictiobus spp.) have been introduced for aquaculture in eastern Europe, Central Asia, Israel, Mexico, Cuba, and Panama (Welcomme 1988). Introductions of the Quillback and Smallmouth Buffalo failed to establish reproductive populations in eastern Europe and Central Asia (Welcomme 1988). The Bigmouth Buffalo (Ictiobus cyprinellus) was introduced successfully into Russia, Cuba, and Panama (Makeyeav 1980). The Black Buffalo (Ictiobus niger) may be established in Central Asia because the population introduced into Cuba originated somewhere in the former Soviet Union (Welcomme 1988). Introductions of Suckers outside their native ranges in North America are mostly associated with baitbucket transfers or accidental stocking with trout (Oncorhynchus spp.) into geographically proximate drainages (e.g., Catostomus spp.) or as forage fish (e.g., Erimyzon spp.; Fuller et al. 1999). Buffalofishes were stocked intentionally in Arizona in 1918 for sportfishing (Minckley 1973) and may have been transplanted subsequently into California by commercial fishers (Moyle 2002). Fuller et al. (1999) comprehensively reviewed catostomid introductions in North America.

459

PHYLOGE NE TIC RELATIONSHIPS

Higher Relationships Before discussing phylogenetic relationships among catostomid lineages, it is worthwhile to briefly summarize previous works on the relationships of Catostomidae within the order Cypriniformes (Carps). A number of anatomical studies compared catostomids with other cypriniform taxa (Ramaswami 1957; see morphology section); Siebert (1987) and G. R. Smith (1992) reviewed historical works commenting on Catostomidae relationships. Wu et al. (1981) presented the first cladistic examination of Cypriniformes relationships, although only Characiformes (Characins) were used as an outgroup. Although Catostomidae and Gyrinocheilidae (Algae Eaters) were depicted as sister-taxa, examination of character distributions indicates that no synapomorphies supported such a relationship. Two synapomorphies, however, did support an unresolved relationship among catostomids, gyrinocheilids, and cobitids (Loaches). Harris & Mayden (2001) resolved the Cobitidae (Loaches) as sister to a clade of Gyrinocheilidae + Catostomidae based on mtDNA 12S and 16S rDNA sequences, supporting Wu et al.’s (1981) hypothesis. One caveat associated with Harris & Mayden’s (2001) analysis is limited taxon sampling among cypriniform fishes. In contrast, both Sawada’s (1982) analysis of the Cobitoidea (Cobitidae + Homalopteridae or Balitoridae, River Loaches) and Siebert’s (1987) analysis of cypriniform relationships yielded the genus Gyrinocheilus as sister to a clade containing Catostomidae + Cobitidae and Homalopteridae. Both studies contain analytical weaknesses, including irreversibility of character evolution (Sawada) and limited number of transformation series examined (Siebert). G. R. Smith (1992) provided the first, all-encompassing phylogenetic analysis of catostomid relationships. In methods, he discussed numerous morphological analyses supporting various cypriniform lineages as potential sister-taxa to catostomids. Despite examining numerous cypriniform taxa, however, only the genera Leptobotia (Cobitidae) and Cyprinus (Cyprinidae) were included as outgroups in his data matrix. Smith’s phylogeny depicts a trichotomy among Leptobotia, Cyprinus, and Catostomidae, undoubtedly due to limited taxon sampling among potential outgroup taxa. Monophyly of catostomids, relative to Leptobotia and Cyprinus, is based on nine, “unique and unambiguous” morphological characters (G. R. Smith 1992:792): concave dorsal edge of opercle (two character

460 FRESHWATER FISHES OF NORTH AMERICA

states given for catostomids); detached cephalic sensory canals; fenestrate basioccipital process; small hyomandibular that articulates only with sphenotic; descending process of second and fourth pleural ribs of Weberian apparatus broadly sutured together; mandibular sensory canals lost; lateral ethmoid that is triradiate in longitudinal section; a porous, minute dermosphenotic; and 18 caudal rays. It would be interesting and informative to determine if these characters continue to diagnose catostomids with the inclusion of additional outgroup taxa. Analyses based on molecular data sets offer contrasting phylogenetic hypotheses of relationships within Cypriniformes. Notably, all molecular analyses resolve Catostomidae as monophyletic, albeit with limited taxon sampling. Clements et al. (2004) examined growth hormone sequences; their analysis resolved a well-supported clade of Cobitidae sister to Catostomidae + Cyprinidae. This latter clade was also proposed by Uyeno and Smith (1972) based on karyotype data. Phylogenetic analyses of reduced (Harris & Mayden 2001; Tang et al. 2005) and complete (Saitoh et al. 2006) mitogenomic sequences consistently resolved Gyrinocheilidae sister to Catostomidae. In contrast, analyses of nuclear genes placed Catostomidae as either sister to remaining Cobitoidea + Cyprinoidea (Chen et al. 2009) or as sister to the Cobitoidea (Mayden et al. 2009); both analyses placed Gyrinocheilidae sister to remaining Cobitoidea.

Intrafamilial Relationships Most early taxonomic and systematic efforts (pre-1900) on catostomids dealt with original descriptions and various contributions to higher-level classifications (reviewed by G. R. Smith 1992). Post-1900 contributions to catostomid classification include Hubbs (1930), who provided a key to eastern North American genera and designated tribes for these genera, and Robins & Raney’s (1956) study of the genera and subgenera of Moxostoma. Nelson (1948, 1949) examined the Weberian apparatus and opercular series in catostomids; he concluded that these structures provided support for the subfamilial and tribal designations in Hubbs (1930). Miller (1959) presented a phylogeny of the Catostomidae that was based largely on Hubbs (1930) and Nelson (1948, 1949). In his discussion on relationships, Miller (1959:199) suggested that the Cycleptinae might be divided into two subfamilies consistent with the disjunct distribution of Cycleptus (North America) and Myxocyprinus (China). He depicted the Catostominae consisting of the Erimyzontini, genera Erimyzon + Minytrema,

sister to Moxostomatini composed of Moxostoma, Thoburnia, Hypentelium, and Lagochila + Catostomini composed of Catostomus, Pantosteus (= Catostomus), Chasmistes, and Xyrauchen. Miller’s hypotheses of relationships received support from Bussjaeger & Briggs (1978), who speculated on evolutionary affinities among catostomids based on bile salt chemistry. Jenkins’s (1970) unpublished dissertation provided a comprehensive review of the taxonomy, classification, morphological variation, and distribution of the Moxostomatini; his was the first work to depict a phylogeny for the species in this tribe. Buth (1978, 1979, 1980) examined species-level relationships among Hypentelium, Thoburnia, and some species of Moxostoma. Ferris & Whitt (1978) constructed a phylogeny of 30 species based on the loss of duplicate gene expression in isozymes. Their Wagner tree placed the Ictiobinae sister to Cycleptinae + Catostominae. Within the Catostominae, they recognized three tribes, Erimyzonini, Moxostomatini, and Catostomini; Moxostoma was paraphyletic with Moxostoma duquesnei (Black Redhorse) sister to Catostomus plebeius (Rio Grande Sucker), Catostomus platyrhyncus (Mountain Sucker), and Catostomus discobolus (Bluehead Sucker). As noted, G. R. Smith (1992) provided the first comprehensive analysis of catostomid relationships based on 64 taxa and 157 morphological, biochemical, and early life history transformation series. Smith’s analysis produced 2 equally parsimonious trees of 852 steps (CI [confidence interval] = 0.35). In his preferred tree (Fig. 13.14) the Ictiobinae was sister to Cycleptinae + Catostominae, although Smith recognized that a limited number of characters supported this relationship. He also recognized the possibility of an Ictiobinae + Catostominae relationship based on a few homoplasious characters. Within the Catostominae, two tribes were recognized, the Catostomini and Moxostomatini. Within the Moxostomatini, Smith’s analysis yielded a paraphyletic Moxostoma grade related to a paraphyletic Scartomyzon grade that, in turn, was related to a trichotomy of Moxostoma ariommum (as Scartomyzon ariommus, Bigeye Jumprock) + Thoburnia + Hypentelium. The second topology produced by this analysis had Moxostoma cervinum (as Scartomyzon cervinus, Blacktip Jumprock) as sister to M. ariommum (as S. ariommus) + Thoburnia + Hypentelium. Harris & Mayden (2001) examined relationships among basal lineages of catostomids based on mtDNA 12S and 16S rDNA gene sequences (Fig. 13.15). These authors examined the influence of five structural classes (short and long stems, bulges, loops, unpaired bases) within the

CATOSTOMIDAE: SUCKERS

461

Catostomini + remaining Moxostomatini clade (Harris & Mayden 2001). Given the uncertain phylogenetic affinities of Erimyzon and Minytrema within the subfamily Catostominae, these taxa were identified as incertae sedis. Within the remaining Moxostomatini, the two species of Scartomyzon were resolved embedded within Moxostoma, questioning the monophyly of both Moxostoma and Scartomyzon if the latter genus is recognized as a distinct taxon. G. R. Smith (1992) noted that both Moxostoma and Scartomyzon formed paraphyletic grades (1992), suggesting that some species currently recognized in these genera may be more closely related to other Moxostoma, Thoburnia + Hypentelium, or form distinct evolutionary lineages. To further examine phylogenetic relationships among the Moxostomatini, Harris et al. (2002) sequenced the entire mitochondrial cytochrome b gene from all species within this tribe and representative taxa from the Catostomini and other catostomid subfamilies. Maximum parsimony analysis of these gene sequences yielded two monophyletic clades: Catostomini (genera Catostomus, Deltistes, and Xyrauchen) + Erimyzonini (genera Erimyzon Figure 13.14. Phylogenetic hypothesis of relationships among catostomid lineages based on morphological, early life history, and biochemical characters (modified from G. R. Smith 1992).

rRNA secondary structure on sequence alignment and partitioned the data into these structural classes to examine rate heterogeneity and nucleotide substitution patterns. These molecular data consistently yielded a monophyletic Catostomidae, Ictiobinae, and Catostomini; the Catostominae was monophyletic in all analyses, except the 12S plus Valine 1:1 weighting analysis; the Moxostomatini was monophyletic in all combined and 16S analyses but was para- or polyphyletic in all analyses of the 12S + Valine data. The Cycleptinae was paraphyletic in all analyses, except the 12S + Valine 1:1 weighting analysis with Myxocyprinus as the basal-most taxon and the genus Cycleptus as either sister to remaining catostomids or sister to Catostominae (as in the majority of analyses). Within the Catostomini, relationships among species examined were not resolved; this is not surprising given the conservative nature of these two genes and the limited number of taxa examined. Phylogenetic affinities of Erimyzon and Minytrema varied depending on the data set and weighting scheme; singly or together these two taxa were either sister to the Catostomini, sister to the Moxostomatini, or basal to a

Figure 13.15. Phylogenetic hypothesis of relationships among basal catostomid lineages based on mtDNA 12S and 16S rDNA gene sequences (redrawn from Harris & Mayden 2001).

462 FRESHWATER FISHES OF NORTH AMERICA

and Minytrema); and Moxostomatini (genera Moxostoma and Scartomyzon) + Thoburniini (genera Hypentelium and Thoburnia). Within the Moxostomatini, the genus Thoburnia was either unresolved or polyphyletic; Thoburnia atripinnis (Blackfin Sucker) was sister to a monophyletic Hypentelium in the maximum likelihood analysis. In turn, these taxa were sister to a monophyletic clade containing the genera Scartomyzon and Moxostoma. The genus Scartomyzon was never resolved as monophyletic but was always recovered as a polyphyletic group embedded within Moxostoma, rendering the latter genus paraphyletic if Scartomyzon continued to be recognized. Relationships among lineages within the Moxostoma and Scartomyzon clade were resolved as a polytomy. Sun et al. (2007) presented a UPGMA (unweighted pair group method with arithmetic mean) tree based on mitochondrial cytochrome b for 17 species of catostomids. Their results are somewhat at odds with previous studies. Although Catostomidae was monophyletic, the genus Minytrema was the basal-most taxon. A paraphyletic Cycleptinae (sensu G. R. Smith 1992) was sister to a monophyletic Ictiobinae; this clade was, in turn, sister to Catostominae, minus the genus Minytrema. Within the Catostominae, the genus Erimyzon was sister to a clade containing a monophyletic Moxostomatini + a non-monophyletic Catostomini, due to the genus Thoburnia being embedded in this tribe. Although there is some concordance between the results of this and previous studies, limited taxon sampling and analytical issues in this paper limit the value of Sun et al.’s (2007) hypothesis. Doosey et al. (2009) used mitochondrial ND 4/5 sequences to infer phylogenetic relationships within the family with more complete taxon sampling than in previous studies. Relationships among basal catostomid lineages differed from previous hypotheses in that a clade consisting of Cycleptinae sister to Myxocyprininae + Ictiobinae was sister to Catostominae. Within Catostominae, the Erimyzonini was sister to a clade containing the Catostomini sister to Moxostomatini + Thoburnini. Species-level groupings recovered within the Moxostomini were similar to those of Harris et al. (2002), although their data provided better resolution to relationships among lineages. Depending on the coding scheme used in the sequence analyses, some species-level relationships received moderate to low bootstrap support in the maximum likelihood analyses. Thus, the most recent published phylogenetic examination of relationships within the Moxostomatini emphasizes the need for additional character data (either molecular or morphological) to clarify relationships among these lineages.

FOSSIL RECORD Suckers are represented in the fossil record of North America and Asia dating from the Eocene (55–35 mya; Smith 1981; Cavender 1986; Sytchevskaya 1986; Chang et al. 2001; Liu & Chang 2009). Fossil catostomids often occur in lacustrine or palustrine sediments in association with fossil Bowfin (Amiidae), Pikes (Esocidae), and Catfishes (Siluriformes) (Cope 1884; Grande et al. 1982; Chang & Maisey 2003). Co-occurrence with minnow (Cyprinidae) fossils is less common and is observed primarily in Miocene-Age deposits (25–5 mya; Cavender 1986). North American fossil deposits from the Eocene have produced a wealth of information on the paleoecology of catostomids. Suckers are often the most abundant fossils at a given locality, and some deposits yielded enough specimens to estimate growth rates and year-class abundances from size-frequency data (Wilson 1984). Paleoichthyological studies have examined large-scale die-offs of freshwater fishes in the western United States. Hypoxia was the most likely cause of death among specimens of the extinct castomid genus †Amyzon, based on the recovery of wellarticulated skeletons with gaping mouths (Barton & Wilson 2005). In contrast, relatively few extinctions occurred during the Late Cenozoic among eastern North American taxa, primarily due to the long-term geological stability of the region and the north-to-south orientation of many drainages that provided access to southern refuges during the glacial advances of the Pleistocene (2.6–0.01 mya; Smith 1981). Fossil catostomids from the Eocene include the genera †Amyzon Cope 1872, †Vasnetzovia Sytchevskaya 1986 (both ictiobines, or carpsuckers and buffalofishes), and †Plesiomyxocyprinus Liu and Chang 2009 (a myxocyprinin in the same lineage as the Chinese Sucker). The genera †Vasnetzovia and †Plesiomyxocyprinus are restricted to East Asia, thus far; the genus †Amyzon occurs in East Asia and western North America. Putative catostomid fossils from the Paleocene Paskapoo Formation of Alberta (Wilson 1980) can only be identified reliably as cyprinoids (Cavender 1991). Fossils of modern Sucker genera are found in Middle to Late Miocene (about 16–5 mya) deposits (Cavender 1986; Chang et al. 2001). Fossils from PleistoceneAge deposits are “osteologically indistinguishable” from extant species (Grande et al. 1982:523). The genus †Amyzon is probably not monophyletic (Wilson 1984; Cavender 1986). In body form, †Amyzon generally resembles a miniaturized form of Carpiodes, having a deeply compressed body and an elongate dorsal fin base.

CATOSTOMIDAE: SUCKERS

This genus is diagnosed by the absence of an intercalar, an anteriorly directed premaxillary process, an intermediate (as opposed to broad or sharp) pterotic ridge, five or more infraorbitals, scales with sharp anterolateral corners, and a short, broad dermethmoid spine (Miller & Smith 1981; G. R. Smith 1992). †Amyzon-like forms existed for ≥20 million years in North America; the most recent fossils are estimated to about 31 million years old (Evernden & James 1964; Cavender 1986). The number of species within †Amyzon depends on interpretation of variation in meristic and morphometric characters, which are notoriously variable among extant Catostomidae. Adding to the taxonomic confusion, many species are described from small or incomplete fossil remains. Some characters originally used for species descriptions are now attributed to clinal variation or sexual dimorphism (Lambe 1906; Wilson 1984). Cope (1872) described the type species, †A. mentale, from the Osino Oil Shales of Eocene-Oligocene Nevada (56–23 mya). Cope (1874, 1875) also described three additional species from the Florissant Formation of Colorado: †A. commune, †A. pendatum, and †A. fusiforme. Of these three species, only †A. commune is still recognized as a valid species (Bruner 1991a). Bruner (1991a) also considered †A. gosiutensis (Grande et al. 1982) from the Eocene Green River Formation in Wyoming to be a junior synonym of †A. aggregatum (Wilson 1977) from British Columbia, despite the great geographic separation between the two localities. †Amyzon brevipinne (Cope 1893) was the first Sucker to be described from the productive fossil beds of Eocene British Columbia. Additional species of †Amyzon remain undescribed from deposits of British Columbia and other areas of North America (Bruner 1991a). These North American fossils are part of a rich, Holarctic †Amyzon fauna of which the most recently described, or redescribed, species are from Asia. Chang et al. (2001) described †A. hunanensis from Hunan Province in southern China and discussed how fossils originally described as minnows, but subsequently identified as Suckers, have expanded the distribution of fossil catostomids in Asia. As yet, it is unclear whether species of †Amyzon ever co-occurred with extant ictiobine genera. The oldest fossils of extant genera ostensibly belong to species of Ictiobus from the Middle Miocene (16–12 mya) strata of South Dakota (Cavender 1986). Putative records of ictiobine fossils from Oligocene (34–23 mya) deposits in Kazakhstan were refuted (but see G. R. Smith 1992). Pliocene-Age (6–2.5 mya) ictiobines include †Ictiobus aguilerai, described from Hidalgo, Mexico (Alvarado-Ortega

463

et al. 2006), as well as fossils of extant species, such as I. bubalus from Oklahoma (G. R. Smith 1962). PleistoceneAge beds have yielded I. cyprinellus from Nebraska (Smith & Lundberg 1972), I. niger from Kansas (Neff 1975), and assemblages of other ictiobines from Texas (Uyeno & Miller 1962; Lundberg 1967), Michigan, and Montana (Smith 1981; Cavender 1986). Fossils of taxa in the subfamily Catostominae are generally more recent in age, but in the western United States a substantial fossil record for Chasmistes confirms their presence by the Late Miocene (12–5.5 mya; Smith 1975; Miller & Smith 1981); Xyrauchen texanus is known from Pliocene-Age deposits (Hoetker & Gobalet 1999). An extensive fossil record exists for Chasmistes, Catostomus, and Deltistes from Plio-Pleistocene-Age deposits (6–0.01 mya) from the western United States (Smith 1981). Fossil-rich deposits containing catostomids include Fossil Lake, Idaho, Glenns Ferry Formation, Utah, Cabbage Patch fauna, Montana, and many others in South Dakota, Nebraska, Oregon, and California (Miller & Smith 1967; Smith 1975; Cavender 1986). Fossils for eastern North American Suckers are generally known from Late Pleistocene deposits (1.8–0.01 mya), but fossil Minytrema melanops and Silver Redhorse (Moxostoma anisurum) were recovered from Early and Middle Pleistocene deposits (2.5–1.8 mya) in Kansas and around the Great Lakes (Cleland 1966; Eshelman 1975). A potentially new species of fossil redhorse (Moxostoma) is known from the Lake Chapala Basin, Jalisco, Mexico (Smith et al. 1975; Miller 1986).

MORPHOLOGY

General Morphology Catostomidae is a diverse family of fishes, exhibiting great variation in body shape and size. In general appearance, Suckers resemble large minnows (Cyprinidae), although large-river species tend to be deep-bodied, almost stocky or chunky in appearance (e.g., the genera Carpiodes and Ictiobus), and stream-dwelling Suckers are more elongate than most North American minnows. Some of the most unusual body shapes in freshwater fishes occur in juvenile Myxocyprinus asiaticus and Xyrauchen texanus, both of which have enlarged nuchal areas that might function as adaptations for living in fast-flowing, large rivers or as an anti-predator defense (see Portz & Tyus 2004). The enlarged nuchal hump of Xyrauchen is modified further into a sharp keel; Xyrauchen literally translates as “razor nape” (Minckley 1973). Interestingly, the nuchal hump of

464 FRESHWATER FISHES OF NORTH AMERICA

Myxocyprinus becomes less pronounced with growth, and in a mature adult, the body somewhat resembles that of Cycleptus (Fig. 13.16). In addition to similar body shape, Suckers and minnows also have a single dorsal fin, a scale-less head, and

cycloid scales. Suckers can be differentiated from North American minnows based on a single dorsal fin with ≥10 fin rays, a more posteriorly placed anal fin, 18 principal caudal fin rays, barbels conspicuously absent from around the mouth, and in most species an inferior, highly protrusible mouth with plicate or papillose lips. Other external features of catostomids, such as pigmentation and meristic and mensural characters, are described and illustrated in many state fish books (Etnier & Starnes 1993; Jenkins & Burkhead 1994; Boschung & Mayden 2004). Catostomids probably originated as large-bodied riverine fishes, and subsequently radiated and adapted to lower-order, higher-gradient streams by becoming increasingly smaller in body size (G. R. Smith 1992). This trend is clearly seen when comparing body sizes of the large river or lake Suckers (genera Carpiodes, Cycleptus, Deltistes, Ictiobus, Myxocyprinus, and Xyrauchen) that can grow to be >60 cm TL (Etnier & Starnes 1993; Baensch & Riehl 1995; Moyle 2002; Boschung & Mayden 2004) with some headwater populations of Hypentelium roanokense, which are ≤14 cm TL (Jenkins & Burkhead 1994), and other dwarf forms of Catostomus (e.g., Salish Sucker, Catostomus sp., about 29 cm FL, Pearson & Healey 2003; Jenny Creek population of the Klamath Smallscale Sucker, Catostomus rimiculus, 21 cm SL, Hohler 1981). Based on weight, the largest Sucker caught in a sport fishery appears to be a 40 kg Ictiobus bubalus taken from Lake Wylie, North Carolina (NCWRC 2010).

Mouth and Lip Morphology

Figure 13.16. (A) Juvenile, (B) semi-adult, and (C) adult Chinese Suckers, Myxocyprinus asiaticus. Juvenile and semi-adult from Yangtze River, Hunan Province, China. Adult photographed in the Shanghai Ocean Aquarium (photographs by and used with permission of N. Khardina, juvenile, and H. Bleher, semi-adult and adult, Aquapress Publishers, Italy, www.aquapress-bleher .com,copyright 2009).

The morphology of the mouth and lips in Suckers is undoubtedly their most distinctive feature. Although the mouth and lips of a few species of minnows in North America might superficially resemble those of catostomids (e.g., cyprinid genera Campostoma and Phenacobius), none have lips so fully developed as those of some Suckers. As with other morphological characters, however, a great deal of variation in lip morphology is exhibited by catostomids. The general tendency, however, is for large-river and lake Suckers (e.g., genera Carpiodes, Chasmistes, Deltistes, Erimyzon, Ictiobus, Minytrema, Xyrauchen) to have less well-developed lips and lip texturing (see following paragraph) than small river– or stream-inhabiting forms; an exception is the genus Cycleptus, species of which frequent large rivers and have big lips and well-developed lip papillae (Burr & Mayden 1999:39). Jenkins & Burkhead (1994) classified the morphology of lip textures in some catostomids into character states.

CATOSTOMIDAE: SUCKERS

They recognized five character states of lip textures: plicate (Fig. 13.17a)—the lips appear to have longitudinal grooves or pleats; subplicate (Fig. 13.17b, lower lip)—the lips are similar to plicate ones, but individual grooves can be divided into smaller sections; plicate-papillose (Fig. 13.17c)—the lip texture is a mix of plicae and papillae; papillose—numerous papillae cover the lips, resembling small, round bumps (Fig. 13.17d) or are raised and elongated resembling villi (see also Rio Grande form of Cycleptus elongatus, Blue Sucker; Burr & Mayden 1999); and semipapillose (Fig. 13.17f)—the plicae are oval or oblong, not columnar, in appearance, and the papillae are not raised. As noted previously, Moxostoma lacerum has unusual lips for a catostomid (Fig. 13.17g). The upper lip is hood like, and the lower lip is cleft into two distinct lobes; lip texturing is lightly papillose with a small number of plicae on the posterior edge of the upper lip (Jordan & Brayton 1877; Etnier & Starnes 1993; Jenkins & Burkhead 1994). The plicae are more pronounced in juveniles than in small adults (Etnier & Starnes 1993). Lip morphology and texturing are related directly to external brain morphology as revealed by an examination of 46 species of catostomids (Miller & Evans 1965). Taste in catostomids occurs either from taste buds on the lips and skin or in the mouth and pharynx. The facial lobe of the brain receives sensory input from taste buds on the

465

lips and skin, and the vagal lobe receives such input from taste buds in the mouth and pharynx. Not surprisingly, species with large, papillose lips, have larger facial lobes, including, e.g., the genera Cycleptus, Hypentelium, Thoburnia, and some Moxostoma and Catostomus (Pantosteus). In contrast, those species with smaller lips and less texturing (e.g., genera Carpiodes, Ictiobus, Erimyzon, various species of Moxostoma and Catostomus) have larger vagal lobes.

Pigmentation and Breeding Tubercles Coloration of Suckers tends to be rather drab when compared with that of other North American freshwater fishes such as Sunfishes, some minnows, and most darters, but even so, a great deal of variation exists among the genera. Colors in adults range from an almost uniform silvery across the body (e.g., Carpiodes spp.) to a dark olive or brown or brassy dorsally, fading to a light yellow or white ventral surface (Etnier & Starnes 1993; Jenkins & Burkhead 1994; Moyle 2002). The genera Cycleptus and Deltistes are dark blue to almost black across the back and sides with a light-colored ventral surface (Moyle 2002; Boschung & Mayden 2004). In contrast to the more uniform body coloration noted above, the genus Hypentelium has dark, dorsal saddles and blotches on the sides (Etnier & Starnes 1993; Boschung & Mayden 2004). The Spotted Sucker and

Figure 13.17. Lip morphology in suckers: (A) plicate, Golden Redhorse, Moxostoma erythrurum; (B) subplicate, Shorthead Redhorse, M. macrolepidotum; (C) plicate-papillose, Rustyside Sucker, Thoburnia hamiltoni; (D) papillose, Northern Hog Sucker, Hypentelium nigricans; (E) papillose, Bigeye Jumprock, M. ariommum; (F) semipapillose, Silver Redhorse, M. anisurum; and (G) Harelip Sucker, M. lacerum (illustrations from Jenkins & Burkhead 1994, and reproduced with permission from the American Fisheries Society).

466

FRESHWATER FISHES OF NORTH AMERICA

A

C

D

B

Blackfin Sucker have dorsal and lateral stripes, although the stripes in the Spotted Sucker are formed by dark spots at the base of the scales (Etnier & Starnes 1993; Boschung & Mayden 2004). In many species, the fins are often colored orange or red and become especially vivid in nuptial individuals. The coloration of juveniles is often quite distinct from that of adults. For example, juvenile Blue Suckers are brownish or brassy colored with clear fins, except for a dark blotch in the caudal fin (Etnier & Starnes 1993). Nuptial body coloration varies from simple intensification of normal life colors (e.g., Cycleptus spp., some Moxostoma spp.) to the development of black or reddish lateral bands (e.g., the genera Catostomus, Chasmistes, Thoburnia, Xyrauchen; Etnier & Starnes 1993; Jenkins & Burkhead 1994; Moyle 2002). Some species (e.g., genera Deltistes and Hypentelium) do not develop nuptial coloration (Etnier & Starnes 1993; Jenkins & Burkhead 1994; Moyle 2002). Catostomids develop distinctive patterns of breeding tubercles, also called pearl or contact organs that are keratonized epidermal structures that function to facilitate contact between spawning individuals (Fig. 13.18). The pat-

Figure 13.18. (A) Head tubercles on a breeding male Robust Redhorse, Moxostoma robustum, from the Oconee River, Georgia (photograph by and used with permission of N. M. Burkhead). (B) Black Buffalo, Ictiobus niger, congregate for spawning after migrating upstream in Citico Creek, Monroe County, Tennessee. Note the small tubercles (light spots) on the male in the center of the photograph (photograph by and used with permission of Dave Herasmitschuk, Freshwaters Illustrated). (C) A female Black Buffalo is sandwiched between two males as the three churn the surface of the water during a spawning bout in Citico Creek (photograph by and used with permission of Kyle Piller). (D) Eggs of the Black Buffalo adhering to a rock in Citico Creek (photograph by and used with permission of Kyle Piller).

tern of breeding tubercles across the body and the extent of their development have been of longstanding interest to naturalists and ichthyologists. Breeding tubercles of catostomids were mentioned by European naturalists as early as 1557 (Wiley & Collette 1970), and these structures were apparently mentioned in ancient Chinese documents (Wiley & Collette 1970 citing Kimura & Tao 1937). Some early works describing breeding tubercles in Suckers include Forbes and Richardson (1909); Fowler (1912), who described and illustrated tubercles on seven Suckers from eastern North America; Reighard (1920); Hubbs (1930); and Branson (1962). Tuberculation patterns are described for every genus in the family in North America (Huntsman 1967; Jenkins 1970; Wiley & Collette 1970; Madsen 1971; Morris & Burr 1982). Tubercle patterns vary a great

CATOSTOMIDAE: SUCKERS

467

deal in Suckers (Figs. 13.18 and 13.19); in fact, there is so much variation that Branson (1962) proposed a hypothesis on the evolution of tubercle patterns in catostomids. Wiley & Collette (1970) provided a comprehensive review of the early literature on breeding tubercles in catostomids, as well as other fishes, and presented information on their structure, function, and evolution. Males of all species develop tubercles about the head, body, and fins. Tuberculation patterns range from small, fine tubercles covering the entire body (or portions of it) and fins (e.g., genera Deltistes; Cycleptus; Ictiobus; Fig. 13.18b) to the large snout tubercles developed by Erimyzon spp. (Fig 13.19). The morphology of these two types of tubercles is quite different. Tubercles on the body and fins of Golden Redhorse (Moxostoma erythrurum) are composed of “mounds of cells formed by epithelial hypertrophy and hyperplasia with keratinization of the tissue” (Wiley & Collette 1970:183). In contrast, the large snout tubercles on male Erimyzon are a “solid keratinized cone supported by vascularized hypertrophied epithelium” (Wiley & Collette 1970:183).

Internal Anatomy Suckers have a physostomus swim bladder (i.e., a duct connects the swim bladder with the gut), like other Cypriniformes. Nelson (1961) compared swim bladder morphology in 52 species of catostomids. Hubbs (1930) used the number of chambers in the swim bladder as a main character in his taxonomic key to catostomid genera from eastern North America. The swim bladder in catostomids consists of two to four chambers. Genera with species having a two-chambered swim bladder include Carpiodes, Catostomus, Cycleptus, Deltistes, Erimyzon, Hypentelium, Ictiobus, Minytrema, Thoburnia, and Xyrauchen. Older literature (e.g., Hubbs 1930) reports the occasional presence of a third chamber in Minytrema; this was disputed by Nelson (1961), who did not find any indication of a third, posterior chamber in his survey of swim bladder morphology. In addition, the same literature reports the swim bladder of Thoburnia as being obsolete or absent. This absence can be attributed to the overall reduced size of the swim bladder, and to the fact that it is often obscured by being embedded in the retroperitoneal tissue in the anterodorsal body cavity (Nelson 1961). Three-chambered swim bladders occur in Moxostoma, except in the Bigeye Jumprock, which has either three or four chambers; when present, the fourth chamber is small and round (Jenkins & Burkhead 1994).

Figure 13.19. (A) Lateral and (B) dorsal views of breeding tubercles on the Creek Chubsucker, Erimyzon oblongus (UAIC 4383.08, male, 154 mm SL) (photograph by ®2010, P. M. Harris).

Osteology Osteological characters of catostomids were used in a number of comparative studies with the primary goal of inferring evolutionary relationships within Catostomidae and among ostariophysans. Edwards (1926) illustrated the skull and associated muscles and ligaments in his examination of mechanisms of jaw protrusion in catostomids (see following subsection). Gregory (1933) also discussed jaw protrusion and illustrated the skull of

468

FRESHWATER FISHES OF NORTH AMERICA

Carpiodes. Ramaswami (1957) illustrated the osteocranium and Weberian apparatus of Catostomus commersonii and Myxocyprinus; he also discussed seven osteological characters that he considered diagnostic of Catostomidae relative to other cypriniform fishes (see phylogenetic relationships section). Weisel (1960) provided a detailed description of the osteocranium and associated myology of the Largescale Sucker (Catostomus macrocheilus). Others described and illustrated the otoliths (Adams 1940), opercular series (Nelson 1949), Weberian apparatus (Nelson 1948; Robins & Raney 1956; Bailey 1959b; Cook 2001; Bird & Hernandez 2007), pharyngeal bones and teeth (Eastman 1977; Engeman et al. 2009), pectoral anatomy (Brousseau 1976), pectoral fin rays (Lundberg & Marsh 1976), and caudal skeleton (Eastman 1980). G. R. Smith (1992) used osteological and other characters to examine phylogenetic relationships among the Catostomidae; he provided illustrations or photographs and character state descriptions for several of these characters.

Functional Morphology of Feeding Like other teleosts, Suckers can protrude their mouths and lips while feeding. In general, the same muscles and bones are involved in opening and closing the mouth (with the exception of M. lacerum, Jenkins 1994). Several characteristics of the bones, muscles, and ligaments of the mouth, however, allow for greater protrusibility than in other cypriniform fishes (described and illustrated in Edwards 1926 and Weisel 1960, and summarized here). First, the premaxillae are loosely attached to the preethmoid via the premaxillary spine and associated ligament. This loose attachment allows for greater rotation and protrusibility when the mouth is open. The extent of mouth protrusion corresponds with the lengths of the premaxillary spine and ligament. In addition, the premaxillae are not attached to the palatines via a ligament, allowing for increased freedom of movement. Second, a cartilaginous rod connects the maxillae with the vomer, creating a moving joint that allows for greater upper jaw movement. Third, the ascending process of the dentary is anterior in position. As the geniohyoideus and sternohyoideus muscles contract, this ascending process moves forward and downward. The ends of the premaxillae are attached to the ends of the maxillae by a ligament that, in turn, is attached to the ascending process of the dentary. Thus, this arrangement has the effect of pulling the premaxillae downward when the dentary is depressed. The overall result is the greater

protrusibility of the mouth in catostomids, relative to that of cyprinids. Manipulation and processing of food items is done by the pharyngeal bones and teeth. The pharyngeal bones are actually greatly enlarged ceratobranchials of the fifth arch (Weisel 1960; Eastman 1977). The pharyngeal teeth are composed of modified dentin (Peyer 1968), the compound that substitutes for enamel in fish teeth (Eastman 1977). Catostomids have a single row of ≥10 teeth; cyprinids have 1–3 rows of pharyngeal teeth with ≤5 teeth/row. Pharyngeal teeth are replaced throughout life. The number of pharyngeal teeth varies considerably among Suckers and is a function of tooth size and feeding ecology (Eastman 1977). Mollusk- and crustaceaneating species (e.g., Copper Redhorse, Moxostoma hubbsi, and River Redhorse, Moxostoma carinatum) have relatively few large, molariform teeth (20–22 and 35– 41, respectively). In contrast species that feed on smallergrained food items such as small crustaceans, insects, and algae, tend to have small, numerous teeth (e.g., carpsuckers, Carpiodes spp., or buffalofishes, Ictiobus spp., ≥175 teeth, Eastman 1977). In addition to large, molariform teeth, M. hubbsi and M. carinatum have large chewing pads associated with the basioccipital bone to aid in crushing mollusks. These pads are crescent shaped so that all the pharyngeal teeth occlude with the pad (Eastman 1977). Eastman (1977) provided a general survey of catostomid pharyngeal bones and teeth, including illustrations and a taxonomic key based on these elements. Food particle selection and retention is accomplished, in part, by the palatal organ, which is a thick, muscular pad in the anterior roof of the pharynx (Eastman 1977; Callan & Sanderson 2003). The surface of the palatal organ is covered with taste buds, the number of which varies from about 50 (M. carinatum) to >130/cm2 (Carpiodes velifer, Eastman 1977). The palatal organ of catostomids contains more striated muscle than that found in cyprinid palatal organs, which consist primarily of adipose tissue (Eastman 1977). Particle selection and retention is probably accomplished by muscular expansion of all, or part, of the palatal organ following stimulation of the taste buds, which results in food particles being held against the pharyngeal arches while other particles are expelled (Sibbing et al. 1986; Sibbing 1988). To our knowledge, no thorough, systematic review of the palatal organ in Suckers is available, and such a review is definitely needed to further understand the functional morphology of feeding in these fishes.

CATOSTOMIDAE: SUCKERS

469

Figure 13.20. (A) Head of Harelip Sucker, Moxostoma lacerum (UMMZ 177435, 117 mm SL), right lateral view, image reversed. (B) Computer-generated x-ray tomographic image of lateral view of left side of head with individual bones colorcoded. See Fink & Humphries (2010) for bone identifications (figure modified from Fink & Humphries 2010 and used with permission of the American Society of Ichthyologists and Herpetologists, copyright 2010). UMMZ = University of Michigan Museum of Zoology.

In an excellent example of the application of new technologies for study of functional morphology, Fink & Humphries (2010) used high-resolution x-ray computed tomography to examine the cranial and oral morphology of M. lacerum (Fig. 13.20). Harelip Suckers apparently fed on snails, based on the occurrence of snail operculi in stomachs of specimens examined (although snail shells were absent, Jenkins 1970, 1994). This diet is somewhat at odds with the morphology of the fifth ceratobranchial bone (pharyngeal tooth-bearing arch), which is not robust enough to crush snail shells. In addition, the pharyngeal teeth are slender (Fink & Humphries 2010:fig. 8), rather than molariform, as found in other snail- or crustacean-eating species (e.g., M. carinatum or M. hubbsi; Jenkins 1970; Eastman 1977). Also, M. lacerum possesses a bony shelf associated with the symphysis of the dentary (lower jaw) that is covered by a pad of keratinized tissue. Apparently, M. lacerum could manipulate snail shells so that it extracted the bodies while expelling the shells (Fink & Humphries 2010). Sadly, such speculation can never be tested because the Harelip Sucker is extinct, being last collected in 1893 (Jenkins & Burkhead 1994).

Intraspecific Variation Species of Suckers often display a great deal of morphological variation across their respective distributions, particularly in western North American species of Catostomus. This variation led to the recognition of a number of subspecies by ichthyologists working in this region during the early 20th century. Given the extensive distributions of some species, attempts to comprehensively examine geographic intraspecific variation are limited. Studies of Catostomus (Smith 1966; Smith et al. 1983),

Cycleptus (Burr & Mayden 1999), and Moxostoma (Robins & Raney 1956, 1957; Jenkins 1970) are notable exceptions. Other studies on morphological variation in catostomids were prompted by concerns over hybridization (e.g., Quist et al. 2009), especially between the lake Suckers (Chasmistes and Deltistes) and Catostomus (Cook 2001; Markle et al. 2005).

GE NE TICS

Karyology Catostomids are allotetraploids and have a zygotic chromosome number of 100, about twice that of most diploid cypriniform fishes (Uyeno & Smith 1972). Based on fossil evidence and lack of multivalents in the meiotic spreads of Erimyzon, a single hybridization event 50 mya is postulated to account for the tetraploid catostomid karyotype (Uyeno & Smith 1972). Electrophoretic studies of isozyme activities and mobilities are consistent with a rapid transition to disomic inheritance of most genes, enabling fixation and sequence divergence of duplicate genes (Ferris 1984). None of 20 catostomid isozyme loci showed evidence supporting tetrasomic inheritance (Allendorf 1975). In addition, isozymes of duplicate genes have different electrophoretic mobilities (Ferris & Whitt 1978), indicating sequence divergence.

Gene Silencing and Duplicate Gene Expression Gene silencing is a prevalent phenomenon in catostomid evolution. Catostomids express 35–65% of their loci in duplicate, averaging 50% retention of ancestral duplicate gene expression with the remaining loci fixed for a null

470 FRESHWATER FISHES OF NORTH AMERICA

allele. Morphologically primitive species tend to express a greater proportion of their loci in duplicate with respect to more derived species (Ferris 1984). Relative duplicate gene expression varies among and within tissues at different stages of development (Ferris 1984). Duplicate isozymes of lactate dehydrogenase B (LDH-B) and glucosephosphate isomerase B (GPI-B) are expressed in almost equal amounts in the embryos of Lake Chubsuckers, Erimyzon sucetta. LDH-B duplicate isozymes, however, are not expressed equally in adults, and only one isozyme of GPI-B is expressed in adult muscle (Shaklee et al. 1974). Two isozymes of creatine kinase B (CK-B) are expressed in the brain, but only one locus is expressed in the heart tissue of most catostomids (Ferris & Whitt 1979). Duplicate gene expression of glucosephosphate isomerase in Moxostoma may provide evidence of reactivation of a silenced locus (Buth 1982). The Greater Jumprock (Moxostoma lachneri) expresses two GPI-A and two GPI-B loci. Other species of Moxostoma retain duplicate expression of GPI-A but express only a single GPI-B locus (Buth 1982). A parsimony analysis of gene-silencing events provided evidence of reactivation of a locus in Minytrema melanops (Ferris & Whitt 1978). Regulatory mutations may play an important role in relative duplicate gene expression in catostomids (Ferris 1984). The less anodally migrating isozyme of LDH-B is less active as a result of regulatory control that reduces the message (Lim & Bailey 1977). Also, the less negative, acidic isozymes have weaker expression than duplicates with a greater net charge (Ferris & Whitt 1979). Natural selection may favor regulatory mutations that reduce relative expression of less stable, acidic isozymes (Ferris 1984). Instances of duplicate gene silencing are used as phylogenetic characters in estimating relationships among catostomids (Ferris & Whitt 1977, 1978; Buth 1978, 1979). Ferris & Whitt (1978) examined duplicate gene expression in representative species of most North American genera. Duplicate gene expression characters also were used to infer phylogenies in Moxostomatini (Buth 1979b), in western North American Catostomini (Crabtree & Buth 1981, 1987), and in Myxocyprinus asiaticus (Tsoi et al. 1989). Buth et al. (1992) used duplicate gene expression and allozyme divergence as species-specific characters to investigate suspected hybridization of the Tahoe Sucker (Catostomus tahoensis) and Cui-ui (Chasmistes cujus) in Pyramid Lake, Nevada.

Genetic Variability Catostomids have an average per-locus heterozygosity of 5% (Ferris & Whitt 1980); loci that retain duplicate expression have a higher average heterozygosity (7.6%) than loci that have been silenced (4.3%). Mean per-locus heterozygosity also differs among genera and species. The morphologically primitive genera Ictiobus and Carpiodes have an average heterozygosity of 9.2% (Ferris & Whitt 1980), but the genera Catostomus, Hypentelium, and Moxostoma typically have average heterozygosities of about 3% (Ferris 1984). But, this pattern does not always hold; the genus Cycleptus has low average heterozygosity, and the Shorthead Redhorse (Moxostoma macrolepidotum) has an average heterozygosity of 10% (Ferris 1984).

Hybridization and Introgression Hybridization in nature among species and genera within Catostomidae is documented extensively (e.g., Hubbs et al. 1943; Hubbs & Miller 1953; Dauble & Buschbom 1981; Scribner et al. 2000). Hybridization is relatively frequent within the tribe Catostomini (Hubbs et al. 1943) and is documented within Ictiobus (Johnson & Minckley 1969) but is apparently less frequent among species of Moxostoma. Nine interspecific combinations of naturally occurring hybrids are known within Catostomus (Hubbs et al. 1943), and intergeneric hybrids between Xyrauchen and Catostomus (Hubbs & Miller 1953; Buth et al. 1987) and Chasmistes and Catostomus (Tranah & May 2006) are also known. Environmental disturbances may be associated with increased rates of hybridization (Nelson 1974), and the effect of artificial impoundment of waterways on hybridization rates among catostomids is an issue of considerable importance in their conservation. Hybridization and introgression can result in a bewildering array of intra- and interspecific morphological variation (Markle et al. 2005), often obscuring taxonomic identities and phylogenetic relationships. These difficulties are particularly apparent in the evaluation of relationships among sympatric Suckers from western North America. The endangered, endemic June Sucker (Chasmistes liorus) and the wider-ranging Utah Sucker (Catostomus ardens) inhabit Utah Lake, Utah. The Suckers of Utah Lake display a gradient of morphologies from individuals apparently adapted to the benthic feeding characteristic of C. ardens to mid-level planktivorous individuals that display characteristic C. liorus morphology to a spectrum

CATOSTOMIDAE: SUCKERS

of intermediate forms (Mock et al. 2006). Based on the morphology of contemporary and museum specimens, a severe drought in the 1930s was postulated to cause extensive hybridization between C. ardens and C. liorus, resulting in the formation of the hybrid lineage C. liorus mictus and the subsequent extinction of C. liorus (Miller & Smith 1981). Cook (2001), however, asserted morphological evidence is insufficient to support a hypothesis of hybridization between these lineages. Similarly, analyses of mitochondrial DNA sequences and amplified fragment length polymorphism data do not support the hypothesis of hybridization between these lineages in Utah Lake (Mock et al. 2006). Fairly recent ecological selection may have contributed to the divergent morphologies of Suckers in Utah Lake (Cook 2001; Mock et al. 2006). The Suckers of the Klamath Basin, Oregon and California, present another example of an apparent disconnect between the phylogenetic signals provided by morphology and genetics that may be the result of hybridization and introgression (Markle et al. 2005). The Klamath Basin is inhabited by three genera of catostomids represented by four native species with a wide range of morphological characteristics. The endangered Shortnose Sucker (Chasmistes brevirostris) and Lost River Sucker (Deltistes luxatus) may be introgressed and may continue to hybridize with the non-imperiled Klamath Largescale Sucker (Catostomus snyderi) and Klamath Smallscale Sucker (Catostomus rimiculus, Miller & Smith 1981). Using a combination of amplified fragment length polymorphism and single strand conformation polymorphism techniques, four of these species and their hybrids were distinguished, but little genetic variation was present among these lineages (Tranah et al. 2003). The Zuni Bluehead Sucker (Catostomus dicobolus yarrowi) displays morphological characters similar to those of the Bluehead Sucker (C. discobolus) with respect to the number of gill rakers and similar to those of the Rio Grande Sucker (Catostomus plebeius) in numbers of vertebrae and dorsal fin rays, jaw morphology, and pigmentation (Smith 1966). Based on morphological data, Smith (1966) hypothesized that C. d. yarrowi is an intergrade resulting from a Late Pleistocene stream capture. The capture led to introgression of C. plebeius characters into the genome of C. discobolus populations inhabiting the headwaters of the Little Colorado River drainage, New Mexico and Arizona. Meristic and morphometric data and variation in 35 allozyme loci were cited as supporting the hypothesis of stream capture and introgressive origin of C. d. yar-

471

rowi (Smith et al. 1983). An expanded reexamination of allozyme data revealed evidence of introgression of C. plebeius into C. d. yarrowi in a single creek in New Mexico but refuted the hypothesis that C. d. yarrowi originated from introgressive hybridization (Crabtree & Buth 1987). Hybridization between Catostomus commersonii and C. macrocheilus is fairly common in three drainages in British Columbia (J. S. Nelson 1968). Nine of 11 lakes examined were occupied by putative F1 hybrids comprising an average of 7% of the individuals examined. Hybridization between the two species does not result in increased embryo mortality, and F1 hybrids are fertile, suggesting the absence of post-reproductive isolating mechanisms. Despite the apparent fertility of hybrids, all hybrid specimens examined were presumed to be F1 progeny (J. S. Nelson 1968). To our knowledge, no one has looked for successful reproduction among the F1 progeny or introgression of these hybrids back into the parental genomes. Ethological isolation is probably the most important mechanism preventing initial hybridization with C. commersonii spawning over shallow gravel substrates and C. macrocheilus spawning over deeper, sandier substrates (J. S. Nelson 1968). Morphologically divergent species within Catostomus may hybridize in areas of sympatry. The Sonora Sucker (Catostomus insignis) and Desert Sucker (Catostomus clarki) are broadly sympatric and hybridize throughout a wide region (Clarkson & Minckley 1988). Catostomus clarki is morphologically adapted for specialized feeding by scraping organic material from solid substrates, but C. insignis is a generalized carnivore (Schreiber & Minckley 1982). Putative F1 hybrids of the two species are intermediate with respect to characteristic morphological traits but display wide variation in feeding behavior, spanning the feeding strategies used by the parental species (Clarkson & Minckley 1988). Anthropogenic environmental disturbances (e.g., dams, species introductions) are associated with the occurrence and frequency of hybridization in Suckers (Cooke et al. 2005). Based on morphological characters, hybridization between Ictiobus bubalus and I. niger (Heard 1958) and between I. cyprinellus and I. bubalus (Johnson & Minckley 1969) is documented in reservoirs. Catostomus commersonii and C. catostomus rarely hybridize under natural conditions (Nelson 1973). They did so in the upper Kananaskis reservoir, Alberta, and the hybridization is associated with introduction of both species to the reservoir, environmental disturbance related to water level fluctuations caused by hydroelectric generation, and high numbers of

472 FRESHWATER FISHES OF NORTH AMERICA

C. catostomus relative to C. commersonii (Nelson 1973). In contrast, frequency of hybridization between C. commersonii and C. macrocheilus in the Williston reservoir, British Columbia, following impoundment of the Peace River was not higher than that of populations inhabiting relatively undisturbed habitats (Nelson 1974).

Molecular Markers A number of biochemical and molecular markers are used in the identification and analysis of genetic variation within and among populations of catostomids. Analyses of mitochondrial DNA sequences, microsatellite DNA markers, amplified fragment length polymorphisms (AFLP), and allozyme data were used to examine taxonomic identity (Tranah et al. 2003; Mock et al. 2006; Tranah & May 2006), potential inbreeding due to small effective population sizes (Lippe et al. 2006), interspecific or intergeneric hybridization (Tranah & May 2006), identification of early life history stages (Wirgin et al. 2004), assessment of genetic variation before captive propagation (Wirgin et al. 2001), and post-propagation population monitoring (Dowling et al. 2005). Microsatellite DNA primers are developed for a variety of catostomid taxa, including the genera Cycleptus (Bessert et al. 2007), Catostomus, Chasmistes, Deltistes (Tranah et al. 2001b; Cardall et al. 2007), and Moxostoma (Lippe et al. 2004). These studies include cross-species amplification profiles for the microsatellite DNA primers, indicating their probable utility in other catostomid taxa. Although some initial cost is associated with their development, microsatellite DNA markers have the potential to address a number of important questions associated with conservation, especially estimates of intra- and interpopulation genetic diversity, gene flow, and the size of the reproductive population. For example, analysis of 21 polymorphic microsatellite loci of Moxostoma hubbsi indicated that populations retained much of their diversity in spite of recent severe population declines and fragmentation throughout the limited range of this imperiled species (Lippe et al. 2006). Although long generation time may have slowed the loss of genetic diversity, simulations indicated that the effective population size of M. hubbsi must remain >400 individuals in order to preserve 90% of the species’ genetic diversity over the next century. One of the difficulties frequently encountered by researchers studying the genetics of polyploid organisms like Suckers is the co-amplification of paralogous loci when using nuclear markers such as microsatellite DNA. Bessert

et al. (2007) used cloning and sequencing to isolate 11 microsatellite DNA loci from their paralogs in Cycleptus elongatus. An alternative approach using AFLP and single strand conformation polymorphism techniques was used in the development of three co-dominant markers for the analyses of genetic identity among Suckers of the Klamath Basin (Tranah et al. 2003). One of several factors threatening endangered C. brevirostris and D. luxatus is the progressive loss of genetic identity through introgression or hybridization with sympatric, non-imperiled Catostomus snyderi and C. rimiculus (Markle et al. 2005). Co-dominant markers specific to these two species (Tranah et al. 2003) are useful in the identification of hybrid individuals and monitoring of introgression among Suckers of the Klamath Basin.

Geographic Genetic Variation Studies of geographic genetic variation in catostomids are limited. Most allozyme studies of geographic variation concentrated on Catostomus (Buth & Crabtree 1982; Ferris et al. 1982; Buth et al. 1987) and Cycleptus (Buth & Mayden 2001). These studies (Ferris et al. 1982; Buth et al. 1987; Buth & Mayden 2001) revealed significant geographic structuring of populations, although this is not always the case (e.g., Buth & Crabtree 1982). Modern molecular techniques have added new insights into our understanding of geographic genetic variation in catostomids, especially by revealing previously unrecognized clades and the influence of historic factors in shaping current lineages (Wirgin et al. 2001; Berendzen et al. 2003; Mock et al. 2006). Only two papers on catostomids are published to date explicitly using phylogeography in the title or abstract. Berendzen et al. (2003) used mitochondrial DNA sequences to examine phylogeographic patterns in Hypentelium. Their results demonstrated concordance between mitochondrial lineages within Hypentelium nigricans (Northern Hog Sucker) and paleohydrological drainage patterns in eastern North America. Using a combination of mitochondrial DNA sequences and microsatellite DNA markers, McPhee et al. (2008) also reported phylogeographic patterns structured by drainage for the Rio Grande Sucker in New Mexico.

PHYSIOLOGY

Environmental pH Tolerances The physiological response of catostomids to extreme environmental acidity or alkalinity is documented in several

CATOSTOMIDAE: SUCKERS

species. Increased anthropogenic nutrient inputs into upper Klamath Lake, south-central Oregon, result in summer and autumn blooms of the blue-green algae, Aphanizomenon flos-aquae, leading to increases in ambient pH values (pH 9.5–10.5) that last several weeks (Falter & Cech 1991). In controlled experiments, Catostomus snyderi and Chasmistes brevirostris displayed sustained loss of equilibrium at pH 10.73 and 9.55, respectively. The pH maximum of C. brevirostris is significantly lower than for sympatric species C. snyderi and the Tui Chub (Siphateles bicolor) and is not correlated significantly with body mass. The comparatively low critical pH maximum of C. brevirostris makes it particularly susceptible to environmental alkalinity during seasonal phytoplankton blooms. Increased ambient pH may be an important factor in C. brevirostris population declines in upper Klamath Lake (Falter & Cech 1991). Physiological stress resulting from exposure to low environmental pH significantly affects ionoregulation in Catostomus commersonii (Fraser & Harvey 1984; Hobe et al. 1984; Hobe & McMahon 1988). Survival in acidic environments is related largely to the ability to prevent loss of sodium ions (Fraser & Harvey 1984). Rapid branchial loss of sodium ions in acidic water (pH 4.0–4.5) leads to decreased plasma osmotic pressure, resulting in decreased extracellular fluid volume and increases in intracellular fluid pressure, hematocrit, hemoglobin concentration, red blood cells, blood viscosity, and blood pressure. Loss of sodium ions in acidic ecosystems is ameliorated by high ambient concentrations of calcium ions, such that low ambient pH is most deleterious to fish populations in soft-water environments (Fraser & Harvey 1984). Catostomus commersonii exposed to acidic soft water exhibit rapid, shallow ventilatory pumping and secrete white mucus that covers the gills and skin, partially reducing branchial ion efflux and gas exchange (Hobe et al. 1984). Acute exposure of C. commersonii to acidic soft water at pH 4.3 results in rapid net influx of protons; whole-body losses of sodium, chloride, calcium, and potassium ions; plasma acidosis from altered respiratory and metabolic function; reduced plasma oxygen partial pressure; increased plasma carbon dioxide partial pressure; increased blood lactate; and increased hemoconcentration (Hobe et al. 1984). Exposure of C. commersonii to acidic soft water at pH 4 for 48 h results in complete mortality (Fraser & Harvey 1984). During a fluctuating ambient pH regime consisting of alternating exposure to soft water at pH 4 and pH 7, the deleterious effects of acid exposure on C. commersonii are not ameliorated by periods of exposure to neutral ambient

473

pH. Fluctuating ambient pH may be more detrimental to C. commersonii inhabiting soft-water ecosystems than gradual, continued exposure to low ambient pH (Hobe & McMahon 1988). Juvenile Robust Redhorse (Moxostoma robustum) exhibit a relatively broad range of pH tolerance in laboratory trials. Juveniles exhibited no mortality over a 96-h period when exposed to soft water with pH values ranging from 4.6 to 9.0. Complete mortality of juvenile M. robustum resulted from exposure to extremely alkaline soft water within the first 3 h at pH ≥10.5, within 55 h at pH 9.9, and within 70 h at pH 9.5. When exposed to extremely low ambient pH, juveniles exhibited complete mortality within 11 h at pH 4.0 and 90% mortality over a 96-h exposure at pH 4.3 (Walsh et al. 1998). Low ambient pH values significantly affect the reproductive life history of C. commersonii. Maturation occurs later and at larger body sizes in C. commersonii from acidic lakes in Ontario with ambient pH ranging from 4.9 to 5.6 than in fish from two lakes with pH values of 6.30 and 6.35 (Trippel & Harvey 1987ab). In females and males the reproductive lifespan of C. commersonii is longer in circumneutral lakes than in lakes with low ambient pH. In contrast to adults in circumneutral lakes, individuals from acidic lakes exhibit increased mortality rates with the onset of sexual maturity. Later onset of sexual maturity and greater reproductive output for a shorter duration before high post-spawning mortality may represent an adaptive life history strategy for C. commersonii under the physiologically stressful conditions of low ambient pH (Trippel & Harvey 1987ab).

Thermal Biology and Metabolism Physiological temperature tolerances and the effects of ambient temperature on respiratory metabolism are known for adults and juveniles in several species of catostomids. In C. commersonii from headwater streams in Missouri, the critical maximum temperature (loss of equilibrium) was 34.9°C, and minimum dissolved oxygen concentration at cessation of opercular movements was 0.98 mg/l (Smale & Rabeni 1995). The hyperthermia tolerance of C. commersonii was lower and the hypoxia tolerance intermediate relative to other fishes collected from the same headwater streams. Critical thermal maxima in adults are known for some species in almost all catostomid genera (Tables 13.2–13.13); the range of critical temperatures for these species is about 30–39°C. To our knowledge, however, the interaction

474

FRESHWATER FISHES OF NORTH AMERICA

Table 13.2. Life history traits of Carpiodes. Life History Trait

Description

References

Strictly freshwater (ppt) Critical thermal maxima Maximum size recorded in length and weight Maximum age

Yes, 0.9–5.81 37–39°C 500 mm TL; 5.45 kg 10–13 years

Age and size at first reproduction Iteroparous or semelparous Fecundity estimates (ovarian counts)

2–3 years; 200–250 mm TL Iteroparous 5,000–200,000

Echelle et al. (1972) Spotila et al. (1979), Mundahl (1990) Trautman (1981) Vanicek (1961), Woodward & Wissing (1976) Trautman (1981), Jester (1972)

Mature egg diameter Egg deposition sites

1.7–2.1 mm Firm gravel bottom, free of vegetation; spawning may occur at night or afternoon April–September; >31°C Slow sections and overflow pools of small streams; 0.3–1.0 m deep 5–13 days at 15–27°C; yolk-sac larva 5–6 mm TL No Crustaceans, chironomids, diatoms, filamentous algae, duckweed Main channels of larger streams and rivers, over sand and gravel, free of vegetation Spawning migrations C. velifer in decline

Range of nesting dates and temperatures Habitat of spawning sites; average water depth Incubation period; larval type at hatching Mean size at hatching and swim-up Parental care Major dietary items (adults) General year-round habitat Migratory or diadromous Imperilment status 1

Behmer (1969), Boschung & Mayden (2004) Behmer (1969), Yeager (1980) Bucholz (1957), Walburg & Nelson (1966), Jester (1972) Woodward & Wissing (1976) Bucholz (1957), Walburg & Nelson (1966) Kay et al. (1994) Yeager (1980) Eastman (1977) Trautman (1981), Boschung & Mayden (2004) Curry & Spacie (1984) Boschung & Mayden (2004)

Salinity at collection locality.

between critical thermal maxima and critical dissolved oxygen minima is known only for two species of Suckers inhabiting the Klamath Basin. Chasmistes brevirostris and C. snyderi have critical thermal maxima between 32 and 33°C (Castleberry & Cech 1992). Both species also have comparable critical dissolved oxygen minima between 11.8 and 12.7 torr (Castleberry & Cech 1992). Because dissolved oxygen levels are often lowest below the thermocline in lakes during summer, benthically oriented species such as C. brevirostris and C. snyderi are particularly susceptible to low dissolved oxygen levels in upper Klamath Lake where concentrations can be ≤5 torr (Scoppettone et al. 1986). Reduced water temperatures from hypolimnetic dam discharges reduce hatching success, development rate, and oxygen consumption of Xyrauchen texanus (Bozek et al. 1990). In laboratory experiments, no eggs incubated at 8°C successfully hatched (Marsh 1985; Bozek et al. 1990). Egg mortality, occurring largely at pre-morula stages of development, was higher at 10°C than at 15 or 20°C (Bozek

et al. 1990). Development rate and oxygen consumption of X. texanus were related positively to temperature. Upper and lower critical thermal maxima of 1-, 2-, and 3-month-old juvenile M. robustum were determined in experiments involving sudden, acute exposure to physiologically stressful conditions. Lower critical thermal maxima were influenced significantly by acclimation temperature and ranged from 5.3°C for 3-month-old juveniles (acclimated at 15°C) to 19.4°C for 2-month-old fish (acclimated at 30°C) (Walsh et al. 1998). Age did not significantly affect upper and lower critical thermal maxima within acclimation temperature groups. Average critical thermal maxima ranged from 34.9°C for fish acclimated at 20°C to 37.16°C for fish acclimated at 30°C. Juvenile M. robustum displayed increased rates of opercular ventilation at and near upper critical thermal maxima with fish acclimated at 20°C displaying higher ventilatory frequency than those acclimated at 30°C. Lower temperatures significantly affect the cardiac output of Catostomus macrocheilus. The critical swimming

CATOSTOMIDAE: SUCKERS

475

Table 13.3. Life history traits of Catostomus. Life History Trait

Description

References

Strictly freshwater (ppt)

Yes, ≤131 32–35°C

Reimers & Bond (1967), Moyle (2002)

Critical thermal maxima

Maximum size recorded in length and weight Maximum age Age and size at first reproduction Iteroparous or semelparous Fecundity estimates (ovarian counts) Egg deposition sites

800 mm TL; weight not given

Mature egg diameter

10–20+ years 2–3 years; 70–150 mm TL Iteroparous 2,000–140,000 Small streams, shoals, or shorelines free from vegetation and silt, 10–150 cm 1.7–3.6 mm

Range of nesting dates and temperatures Habitat of spawning sites; average water depth Incubation period; larval type at hatching

March–July; 10–15°C Rocky areas of small streams or shorelines of lakes, twilight hours 4–15 days, 10–21°C; yolk-sac larva

Mean size at hatching and swim-up Parental care Major dietary items (adults)

8–12 mm None Diatoms, dipteran larvae, mollusks, cladocerans, copepods, detritus, fish eggs Cool, well-oxygenated streams, rivers, and lakes Spawning migrations 22 species Vulnerable, Threatened, or Endangered

General year-round habitat Migratory or diadromous Imperilment status 1

Reutter & Herdendorf (1976), Castleberry & Cech (1992), Smale & Rabeni (1995) Page & Burr (1991) Hauser (1969), Dauble (1986) Hauser (1969), Dauble (1986) Geen et al. (1966), Bailey (1969) Stewart (1926), Vessel & Eddy (1941) Bailey (1969), Curry & Spacey (1984) Stewart (1926), Fish (1932), Fuiman & Witman (1979), Dauble (1986) Raney (1943), Dauble (1986) Curry & Spacey (1984), Raney (1943) Raney & Webster (1942), Long & Ballard (1976) Fuiman & Whitman (1979) Dauble (1986) Page & Burr (1991) Werner (1979), Dauble (1986) Jelks et al. (2008)

Salinity at collection locality.

speed, maximum cardiac output, and scope of cardiac output were significantly lower in swimming C. macrocheilus at 5°C relative to values at 10°C but were not different within the range of 10–16°C (Kolok et al. 1993). Cardiac and swimming performance of C. macrocheilus is significantly reduced at temperatures 31°C 495 mm TL; weight not given

Reutter & Herdendorf (1976) Lambou (1961)

Iteroparous or semelparous Fecundity estimates (ovarian counts) Mature egg diameter Egg deposition sites

Iteroparous 19,600–51,000 2.3–2.6 mm hardened Riffles and shoals above pools, in depressions behind large rocks March–May; 12–20°C Riffles and shoals above pools

Boschung & Mayden (2004) White & Haag (1977), Boschung & Mayden (2004) Kay et al. (1994) White (1974) Boschung & Mayden (2004) McSwain & Gennings (1972), Hogue & Buchanan (1977), Pflieger (1997) Kay et al. (1994) Hogue & Buchanan (1977), Pflieger (1997)

10+ years 1–4 years; 150–300 mm

Range of nesting dates and temperatures Habitat of spawning sites; average water depth Incubation period; larval type at hatching Mean size at hatching and swim-up Parental care Major dietary items (adults) General year-round habitat Migratory or diadromous

4.5–12 days, 14–20°C; yolk-sac larva

Hogue & Buchanan (1977), Pflieger (1997)

6 mm, 10 mm None Copepods, cladocerans, chironomid larvae Generalist; streams, rivers, and lakes No

Kay et al. (1994)

Imperilment status

Localized population declines

These migrations can range from a few hundred meters (e.g., Hypentelium, Matheney & Rabeni 1995) to several hundred kilometers (e.g., Cycleptus, Ictiobus; Fig. 13.18c; Hesse et al. 1982; Mettee 2000; Peterson et al. 2000). Migratory behavior associated with reproduction is documented in Catostomus (Olson & Scidmore 1963; Werner 1979; Kennen et al. 1994; McKinney et al. 1999; Douglas & Douglas 2000; Doherty et al. 2010), Carpiodes (Madsen 1971; Parker & Franzin 1991; Bonneau & Scarnecchia 2002), Chasmistes (Scoppettone et al. 1983; Scoppettone & Vinyard 1991), Cycleptus, Deltistes (Golden 1969), Hypentelium, Ictiobus (Bulow et al. 1988), Minytrema (McSwain & Gennings 1972), Moxostoma (Bulow et al. 1988; Parker & Franzin 1991; Cook & Bunt 1999), and Xyrauchen (Tyus 1987; Tyus & Karp 1990). Cycleptus meridionalis (Southeastern Blue Sucker) demonstrates a remarkable homing tendency to spawning sites with individuals returning year after year to a single brush pile to spawn (S. Mettee, pers. comm.). Increased water flow due to snowmelt or rain, rather than temperature, is the primary cue triggering spawning migra-

White & Haag (1977) Boschung & Mayden (2004) Kay et al. (1994), Boschung & Mayden (2004) Boschung & Mayden (2004)

tions in many Suckers (Lucas & Baras 2001; see also spawning season and conditions subsection).

Spawning One of the earliest notes on spawning in Suckers is by Reighard (1904), who presents a brief account of spawning behavior in Ictiobus niger (Fig. 13.18bc). Later, Reighard (1920) provided detailed descriptions of spawning behavior and sexual dimorphism in Catostomus commersonii, Pealip Redhorse (Moxostoma pisolabrum), and Northern Hog Sucker (Hypentelium nigricans). Spawning behavior is surprisingly similar for most species of catostomids, and the following summary of generalized spawning behavior is from Reighard (1920) and Page & Johnston (1990). Suckers generally spawn in the spring over rubble, gravel, or coarse sand substrates (Tables 13.2–13.13; Figs. 13.18d and 13.22); Erimyzon oblongus will also spawn over vegetation. Stream- and river-dwelling Suckers spawn in shallow riffles with moderate to fast currents. Lake-

CATOSTOMIDAE: SUCKERS

483

Table 13.11. Life history traits of Moxostoma. Life History Trait

Description

References

Strictly freshwater (ppt) Critical thermal maxima Maximum size recorded in length and weight Maximum age Age and size at first reproduction Iteroparous or semelparous Fecundity estimates (ovarian counts) Egg diameter (mature ova) Egg deposition sites

Yes, 4–61 35.1–35.4°C 76 cm TL; 8 kg

Walsh et al. (1998) Reash et al. (2000) RRCC (2010)

8–30+ years 2–5 years; size varies Iteroparous 1,000–44,000 2.6–4.4 mm Shallow riffles or deep runs over gravel 9–31°C Shallow riffles or deep runs over gravel 3–9 days at 14.4–22°C; yolk-sac larva 7.7–11 mm hatching None Varies among species Small to large rivers; some species adapt to reservoirs Spawning migrations

Meyer (1962), Mongeau et al. (1992) Meyer (1962) Kay et al. (1994) Jenkins & Burkhead (1994) Kay et al. (1994) Etnier & Starnes (1993), Jenkins & Burkhead (1994), Boschung & Mayden (2004) Jenkins & Burkhead (1994), Kay et al. (1994) Etnier & Starnes (1993), Jenkins & Burkhead (1994), Boschung & Mayden (2004) Kay et al. (1994) Kay et al. (1994)

Range of nesting dates and temperatures Habitat of spawning sites; average water depth Incubation period; larval type at hatching Mean size at hatching and swim-up Parental care Major dietary items (adults) General year-round habitat Migratory or diadromous Imperilment status 1

Seven species Vulnerable to Endangered, one species Extinct

Trautman (1981), Boschung & Mayden (2004) Trautman (1981), Etnier & Starnes (1993), Boschung & Mayden (2004) Etnier & Starnes (1993), Jenkins & Burkhead (1994), Boschung & Mayden (2004) Jelks et al. (2008)

100% juvenile mortality when acclimated in soft water.

dwelling Suckers either migrate upstream to spawn in riffles within tributaries or spawn over coarse substrates along shallow lake shorelines. Ictiobus cyprinellus and I. niger typically use inundated floodplains (a behavior not observed in I. bubalus) but may spawn along shorelines or ascend streams (Fig. 13.18bcd) if floodplains are unavailable (Yeager 1936; Johnson 1963; Trautman 1981). In streams, male Suckers congregate over spawning beds, while females occupy deeper pools upstream of the riffle. When ready to spawn the female drifts tail-first downstream into the spawning area and is approached from either side by >2 males (Fig. 13.18c and Fig. 13.22). Males press against the female, and the fish vibrate rapidly while releasing their gametes. A single spawning act generally lasts 800,000 eggs female−1 year −1 (Kay et al. 1994). In various Iowa lakes, populations averaged about 400,000 eggs female−1 year−1 (Carlander 1969). Carpiodes carpio (288 g, average weight) from the Des Moines River, Iowa, produced between 4,430 and 154,000 ova/female with most females producing 400 mm (Braaten et al. 1999). Species of Cycleptus become reproductively mature at ages 3–11, when individuals are 30–60 cm TL (Kay et al. 1994; Peterson et al. 1999). Maximum age of C. meridionalis based on increment analysis of the opercle is 31 for females and 33 for males (Peterson & Nicholson 1997; Peterson et al. 1999). These maximum age estimates probably also apply to C. elongatus (Burr & Mayden 1999), but to our knowledge, estimates of the age of this species based on opercular bones are unavailable.

The western North American lake Suckers (genera Chasmistes and Deltistes) are among the longest-lived catostomids. Chasmistes cujus lives ≥40 years, reaching maturity at ages 6–12. In a study of reproductive ecology in C. cujus, females at age 44 remained highly fecund and had an estimated reproductive life of 29 years (opercle bone aging; Scoppettone et al. 2000). In the Klamath Basin, California and Oregon, water temperature and dissolved oxygen have substantial effects on growth rates of D. luxatus and C. brevirostris (Terwilliger et al. 2003). The lifespans of these two species may reach 43 and 33 years, respectively (opercle bone aging; Scoppettone & Vinyard 1991; Markle & Cooperman 2001). Deltistes luxatus may attain 91 cm TL; C. brevirostris may reach 64 cm TL (Moyle 2002). Pond-reared X. texanus reach maturity at ages 2–6 with males usually reaching maturity more rapidly than females (Minckley et al. 1991). Growth is rapid when food is available and is highest in the population in Lake Mead, Arizona and Nevada (Papoulias & Minckley 1989; Ruppert et al. 1999). A mark-recapture study of the Lake Mead population estimated mean annual growth at 18.7 mm between July and the following June (Ruppert et al. 1999). Minytrema melanops and Erimyzon spp. mature on average at ages 2–4 and 14–30 cm TL (Jackson 1957; Pfleiger 1997). Lifespans of Erimyzon spp. and Minytrema are 5–6 years and >10 years, respectively (Boschung & Mayden 2004). Within the redhorse and jumprock Suckers (Moxostoma), the Copper Redhorse (M. hubbsi) is the largest (>70 cm TL) and most long-lived (>30 years) species. The Blacktip Redhorse (M. cervinum) may be the smallest species in the genus with adults averaging 70–165 mm SL (Jenkins & Burkhead 1994); no age estimates appear to be available for this species. Small, headwater species mature more quickly than large, riverine Suckers. Species of Hypentelium and Thoburnia mature within 1–3 years at sizes between 7 and 17 cm TL (Raney & Lachner 1946ab; Carlander 1969; Scott & Crossman 1973; Trautman 1981). Though headwater species mature more quickly than larger Suckers such as Moxoxtoma, their lifespans are considerably shorter. In New York streams, H. nigricans individuals may live ≤10 years (Raney & Lachner 1946b). Western species of Catostomus with limited geographic ranges are generally characterized by short lifespan, small size, and early maturity in comparison to more widespread congeners. In Montana, C. platyrhynchus reach means of 93, 116, and 131 mm TL at ages 1, 2, and 3, respectively (Hauser 1969). All individuals matured at ages 3–5.

CATOSTOMIDAE: SUCKERS

Minimum size at maturity is 127 mm TL for females and 115 mm TL for males (Hauser 1969; Wydoski & Wydoski 2002). Individuals from a population in Lost Creek Reservoir, Utah, grew faster than those in Montana. Maximum size for C. platyrhynchus was 220 mm TL for females and 196 mm TL for males. The oldest observed individuals were age 6 (Wydoski & Wydoski 2002). Catostomus plebeius and C. microps live 50% of the fish biomass (Lalancette 1977; Trippel & Harvey 1987a; Chen & Harvey 1999). Species of Carpiodes and Ictiobus are consistently among the most abundant fishes sampled in large rivers (Braaten & Guy 1999). Species of Moxostoma also occur in high abundance; they comprised >25% of total fish biomass in the Des Moines River, Iowa (Meyer 1962). Other studies reveal species occurring at lower natural densities, such as Erimyzon oblongus, averaging 8 adults/ ha (Wagner & Cooper 1963). For most Suckers, especially western species, larval development is intimately associated with flow regime. Young-of-the-year spend the first weeks of life drifting downstream from the spawning area, eventually reaching the basin, lake, or reservoir where they complete development. A few individuals may drift into backwaters or oxbows and return to the main stem after a few weeks. Drift primarily occurs at night, which may reduce the risk of predation. Drift may begin after initial feeding, as inferred from observations of C. brevirostris and D. luxatus (Cooperman & Markle 2003b). In the Gila River, larval drift samples contained ≤95% catostomids (Bestgen et al. 1987; Sublette et al. 1990). In the Smith and Van Duzen Rivers, California, however, suckers made up a much smaller proportion of larval drift (White & Harvey 2003). Once juveniles are large enough to swim against the current, most move to deeper waters in the lake but may not associate with adult fish (Scoppettone et al. 1983). Among western Suckers, Catostomus warnerensis demonstrates markedly different larval ecology. Young of this species immediately avoid drift by inhabiting refugia in the substrate. In field experiments, larvae resisted downstream transport by using available cover to limit the distance and duration of drift events (Kennedy & Vinyard 1997). This behavior in C. warnerensis larvae may have evolved in response to unreliable lake habitat in Warner Valley during the Pleistocene because of climatic fluctuations.

Fish Kills, Diseases, and Die- Offs Large, highly productive fishes that are not easily caught on hook and line, such as Suckers, have been dubbed “trash fish” by some anglers and, historically, by resource managers. Early efforts to increase sportfish production included widespread use of piscicides (fish poisons) in hopes of reducing competition from Suckers and other putative trash species. The infamous Green River Fish Con-

495

trol Project is among the few deliberate fish kills that received a quantitative evaluation (Binns 1967). Initiated in 1962, this project dumped about 79,500 l of rotenone (a piscicide) into >700 km of the Green River, Wyoming. The combined impacts of rotenone application and the impoundment of Flaming Gorge Reservoir, just downstream, greatly reduced native non-gamefishes, including three species that would later be protected under the U.S. Endangered Species Act. Federal wildlife agencies, urged by nongovernment conservation groups and public media, quickly contested the use of piscicide on native fish populations. Unfortunately, populations of X. texanus and numerous cyprinids had already been drastically impacted. With few exceptions (such as Ictiobus cyprinellus), catostomids are relatively sensitive to low oxygen concentrations and pH extremes (Becker 1983). With increasing human disturbance, the frequency of large-scale fish die-offs has increased, especially in isolated western drainages (Perkins et al. 2000). Endorheic lakes in the western United States are subject to large hydrologic fluctuations, which may be natural or anthropogenic in origin. Upper Klamath Lake, Oregon, has experienced dramatic die-offs of Chasmistes brevirostris and Deltistes luxatus. Large-scale fish kills in upper Klamath Lake and similar watersheds may be attributable to extreme fluctuations in cyanobacteria populations, disease, dissolved oxygen decrease, temperature increase, or any combination of these factors (Scoppettone & Vinyard 1991; Perkins et al. 2000). Manmade reservoirs and developing agriculture synergistically augment these risks by increasing nutrient load and reducing dissolved oxygen levels in the watershed. In Klamath Basin, a disease known as columnaris (caused by Flavobacterium columnare) is much more prevalent among Suckers when they are exposed to un-ionized ammonia, low dissolved oxygen, temperature and pH extremes, or combinations of these factors (Marin & Saiki 1999; Markle & Cooperman 2001). Natural die-offs also occur during prolonged drought; however, most native Suckers have evolved behavioral and life history traits that minimize drought mortality. Die-offs also are associated with extreme floods, although these are normally localized events from which catostomid populations quickly recover (Greenfield et al. 1970; Moyle 2002).

Parasitism Catostomids are hosts to a variety of metazoan ectoparasites and endoparasites, many of which have minimal or unknown impacts on host metabolism and immunological

496

FRESHWATER FISHES OF NORTH AMERICA

health. In addition to parasite diversity, relative incidence of parasitism in Suckers may exceed that of all other sympatric fishes. In a survey of fish parasites of southeastern Washington, C. macrocheilus and C. columbianus had the highest rate of infection (91.7%) among all species examined (Griffith 1953). Hoff man (1998) provided a comprehensive list of metazoans known to parasitize catostomids, which includes platyhelminths, nematodes, acanthocephalans, annelids, and arthropods. Catostomid parasites have been characterized primarily in abundant and widespread taxa such as Catostomus and Ictiobus. The intensity of catostomid parasite infestations often follows a seasonal pattern, as shown in the copepod, Lernaea cyprinacea, the trematode, Triganodistomum attenuatum, and the cestode, Glaridacris catostomi (Whitaker & Schlueter 1975; Muzzall 1980b; Marcogliese 1991). Parasite diversity and host organ specificity may differ markedly between sympatric Suckers. In Lake Superior, a comparison of parasites in C. commersonii and C. catostomus found 17 and 8 species, respectively (Hogue et al. 1993). Differences in parasite diversity among sympatric species may be due to host niche breadth (Barton 1980) or temperature limitations as mediated by host habitat preference (Becker 1983). Some catostomid parasites demonstrate strong microhabitat preference that leads to regional specificity within host organs. For example, Neoechinorhynchus crassus prefers areas of the gastrointestinal system where aminopeptidase levels are highest (Uglem & Beck 1972). This species occurs in the anterior portion of the gut in C. macrocheilus and C. catostomus, but prefers the posterior region of the gut in C. commersonii (Hogue et al. 1993). Similar patterns are observed in Triganodistomum attenuatum (Muzzall 1980a). Suckers are hosts for freshwater Mussel (Unionoidea) larvae, known as glochidia, that attach to gill or fin membranes. The degree of host specificity for most glochidiaSucker associations is usually unknown. Some Suckers (as well as other fishes), however, serve as hosts for host generalist mussels (e.g., Anodonta implicata, Alewife Floater, Lampsilis reeveiana brevicula, Ozark Broken Ray, and Pyganodon cataracta, Eastern Floater, host C. commersonii, Davenport & Warmuth 1965; OSUMD 2010; Margaritifera falcata, Western Pearlshell, host Catostomus tahoensis, Murphy 1942). Similarly, the genus Hypentelium serves as host to at least four host generalist mussel species (e.g., Alasmidonta undulata, Triangle Floater, OSUMD 2010; L. r. brevicula, OSUMD 2010; Lasmigona costata, Flutedshell, OSUMD 2010; Strophitus spp., Haag & Warren 1997; Williams et al. 2008). From a conservation perspective,

H. nigricans may be one of the most important host species, being the only documented host for the endangered, host specialist Cumberland Elktoe, Alasmidonta atropurpurea (Gordon & Layzer 1993). During freshwater life history stages, Lampreys (Petromyzontidae) can parasitize catostomids, preferring larger species when available. For example, I. bubalus are parasitized preferentially by Chestnut Lampreys (Ichthyomyzon castaneus) and Ohio Lampreys (Ichthyomyzon bdellium) (Metee et al. 1996). Lampreys will, however, attach to catostomids >36 g in the laboratory (Parker & Lennon 1956).

Haff Disease Haff disease was first diagnosed in patients from the Koenigsberger Haff shore of the Baltic Sea. Discovered in 1924, the disease is characterized by unexplained rhabdomyolysis (muscle cell destruction) following consumption of Burbot (L. lota), Eels, or Pikes from Europe, and species of Ictiobus from the eastern United States (Centers for Disease Control and Prevention 1998). Once muscle cell membranes are destroyed, cell contents are released into the bloodstream, causing widespread complications. Signs of Haff disease appear between 30 min and 18 h after consumption, and include severe muscular rigidity, vomiting, and coffee-colored urine. Patients complain of muscle pain, dry mouth, pain to light touch, painful breathing, and whole-body numbness. Laboratory tests reveal elevated levels of creatine kinase and myoglobin in patients’ blood (Buchholz et al. 2000). Major symptoms fade after 36 h of treatment, but muscular weakness and soreness may persist for >6 months. Fever is not observed, nor is splenomegaly, hepatomegaly, or neurological anomaly. The etiology of Haff disease is unknown, but in absence of fever the causal agent is deemed a noninfectious toxin (Centers for Disease Control and Prevention 1998). Hexane-soluble, neutral lipids were extracted from leftover fish tissue and, when fed to lab mice, caused redbrown urine and behavioral changes consistent with muscle impairment. Historically, human case fatality is 1% (Zu 1939).

CONSERVATION Although the 74 species of catostomids in North America represent only about 8% of the species diversity in North American freshwater fishes (Burr & Mayden 1992; Jen-

CATOSTOMIDAE: SUCKERS

kins & Burkhead 1994), about 35% of catostomid taxa are either Endangered, Threatened, or of Special Concern, and three, June Sucker (C. liorus liorus), Snake River Sucker (Chasmistes muriei), and Harelip Sucker (M. lacerum), are extinct (Miller et al. 1989; Warren & Burr 1994; Nelson et al. 2004; Burkhead 2012). Warren & Burr (1994:14) noted that “Suckers . . . are imperiled disproportionately relative to their representation in the total freshwater native fish fauna.” At present, eight species and two subspecies are considered endangered in North America based on official conservation status lists: Salish Sucker, Catostomus catostomus ssp.; Sonora Sucker, Catostomus insignis; Modoc Sucker, Catostomus microps; Shortnose Sucker, Chasmistes brevirostris; Cui-ui, Chasmistes cujus; June Sucker, Chasmistes liorus mictus; Lost River Sucker, Deltistes luxatus; Lake Chubsucker, Erimyzon sucetta; Copper Redhorse, Moxostoma hubbsi; and Razorback Sucker, Xyrauchen texanus (SEDESOL 1994; USFWS 2010b; COSEWIC 2011). Caveats to the above list are that the Sonora Sucker is known from only two localities in Mexico, hence the inclusion of this species on the SEDESOL list (Miller et al. 2005), and that the Lake Chubsucker is rare in Canada and hence on the Canadian list (COSEWIC 2011). For the Sonora Sucker, both Mexican localities are headwater streams in the Santa Cruz and San Pedro River basins that flow into Arizona, where the species is predominantly found (Minckley 1973). The Lake Chubsucker is not imperiled over its broad range in the United States (Warren et al. 2000; Jelks et al. 2008). Interestingly, eight of these species are from western North America and are restricted in distribution to single drainages or basins with the exception of the Razorback Sucker from the Colorado River basin. Other evaluations by ichthyologists and fisheries biologists of Threatened and Endangered species status based on best available information, however, indicate that an additional 35 taxa are imperiled in part, or all, of their distributions (Table 13.14). Surprisingly, only six species (Catostomus warnerensis, C. liorus, D. luxatus, Moxostoma congestum, Myxocyprinus asiaticus, and X. texanus) are featured in the Threatened Fishes of the World series published in Environmental Biology of Fishes (Williams 1995; Marsh 1996; Whitney & Belk 2000; Gao et al. 2008; Bean et al. 2009; Evans et al. 2009). Threats to Suckers (like those affecting other North American freshwater fishes) include dams, diverting water for agricultural purposes, pollution, habitat degradation, introduced species, and a negative image by some fishery managers and the public regarding ecological

497

roles and interactions of Suckers with more “desirable” fishes (Holey et al. 1979; Minckley & Douglas 1991; Minckley et al. 1991; Warren & Burr 1994; Warren et al. 2000; Cooke et al. 2005). Cooke et al. (2005) reviewed the general conservation status of catostomids and presented several regional case studies highlighting threats to these species. Interestingly, they concluded that imperiled Suckers are often faced with multiple threats throughout their life history that have a synergistic effect, magnifying the susceptibility of these species to continuing, long-term threats. Suckers from the Klamath Basin of southern Oregon and northern California are an excellent example of synergistic effects from multiple threats (see also ecology section). Four species of Suckers inhabit this basin; two species, Shortnose Sucker and Lost River Sucker, occur primarily in upper Klamath Lake, but spawn in tributary streams and rivers or in springs or other areas along the shore of the lake. Both species are long-lived, obligate lake dwellers; sexual maturity occurs between 4 and 9 years of age (Cooperman & Markle 2003a). Historically, large spawning migrations of both species occurred in the Sprague River; these migrations were large enough to support both indigenous and commercial fisheries (Cope 1879; Gilbert 1898). Sport fisheries for both species occurred until their listing as endangered under the U.S. Endangered Species Act (Moyle 2002). In 1914, construction of a dam on the Sprague River effectively blocked access to about 90% of spawning habitat (Cooke et al. 2005). Subsequent development of a federal irrigation project in the basin led to fluctuating lake levels that isolated or dried lakeshore spawning habitats and larval and juvenile Sucker nursery grounds (Markle & Dunsmoor 2007). In addition, larval and juvenile Suckers can be entrained in the primary irrigation canal for this project, carrying them away from the lake. Eutrophication of the lake from increasing anthropogenic activities has negatively affected water quality; pH levels >10.0 and dissolved oxygen concentrations 1,100 kg/ha and average about 560 kg/ha (Borgstrom 1978). When raised in monoculture, buffalofish stocks are often less expensive to raise than other commercially important fishes because they can be successfully grown without the addition of supplemental feed; when raised in polyculture with Channel Catfish, buffalofishes feed on excess Catfish feed and naturally occurring zooplankton (McGeachin 1993). Buffalofishes reach marketable size from 1–3 kg (McGeachin 1993). The harvest of buffalofishes peaked in 1982 with >14,500 mt captured but declined dramatically in the late 1980s to about 1,800 mt/ year harvested from the 1990s to 2004 (FAO 2006).

Baitfish Because small Suckers are an important natural source of forage for large predatory fishes (Robison & Buchanan 1988), anglers prize them as excellent baitfish. Several

species of Suckers are raised in intense aquaculture facilities in North America for use as baitfish. In areas with coldwater fisheries Catostomus commersonii are successfully cultured as baitfish (Bandeen & Leatherland 1997). White Sucker fry also are raised for use as feed for cultured esocids such as Northern Pike, E. lucius, and Muskellunge, E. masquinongy (Westers & Stickney 1993). Chubsuckers (Erimyzon spp.) are locally important baitfish to anglers in several warm-water regions (Davis 1993). Both White Suckers and chubsuckers are most commonly intensively cultured in fertilized, earthen ponds without the addition of exogenous feed (Davis 1993). Nevertheless, aquaculturists have had difficulty in growing juvenile White Suckers >1 g of body weight, possibly because of nutritional deficiencies in the natural diets available to juvenile Suckers in ponds under dense stocking conditions (Bandeen 1995). Providing juvenile White Suckers with an exogenous feed originally formulated for salmonids that is high in lipids and protein and low in carbohydrates improved their growth rates (Bandeen & Leatherland 1997).

Ecological Indicators In addition to the direct use of Suckers as human food or baitfish, Sucker populations may be of considerable economic importance as ecological indicators of the health of ecosystems impacted by low-level contamination. Populations of indigenous, widely distributed fish species with known life history characteristics that are suitable for laboratory research may provide valuable information about ecosystem-level effects of low-level contamination by serving as environmental sentinels (Munkittrick & Dixon 1989). Suckers are particularly well suited to study as ecological indicators of low-level contamination for several reasons. As benthic foragers, Suckers are exposed directly to contaminated sediments, reducing lag time between a contamination event and the detection of its ecological effects (Munkittrick & Dixon 1989). Many species of Sucker mature quickly and are common and easily collected, facilitating the study of relatively short-term population changes. As a consequence, Suckers are ideal indicator species for monitoring environmental quality.

LITERATURE GUIDE Although a comprehensive book has never been dedicated to the Catostomidae, many of the regional faunal guides

CATOSTOMIDAE: SUCKERS

provide a wealth of information on the biology, distribution, local importance, and evolutionary relationships of Suckers. Three books, in particular, provide exceptional information and illustrations of eastern North American Suckers: Freshwater Fishes of Virginia (Jenkins & Burkhead 1994), The Fishes of Tennessee (Etnier & Starnes 1993), and Fishes of Alabama (Boschung & Mayden 2004). For western Suckers, we recommend Inland Fishes of California (Moyle 2002), The Fishes of New Mexico (Sublette et al. 1990), and Freshwater Fishes of Mexico (Miller et al. 2005). The field guide by Page & Burr (1991, revised 2011) provides useful information on identification, distribution, and habitat for species of Suckers north of Mexico. Bruner

501

(1991b) compiled a bibliography for the family. Smith’s (1992) chapter on the phylogeny and biogeography of catostomids, which is frequently cited in this chapter, provides an excellent review of the historical literature associated with the taxonomy and systematics of the family.

Acknowledgments The authors thank Mel Warren and Brooks Burr for the invitation to participate in this volume and for Mel’s editorial efforts on this chapter. This material is based upon work supported, in part, by the National Science Foundation under Grant No. 0431263 (PMH).

This page intentionally left blank

Literature Cited

Able, K. W., and M. P. Fahay. 2010. Ecology of Estuarine Fishes: Temperate Waters of the Western North Atlantic. Johns Hopkins University Press, Baltimore, Maryland. Abou-Seedo, F. S., and I. C. Potter. 1979. The estuarine phase in the spawning run of the River Lamprey Lampetra fluviatilis. Journal of Zoology, London 188:5–25. Abramoff, P., R. M. Darnell, and J. S. Balsano. 1968. Electrophoretic demonstration of the hybrid origin of the gynogenetic teleost Poecilia formosa. American Naturalist 102:555–558. Acuña, S., D-F. Deng, P. Lehman, and S. Teh. 2012. Sublethal dietary effects of Microcystis on Sacramento Splittail, Pogonichthys macrolepidotus. Aquatic Toxicology 110–111:1–8. Adámek, Z., I. Sukop, P. M. Rendón, and J. Kouril. 2003. Food competition between 2+ Tench (Tinca tinca L.), Common Carp (Cyprinus carpio L.) and Bigmouth Buffalo (Ictiobus cyprinellus Val.) in pond polyculture. Journal of Applied Ichthyology 19:165–169. Adams, C. C., and T. L. Hankinson. 1928. The ecology and economics of Oneida Lake Fishes. Roosevelt Wildlife Annals 1:235–548. Adams, L. A. 1940. Some characteristic otoliths of American Ostariophysi. Journal of Morphology 66:497–527. Adams, L. A. 1942. Age determination and rate of growth in Polyodon spathula, by means of the growth rings of the otoliths and dentary bone. The American Midland Naturalist 28:617–630. Adams, P. B., C. Grimes, J. E. Hightower, S. T. Lindley, M. L. Moser, and M. J. Parsley. 2007. Population status of North American Green Sturgeon, Acipenser medirostris. Environmental Biology of Fishes 79:339–356. Adams, S. B., and M. L. Warren, Jr. 2005. Recolonization by warmwater fishes and crayfishes after severe drought in upper Coastal Plain hill streams. Transactions of the American Fisheries Society 134:1173–1192. Adams, S. R., G. L. Adams, and J. J. Hoover. 2003a. Oral grasping: a distinctive behavior of cyprinids for maintaining station in flowing water. Copeia 2003:851–857. Adams, S. R., G. L. Adams, and G. R. Parsons. 2003b. Critical swimming speed and behavior of juvenile Shovelnose Sturgeon and Pallid Sturgeon. Transactions of the American Fisheries Society 132:392–397.

Adams, S. R., J. J. Hoover, and K. J. Killgore. 1999a. Swimming endurance of juvenile Pallid Sturgeon (Scaphirhynchus albus). Copeia 1999:802–807. Adams, S. R., J. J. Hoover, and K. J. Killgore. 2000a. Swimming endurance of the Topeka Shiner, an imperiled midwestern minnow. American Midland Naturalist 144:178–186. Adams, S. R., T. M. Keevin, K. J. Killgore, and J. J. Hoover. 1999b. Stranding potential of young fishes subjected to simulated vessel-induced drawdown. Transactions of the American Fisheries Society 128:1230–1234. Adams, S. R., G. R. Parsons, J. J. Hoover, and K. J. Killgore. 1997. Observations of swimming ability in Shovelnose Sturgeon (Scaphirhynchus platorynchus). Journal of Freshwater Ecology 12:631–633. Aday, D. D., D. P. Philipp and D. H. Wahl. 2006. Sex-specific life history patterns in Bluegill (Lepomis macochirus): interacting mechanisms influence individual body size. Oecologia 147:31–38. Aday, D. D., D. H. Wahl, and D. P. Philipp. 2003. A mechanism for social inhibition of sexual maturation in Bluegill. Journal of Fish Biology 62:486–490. Aeschlimann, P. B., M. A. Häberli, T. B. H. Reusch, T. Boehm, and M. Milinski. 2003. Female Sticklebacks (Gasterosteus aculeatus) use self-reference to optimize Mhc allele number during mate selection. Behavioral Ecology and Sociobiology 54:119–126. Afzelius, B. A., and S. D. Mims. 1995. Sperm structure of the Bowfin, Amia calva L. Journal of Submicroscopic Cytology and Pathology 27:291–294. Agassiz, A. 1879. The development of Lepidosteus. Proceedings of the American Academy of Arts and Sciences. 14:65–76. AGFC (Arkansas Game and Fish Commission). 2007. Fishing Guidebook. Arkansas Game and Fish Commission, Little Rock. Aguilera, C., R. Mendoza, G. Rodríguez, and G. Márquez. 2002. Morphological description of Alligator Gar and Tropical Gar larvae, with an emphasis on growth indicators. Transactions of the American Fisheries Society 131:899–909. Ahlgren, M. O. 1990a. Diet selection and the contribution of detritus to the diet of the juvenile White Sucker (Catostomus commersoni). Canadian Journal of Fisheries and Aquatic Sciences 47:41–48.

504

LITERATURE CITED

Ahlgren, M. O. 1990b. Nutritional significance of facultative detritivory to the juvenile White Sucker (Catostomus commersoni). Canadian Journal of Fisheries and Aquatic Sciences 47:49–54. Ahlgren, M. O. 1996. Selective ingestion of detritus by a north temperate omnivorous fish, the juvenile White Sucker (Catostomus commersoni). Environmental Biology of Fishes 46:375–381. Aida, K., K. Tsukamoto, and K. Yamauchi (eds.). 2003. Eel Biology. Springer, Tokyo. Aieta A. E., and K. Oliveira. 2009. Distribution, prevalence, and intensity of the swim bladder parasite Anguillicola cyassus in New England and eastern Canada. Diseases of Aquatic Organisms 84:229–235. Alam, S. K., M. S. Brim, G. A. Carmody, and F. M. Parauka. 2000. Concentrations of heavy and trace metals in muscle and blood of juvenile Gulf Sturgeon (Acipenser oxyrinchus desotoi) from the Suwannee River, Florida. Journal of Environmental Science and Health, Part A. 35:645–660. Albanese, B., P. L. Angermeier, and C. Gowan. 2003. Designing mark-recapture studies to reduce effects of distance weighting on movement distance distributions of stream fishes. Transactions of the American Fisheries Society 132:925–939. Albanese, B., P. L. Angermeier, and J. T. Peterson. 2009. Does mobility explain variation in colonisation and population recovery among stream fishes? Freshwater Biology 54:1444–1460. Albert, V., B. Jónsson, and L. Bernatchez. 2006. Natural hybrids in Atlantic Eels (Anguilla anguilla, A. rostrata): evidence for successful reproduction and fluctuating abundance in space and time. Molecular Ecology 15:1903–1916. Aldrich, D. V. 1965. Observations on the ecology and life cycle of Prochristianella penaei, Kruse (Cestoda: Trypanorhyncha). The Journal of Parasitology 51:370–376. Alexander, C. M., A. I. Myhr, III, and J. L. Wilson. 1985. Harvest potential of Paddlefish stocks in Watts Bar Reservoir, Tennessee. Proceedings of the Annual Conference of the Southeastern Association of Fish and Wildlife Agencies 39:45–55. Alexander, C. M., and D. C. Peterson. 1982. Feasibility of a commercial Paddlefish harvest from Norris Reservoir, Tennessee. Proceedings of the Annual Conference of the Southeastern Association of Fish and Wildlife Agencies 36:202–212. Alexander, M. L. 1914. The paddle-fish (Polyodon spathula). (Commonly called “spoonbill cat.”). Transactions of the American Fisheries Society 44:73–78. Alexander, M. L. 1915. More about the paddle-fish (Polyodon spathula). Transactions of the American Fisheries Society 45:34–39. Alexander, R. D. 1975. Natural selection and specialized chorusing behavior in acoustic insects, p. 35–77. In Insects, Science, and Society. D. Pimental (ed.). Academic Press, New York. Alexander, R. M. 1966. The functions and mechanisms of the protrusible upper jaws of two species of cyprinid fish. Journal of Zoology 149:288–296. Ali, M. A., and M. Anctil. 1974. Retinomotor responses and isolated photoreceptors in Amia calva (Holostei: Amiidae). Copeia 1974:379–386. Allan, J. R. 1986. The influence of species composition on behavior in mixed-species cyprinid shoals. Journal of Fish Biology 29, Supplement A:97–106. Allen, D. M., W. S. Johnson, and V. Ogburn-Matthews. 1995. Trophic relationships and seasonal utilization of salt-marsh creeks by zooplanktivorous fishes. Environmental Biology of Fishes 42:37–50.

Allen, G. R. 1989. Freshwater Fishes of Australia. T. F. H. Publications, Inc., Neptune City, New Jersey. Allen, G. R., S. H. Midgley, and M. Allen. 2002. Field Guide to Freshwater Fishes of Australia. Western Australian Museum, Perth. Allen, P. J., and J. J. Cech, Jr. 2007. Age/size effects on juvenile Green Sturgeon, Acipenser medirostris, oxygen consumption, growth, and osmoregulation in saline environments. Environmental Biology of Fishes 79:211–229. Allen, P. J., B. Hodge, I. Werner, and J. J. Cech, Jr. 2006. Effects of ontogeny, season, and temperature on the swimming performance of juvenile Green Sturgeon (Acipenser medirostris). Canadian Journal of Fisheries and Aquatic Sciences 63:1360–1369. Allen, W. F. 1911. Notes on the breeding season and young of Polyodon spathula. Journal of the Washington Academy of Science 10:280–282. Allendorf, F. W. 1975. Genetic variability in a species possessing extensive gene duplication: genetic interpretation of duplicate loci and examination of variation in populations of Rainbow Trout. Unpubl. Ph.D. diss., University of Washington, Seattle. Allendorf, F. W., R. F. Leary, P. Spruell, and J. K. Wenburg. 2001. The problems with hybrids: setting conservation guidelines. Trends in Ecology & Evolution 16:613–622. Allendorf, F. W., and R. F. Leary. 1988. Conservation and distribution of genetic variation in a polytypic species, the Cutthroat Trout. Conservation Biology 2:170–184. Allendorf, F. W., and R. S. Waples. 1996. Conservation and genetics of salmonid fishes, p. 238–280. In Conservation Genetics: Case Histories from Nature. J. C. Avise and J. L. Hamrick (eds.). Chapman and Hall, New York. Allis, E. P., Jr. 1920. The branches of the branchial nerves of fishes, with special reference to Polyodon spathula. The Journal of Comparative Neurology 32:137–153. Altinok, I., S. M. Galli, and F. A. Chapman. 1998. Ionic and osmotic regulation capabilities of juvenile Gulf of Mexico Sturgeon, Acipenser oxyrinchus de sotoi. Comparative Biochemistry and Physiology Part A 120:609–616. Alvarado-Ortega, J., O. Carranza-Castañeda, and G. AlvarezReyes. 2006. A new fossil species of Ictiobus (Teleostei: Catostomidae) from Pliocene lacustrine sediments near Tula de Allende, Hidalgo, Mexico. Journal of Paleontology 80:993–1008. Álvarez del Villar, J. 1966. Ictiología michoacana. IV. Contribución al conocimiento biológico y sistemático de las lampreas de Jacona, Mich., México. Anales de la Escuela Nacional de Ciencias Biológicas México13:107–144. Amarasinghe, U. S., and R. L. Welcomme. 2002. An analysis of fish species richness in natural lakes. Environmental Biology of Fishes 65:327–339. Amberg, J. J., R. Goforth, T. Stefanavage, and M. S. Sepúlveda. 2010. Sexually dimorphic gene expression in the gonad and liver of Shovelnose Sturgeon (Scaphirhynchus platorynchus). Fish Physiology and Biochemistry 36:923–932. Amemiya, C. T., and J. R. Gold. 1988. Chromosomal NORs as taxonomic and systematic characters in North American cyprinid fishes. Genetica 76:81–90. Amemiya, C. T., and J. R. Gold. 1990. Chromosomal nor phenotypes of 7 species of North American Cyprinidae, with comments on cytosystematic relationships of the Notropis volucellus species-group, Opsopoeodus emiliae, and the genus Pteronotropis. Copeia 1990:68–78.

LITERATURE CITED

Amemiya, C. T., P. K. Powers, and J. R. Gold. 1992. Chromosomal evolution in North American cyprinids, p. 515–550. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Amemiya, Y., and J. H. Youson. 2004. Primary structure of stanniocalcin in two basal Actinopterygii. General and Comparative Endocrinology 135:250–257. Amesbury, E., and F. F. Snelson, Jr. 1997. Spine replacement in a freshwater population of the Atlantic Stingray, Dasyatis sabina. Copeia 1997:220–223. Amin, O. M. 1989. Abnormalities in some helminth parasites of fish. Transactions of the American Microscopical Society 108:27–39. Ammerman, L. K., and D. C. Morizot. 1989. Biochemical genetics of endangered Colorado Squawfish populations. Transactions of the American Fisheries Society 118:435–440. Anders, P. J., D. L. Richards, and M. S. Powell. 2002. The first endangered White Sturgeon population: repercussions in an altered large river-floodplain ecosystem, p. 67–82. In Biology, Management, and Protection of North American Sturgeon. W. Van Winkle, P. Anders, D. H. Secor, and D. Dixon (eds.). American Fisheries Society Symposium 28, Bethesda, Maryland. Anderson, E. 1953. Introgressive hybridization. Biological Reviews 28:280–307. Anderson, K. A., R. A. Rountree, and F. Juanes. 2008. Soniferous fishes in the Hudson River. Transactions of the American Fisheries Society 137:616–626. Andreasen, J. K. 1975. Systematics and status of the family Catostomidae in southern Oregon. Unpubl. Ph.D. diss., Oregon State University, Corvallis. Andreasen, L. 1999. Captive propagation as a recovery tool for North American Sturgeon, p. 121–129. In Proceedings of the Symposium on the Harvest, Trade and Conservation of North American Paddlefish and Sturgeon, May 7–8, 1998, Chattanooga, TN. D. F. Williamson, G. W. Benz, and C. M. Hoover (eds.). TRAFFIC North America / World Wildlife Fund, Washington, D.C. Angermeier, P. L. 1992. Predation by rock bass on other stream fishes: experimental effects of depth and cover. Environmental Biology of Fishes 34:171–180. Angermeier, P. L., and I. J. Schlosser. 1989. Species-area relationships for stream fishes. Ecology 70:1450–1462. Angermeier, P. L., and M. R. Winston. 1998. Local vs. regional influences on local diversity in stream fish communities. Ecology 79:911–927. Angradi, T. R., J. S. Spaulding, and E. D. Koch. 1991. Diel food utilization by the Virgin River Spinedace, Lepidomeda mollispinis mollispinis, and Speckled Dace, Rhinichthys osculus, in Beaver Dam Wash, Utah. The Southwestern Naturalist 36:158–170. Angus, R. A. 1989. A genetic overview of poeciliid fishes, p. 51–68. In Ecology and Evolution of Livebearing Fishes (Poeciliidae). G. K. Meffe and F. F. Sneldon, Jr. (eds.). Prentice-Hall, New Jersey. Angus, R. A., and R. J. Schultz. 1979. Clonal diversity in the unisexual fish Poeciliopsis monacha-lucida: a tissue graft analysis. Evolution 33:27–40. Ankley, G. T., K. M. Jensen, M. D. Kahl, J. J. Korte, and E. A. Makynen. 2001. Description and evaluation of a short-term reproduction test with the Fathead Minnow (Pimephales promelas). Environmental Toxicology and Chemistry 20:1276–1290.

505

Ankley, G. T., and D. L. Villeneuve. 2006. The Fathead Minnow in aquatic toxicology: past, present and future. Aquatic Toxicology 78:91–102. Annett, C. A. 1998. Hunting behavior of Florida Largemouth Bass (Micropterus salmoides floridanus), in a channelized river. Environmental Biology of Fishes 53:75–87. Aoyama, J. 2003. Origin and evolution of the Freshwater Eels, p. 19–29. In Eel Biology. K. Aida, K. Tsukamoto, and K. Yamauchi (eds.). Springer, Tokyo. Aoyama, J. 2009. Life history and evolution of migration in catadromous Eels (genus Anguilla), Aqua-BioScience Monographs 2:1–42. Aoyama, J., M. Nishida, and K. Tsukamoto. 2001. Molecular phylogeny and evolution of the Freshwater Eel (genus Anguilla). Molecular Phyolgenetics and Evolution 20:450–459. Aoyama, J., and K. Tsukamoto. 1997. Evolution of the Freshwater Eels. Naturwissenschaften 84:17–21. Applebaum, S. L., and A. Cruz. 2000. The role of mate-choice copying and disruption effects in mate preference determination of Limia perugiae (Cyprinodontiformes, Poeciliidae). Ethology 106:933–944. Applegate, V. C. 1950. Natural history of the Sea Lamprey, Petromyzon marinus, in Michigan. United States Department of the Interior. Special Scientific Report—Fisheries No. 55:1–237. Applegate, V. C., and C. L. Brynildson. 1952. Downstream movement of recently transformed Sea Lampreys, Petromyzon marinus, in the Carp Lake River, Michigan. Transactions of the American Fisheries Society 81:275–290. Appy, R. G., and R. C. Anderson. 1981. The parasites of Lampreys, p. 1–42. In The Biology of Lampreys. Vol. 3. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Argent, D. G., W. G. Kimmel, R. Lorson, P. McKeown, D. M. Carlson, and M. Clancy. 2009. Paddlefish restoration to the upper Ohio and Allegheny River systems, p. 397–409. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Arias-Rodriguez, L., S. Paramo-Delgadillo, and W. M. ContrerasSanchez. 2009. Cariotipo del Pejelagarto Tropical Atractosteus tropicus (Lepisosteiformes: Lepisosteidae) y variación cromosómica en sus larvas y adultos. Revista de Biologia Tropical 57:529–539. Arndt, G. M., J. Gessner, E. Anders, S. Spratte, J. Filipiak, L. Debus, and K. Skora. 2000. Predominance of exotic and introduced species among Sturgeons captured from the Baltic and North Seas and their watersheds, 1981–1999. Boletín Instituto Español de Oceanografía 16:29–36. Arndt, G. M., J. Gessner, and C. Raymakers. 2002. Trends in farming, trade and occurrence of native and exotic Sturgeons in natural habitats in Central and Western Europe. Journal of Applied Ichthyology 18:444–448. Arratia, G. 1997. Basal teleosts and teleostean phylogeny. Paleoichthyologica 7:1–168. Arratia, G. 1999. The monophyly of Teleostei and stem-group teleosts. Consensus and disagreements, p. 265–334. In Mesozoic Fishes 2—Systematics, and Fossil Record. G. Arratia and H.-P. Schultze (eds.). Verlag Dr. Friedrich Pfeil, München.

506

LITERATURE CITED

Arratia, G. 2004. Mesozoic halecostomes and the early radiation of teleosts, p. 279–315. In Mesozoic Fishes 3—Systematics, Paleoenvironments and Biodiversity. G. Arratia and A. Tintori, editors. Verlag Dr. Friedrich Pfeil, München. Arrington, D. A., K. O. Winemiller, W. F. Loftus, and S. Akin. 2002. How often do fishes “run on empty”? Ecology 83:2145–2151. Artyukhin, E. N. 1995. On biogeography and relationships within the genus Acipenser. The Sturgeon Quarterly 3:6–8. Artyukhin, E. N. 2006. Morphological phylogeny of the order Acipenseriformes. Journal of Applied Ichthyology 22 (Supplement 1):66–69. Ash, L. R. 1962. Development of Gnathostoma procyonis Chandler, 1942, in the first and second intermediate hosts. The Journal of Parasitology 48:298–305. ASMFC (Atlantic States Marine Fisheries Commission). 2000. Interstate fishery management plan for the American Eel. ASMFC Fishery Management Report No. 36. Aspbury, A. S., J. M. Coyle, and C. R. Gabor. 2010b. Effect of predation on male mating behaviour in a unisexual-bisexual mating system. Behaviour 147:53–63. Aspbury, A. S., C. M. Espinedo, and C. R. Gabor. 2010a. Lack of species discrimination based on chemical cues by male Sailfin Mollies, Poecilia latipinna. Evolutionary Ecology 24:69–82. Aspbury, A. S., and C. R. Gabor. 2004. Differential sperm priming by male Sailfin Mollies (Poecilia latipinna): effects of female and male size. Ethology 110:193–202. Aspinwall, N., D. Carpenter, and J. Bramble. 1993. The ecology of hybrids between the Peamouth, Mylocheilus caurinus, and the Redside Shiner, Richardsonius balteatus, at Stave Lake, British Columbia, Canada. Canadian Journal of Zoology 71:83–90. Auer, N. A. 1982. Identification of larval fishes of the Great Lakes basin with emphasis on the Lake Michigan drainage. Great Lakes Fishery Commission, Special Publication 82–3. Auer, N. A. 1996. Importance of habitat and migration to Sturgeons with emphasis on Lake Sturgeon. Canadian Journal of Fisheries and Aquatic Sciences 53 (Supplement 1):152–160. Auer, N. A. 1999. Population characteristics and movements of Lake Sturgeon in the Sturgeon River and Lake Superior. Journal of Great Lakes Research 25:282–293. Auer, N. A. 2004. Conservation, p. 252–276. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. H. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. Auer, N. A., and E. A. Baker. 2002. Duration and drift of larval Lake Sturgeon in the Sturgeon River, Michigan. Journal of Applied Ichthyology 18:557–564. Austad, S. N. 1984. A classification of alternative reproductive behaviors and methods for field-testing ESS models. American Zoologist 24:309–319. Avise, J. C. 1977. Genic heterozygosity and rate of speciation. Paleobiology 3:422–432 Avise, J. C. 1992. Molecular population structure and the biogeographic history of a regional fauna: a case history with lessons for conservation biology. Oikos 63:62–76. Avise, J. C. 2001. Cytonuclear genetic signatures of hybridization phenomena: rationale, utility, and empirical examples from fishes and other aquatic animals. Reviews in Fish Biology and Fisheries 10:253–263.

Avise, J. C. 2003. Catadromous Eels of the North Atlantic: a review of molecular genetic findings relevant to natural history, population structure, speciation, and phylogeny, p. 31–48. In Eel Biology. K. Aida, K. Tsukamoto, and K. Yamauchi (eds.). Springer, Tokyo. Avise, J. C., J. Arnold, R. M. Ball, E. Bermingham, T. Lamb, J. E. Neigel, C. A. Reeb, and N. C. Saunders. 1987. Intraspecific phylogeography: the mitochondrial DNA bridge between population genetics and systematics. Annual Review of Ecology and Systematics 18:489–522. Avise, J. C., and F. J. Ayala. 1976. Genetic differentiation in speciose versus depauperate phylads: evidence from the California minnows. Evolution 30:46–58. Avise, J. C., G. S. Helfman, N. C. Saunders, and L. S. Hales. 1986. Mitochondrial DNA differentiation in North Atlantic Eels: population genetic consequences of an unusual life history pattern. Proceedings of the National Academy of Sciences 83:4350–4354. Avise, J. C., W. S. Nelson, J. Arnold, R. K. Koehn, G. C. Williams, and V. Thorsteinsson. 1990. The evolutionary genetic status of Icelandic Eels. Evolution 44:1254–1262. Avise, J. C., J. J. Smith, and F. J. Ayala. 1975. Adaptive differentiation with little genic change between two native California minnows. Evolution 29:411–426. Avise, J. C., J. C. Trexler, J. Travis, and W. S. Nelson. 1991. Poecilia mexicana is the recent female parent of the unisexual fish P. formosa. Evolution 45:1530–1533. Ayres, W. O. 1854. Description of new species of fishes from California. Daily Placer Times and Transcript 5. Baensch, H. A., and R. Riehl. 1995. Aquarienn Atlas. Band 4. Mergus Verlag GmbH, Verlad für Natur- und Heimtierkunde, Melle, Germany. Baerends, G. P., R. Brouwer, and H. T. J. Waterbolk. 1955. Ethological studies on Lebistes reticulates (Peters). I. An analysis of the male courtship pattern. Behaviour 8:249–335. Bahr, K. 1933. Das Flüssneunauge (Lampetra fluviatilis) als Urheber von Fischverletzungen. Mitteilungen des Deutschen Seefischerei-Vereins 49:3–8. Bailey, J. E. 1952. Life history and ecology of the Sculpin, Cottus bairdi punctulatus, in southwestern Montana. Copeia 1952:243–255. Bailey, M. M. 1969. Age, growth, and maturity of the Longnose Sucker, Catostomus catostomus, of western Lake Superior. Journal of the Fisheries Research Board of Canada 26:1289–1299. Bailey, R. M. 1951. A check list of fishes of Iowa with keys for identification, p. 187–238. In Iowa Fish and Fishing. 1st edition. J. R. Harlan and E. B. Speaker (eds.). Iowa State Conservation Commission, Des Moines. Bailey, R. M. 1956. A revised list of the fishes of Iowa with keys for identification, p. 327–377. In Iowa Fish and Fishing. 3rd edition. J. R. Harlan and E. B. Speaker (eds.). Iowa State Conservation Commission, Des Moines. Bailey, R. M. 1959a. Distribution of the American cyprinid fish, Notropis anogenus. Copeia 1959:119–123. Bailey, R. M. 1959b. A new catostomid fish, Moxostoma (Thoburnia) atripinne, from the Green River drainage, Kentucky and Tennessee. Occasional Papers of the Museum of Zoolology University of Michigan 599:1–19.

LITERATURE CITED

Bailey, R. M., and F. B. Cross. 1954. River Sturgeons of the American genus Scaphirhynchus: characters, distribution, and synonymy. Papers of the Michigan Academy of Science, Arts, and Letters 39:169–208. Bailey, R. M., and G. R. Smith. 1981. Origin and geography of the fish fauna of the Laurentian Great Lakes basin. Canadian Journal of Fisheries and Aquatic Sciences 38:1539–1561. Bain, M. B. 1997. Atlantic and Shortnose Sturgeons of the Hudson River: common and divergent life history attributes. Environmental Biology of Fishes 48:347–358. Baird, D., and R. E. Ulanowicz. 1989. The seasonal dynamics of the Chesapeake Bay ecosystem. Ecological Monographs 59:329–364. Baird, R. C. 1968. Aggressive behavior and social organization in Molliensia latipinna (Le Sueur). Texas Journal of Science 20:157–176. Bajer, P. G., and M. L. Wildhaber. 2007. Population viability analysis of lower Missouri River Shovelnose Sturgeon with initial application to the Pallid Sturgeon. Journal of Applied Ichthyology 23:457–464. Bajkov, A. 1930. Fishing industry and fisheries investigations in the prairie provinces. Transactions of the American Fisheries Society 60:215–237. Baker, J. A., W. A. Cresko, S. A. Foster, and D. C. Heins. 2005. Life-history differentiation of benthic and limnetic ecotypes in a polytypic population of Threespine Stickleback (Gasterosteus aculeatus). Evolutionary Ecology Research 7:121–131. Baker, J. A., and S. A. Foster. 1994. Observations on a foraging association between two freshwater stream fishes. Ecology of Freshwater Fish 3:137–139. Baker, J. A., and D. C. Heins. 1994. Reproductive life history of the North American madtom catfish, Noturus hildebrandi (Bailey & Taylor 1950), with a review of data for the genus. Ecology of Freshwater Fish 3:167–175. Baker, J. A., K. J. Killgore, and R .L. Kasul. 1991. Aquatic habitats and fish communities in the lower Mississippi River. Aquatic Sciences 53:313–356. Baker, J. A., and S. T. Ross. 1981. Spatial and temporal resource utilization by southeastern cyprinids. Copeia 1981:178–189. Balfour, F. M., and W. N. Parker. 1882. On the structure and development of Lepidosteus. Philosophical Transactions of the Royal Society of London 173:359–442. Ballantyne, P. K., and P. W. Colgan. 1978a. Sound production during agonistic and reproductive behaviour in the Pumpkinseed (Lepomis gibbosus), the Bluegill (L. macrochirus), and their hybrid Sunfish I. Context. Biology of Behaviour 3:113–135. Ballantyne, P. K., and P. W. Colgan. 1978b. Sound production during agonistic and reproductive behaviour in the Pumpkinseed (Lepomis gibbosus), the Bluegill (L. macrochirus), and their hybrid Sunfish II. Recipients. Biology of Behaviour 3:207–220. Ballantyne, P. K., and P. W. Colgan. 1978c. Sound production during agonistic and reproductive behaviour in the Pumpkinseed (Lepomis gibbosus), the Bluegill (L. macrochirus), and their hybrid Sunfish III. Response. Biology of Behaviour 3:221–232. Ballard, W. W., and R. G. Needham. 1964. Normal embryonic stages of Polyodon spathula (Walbaum). Journal of Morphology 114:465–478. Baltz, D. M., and P. B. Moyle. 1983. Segregation by species and size classes of Rainbow Trout, Salmo gairdneri, and Sacramento

507

Sucker, Catostomus occidentalis, in three California streams. Environmental Biology of Fishes 10:101–110. Baltz, D. M., B. Vondracek, L. R. Brown, and P. B. Moyle. 1991. Seasonal changes in microhabitat selection by Rainbow Trout in a small stream. Transactions of the American Fisheries Society 120:166–176. Banarescu, P. 1964. Fauna republicii populare Romine, volume 13, Pisces—Osteicthyes. Editura Academiei Republicii Populare Romine, Bucharest, Romania. Bandeen, J. 1995. Studies on the growth physiology of captive juvenile White Suckers (Catostomus commersoni) fed commercial fish diets. Unpubl. Master’s thesis, University of Guelph, Guelph, Ontario. Bandeen, J., and J. F. Leatherland. 1997. Evaluation of commercial catfish, tilapia and salmonid diets for growth promotion of White Suckers. Aquaculture International 5:315–326. Bandoli, J. H. 1997. Factors influencing reproductive success in male Spottail Darters (Etheostoma squamiceps, Pisces, Percidae). Proceedings of the Indiana Academy of Science 106:145–157. Bandoli, J. H. 2002. Brood defense and filial cannibalism in the Spottail Darter (Etheostoma squamiceps): the effects of parental status and prior experience. Behavioral Ecology and Sociobiology 51:222–226. Bandoli, J. H. 2006. Male Spottail Darters (Etheostoma squamiceps) do not use chemical or positional cues to discriminate between sired and foster eggs. Behavioral Ecology and Sociobiology 59:606–613. Bangham, R. V. 1955. Studies on fish parasites of Lake Huron and Manitoulin Island. American Midland Naturalist 53:184–194. Banish, N. P., J. T. Peterson, and R. F. Thurow. 2008. Physical, biotic, and sampling influences on diel habitat use by streamdwelling Bull Trout. North American Journal of Fisheries Management 28:176–187. Barbaro, K. C., M. S. Lira, M. B. Malta, S. L. Soares, D. G. Neto, J. L. C. Cardoso, M. L. Santoro, and V. Haddad, Jr. 2007. Comparative study on extracts from the tissue covering the stingers of the freshwater (Potamotrygon falkneri) and marine (Dasyatis guttara) Stingrays. Toxicon 50:676–687. Barber, W. E., and W. L. Minckley. 1966. Fishes of Aravaipa Creek, Graham and Pinal Counties, Arizona. Southwestern Naturalist 11:315–324. Barber, W. E., and W. L. Minkley. 1983. Feeding ecology of a southwestern cyprinid fish, the Spikedace, Meda fulgida Girard. The Southwestern Naturalist 28:33–40. Barber, W. E., D. C. Williams, and W. L. Minckley. 1970. Biology of the Gila Spikedace, Meda fulgida, in Arizona. Copeia 1970:9–18. Barbin, G. P., and W. H. Krueger. 1994. Behavior and swimming performance of elvers of the American Eel, Anguilla rostrata, in an experimental flume. Journal of Fish Biology 45:111–121. Barbin, G. P., and J. D. McCleave. 1997. Fecundity of the American Eel, Anguilla rostrata at 45 degree N in Maine, U.S.A. Journal of Fish Biology 51:840–847. Barbin, G. P., S. J. Parker, and J. D. McCleave. 1998. Olfactory clues play a critical role in the estuarine migration of silverphase American Eels. Environmental Biology of Fishes 53:283–291.

508

LITERATURE CITED

Barbour, C. D. 1973. A biogeographical history of Chirostoma (Pisces: Atherinidae): a species flock from the Mexican Plateau. Copeia 1973:533–556. Barbour, C. D., and J. H. Brown. 1974. Fish species diversity in lakes. The American Naturalist 108:473–489. Barbour, C. D., and R. R. Miller. 1978. A revision of the Mexican cyprinid fish genus Algansea. Miscellaneous Publications of the Museum of Zoology. University of Michigan:1–72. Bardack, D., and E. S. Richardson, Jr. 1977. New agnathous fishes from the Pennsylvanian of Illinois. Fieldiana: Geology 33:489–510. Bardack, D., and R. Zangerl. 1971. Lampreys in the fossil record, p. 67–84. In The Biology of Lampreys. Vol. 1. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Bardi, R. W., Jr., F. A. Chapman, and F. T. Barrows. 1998. Feeding trials with hatchery-produced Gulf of Mexico Sturgeon larvae. The Progressive Fish-Culturist 60:25–31. Barila, T. Y., and J. R. Stauffer, Jr. 1980. Temperature behavioral responses of the American Eel Anguilla rostrata (LeSueur), from Maryland. Hydrobiologia 74:49–51. Barko, V. A., M. W. Palmer, D. P. Herzog, and B. S. Ickes. 2004. Influential environmental gradients and spatiotemporal patterns of fish assemblages in the unimpounded upper Mississippi River. American Midland Naturalist 152:369–385. Barrientos-Villalobos, J., and A. Espinosa De Los Monteros. 2008. Genetic variation and recent population history of the Tropical Gar, Atractosteus tropicus Gill (Pisces: Lepisosteidae). Journal of Fish Biology 73:1919–1936. Barry, P. M., R. F. Carline, D. G. Argent, and W. G. Kimmel. 2007. Movement and habitat use of stocked juvenile Paddlefish in the Ohio River system, Pennsylvania. North American Journal of Fisheries Management 27:1316–1325. Barse, A. M., and D. H. Secor. 1999. An exotic nematode parasite of the American Eel. Fisheries 24:6–10. Bart, H. L., Jr. 1989. Fish habitat association in an Ozark stream. Environmental Biology of Fishes 24:173–186. Bart, H. L. Jr., and L. M. Page. 1991. Morphology and adaptive significance of fin knobs in egg-clustering darters. Copeia 1991:80–86. Bartels, H., and I. C. Potter. 2004. Cellular composition and ultrastructure of the gill epithelium of larval and adult Lampreys: implications for osmoregulation in fresh and seawater. The Journal of Experimental Biology 207:3447–3462. Bartels, H., A. Schmiedl, J. Rosenbruch, and I. C. Potter. 2009. Exposure of the gill epithelial cells of larval Lampreys to an iondeficient environment: a stereological study. Journal of Electron Microscopy 58:253–260. Bartley, D. M., K. Rana, and A. J. Immink. 2001. The use of interspecific hybrids in aquaculture and fisheries. Reviews in Fish Biology and Fisheries 10:325–337. Barton, B. A. 1980. Spawning migrations, age and growth and summer feeding of White and Longnose Suckers in an irrigation reservoir. Canadian Field Naturalist 94:300–304. Barton, B. A., A. B. Rahn, G. Feist, H. Bollig, and C. B. Schreck. 1998. Physiological stress responses of the freshwater chondrostean Paddlefish (Polyodon spathula) to acute physical disturbances. Comparative Biochemistry and Physiology Part A 120:355–363. Barton, D. G., and M. V. H. Wilson. 2005. Taphonomic variations in Eocene fish-bearing varves at Horsefly, British Columbia, reveal 10,000 years of environmental change. Canadian Journal of Earth Science 42:137–149.

Barton, N. H., and G. M. Hewitt. 1985. Analysis of hybrid zones. Annual Review of Ecology and Systematics 16:113–148. Bartsch, M. R., L. A. Bartsch, and S. Gutreuter. 2005. Strong effects of predation by fishes on an invasive macroinvertebrate in a large floodplain river. Journal of the North American Benthological Society 24:168–177. Basolo, A. L. 1990. Female preference for male sword length in the Green Swordtail (Pisces: Poeciliidae). Animal Behaviour 40:332–338. Basolo, A. L. 1995a. A further examination of a preexisting bias favoring a sword in the genus Xiphophorus. Animal Behaviour 50:365–375. Basolo, A. L. 1995b. Phylogenetic evidence for the role of a preexisting bias in sexual selection. Proceedings of the Royal Society of London B 259:307–311. Basolo, A. L., and G. Alcaraz. 2003. The turn of the sword: length increases male swimming costs in swordtails. Proceedings of the Royal Society of London B 270:1631–1636. Bassista, T. P., and K. J. Hartman. 2005. Reproductive biology and egg mortality of Bay Anchovy, Anchoa mitchilli, in the Hudson River Estuary. Environmental Biology of Fishes 73:49–59. Bath, D. W., J. M. O’Conner, J. B. Alber, and L. G. Arvidson. 1981. Development and identification of larval Atlantic Sturgeon (Acipenser oxyrhynchus) and Shortnose Sturgeon (A. brevirostrum) from the Hudson River estuary, New York. Copeia 1981:711–717. Battle, H. I., and W. M. Sprules. 1960. A description of the semibuoyant eggs and early developmental stages of the Goldeye, Hiodon alosoides (Rafinesque). Journal of the Fisheries Research Board of Canada 17:245–266. Bauer, O. N., O. N. Pugachev, and V. N. Voronin. 2002. Study of parasites and diseases of Sturgeons in Russia: a review. Journal of Applied Ichthyology 18:420–429. Baxter, E. W. 1956. Observations on the buccal glands of Lampreys (Petromyzonidae). Proceedings of the Zoological Society of London 127:95–118. Baxter, R. M. 1977. Environmental effects of dams and impoundments. Annual Review of Ecology and Systematics 8:255–283. Beach, H. 1902. The paddle fish (Polyodon spathula). Bulletin of Wisconsin Natural History Society 2:85–86. Beamesderfer, R. C. P., T. A. Rien, and A. A. Nigro. 1995. Differences in the dynamics and potential production of impounded and unimpounded White Sturgeon populations in the lower Columbia River. Transactions of the American Fisheries Society 124:857–872. Beamesderfer, R. C. P., M. L. Simpson, and G. J. Kopp. 2007. Use of life history information in a population model for Sacramento Green Sturgeon. Environmental Biology of Fishes 79:215–337. Beamish, F. W. H. 1973. Oxygen consumption of Petromyzon marinus in relation to body weight and temperature. Journal of the Fisheries Research Board of Canada 30:1367–1370. Beamish, F. W. H. 1974. Swimming performance of adult Sea Lamprey, Petromyzon marinus, in relation to weight and temperature. Transactions of the American Fisheries Society 103:355–358. Beamish, F. W. H. 1978. Swimming capacity, p. 107–187. In Fish Physiology, Vol. 7, W. S. Hoar and D. J. Randall (eds.). Academic Press, New York. Beamish, F. W. H. 1979. Migration and spawning energetics of the anadromous Sea Lamprey, Petromyzon marinus. Environmental Biology of Fishes 4:3–7.

LITERATURE CITED

Beamish, F. W. H. 1980a. Osmoregulation in juvenile and adult Lampreys. Canadian Journal of Fisheries and Aquatic Sciences 37:1739–1750. Beamish, F. W. H. 1980b. Biology of the North American anadromous Sea Lamprey, Petromyzon marinus. Canadian Journal of Fisheries and Aquatic Sciences 37:1924–1943. Beamish, F. W. H. 1993. Environmental sex determination in Southern Brook Lamprey, Ichthyomyzon gagei. Canadian Journal of Fisheries and Aquatic Sciences 50:1299–1307. Beamish, F. W. H., and L. S. Austin. 1985. Growth of the Mountain Brook Lamprey, Ichthyomyzon greeleyi Hubbs and Trautman. Copeia 1985:881–890. Beamish, F. W. H., and J.-A. Jebbink. 1994. Abundance of Lamprey larvae and physical habitat. Environmental Biology of Fishes 39:209–214. Beamish, F. W. H., and M. Legrow. 1983. Bioenergetics of the Southern Brook Lamprey, Ichthyomyzon gagei. Journal of Animal Ecology 52:575–590. Beamish, F. W. H., and I. C. Potter. 1975. The biology of the anadromous Sea Lamprey (Petromyzon marinus) in New Brunswick. Journal of Zoology, London 177:57–72. Beamish, F. W. H., I. C. Potter, and E. Thomas. 1979. Proximate composition of the adult anadromous Sea Lamprey, Petromyzon marinus, in relation to feeding, migration and reproduction. Journal of Animal Ecology 48:1–19. Beamish, F. W. H., P. D. Strachan, and E. Thomas. 1978. Osmotic and ionic performance of the anadromous Sea Lamprey, Petromyzon marinus. Comparative Biochemistry and Physiology 60A:435–443. Beamish, F. W. H., and E. J. Thomas. 1983. Potential and actual fecundity of the “paired” Lampreys, Ichthyomyzon gagei and I. castaneus. Copeia 1983:367–374. Beamish, R. J. 1973. Determination of age and growth of populations of the White Sucker (Catostomus commersoni) exhibiting a wide range in size at maturity. Journal of the Fisheries Research Board of Canada 30:607–616. Beamish, R. J. 1980. Adult biology of the River Lamprey (Lampetra ayresi) and the Pacific Lamprey (Lampetra tridentata) from the Pacific coast of Canada. Canadian Journal of Fisheries and Aquatic Sciences 37:1906–1923. Beamish, R. J. 2001. Updated status of the Vancouver Island Lake Lamprey, Lampetra macrostoma, in Canada. Canadian FieldNaturalist 115:127–130. Beamish, R. J., and E. J. Crossman. 1977. Validity of the subspecies designation for the Dwarf White Sucker (Catostomus commersoni utawana). Journal of the Fisheries Research Board of Canada 34:371–378. Beamish, R. J., and C.-E. M. Neville. 1992. The importance of size as an isolating mechanism in Lampreys. Copeia 1992:191–196. Beamish, R. J., and C.-E. M. Neville. 1995. Pacific salmon and Pacific herring mortalities in the Fraser River plume caused by River Lamprey (Lampetra ayresi). Canadian Journal of Fisheries and Aquatic Sciences 52:644–650. Beamish, R. J., and T. G. Northcote. 1989. Extinction of a population of anadromous parasitic Lamprey, Lampetra tridentata, upstream of an impassable dam. Canadian Journal of Fisheries and Aquatic Sciences 46:420–425. Beamish, R. J., and T. Uyeno. 1978. Karyotype of Hiodon tergisus and DNA values of Hiodon tergisus and Hiodon alosoides. Chromosome Information Ser vice 24:5–7.

509

Beamish, R. J., and R. E. Withler. 1986. A polymorphic population of Lampreys that may produce parasitic and nonparasitic varieties, p. 31–49. In Indo-Pacific Fish Biology: Proceedings of the Second International Conference on Indo-Pacific Fishes. T. Uyeno, R. Arai, T. Taniuchi, and K. Matsuura (eds.). Ichthyological Society of Japan, Tokyo. Bean, P. T., M. G. Bean, and T. H. Bonner. 2009. Threatened fishes of the world: Moxostoma congestum (Baird and Girard, 1854) (Catostomidae). Environmental Biology of Fishes 85:173–174. Bean, T. H. 1884. Cata logue of the collection of fishes exhibited by the United States National Museum, p. 387–510 (Pt. F). In Descriptive Cata logue of the Collections Sent from the United States to the International Fisheries Exhibition, London, 1883, Constituting a Report upon the American Section. Prepared under the direction of G. Brown Goode. Bulletin United States National Museum 27:1–1279. Beard, D. C. 1878. The paddle fish of the Mississippi. Scientific American 39:391. Beard, J. 1889. On the early development of Lepidosteus osseus— preliminary notice. Proceedings of the Royal Society of London 46:108–118. Bearden, C. M. 1965. Elasmobranch fishes of South Carolina. Contribution of Bears Bluff Laboratory Number 42, Wadmalaw Island, South Carolina. Beatty, D. D. 1975. Visual pigments of the American Eel, Anguilla rostrata. Vision Research 15:771–776. Becker, D., N. Galili, and G. Degani. 1992. Gcms-identified steroids and steroid glucoronides in gonads and holding water of Trichogaster trichopterus (Anabantidae, Pallas 1770). Comparative Biochemistry and Physiology B 103:15–19. Becker, G. C. 1983. Fishes of Wisconsin. University of Wisconsin Press, Madison. Beecher, H. A., W. C. Hixson, and T. S. Hopkins. 1977. Fishes of a Florida oxbow lake and its parent river. Florida Scientist 40:140–148. Beehler, B. M., and M. S. Foster. 1988. Hotshots, hotspots, and female preference in the organization of lek mating systems. American Naturalist 131:203–219. Begon, M., J. L. Harper, and C. R. Townsend. 1996. Ecology, individuals, populations and communities. 3rd edition. Blackwell Science, Cambridge, Massachusetts. Behmer, D. J. 1969. A method of estimating fecundity; with data of River Carpsucker, Carpoides carpio. Transactions of the American Fisheries Society 98:520–523. Beitinger, T. L. and W. A. Bennett. 2000. Quantification of the role of acclimation temperature in temperature tolerance of fishes. Environmental Biology of Fishes 58:277–288. Beitinger, T. L., W. A. Bennett, and R. W. McCauley. 2000. Temperature tolerances of North American freshwater fishes exposed to dynamic changes in temperature. Environmental Biology of Fishes 58:237–275. Beitinger, T. L., and L. C. Fitzpatrick. 1979. Physiological and ecological correlates of preferred temperature in fish. American Zoologist 19:319–329. Beitinger, T. L., J. J. Magnuson, W. H. Neill, and W. R. Shaffer. 1975. Behavioural thermoregulation and activity patterns in the Green Sunfish, Lepomis cyanellus. Animal Behaviour 23:222–229. Belanger, A. J., W. J. Arbuckle, L. D. Corkum, D. B. Gammon, W. Li, A. P. Scott and B. S. Zielinski. 2004. Behavioural and electrophysiological responses by reproductive female Neogobius

510

LITERATURE CITED

melanostomus to odours released by conspecific males. Journal of Fish Biology 65:933–946. Belk, M. C. 1998. Age and growth of June Sucker (Chasmistes liorus) from otoliths. Great Basin Naturalist 58:390–392. Belk, M. C., and L. S. Hales, Jr. 1993. Predation-induced differences in growth and reproduction of Bluegills (Lepomis macrochirus). Copeia 1993:1034–1044. Bell, M. A., W. E. Aguirre, and N. J. Buck. 2004. Twelve years of contemporary armor evolution in a Threespine Stickleback population. Evolution 58:814–824. Bell, M. A., and S. A. Foster. 1994. Introduction to the evolutionary biology of the Threespine Stickleback, p. 1–27. In The Evolutionary Biology of the Threespine Stickleback. M. A. Bell and S. A. Foster (eds.). Oxford University Press, New York. Belle-Isles, J.-C., D. Cloutier, and J. G. FitzGerald. 1990. Female cannibalism and male courtship tactics in Threespine Sticklebacks. Behavioral Ecology and Sociobiology 26:363–368. Bemis, W. E., E. K. Findeis, and L. Grande. 1997. An overview of Acipenseriformes. Environmental Biology of Fishes 48:25–71. Bemis, W. E., and L. Grande. 1992. Early development of the actinopterygian head. I. External development and staging of the Paddlefish Polyodon spathula. Journal of Morphology 213:47–83. Bemis, W. E., and L. Grande. 1999. Development of the median fins of the North American Paddlefish (Polyodon spathula), and a reevaluation of the lateral fin-fold hypothesis, p. 41–68. In Mesozoic Fishes 2—Systematics and Fossil Record. G. Arratia and H.-P. Schultze (eds.). Verlag Dr. Friedrich Pfeil, München, Germany. Bemis, W. E., and B. Kynard. 1997. Sturgeon rivers: an introduction to acipenseriform biogeography and life history. Environmental Biology of Fishes 48:167–183. Benner, J. S., J. C. Ridge, and R. J. Knecht. 2009. Timing of postglacial reinhabitation and ecological development of two New England, USA, drainages based on trace fossil evidence. Palaeogeography, Palaeoclimatology, Palaeoecology 272:212–231. Benner, J. S., J. C. Ridge, and N. K. Taft. 2008. Late Pleistocene freshwater fish (Cottidae) trackways from New England (USA) glacial lakes and a reinterpretation of the ichnogenus Broomichnium Kuhn. Palaeo 260:375–388. Bennett, D. K. 1979. Three Late Cenozoic fish faunas from Nebraska. Transactions of the Kansas Academy of Sciences 82:146–177. Benson, A. C., T. M. Sutton, R. F. Elliot, and T. G. Meronek. 2006. Biological attributes of age-0 Lake Sturgeon in the lower Peshtigo River, Wisconsin. Journal of Applied Ichthyology 22:103–108. Benson, B. J., and J. J. Magnuson. 1992. Spatial heterogeneity of littoral fish assemblages in lakes: relation to species diversity and habitat structure. Canadian Journal of Fisheries and Aquatic Sciences 49:1493–1500. Benson, R. L., S. Turo, and B. W. McCovey, Jr. 2007. Migration and movement patterns of Green Sturgeon (Acipenser medirostris) in the Klamath and Trinity rivers, California, USA. Environmental Biology of Fishes 79:269–279. Berejikian, B. A., R. J. F. Smith, E. P. Tezak, S. L. Schroder, and C. M. Knudsen. 1999. Chemical alarm signals and complex hatchery rearing habitats affect antipredator behavior and survival of Chinook Salmon (Oncorhynchus tshawytscha) juveniles. Canadian Journal of Fisheries and Aquatic Sciences 56:830–838. Berenbrink, M., P. Koldkjaer, O. Kepp, and A. R. Cossins. 2005. Evolution of oxygen secretion in fishes and the emergence of a complex physiological system. Science 307:1752–1757.

Berendzen, P. B., A. M. Simons, and R. M. Wood. 2003. Phylogeography of the Northern Hog Sucker, Hypentelium nigricans (Teleostei: Cypriniformes): genetic evidence for the existence of the ancient Teays River. Journal of Biogeography 30: 1139–1152. Berendzen, P. B., A. M. Simons, R. M. Wood, T. E. Dowling, and C. L. Secor. 2008a. Recovering cryptic diversity and ancient drainage patterns in eastern North America: historical biogeography of the Notropis rubellus species group (Teleostei: Cypriniformes). Molecular Phylogenetics and Evolution 46:721–737. Berendzen, P. B., T. Gamble, and A. M. Simons. 2008b. Phylogeography of the Bigeye Chub Hybopsis amblops (Teleostei: Cypriniformes): Early Pleistocene diversification and postglacial range expansion. Journal of Fish Biology 73:2021–2039. Berg, J. J., M. S. Allen, and K. J. Sulak. 2007. Population assessment of the Gulf of Mexico Sturgeon in the Yellow River, Florida, p. 365–379. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Berg, J. K. 1999. Final report of the River Otter research project on the Upper Colorado River Basin in and adjacent to Rocky Mountain National Park, Colorado. United States National Park Ser vice. Rocky Mountain National Park, Estes Park, Colorado. Berg, L. S. 1940. Classification of fishes both recent and fossil. Trudy Zoologicheskogo Instituta 5:87–345 (English translation, 1947, J. W. Edwards, Ann Arbor, Michigan, p. 346–517). Bergstedt, R. A., R. B. McDonald, M. B. Twohey, K. M. Mullett, R. J. Young, and J. W. Heinrich. 2003. Reduction in Sea Lamprey hatching success due to release of sterilized males. Journal of Great Lakes Research 29 (Suppl. 1):435–444. Bergstedt, R. A., and J. G. Seelye. 1995. Evidence for lack of homing by Sea Lampreys. Transactions of the American Fisheries Society 124:235–239. Bergstedt, R. A., and W. D. Swink. 1995. Seasonal growth and duration of the parasitic life stage of the landlocked Sea Lamprey (Petromyzon marinus). Canadian Journal of Fisheries and Aquatic Sciences 52:1257–1264. Bermingham, E., and J. C. Avise. 1986. Molecular zoogeography of freshwater fishes in the southeastern United States. Genetics 113:939–965. Bermingham, E., T. Lamb, and J. C. Avise. 1986. Size polymorphism and heteroplasmy in the mitochondrial DNA of lower vertebrates. The Journal of Heredity 77:249–252. Bernatchez, L., and C. Landry. 2003. MCH studies in nonmodel vertebrates: what have we learned about natural selection in 15 years? Journal of Evolutionary Biology 16:363–377. Bernatchez, L., J. A. Vuorinen, R. A. Bodaly, and J. J. Dodson. 1996. Genetic evidence of reproductive isolation and multiple origins of sympatric trophic ecotypes of whitefish (Coregonus). Evolution 50:624–635. Bernau, N. A., R. L. Puzdrowski, and R. B. Leonard. 1991. Identification of the midbrain locomotor region and its relation to descending locomotor pathways in the Atlantic Stingray, Dasyatis sabina. Brain Research 557:83–94. Berra, T. M. 2001. Freshwater Fish Distribution. Academic Press, San Diego, California. Berra, T. 2007. Freshwater Fish Distribution. University of Chicago Press, Chicago, Illinois.

LITERATURE CITED

Berra, T., and P. Petry. 2006. Fish assemblage of Cedar Fork Creek, Ohio, unchanged for 28 years. Ohio Journal of Science 106:98–102. Bertin, L. 1956. Eels, a Biological Study. Cleaver-Hume, London. Bertolo, A., and M. Pierre. 2005. The relationship between piscivory and growth of White Sucker (Catostomus commersoni) and Yellow Perch (Perca flavescens) in headwater lakes of the Canadian Shield. Canadian Journal of Fisheries and Aquatic Sciences 62:2706–2715. Bessert, M. L., C. Sitzman, and G. Ortı. 2007. Avoiding paralogy: diploid loci for allotetraploid Blue Sucker fish (Cycleptus elongatus, Catostomidae). Conservation Genetics 8:995–998. Best, A. C. G., and J. A. C. Nicol. 1979. On the eye of the Goldeye Hiodon alosoides (Teleostei: Hiodontidae). Journal of Zoology (London) 188:309–332. Bestgen, K. R. 1996. Growth, survival, and starvation resistance of Colorado squawfish larvae. Environmental Biology of Fishes 46:197–209. Bestgen, K. R., D. W. Beyers, J. A. Rice, and G. B. Haines. 2006. Factors affecting recruitment of young Colorado Pikeminnow: synthesis of predation experiments, field studies, and individualbased modeling. Transactions of the American Fisheries Society 135:1722–1742. Bestgen, K. R., and D. L. Propst. 1996. Redescription, geographic variation, and taxonomic status of Rio Grande Silvery Minnow Hybognathus amarus (Girard, 1856). Copeia 1996:41–55. Bestgen, K. R., D. L. Propst, and C. W. Painter. 1987. Transport ecology of larval fishes in the Gila River, New Mexico. Proceedings of the Desert Fishes Council 16–18:174. Betancur-R., R., R. E. Broughton, E. O. Wiley, K. Carpenter, J. A. López, C. Li, N. I. Holcroft, D. Arcila, M. Sanciangco, J. C. Cureton, F. Zhang, T. Buser, M. A. Campbell, J. A. Ballesteros, A. Roa-Varon, S. Willis, W. C. Borden, T. Rowley, P. C. Reneau, D. J. Hough, G. Lu, T. Grande, G. Arratia, and G. Ortí. 2013. The tree of life and a new classification of bony fishes. PLOS Currents Tree of Life. 2013 April 18. Edition 1. doi:10.1371/currents. tol.53ba26640df0ccaee75bb165c8c26288. Bettoli, P. W., M. Casto-Yerty, G. D. Scholten, and E. J. Heist. 2009a. Bycatch of the endangered Pallid Sturgeon (Scaphirhynchus albus) in a commercial fishery for Shovelnose Sturgeon (Scaphirhynchus platorynchus). Journal of Applied Ichthyology 25:1–4. Bettoli, P. W., J. A. Kerns, and G. D. Scholten. 2009b. Status of Paddlefish in the United States, p. 23–38. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Bettoli, P. W., and G. D. Scholten. 2006. Bycatch rates and initial mortality of Paddlefish in a commercial gillnet fishery. Fisheries Research 77:343–347. Bettoli, P. W., G. D. Scholten, and W. C. Reeves. 2007. Protecting Paddlefish from overfishing: a case history of the research and regulatory process. Fisheries 32:390–397. Beugly, J., and M. Pyron. 2010. Temporal and spatial variation in the long-term functional organization of fish assemblages in a large river. Hydrobiologia 654:215–226. Bevelander, G. 1934. The gills of Amia calva specialized for respiration in an oxygen deficient habitat. Copeia 1934:123–127. Beyer, J. M., G. Lucchetti, and G. Gray. 1988. Digestive tract evacuation in northern squawfish (Ptychocheilus oregonensis). Canadian Journal of Fisheries and Aquatic Sciences 45:548–553.

511

Beyers, D. W., and M. S. Farmer. 2001. Effects of copper on olfaction of Colorado Pikeminnow. Environmental Toxicology and Chemistry 20:907–912. Bielawski, J. P., A. Brault, and J. R. Gold. 2002. Phylogenetic relationships within the genus Pimephales as inferred from ND4 and ND4L nucleotide sequences. Journal of Fish Biology 61:293–297. Bielawski, J. P., and J. R. Gold. 1996. Unequal synonymous substitution rates within and between two protein-coding mitochondrial genes. Molecular Biology and Evolution 13:889–892. Bielawski, J. P., and J. R. Gold. 2001. Phylogenetic relationships of cyprinid fishes in subgenus Notropis inferred from nucleotide sequences of the mitochondrially encoded cytochrome b gene. Copeia 2001:656–667. Bielawski, J. P., and J. R. Gold. 2002. Mutation patterns of mitochondrial H- and L-strand DNA in closely related cyprinid fishes. Genetics 161:1589–1597. Bietz, B. F. 1981. Habitat availability, social attraction and nest distribution patterns in Longear Sunfish (Lepomis megalotis peltastes). Environmental Biology of Fishes 6:193–200. Bigelow, H. B., and W. C. Shroeder. 1953a. Fishes of the Western North Atlantic; Sawfishes, Guitarfishes, Skates, Rays, and chimaeroids. Sears Foundation Marine Research 1(2):1–588. Bigelow, H. B., and W. C. Schroeder. 1953b. Fishes of the Gulf of Maine. Fishery Bulletin of the Fish and Wildlife Service 53:1–577. Billard, R. 1978. Changes in structure and fertilizing ability of marine and freshwater fish spermatozoa diluted in various salinities. Aquaculture 14:187–198. Billard, R., and G. Lecointre. 2001. Biology and conservation of Sturgeon and Paddlefish. Reviews in Fish Biology and Fisheries 10:355–392. Billman, E. J., and M. Pyron. 2005. Evolution of form and function: morphology and swimming per formance in North American minnows. Journal of Freshwater Ecology 20:221–232. Billman, E. J., E. J. Wagner, and R. E. Arndt. 2010. Effects of temperature on the survival and growth of age-0 Least Chub (Iotichthys phlegethontis). Western North American Naturalist 66:434–440. Binder, T. R., and D. G. McDonald. 2007. Is there a role for vision in the behaviour of Sea Lampreys (Petromyzon marinus) during their upstream spawning migration? Canadian Journal of Fisheries and Aquatic Sciences 64:1403–1412. Binder, T. R., and D. G. McDonald. 2008a. The role of dermal photoreceptors during the Sea Lamprey (Petromyzon marinus) spawning migration. Journal of Comparative Physiology A 194:921–928. Binder, T. R., and D. G. McDonald. 2008b. The role of temperature in controlling diel activity in upstream migrant Sea Lampreys (Petromyzon marinus). Canadian Journal of Fisheries and Aquatic Sciences 65:1113–1121. Binns, N. A. 1967. Effects of rotenone on the fauna of the Green River, Wyoming. Wyoming Game and Fish Commision, Fisheries Technical Bulletin:1–114. Bird, D. J., P. L. Lutz, and I. C. Potter. 1976. Oxygen dissociation curves of the blood of larval and adult Lampreys (Lampetra fluviatilis). The Journal of Experimental Biology 65:449–458. Bird, D. J., and I. C. Potter. 1979a. Metamorphosis in the paired species of Lampreys, Lampetra fluviatilis (L.) and Lampetra planeri (Bloch). 1. A description of the timing and stages. Zoological Journal of the Linnean Society 65:127–143. Bird, D. J., and I. C. Potter. 1979b. Metamorphosis in the paired species of Lampreys, Lampetra fluviatilis (L.) and Lampetra planeri (Bloch). 2. Quantitative data for body proportions, weights,

512

LITERATURE CITED

lengths and sex ratios. Zoological Journal of the Linnean Society 65:145–160. Bird, N. C., and L. P. Hernandez. 2007. Morphological variation in the Weberian apparatus of Cypriniformes. Journal of Morphology 268:739–757. Biro, P. A., J. R. Post, and E. A. Parkinson. 2003. Population consequences of a predator-induced habitat shift by trout in whole-lake experiments. Ecology 84:691–700. Birstein, V. J. 1993. Sturgeons and Paddlefishes: threatened fishes in need of conservation. Conservation Biology 7:773–787. Birstein, V. J., and W. E. Bemis. 1997. How many species are there within the genus Acipenser? Environmental Biology of Fishes 148:157–163. Birstein, V. J., W. E. Bemis, and J. R. Waldman. 1997a. The threatened status of acipenseriform species: a summary. Environmental Biology of Fishes 48:427–435. Birstein, V. J., and R. DeSalle. 1998. Molecular phylogeny of Acipenserinae. Molecular Phylogenetics and Evolution 9:141–155. Birstein, V. J., P. Doukakis, and R. DeSalle. 1999. Molecular phylogeny of Acipenserinae and black caviar species identification. Journal of Applied Ichthyology 15:12–16. Birstein, V. J., P. Doukakis, and R. DeSalle. 2002. Molecular phylogeny of Acipenseridae: nonmonophyly of Scaphirhynchinae. Copeia 2002:287–301. Birstein, V. J., R. Hanner, and R. DeSalle. 1997b. Phylogeny of Acipenseriformes: cytogenic and molecular approaches. Environmental Biology of Fishes 48:127–155. Birstein, V. J., J. R. Waldman, and W. E. Bemis (eds.). 1997c. Sturgeon Biodiversity and Conservation. Kluwer Academic Publishers, Dordrecht, The Netherlands. Bisazza, A., L. Facchin, R. Pignatti, and G. Vallortigara. 1998. Lateralization of detour behaviour in poeciliid fish: the effect of species, gender and sexual motivation. Behavioural Brain Research 91:157–164. Bisazza, A., A. Marconato, and G. Marin. 1989. Male mate preferences in the mosquitofish Gambusia holbrooki. Ethology 83:335–343. Bisazza, A., and G. Marin. 1988. Sexual selection and sexual dimorphism in the poeciliid fish Gambusia affinis Holbrooksi Grd. Monitore Zoologico Italiano 22:530–531. Bisazza, A., and G. Marin. 1991. Male size and female mate choice in the Eastern Mosquitofish (Gambusia holbrooki: Poeciliidae). Copeia 1991:730–735. Bisazza, A., and G. Marin. 1995. Sexual selection and sexual size dimorphism in the Eastern Mosquitofish, Gambusia holbrooki (Pisces: Poeciliidae). Ethology, Ecology and Evolution 7:169–183. Bisazza, A., and A. Pilastro. 1997. Small male mating advantage and reversed sexual dimorphism in poeciliid fishes. Journal of Fish Biology 50:397–406. Bisazza, A., G. Vaccari, and A. Pilastro. 2001. Female mate choice in a mating system dominated by male sexual coercion. Behavioral Ecology 12:59–64. Bishop, F. G. 1974. Observations on the fish fauna of the Peace River in Alberta. Alberta Fish and Wildlife Report. 30 p. Bishop, R. E., J. J. Torres, and R. E. Crabtree. 2000. Chemical composition and growth indices in leptocephalus larvae. Marine Biology 137:205–214. Bishop, R. E., and J. J. Torres. 2001. Leptocephalus energetics: assembly of the energetics equation. Marine Biology 138:1093–1098.

Bjerselius, R., W. Li, and P. W. Sorensen. 1996. Spermiated male Sea Lamprey release a potent sex pheromone, p. 271. In Proceedings of the Fifth International Symposium on the Reproductive Physiology of Fish. P. Thomas and F. Goetz (eds.). Austin, Texas. Bjerselius, R., K. H. Olsén, and W. Zheng. 1995. Behavioural and endocrinological responses of mature male Goldfish to the sex pheromone 17α,20β-Dihydroxy-4-pregnen-3-one in the water. Journal of Experimental Biology 198:747–754. Blacklidge, K. H., and C. A. Bidwell. 1993. Three ploidy levels indicated by genome quantification in Acipenseriformes of North America. Journal of Heredity 84:427–430. Blackwell, B. G., B. R. Murphy, and V. M. Pitman. 1995. Suitability of food resources and physiochemical parameters in the lower Trinity River, Texas for Paddlefish. Journal of Freshwater Ecology 10:163–175. Blair, J. E., and S. B. Hedges. 2005. Molecular phylogeny and divergence times of deuterostome animals. Molecular Biology and Evolution 22:2275–2284. Blanchfield, P. J., and M. S. Ridgway. 1999. The cost of peripheral males in a Brook Trout mating system. Animal Behaviour 57:537–544. Blinn, D. W., C. Runck, D. A. Clark, and J. N. Rinne. 1993. Effects of Rainbow Trout predation on Little Colorado Spinedace. Transactions of the American Fisheries Society 122:139–143. Blinn, D. W., J. White, T. Pradetto, and J. O’Brien. 1998. Reproductive ecology and growth of a captive population of Little Colorado Spinedace (Lepidomeda vittata: Cyprinidae). Copeia 1998:1010–1015. Blonder, B. I., and W. S. Alevizon. 1988. Prey discrimination and electroreception in the Stingray Dasyatis sabina. Copeia 1988:33–36. Blouw, D. M., and D. W. Hagen. 1981. Ecology of the Fourspine Stickleback, Apeltes quadracus, with respect to a polymorphism for dorsal spine number. Canadian Journal of Zoology 59:1677—1692. Blum, M. J., D. A. Neely, P. M. Harris, and R. L. Mayden. 2008. Molecular systematics of the cyprinid genus Campostoma (Actinopterygii: Cypriniformes): disassociation between morphological and mitochondrial differentiation. Copeia 2008:360–369. Bogue, M. B. 2000. Fishing the Great Lakes: An Environmental History, 1783–1933. University of Wisconsin Press, Madison, Wisconsin. Böhlke, J. E., and C. C. G. Chaplin. 1968. Fishes of the Bahamas and adjacent tropical waters. Livingston Publishing Company, Wyneewood, Pennsylvania. Boiko, N. E., R. A. Grigoryan, and A. S. Chichachev. 1993. Olfactory imprinting in juveniles of Russian Sturgeon, Acipenser gueldenstaedtii. Journal of Evolutionary Biochemistry and Physiology 29:509–514 (in Russian). Bolger, T., and P. L. Connolly. 1989. The selection of suitable indices for the measurement and analysis of fish condition. Journal of Fish Biology 34:171–182. Bolker, J. A. 2004. Embryology, p. 134–146. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. H. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. Bolnick, D. I. 2004. Can intraspecific competition drive disruptive selection? An experimental test in natural populations of Sticklebacks. Evolution 58:608–618. Bolnick, D. I., and O. L. Lau. 2008. Predictable patterns of disruptive selection in Stickleback in postglacial lakes. The American Naturalist 172:1–11.

LITERATURE CITED

Bond, D. D., and J. J. Long. 1984. Status of the Oregon Chub, Hybopsis crameri. Proceedings of the annual conference of the Western Association of Fish and Wildlife Agencies 64:483–486. Bonham, K. 1940. Food of Gars in Texas. Transactions of the American Fisheries Society 70:356–362. Bonneau, J. L., and D. L. Scarnecchia. 1998. Seasonal and diel changes in habitat use by juvenile Bull Trout (Salvelinus confluentus) and Cutthroat Trout (Oncorhynchus clarki) in a mountain stream. Canadian Journal of Zoology 76:783–790. Bonneau, J. L., and D. L. Scarnecchia. 2002. Spawning-season homing of Common Carp and River Carpsucker. Prairie Naturalist 34:13–20. Bonner, T. H., and G. R. Wilde. 2000. Changes in the Canadian River fish assemblage associated with reservoir construction. Journal of Freshwater Ecology 15:189–198. Bonnot, T. W., M. L. Wildhaber, J. J. Millspaugh, A. J. DeLonay, R. B. Jacobson, and J. L. Bryan. 2011. Discrete choice modeling of Shovelnose Sturgeon habitat selection in the lower Missouri River. Journal of Applied Ichthyology 27:291–300. Bonvillain, C. P., A. M. Ferrara, and Q. C. Fontenot. 2008. Relative abundance and biomass estimate of a Spotted Gar population in a seasonally connected large river floodplain lake. Proceedings of the Annual Conference of the Southeastern Association of Fish and Wildlife Agencies 62:177–182. Boone, E. A., Jr., and T. J. Timmons. 1995. Density and natural mortality of Paddlefish, Polyodon spathula, in an unfished Cumberland River subimpoundment, South Cross Creek Reservoir, Tennessee. Journal of Freshwater Ecology 10:421–431. Boreman, J. 1997. Sensitivity of North American Sturgeons and Paddlefish to fishing mortality. Environmental Biology of Fishes 48:399–405. Borgstrom, G. 1978. The contribution of freshwater fish to human food, p. 469–491. In Ecology of Freshwater Fish Production. S. D. Gerking (ed.). John Wiley & Sons, New York. Borowsky, R. 1973. Social control of adult size in male Xiphophorus variegatus. Nature 254:332–335. Borowsky, R. 1978. Social inhibition of maturation in natural populations of male Xiphophorus variegatus (Pisces: Poeciliidae). Science 201:933–935. Borowsky, R., and K. D. Kallman. 1976. Patterns of mating in natural populations of Xiphophorus (Pisces: Poeciliidae). I. X. maculates from Belize and Mexico. Evolution 30:693–706. Borowsky, R., and J. Khouri. 1976. Patterns of mating in natural populations of Xiphophorus. II. X. variatus from Tamaulipas, Mexico. Copeia 1976:727–734. Boschung, H., and R. L. Mayden. 2004. Fishes of Alabama. Smithsonian Institution Press, Washington, D.C. Bottrell, C. E., R. H. Ingersol, and R. W. Jones. 1964. Notes on the embryology, early development, and behavior of Hybopsis aestivalis tetranemus (Gilbert). Transactions of the American Microscopical Society 83:391–399. Boucher, D. H., S. James, and K. H. Keeler. 1982. The ecology of mutualism. Annual Review of Ecology and Systematics 13:315–347. Boudreaux, P., A. Ferrara, and Q. Fontenot. 2007a. Chloride inhibition of nitrite uptake for non-teleost Actinoperygiian fishes. Comparative Biochemistry and Physiology Part A 147:420–425. Boudreaux, P., A. Ferrara, and Q. Fontenot. 2007b. Acute toxicity of ammonia to Spotted Gar, Lepisosteus oculatus, Alligator Gar, Atractosteus spatula, and Paddlefish, Polyodon spathula. World Aquaculture Society 38:322–325.

513

Boughman, J. W. 2001. Divergent sexual selection enhances reproductive isolation in Sticklebacks. Nature 411:944–947. Boutilier, R. G., R. A. Ferguson, R. P. Henry, and B. L. Tufts. 1993. Exhaustive exercise in the Sea Lamprey (Petromyzon marinus): relationship between anaerobic metabolism and intracellular acidbase balance. The Journal of Experimental Biology 178:71–88. Bowfin Anglers’ Group. 2011. Bowfin Anglers’ Group. Available from http://www.bowfinanglers.com; as of 24 October 2011. Bowman, M. L. 1970. Life history of the Black Redhorse, Moxostoma duquesnei (Le Seur), in Missouri. Transactions of the American Fisheries Society 99:546–559. Bowman, T. E., S. A. Grabe, and J. H. Hecht. 1977. Range extension and new hosts for the cymothoid isopod Anilocra acuta. Chesapeake Science 18:390–393. Boydstun, C., P. Fuller, and J. D. Williams. 1995. Nonindigenous fish, p. 431–433. In Our living resources: a report to the nation on the distribution, abundance, and health of US plants, animals and ecosystems. E. T. LaRoe, G. S. Farris, C. E. Puckett, P. D. Doran, and M. H. Mac (eds.). Department of the Interior, National Biological Ser vice, Washington, D.C. Boysen, K. A., and J. J. Hoover. 2009. Swimming performance of juvenile White Sturgeon (Acipenser transmontanus): training and probability of entrainment due to dredging. Journal of Applied Ichthyology 25(Supplement 2):54–59. Bozek, M. A., T. M. Burri, and R. V. Frie. 1999. Diets of Muskellunge in northern Wisconsin lakes. North American Journal of Fisheries Management 19:258–270. Bozek, M. A., L. J. Paulson, and G. R. Wilde. 1990. Effects of ambient Lake Mohave temperatures on development, oxygen consumption, and hatching success of the Razorback Sucker. Environmental Biology of Fishes 27:255–263. Bozeman, E. L., G. S. Helfman, and T. Richardson. 1985. Population size and home range of American Eels in a Georgia tidal creek. Transactions of the American Fisheries Society 114:821–825. Braaten, P. J., M. R. Doeringsfeld, and C. S. Guy. 1999. Comparison of age and growth estimates for River Carpsuckers using scales and dorsal fin ray sections. North American Journal of Fisheries Management 19:786–792. Braaten, P. J., and D. B. Fuller. 2007. Growth rates of young-ofyear Shovelnose Sturgeon in the upper Missouri River. Journal of Applied Ichthyology 23:506–515. Braaten, P. J., D. B. Fuller, L. D. Holte, R. D. Lott, W. Viste, T. F. Brandt, and R. G. Legare. 2008. Drift dynamics of larval Pallid Sturgeon and Shovelnose Sturgeon in a natural side channel of the upper Missouri River, Montana. North American Journal of Fisheries Management 28:808–826. Braaten, P. J., D. B. Fuller, and R. D. Lott. 2009a. Spawning migrations and reproductive dynamics of Paddlefish in the upper Missouri River basin, Montana and North Dakota, p. 103–122. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Braaten, P. J., D. B. Fuller, R. D. Lott, and G. R. Jordan. 2009b. An estimate of the historic population size of adult Pallid Sturgeon in the upper Missouri River basin, Montana and North Dakota. Journal of Applied Ichthyology 25 (Supplement 2):2–7. Braaten, P. J., D. B. Fuller, R. D. Lott, M. P. Ruggles, and R. J. Holm. 2010. Spatial distribution of drifting Pallid Sturgeon larvae in the Missouri River inferred from two net designs and

514

LITERATURE CITED

multiple sampling locations. North American Journal of Fisheries Management 30:1062–1074. Braaten, P. J., D. B. Fuller, and N. D. McClenning. 2007. Diet composition of larval and young-of-year Shovelnose Sturgeon in the upper Missouri River. Journal of Applied Ichthyology 23:516–520. Braaten, P. J., and C. S. Guy. 1999. Relations between physicochemical factors and abundance of fishes in tributary confluences of the lower channelized Missouri River. Transactions of the American Fisheries Society 128:1213–1221. Bradbury, J. W. 1981. The evolution of leks, p. 138–169. In Natural Selection and Social Behavior: Recent Research and New Theory. R. D. Alexander and D. W. Tinkle (eds.). Chiron Press, New York. Bradley, J. L. 1996. Prey energy content and selection, habitat use and daily ration of the Atlantic Stingray, Dasyatis sabina. Unpubl. Master’s thesis, Florida Institute of Technology, Melbourne. Brady, T., M. Hove, C. Nelson, R. Gordon, D. Hornbach, and A. Kapuscinski. 2004. Suitable host fish species determined for Hickorynut and Pink Heelsplitter. Ellipsaria 6:14–15. Braekevelt, C. R. 1982a. Photoreceptor fine structure in the Goldeye (Hiodon alosoides) (Teleost). Anatomy and Embryology 165:177–192. Braekevelt, C. R. 1982b. Fine structure of the retinal epithelium and retinal tapetum lucidum of the Goldeye Hiodon alosoides. Anatomy and Embryology 164:287–302. Brainerd, E. L. 1994. The evolution of lung-gill bimodal breathing and the homology of vertebrate respiratory pumps. American Zoologist 34:289–299. Brainerd, E. L. 1997. Efficient fish not faint-hearted. Nature 389:229–230. Bramblett, R. G., and R. G. White. 2001. Habitat use and movement of Pallid and Shovelnose Sturgeon in the Yellowstone and Missouri rivers in Montana and North Dakota. Transactions of the American Fisheries Society 130:1006–1025. Brander, K. M. 2007. Global fish production and climate change. Proceedings of the National Academy of Sciences 104:19709–19714. Branson, B. A. 1962. Observation on the distribution of nuptial tubercles in some catostomid fish. Transactions of the Kansas Academy of Science 64:360–372. Breder, C. M. 1926. The locomotion of fishes. Zoologica 4:159–297. Breder, C. M. 1936. The reproductive habits of the North American Sunfishes (family Centrarchidae). Zoologica 21:1–47. Breder, C. M., and L. A. Krumholz. 1941. On the uterine young of Dasyatis sabina (LeSueur) and Dasyatis hastatus (DeKay). Zoologica 26:49–53. Breder, C. M., and D. E. Rosen. 1966. Modes of Reproduction in Fishes. T. F. H. Publications, Neptune City, New Jersey. Breitburg, D. L. 1994. Behavioral response of fish larvae to low dissolved oxygen concentrations in a stratified water column. Marine Biology 120:615–625. Brennan, J. S., and G. M. Cailliet. 1989. Comparative agedetermination techniques for White Sturgeon in California. Transactions of the American Fisheries Society 118:296–310. Brett, B. L. H., and D. J. Grosse. 1982. A reproductive pheromone in the Mexican poeciliid fish, Poecilia chica. Copeia 1982:219–223.

Brett, J. R. 1971. Energetic responses of salmon to temperature: a study of some thermal relations in the physiology and freshwater ecology of Sockeye Salmon (Oncorhynchus nerka). American Zoologist 11:91–113. Briand, C., D. Fatina, and A. Legault. 2002. Role of Eel odour on the efficiency of an Eel, Anguilla anguilla, ladder and trap. Environmental Biology of Fishes 65:473–477. Briggs, J. C. 1986. Introduction to the zoogeography of North American fishes, p. 1–16. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. Briggs, J. C. 1987. Biogeography and plate tectonics. Developments in Palaeontology and stratigraphy, 10. Elsevier Science Publishing Company, New York. Brinkman, E. L. 2008. Contributions to the life history of Alligator Gar, Atractosteus spatula (Lacépède), in Oklahoma. Unpubl. Master’s thesis, Oklahoma State University, Stillwater. Briolay, J., N. Galtier, R. M. Brito, and Y. Bouvet. 1998. Molecular phylogeny of Cyprinidae inferred from cytochrome b DNA sequences. Molecular Phylogenetics and Evolution 9:100–108. Britz, R. 2004. Egg structure and larval development of Pantodon buchholzi (Teleostei: Osteoglossomorpha), with a review of data on reproduction and early life history in other osteoglossomorphs. Ichthyological Explorations of Freshwaters 15:209–224. Brodeur, P., P. Magnan, and M. Legault. 2001. Response of fish communities to different levels of White Sucker (Catostomus commersoni) biomanipulation in five temperate lakes. Canadian Journal of Fisheries and Aquatic Sciences 58:1998–2010. Bronte, C. R., and D. W. Johnson. 1985. Growth of Paddlefish in two mainstem reservoirs with reference to commercial harvest. Transactions of the Kentucky Academy of Science 46:28–32. Bronzi, P., H. Rosenthal, G. Arlati, and P. Williot. 1999. A brief overview on the status and prospects of Sturgeon farming in Western and Central Europe. Journal of Applied Ichthyology 15:224–227. Brooks, D. R. 1985. Historical ecology: a new approach to studying the evolution of ecological associations. Annals of the Missouri Botanical Garden 72:660–680. Brooks, D. R., and D. A. McLennan. 1991. Phylogeny, ecology, and behavior, a research program in comparative biology. University of Chicago Press, Chicago, Illinois. Brooks, D. R., and D. A. McLennan. 2002. The Nature of the Organism: An Evolutionary Voyage Through Space and Time. University of Chicago Press, Chicago, Illinois. Brooks, D. R., and E. O. Wiley. 1988. 2nd edition. Evolution as Entropy: Toward a Unified Theory of Biology. University of Chicago Press, Chicago, Illinois. Brouder, M. J. 1999. Relationship between length of Roundtail Chub and infection intensity of Asian fish tapeworm Bothriocephalus acheilognathi. Journal of Aquatic Animal Health 11:302–304. Brouder, M. J., and T. L. Hoff nagle. 1997. Distribution and prevalence of the Asian fish tapeworm, Bothriocephalus acheilognathi, in the Colorado River and tributaries, Grand Canyon, Arizona, including two new records. Journal of the Helminthological Society of Washington 64:219–226. Brouder, M. J., D. D. Rogers, and L. A. Avenertti. 2000. Life History and ecology of the Roundtail Chub Gila robusta, from two

LITERATURE CITED

streams in the Verde River basin. Technical Guidance Bulletin 3:1–16. Broughton, R. E., and T. E. Dowling. 1994. Length variation in mitochondrial DNA of the minnow Cyprinella spiloptera. Genetics 138:179–190. Broughton, R. E., and J. R. Gold. 2000. Phylogenetic relationships in the North American cyprinid genus Cyprinella (Actinopterygii: Cyprinidae) based on sequences of the mitochondrial ND2 and ND4L genes. Copeia 2000:1–10. Broughton, R. E., G. J. P. Naylor, and T. E. Dowling. 1998. Conflicting phylogenetic patterns caused by molecular mechanism in mitochondrial DNA sequences. Systematic Biology 47:696–701. Broussard, N. M. 2009. Stage specific potency and phylogenetic sensitivity of Gar toxin. Unpubl. Master’s thesis, Nicholls State University, Thibodaux, Louisiana. Brousseau, R. A. 1976. The pectoral anatomy of selected Ostariophysi. II. The Cypriniformes and Siluriformes. Journal of Morphology 150:79–116. Brown, B. A. 1984. Comparative life histories of some species of redhorse, subgenus Moxostoma, genus Moxostoma. Unpubl. Master’s thesis, Indiana State University, Terre Haute. Brown, G. E., J. C. Adrian, Jr., M. G. Lewis, and J. M. Tower. 2002a. The effects of reduced pH on chemical alarm signaling in ostariophysan fishes. Canadian Journal of Fisheries and Aquatic Sciences 59:1331–1338. Brown, G. E., J. C. Adrian, Jr., N. T. Naderi, M. C. Harvey, and J. M. Kelly. 2003. Nitrogen oxides elicit antipredator responses in juvenile Channel Catfish, but not in Convict Cichlids or Rainbow Trout: conservation of the ostariophysan alarm pheromone. Journal of Chemical Ecology 29:1781–1796. Brown, G. E., J. C. Adrian, Jr., T. Patton and D. P. Chivers. 2001c. Fathead Minnows learn to recognize predator odour when exposed to concentrations of artificial alarm pheromone below their behavioural response threshold. Canadian Journal of Zoology 79:2239–2245. Brown, G. E., J. C. Adrian, Jr., and M. I. Shih. 2001b. Behavioural response of Fathead Minnows to hypoxanthine-3-N-oxide at varying concentrations. Journal of Fish Biology 58:1465–1470. Brown, G. E., J. C. Adrian, Jr., E. Smyth, H. Leet, and S. Brennan. 2000. Ostariophysan alarm pheromones: laboratory and field tests of the functional significance of nitrogen oxides. Journal of Chemical Ecology 26:139–154. Brown, G. E., and S. Brennan. 2000. Chemical alarm signals in juvenile Green Sunfish (Lepomis cyanellus, Centrarchidae). Copeia 2000:1079–1082. Brown, G. E., and J. A. Brown. 1992. Do Rainbow Trout and Atlantic Salmon discriminate kin? Canadian Journal of Zoology 70:1636–1640. Brown, G. E., and J. A. Brown. 1993a. Do kin always make better neighbours? The effects of territory quality. Behavioral Ecology and Sociobiology 33:225–231. Brown, G. E., and J. A. Brown. 1993b. Social dynamics in salmonid fishes; do kin make better neighbours? Animal Behaviour 45:863–871. Brown, G. E., and J. A. Brown. 1996. Does kin-based territorial behaviour increase kin-based foraging in juvenile salmonids? Behavioral Ecology 1:24–29. Brown, G. E., J. A. Brown, and A. M. Crosbie. 1993. Phenotype matching in juvenile Rainbow Trout. Animal Behaviour 46:1223–1225.

515

Brown, G. E., J. A. Brown, and W. R. Wilson. 1996. The effects of kinship on the growth of juvenile Arctic Char. Journal of Fish Biology 48:313–320. Brown, G. E., D. P. Chivers, and R. J. F. Smith. 1995a. Fathead Minnows avoid conspecific and heterospecific alarm pheromones in the feces of Northern Pike. Journal of Fish Biology 47:387–393. Brown, G. E., D. P. Chivers, and R. J. F. Smith. 1995b. Localized defecation by pike: a response to labelling by cyprinid alarm pheromone? Behavioral Ecology and Sociobiology 36:105–110. Brown, G. E., D. P. Chivers, and R. J. F. Smith. 1997. Differential learning rates of chemical versus visual cues of a northern pike by Fathead Minnows in a natural habitat. Environmental Biology of Fishes 40:89–96. Brown, G. E., D. L. Gershaneck, D. L. Plata, and J. L. Golub. 2002b. Ontogenetic changes in response to heterospecific alarm cues by juvenile Largemouth Bass are phenotypically plastic. Behaviour 139:913–927. Brown, G. E., and J.-G. Godin. 1997. Anti-predator responses to conspecific and heterospecific skin extracts by Threespine Sticklebacks: alarm pheromones revisited. Behaviour 134:1123–1134. Brown, G. E., J. L. Golub, and D. L. Plata. 2001a. Attack cone avoidance during predator inspection visits by wild Finescale Dace (Phoxinus neogaeus): the effects of predator diet. Journal of Chemical Ecology 27:1657–1666. Brown, G. E., M. C. Harvey, A. O. H. C. Leduc, M. C. O. Ferrari, and D. P. Chivers. 2009. Social context, competitive interactions and the dynamic nature of antipredator responses of juvenile Rainbow Trout Oncorhynchus mykiss. Journal of Fish Biology 75:552–562. Brown, G. E., V. J. LeBlanc, and L. E. Porter. 2001b. Ontogenetic changes in the response of Largemouth Bass (Micropterus salmoides, Centrarchidae, Perciformes) to heterospecific alarm pheromones. Ethology 107:401–414. Brown, G. E., and R. J. F. Smith. 1994. Fathead Minnows use chemical cues to discriminate natural shoalmates from unfamiliar conspecifics. Journal of Chemical Ecology 20:3051–3061. Brown, G. E., and R. J. F. Smith. 1996. Foraging trade-offs in Fathead Minnows (Pimephales promelas, Osteichthyes, Cyprinidae): acquired predator recognition in the absence of an alarm pheromone. Ethology 102:776–785. Brown, G. E., and R. J. F. Smith. 1997. Conspecific skin extract elicits anti-predator behaviour in juvenile Rainbow Trout (Oncorhynchus mykiss). Canadian Journal of Zoology 75: 1916–1922. Brown, G. E., and R. J. F. Smith. 1998. Acquired predator recognition in juvenile Rainbow Trout (Oncorhynchus mykiss): conditioning hatchery reared fish to recognize chemical cues of a predator. Canadian Journal of Fisheries and Aquatic Sciences 55:611–617. Brown, J. H. 1995. Macroecology. University of Chicago Press, Chicago, Illinois. Brown, J. H., B. J. Fox, and D. A. Kelt. 2000. Assembly rules: desert rodent communities are structured at scales from local to continental. The American Naturalist 156:314–321. Brown, J. R., K. Beckenbach, A. T. Beckenbach, and M. J. Smith. 1996. Length variation, heteroplasmy and sequence divergence in the mitochondrial DNA of four species of Sturgeon (Acipenser). Genetics 142:525–535.

516

LITERATURE CITED

Brown, K. L. 1985. Demographic and genetic characteristics of dispersal in the mosquitofish, Gambusia affinis (Pisces: Poeciliidae). Copeia 1985:597–612. Brown, K. L. 1987. Colonization by mosquitofish (Gambusia affinis) of a Great Plains River basin. Copeia 1987:336–351. Brown, L. 1981. Patterns of female choice in Mottled Sculpins (Cottidae, Teleostei). Animal Behaviour 29:375–382. Brown, L., and J. F. Downhower. 1982. Polygamy in the Mottled Sculpins (Cottus bairdi) of southwestern Montana (Pisces: Cottidae). Canadian Journal of Zoology 60:1973–1980. Brown, L., and J. F. Downhower. 1983. Constraints on female choice in the Mottled Sculpin, p. 39–54. In Social Behavior of Female Vertebrates. S. Wasser (ed.), Academic Press, New York. Brown, L. R., and P. B. Moyle. 1991. Changes in habitat and microhabitat partitioning within an assemblage of stream fishes in response to predation by Sacramento squawfish (Ptychocheilus grandis). Canadian Journal of Fisheries and Aquatic Sciences 48:849–856. Brown, L. R., S. D. Chase, M. G. Mesa, R. J. Beamish, and P. B. Moyle (eds.). 2009. Biology, Management, and Conservation of Lampreys in North America. American Fisheries Society, Symposium 72, Bethesda, Maryland. Brown, R. W. 1956. Composition of Scientific Words. Smithsonian Institution Press, Washington. Bruch, R. M. 1999. Management of Lake Sturgeon on the Winnebago system—long term impacts of harvest and regulations on population structure. Journal of Applied Ichthyology 15:142–152. Bruch, R. M., and F. P. Binkowski. 2002. Spawning behavior of Lake Sturgeon (Acipenser fulvescens). Journal of Applied Ichthyology 18:570–579. Bruch, R. M., G. Miller, and M. J. Hansen. 2006. Fecundity of Lake Sturgeon (Acipenser fulvescens, Rafinesque) in Lake Winnebago, Wisconsin, USA. Journal of Applied Ichthyology 22 (Supplement 1):116–118. Bruner, J. C. 1991a. Comments on the genus Amyzon (Family Catostomidae). Journal of Paleontology 65:678–686. Bruner, J. C. 1991b. Bibliography of the family Catostomidae (Cypriniformes). Natural History Occasional Paper No. 14, Provincial Museum of Alberta, Edmonton, Alberta, Canada. Bruno, J. F., J. J. Stachowicz, and M. D. Bertness. 2003. Inclusion of facilitation into ecological theory. Trends in Ecology and Evolution 18:119–125. Bryan, J. L., M. L. Wildhaber, D. M. Papoulias, A. J. DeLonay, D. E. Tillitt, and M. L. Annis. 2007. Estimation of gonad volume, fecundity, and reproductive stage of Shovelnose Sturgeon using sonography and endoscopy with application to the endangered Pallid Sturgeon. Journal of Applied Ichthyology 23:411–419. Bryan, M. B., D. Zalinski, K. B. Filcek, S. Libants, W. Li, and K. T. Scribner. 2005. Patterns of invasion and colonization of the Sea Lamprey (Petromyzon marinus) in North America as revealed by microsatellite genotypes. Molecular Ecology 14:3757–3773. Bryan, S. D. 1999. Threatened fishes of the world: Lepidomeda vittata Cope, 1874 (Cyprinidae). Environmental Biology of Fishes 55:226. Bryant, P. B. 1987. A study of the alarm system in selected fishes of northern Mississippi. Unpubl. Master’s thesis, University of Mississippi, Oxford. Bryer, P. J., R. S. Mirza, and D. P. Chivers. 2001. Chemosensory assessment of predation risk by Slimy Sculpins (Cottus cogna-

tus): responses to alarm, disturbance, and predator cues. Journal of Chemical Ecology 27:533–546. Buchholz, U., E. Mouzin, R. Dickey, R. Moolenaar, N. Sass, and L. Mascola. 2000. Haff disease: from the Baltic Sea to the U.S. shore. Emerging Infectious Diseases 6:192–195. Bucholz, M. 1957. Age and growth of River Carpsucker in Des Moines River, Iowa. Proceedings of the Iowa Academy of Sciences 64:589–600. Buckel, J. A., D. O. Conover, N. D. Steinberg, and K. A. McKown. 1999. Impact of age-0 Bluefish (Pomatomus saltatrix) predation on age-0 fishes in the Hudson River Estuary: evidence for density-dependent loss of juvenile Striped Bass (Morone saxatilis). Canadian Journal of Fisheries and Aquatic Science 56:275–287. Buckley, J., and B. Kynard. 1981. Spawning and rearing of Shortnose Sturgeon from the Connecticut River. The Progressive Fish-Culturist 43:74–76. Buckley, J., and B. Kynard. 1985. Habitat use and behavior of prespawning and spawning Shortnose Sturgeon, Acipenser brevirostrum, in the Connecticut River, p. 111–117. In North American Sturgeons: Biology and Aquaculture Potential. F. P. Binkowski and S. I. Doroshov (eds.). Dr. W. Junk Publishers, Dordrecht, The Netherlands. Buddington, R. K., and J. P. Christofferson. 1985. Digestive and feeding characteristics of the chondrosteans, p. 31–41. In North American Sturgeons: Biology and Aquaculture Potential. F. P. Binkowski and S. I. Doroshov (eds.). Dr. W. Junk Publishers, Dordrecht, The Netherlands. Buettner, M. E., and G. G. Scoppettone. 1990. Life history and status of catostomids in Upper Klamath Lake, Oregon. United States Fish and Wildlife Ser vice Report, Reno, Nevada. Bufalino, A. P., and R. L. Mayden. 2010a. Phylogenetic relationships of North American phoxinins (Actinopterygii: Cypriniformes: Leuciscidae) as inferred from S7 nuclear DNA sequences. Molecular Phylogenetics and Evolution 55:143–152. Bufalino, A. P., and R. L. Mayden. 2010b. Molecular phylogenetics of North American phoxinins (Actinopterygii: Cypriniformes: Leuciscidae) based on RAG1 and S7 nuclear DNA sequence data. Molecular Phylogenetics and Evolution 55:274–283. Buisson, L., G. Grenouillet, N. Casajua, and S. Lek. 2010. Predicting the potential impacts of climate change on stream fish assemblages, p. 327–346. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73, Bethesda, Maryland. Bullard, S. A., and K. Jensen. 2008. Blood flukes (Digenea: Aporocotylidae) of Stingrays (Myliobatiformes: Dasyatidae): Orchispirium heterovitellatum from Himmantura imbricata in the Bay of Bengal and a new genus and species of Aporocotylidae from Dasyatis sabina in the northern Gulf of Mexico. Journal of Parasitology 94:1311–1321. Bullard, S. A., S. D. Snyder, K. Jensen, and R. M. Overstreet. 2008. New genus and species of Aporocotylidae (Digenea) from a basal actinopterygian, the American Paddlefish, Polyodon spathula, (Acipenseriformes; Polyodontidae) from the Mississippi Delta. Journal of Parasitology 94:487–495. Büllesbach, E. E., C. Schwabe, and E. R. Lacey. 1997. Identification of a glycosylated relaxin-like molecule from the male Atlantic Stingray, Dasyatis sabina. Biochemistry 36:10735–10741.

LITERATURE CITED

Bullock, T. H., D. A. Bodznick, and R. G. Northcutt. 1983. The phylogenetic distribution of electroreception: evidence for convergent evolution of a primitive vertebrate sense modality. Brain Research Reviews 6:25–46. Bulow, F. J., M. A. Webb, W. D. Crumby, and S. S. Quisenberry. 1988. Effectiveness of a fish barrier dam in limiting movement of rough fishes from a reservoir into a tributary stream. North American Journal of Fisheries Management 8:273–275. Burger, J., K. F. Gaines, and M. Gochfeld. 2001a. Ethnic differences in risk from mercury among Savannah River fishermen. Risk Analysis 21:533–544. Burger, J., K. F. Gaines, J. D. Peles, W. L. Stephens, Jr., C. S. Boring, I. L. Brisbin, Jr., J. Snodgrass, A. L. Bryan, Jr., M. H. Smith, and M. Gochfeld. 2001b. Radiocesium in fish from the Savannah River and Steel Creek: potential food chain exposure to the public. Risk Analysis 21:545–559. Burger, J., W. Stephens, C. S. Boring, M. Kuklinski, J. W. Gibbons, and M. Gochfeld. 1999. Ethnic and socioeconomic differences in exposure from fish caught along the Savannah River. Risk Analysis 19:427–438. Burgess, G. H. 1980. Anchoa mitchilli (Valenciennes), Bay Anchovy, p. 73. In Atlas of North American Freshwater Fishes. D. S. Lee, C. Gilbert, C. Hocutt, R. Jenkins, and McAllister (eds.). North Carolina State Museum of Natural History, Raleigh. Burggren, W., J. Dunn, and K. Barnard. 1979. Brachial circulation and gill morphometrics in the Sturgeon Acipenser transmontanus, an ancient chondrosteian fish. Canadian Journal of Zoology 57:2160–2170. Burggren, W. W. 1978. Gill ventilation in the Sturgeon, Acipenser transmontanus: unusual adaptations for bottom dwelling. Respiratory Physiology 34:153–170. Burggren, W. W., and W. E. Bemis. 1992. Metabolism and ram ventilation in juvenile Paddlefish, Polyodon spathula (Chondrostei: Polyodontidae). Physiological Zoology 65:515–539. Burggren, W. W, and D. J. Randall. 1978. Oxygen uptake and transport during hypoxic exposure in the Sturgeon Acipenser transmontanus. Respiratory Physiology 34:171–183. Burke, J. S., and J. S. Ramsey. 1995. Present and recent historic habitat of the Alabama Sturgeon, Scaphirhynchus suttkusi Williams and Clemmer, in the Mobile Basin. Bulletin Alabama Museum of Natural History 17:17–24. Burkhardt, D. A., J. Gottesman, J. S. Levine, and E. F. MacNichol, Jr. 1983. Cellular mechanisms for color-coding in holostean retinas and the evolution of color vision. Vision Research 23:1031–41 Burkhead, N. M. 1980. The life history of the Central Stoneroller minnow, Campostoma a. anomalum (Rafinesque), in five streams in east Tennessee. Tennessee Wildlife Resources Agency Technical Report 80-50, Nashville. Burkhead, N. M., and B. H. Bauer. 1983. An intergeneric cyprinid hybrid, Hybopsis monacha × Notropis galacturus, from the Tennessee river drainage. Copeia 1983:1074–1077. Burkhead, N. M., and H. L. Jelks. 2001. Effects of suspended sediment on the reproductive success of the Tricolor Shiner, a crevice-spawning minnow. Transactions of the American Fisheries Society 130:959–968. Burleson, M. L., B. N. Shipman, and N. J. Smatresk. 1998. Ventilation and acid-base recovery following exhausting activity in an air-breathing fish. The Journal of Experimental Biology 201:1359–1368.

517

Burnard, D., R. E. Gozlan, and S. W. Griffiths. 2008. The role of pheromones in freshwater fishes. Journal of Fish Biology 73:1–16. Burness, G., S. J. Casselman, A. I. Schulte-Hostedde, C. D. Moyes, and R. Montgomerie. 2004. Sperm swimming speed and energetics vary with sperm competition risk in Bluegill (Lepomis macrochirus). Behavioral Ecology and Sociobiology 56:65–70. Burns, T. A., D. T. Stalling, and W. Goodger. 1981. Gar ichthyootoxin—its effect on crayfish, with notes on Bluegill Sunfish. The Southwestern Naturalist 25:513–515. Burr, B. M. 1980. Polyodon spathula (Walbaum) Paddlefish, p. 45– 46. In Atlas of North American Freshwater Fishes. D. S. Lee, C. R. Gilbert, C. H. Hocutt, R. E. Jenkins, D. E. McAllister, and J. R. Stauffer, Jr. (eds.). North Carolina State Museum of Natural History, Raleigh. Burr, B. M., and R. C. Heidinger. 1983. Reproductive behavior of the Bigmouth Buffalo Ictiobus cyprinellus in Crab Orchard Lake, Illinois. American Midland Naturalist 110:220–221. Burr, B. M., and R. L. Mayden. 1992. Phylogenetics and North American freshwater fishes, p. 18–75. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Burr, B. M., and R. L. Mayden. 1999. A new species of Cycleptus (Cypriniformes: Catostomidae) from Gulf Slope Drainages of Alabama, Mississippi, and Louisiana, with a review of the distribution, biology, and conservation status of the genus. Bulletin of the Alabama Museum of Natural History 20:20–58. Burr, B. M., and M. A. Morris. 1977. Spawning behavior of the Shorthead Redhorse, Moxostoma macrolepidotum, in Big Rock Creek, Illinois. Transactions of the American Fisheries Society 106:80–82. Burr, B. M., and P. W. Smith. 1976. Status of the Largescale Stoneroller, Campostoma oligolepis. Copeia 1976:521–531. Burr, B. M., and M. L. Warren, Jr. 1986a. Status of the Bluehead Shiner (Notropis hubbsi) in Illinois. Transactions of the Illinois State Academy of Science 79:129–136. Burr, B. M., and M. L. Warren, Jr. 1986b. A Distributional Atlas of Kentucky Fishes. Kentucky Nature Preserves Commission Scientific and Technical Series 1, Frankfort, Kentucky, USA. Burr, B. M., and M. L. Warren, Jr. 1995. Freshwater Fishes of Virginia (book review). Copeia 1995:259–252. Burr, J. G. 1931. Electricity as a means of garfish and carp control. Transactions of the American Fisheries Society 61:174–182. Bussing, W. A. 1998. Peces de las Aguas Continentals de Costa Rica. 2nd edition. Universidad de Costa Rica, San Jose. Bussjaeger, C., and T. Briggs. 1978. Phylogenetic implications of bile salts in some catostomid fishes. Copeia 1978:533–535. Buth, D. G. 1978. Biochemical systematics of the Moxostomatini (Cypriniformes, Catostomidae). Unpubl. Ph.D. diss., University of Illinois, Urbana. Buth, D. G. 1979a. Biochemical systematics of the cyprinid genus Notropis—I. The subgenus Luxilus. Biochemical Systematics and Ecology 7:69–79. Buth, D. G. 1979b. Duplicate gene expression in tetraploid fishes of the tribe Moxostomatini (Cypriniformes, Catostomidae). Comparative Biochemistry and Physiology 63B:7–12. Buth, D. G. 1980. Evolutionary genetics and systematic relationships in the catostomid genus Hypentelium. Copeia 1980: 280–290.

518

LITERATURE CITED

Buth, D. G. 1982. Glucosephosphate-isomerase expression in the tetraploid fish, Moxostoma lachneri (Cypriniformes, Catostomidae): evidence for a “retetraploidization”? Genetica 57:171–175. Buth, D. G., and C. B. Crabtree. 1982. Genetic variability and population structure of Catostomus santaanae in the Santa Clara drainage. Copeia 1982:439–444. Buth, D. G., T. R. Haglund, and W. L. Minckley. 1992. Duplicate gene expression and allozyme divergence diagnostic for Catostomus tahoensis and the endangered Chasmistes cujus in Pyramid Lake, Nevada. Copeia 4:935–941. Buth, D. G., and R. L. Mayden. 2001. Allozymic and isozymic evidence for polytypy in the North American catostomid genus Cycleptus. Copeia 2001:899–906. Buth, D. G., R. W. Murphy, and L. Ulmer. 1987. Population differentiation and introgresseve hybridization of the Flannelmouth Sucker and of hatchery and native stocks of the Razorback Sucker. Transactions of the American Fisheries Society 116: 103–110. Butler, V. L. 1996. Tui Chub taphonomy and the importance of marsh resources in the western Great Basin of North America. American Antiquity 61:699–717. Butler, V. L. 2001. Fish faunal remains, p. 271–280. In Archaeological Survey and Excavations in the Carson Desert and Stillwater Mountains, Nevada. R. L. Kelly (ed.). University of Utah Anthropological Papers No 123, Salt Lake City, Utah. Buynak, G. L., and H. W. Mohr, Jr. 1978. Larval development of the northern hogsucker (Hypentelium nigricans), from the Susquehanna River. Transactions of the American Fisheries Society 107:595–599. Buynak, G. L. and H. W. Mohr, Jr. 1980a. Larval development of stoneroller, Cutlips Minnow, and River Chub with diagnostic keys, including four additional cyprinids. The Progressive Fish-Culturalist 42:127–135. Buynak, G. L. and H. W. Mohr, Jr. 1980b. Larval development of Golden Shiner and Comely Shiner from northeastern Pennsylvania. Progressive Fish-Culturist 42:206–211. Byers, S., and G. L. Vinyard. 1990. The effects on the plankton community of filter-feeding Sacramento Blackfish, Orthodon microlepidotus. Oecologia 83:352–357. Cahn, A. R. 1927. An ecological study of the southern Wisconsin fishes. The Brook Silverside (Labidesthes sicculus) and the Cisco (Leucichthys artedi) in their relations to the region. Illinois Biological Monograph 11:1–151. Cain, P. and S. Malwal. 2002. Landmark use and development of navigation behavior in the weakly electric fish Gnathonemus petersii (Mormyridae; Teleostei). Journal of Experimental Biology 205:3915–3923. Cairns, D. K., V. Tremblay, F. Carson, J. M. Casselman, G. Verreault, B. M. Jessop, Y. de Lafontaine, F. G. Bradford, R. Verdon, P. Dumont, Y. Maihot, J. Zhu, A. Mathers, K. Oliveira, K. Benhalima, J. Dietrich, J. A. Hallet, and M. Lagacé. 2008. American Eel abundance indicators in Canada. Canadian Data Reports of Fisheries and Aquatic Science No. 1207. Cake, M. H., I. C. Potter, G. W. Power, and M. Tajbakhsh. 1992. Digestive enzyme activities and their distribution in the alimentary canal of larvae of the three extant Lamprey families. Fish Physiology and Biochemistry 10:1–10. Calhoun, S. W., E. G. Zimmerman, and T. L. Beitinger. 1982. Stream regulation alters acute temperature preferenda of Red

Shiners, Notropis lutrensis. Canadian Journal of Fisheries and Aquatic Sciences 39:360–363. Callan, W. T., and S. L. Sanderson. 2003. Feeding mechanisms in carp: crossflow filtration, palatal protrusions and flow reversals. The Journal of Experimental Biology 206:883–92. Campana, S. E. 2001. Accuracy, precision and quality control in age determination, including a review of the use and abuse of age validation methods. Journal of Fish Biology 59:197–242. Campbell, J. G., and L. R. Goodman. 2004. Acute sensitivity of juvenile Shortnose Sturgeon to low dissolved oxygen concentrations. Transactions of the American Fisheries Society 133:772–776. Campbell, R. R. 1992. Rare and endangered fishes and marine mammals of Canada: COSEWIC fish and marine mammal subcommittee status reports VIII. Canadian Field Naturalist 106:1–6. Campton, D. E., A. L. Bass, F. A. Chapman, and B. W. Bowen. 2000. Genetic distinction of Pallid, Shovelnose, and Alabama Sturgeon: emerging species and the US Endangered Species Act. Conservation Genetics 1:17–32. Canadian Eel Working Group. 2009. American Eel management plan. Fisheries and Oceans Canada, Ontario Ministry of Natural Resources, Ministère des Ressources naturelles et de la Faune du Québec. 39 p. Candolin, U. 1999. The relationship between signal quality and physical condition: is sexual signalling honest in the threespined Stickleback? Animal Behaviour 58:1261–1267. Candolin, U. 2000a. Male-male competition ensures honest signalling of male parental ability in the three-spined Stickleback (Gasterosteus aculeatus). Behavioral Ecology and Sociobiology 49:57–61. Candolin, U. 2000b. Increased signaling effort when survival prospects decrease: male-male competition ensures honesty. Animal Behaviour 60:417–422. Candolin, U. 2009. Population responses to anthropogenic disturbance: lessons from three-spined Sticklebacks Gasterosteus aculeatus in eutrophic habitats. Journal of Fish Biology 75:2108–2121. Candolin, U., J. Engström-Öst, and T. Salesto. 2008. Humaninduced eutrophication enhances reproductive success through effects on parenting ability in Sticklebacks. Oikos 117:459– 465. Capagna, C. G., and J. J. Cech. 1981. Gill ventilation and respiratory efficiency of Sacramento Blackfish, Orthodon microlepidotus Ayres, in hypoxic environments. Journal of Fish Biology 19: 581–591. Cardall, B. L., L. S. Bjerregaard, and K. E. Mock. 2007. Microsatellite markers for the June Sucker (Chasmistes liorus mictus), Utah Sucker (Catostomus ardens), and five other catostomid fishes of western North America. Molecular Ecology Notes 7:457–460. Cardwell, J. R., J. G. Dulka, and N. E. Stacey. 1992. Acute olfactory sensitivity to prostaglandins but not to gonadal steroids in two sympatric species of Catostomus (Pisces: Cypriniformes). Canadian Journal of Zoology 70:1897–1903. Carl, G. C., W. A. Clemens, and C. C. Lindsey. 1967. The FreshWater Fishes of British Columbia. British Columbia Provincial Museum, Victoria. Carlander, K. D. 1969. Handbook of freshwater fishery biology. Vol 1. Life history data on freshwater fishes of the United States and

LITERATURE CITED

Canada, exclusive of the Perciformes. Iowa State University Press, Ames, Iowa. Carlander, K. D. 1969. Handbook of Freshwater Fishery Biology. Volume One. The Iowa State University Press, Ames. Carlson, D. M., and P. S. Bonislawsky. 1981. The Paddlefish (Polyodon spathula) fisheries of the midwestern United States. Fisheries 6:17–27. Carlson, D. M., M. K. Kettler, S. E. Fisher, and G. S. Whitt. 1982. Low genetic variability in Paddlefish populations. Copeia 1982:721–725. Carlson, D. M., W. L. Pflieger, L. Trial, and P. S. Haverland. 1985. Distribution, biology and hybridization of Scaphirhynchus albus and S. platorynchus in the Missouri and Mississippi rivers. Environmental Biology of Fishes 14:51–59. Carlson, R. L., P. C. Wainwright, and T. J. Near. 2009. Relationship between species co-occurrence and rate of morphological change in Percina darters (Percidae: Etheostomatinae). Evolution 63:767–778. Carlton, J. T., J. K. Thompson, L. E. Schemel, and F. H. Nichols. 1990. Remarkable invasion of San Francisco Bay (California, USA) by the Asian clam, Potamocorbula amurensis. 1. Introduction and dispersal. Marine Ecology Progress Series 66:81–94. Caroffino, D. C., T. M. Sutton, and D. J. Daugherty. 2009. Assessment of the vertical distribution of larval Lake Sturgeon drift in the Peshtigo River, Wisconsin, USA. Journal of Applied Ichthyology 25 (Supplement 2):14–17. Caron, F., D. Hatin, and R. Fortin. 2002. Biological characteristics of adult Atlantic Sturgeon (Acipenser oxyrinchus) in the St Lawrence River estuary and the effectiveness of management rules. Journal of Applied Ichthyology 18:580–585. Carnes, W. C. 1958. Contributions to the biology of the eastern Creek Chubsucker, Erimyzon oblongus oblongus (Mitchill). Unpubl. Master’s thesis, North Carolina State College, Raleigh. Carney, D. A., and L. M. Page. 1990. Meristic characteristics and zoogeography of the genus Ptychocheilus (Teleostei: Cyprinidae). Copeia 1990:171–181. Carpenter, C. C. 1975. Functional aspects of the notochordal appendage of young-of-the-year Gar (Lepisosteus). Proceedings of the Oklahoma Academy of Science 55:57–64. Carpenter, G. C. 1995. Notochordal filament in young Gar: morphological characteristics. The Southwestern Naturalist 40:427–428. Carpenter, J. 2005. Competition for food between an introduced crayfish and two fishes endemic to the Colorado River basin. Environmental Biology of Fishes 72:335–342. Carr, M. H. 1942. The breeding habits, embryology, and larval development of the Largemouth Black Bass in Florida. Proceedings of the New England Zoology Club 20:43–77. Carr, M. H. 1946. Notes on the breeding habits of the eastern stumpknocker, Lepomis punctatus punctatus (Cuvier). Journal of Florida Academy of Sciences 9:101–106. Carr, W. E. S., and C. A. Adams. 1973. Food habits of juvenile marine fishes occupying seagrass beds in the estuarine zone near Crystal River, Florida. Transactions of the American Fisheries Society 102:511–540. Carr, S. H., F. Tatman, and F. A. Chapman. 1996. Observations on the natural history of the Gulf of Mexico Sturgeon (Acipenser oxyrinchus de sotoi Vladykov 1955) in the Suwannee River, southeastern United States. Ecology of Freshwater Fish 5:169–174.

519

Carroll, A. M., and P. C. Wainwright. 2003. Functional morphology of prey capture in the Sturgeon, Scaphirhynchus albus. Journal of Morphology 256:270–284. Carton, A. G., and J. C. Montgomery 2003. Evidence of a rheotactic component in the odour search behaviour of Freshwater Eels. Journal of Fish Biology 62:501–516. Carvalho, M. R. de. 1996. Higher-level elasmobranch phylogeny, basal squaleans, and paraphyly, p. 35–62. In Interrelationships of Fishes. M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson (eds.). Academic Press, San Diego, California. Carvalho, M. R. de. 2004. Freshwater Stingrays of the Green River formation of Wyoming (early Eocene) with the descriptions of a new genus and species and an analysis of its phylogenetic relationships (Chondrichthyes: Myliobatiformes). Bulletin of the American Museum of Natural History 284:1–136. Carveth, C. J., A. Widmer, S. A. Bonar, and W. Matter. 2004. Estimation of acute upper lethal water temperature tolerances of native Arizona fishes. Report to the Water Resources Research Center, University of Arizona, Tucson, Arizona. Carveth, C. J., A. M. Widmer, and S. A. Bonar. 2006. Comparison of upper thermal tolerances of native and nonnative fish species in Arizona. Transactions of the American Fisheries Society 135:1433–1440. Case, G. R. 1994. Fossil fish remains from the Late Paleocene Tuscahoma and Early Eocene Bashi Formations of Meridian, Lauderdale County, Mississippi. Palaeontographica 230: 97–138. Cashner, M. F. 2004. Are Spotted Bass (Micropterus punctulatus) attracted to Schreckstoff ? A test of the predator attraction hypothesis. Copeia 2004:592–598. Cashner, M. F., K. R. Piller, and H. L. Bart. 2011. Phylogenetic relationships of the North American cyprinid subgenus Hydrophlox. Molecular Phylogenetics and Evolution 59:725–735. Cashner, R. C., W. J. Matthews, E. Marsh-Matthews, P. J. Unmack, and F. M. Cashner. 2010. Recognition and redescription of a distinctive stoneroller from the southern Interior Highlands. Copeia 2010:300–311. Casselman, J. M. 2003. Dynamics of resources of the American Eel, Anguilla rostrata: declining abundance in the 1990s, p. 255–274. In Eel Biology. K. Aida, K. Tsukamoto, and K. Y. Yamauchi (eds.). Springer, Tokyo. Casselman, J. M. and D. K. Cairns (eds.). 2009. Eels at the Edge: Science, Status, and Conservation Concerns. American Fisheries Society Symposium 58. American Fisheries Society, Bethesda, Maryland. Casselman, S. J., and R. Montgomerie. 2004. Sperm traits in relation to male quality in colonial spawning Bluegill. Journal of Fish Biology 64:1700–1711. Castillo-Rivera, M., G. Moreno, and R. Iniestra. 1994. Spatial, seasonal, and diel variation in abundance of the Bay Anchovy, Anchoa mitchilli (Teleostei: Engraulidae), in a tropical coastal lagoon of Mexico. The Southwestern Naturalist 39:263–268. Castleberry, D. T., and J. J. Cech. 1992. Critical thermal maxima and oxygen minima of five fishes from the Upper Klamath Basin. California Fish and Game 78:145–152. Castonguay, M., P. V. Hodson, C. M. Couillard, M. J. Eckersley, J. -D. Dutil, and G. Verrault. 1994. Why is recruitment of the American Eel, Anguilla rostrata, declining in the St. Lawrence River and the Gulf? Canadian Journal of Fisheries and Aquatic Science 51:479–488.

520 LITERATURE CITED

Castonguay, M., and J. D. McCleave. 1987. Vertical distributions, diel and ontogenetic vertical migrations and net avoidance of leptocephali of Anguilla and other common species in the Sargasso Sea. Journal of Plankton Research 9:195–214. Castro, L. R., and R. K. Cowen. 1991. Environmental factors affecting the early life history of Bay Anchovy, Anchoa mitchilli in Great South Bay, New York. Marine Ecology Progress Series 76:235–247. Cavender, T. 1966. Systematic position of the North American Eocene fish, “Leuciscus” rosei Hussakof. Copeia 1966:311–320. Cavender, T. M. 1986. Review of the fossil history of North American freshwater fishes, p. 699–724. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt & E. O. Wiley (eds.). John Wiley & Sons, New York. Cavender, T. M. 1991. The fossil record of the Cyprinidae, p. 34–54. In Cyprinid Fishes, Systematics, Biology and Exploitation. I. J. Winfield and J. S. Nelson (eds.). Chapman and Hall, New York. Cavender, T. M., and M. M. Coburn. 1992. Phylogenetic relationships of North American Cyprinidae, p. 293–327. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Cech, J. J., Jr., and C. E. Crocker. 2002. Physiology of Sturgeon: effects of hypoxia and hypercapnia. Journal of Applied Ichthyology 18:320–324. Cech, J. J., S. J. Mitchell, and J. M. Massingill. 1979. Respiratory adaptations of Sacramento Blackfish, Orthodon microlepidotus (Ayres), for hypoxia. Comparative Biochemistry and Physiology 63A:411–415. Cech, J. J., Jr., and S. I. Doroshov. 2004. Environmental requirements, preferences, and tolerance limits of North American Sturgeons, p. 73–86. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. H. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. Centers for Disease Control and Prevention. 1998. Haff disease associated with eating buffalo fish—United States, 1997. Morbidity and Mortality Weekly Report 47:1091–3. Cerri, R. D., and D. F. Fraser. 1983. Predation and risk in foraging minnows: balancing conflicting demands. The American Naturalist 121:552–561. Chambers, K. E., R. McDaniell, J. D. Raincrow, M. Deshmukh, P. F. Stadler, and C. H. Chiu. 2009. Hox cluster duplication in the basal teleosts Hiodon alosoides (Osteoglossomorpha). Theory in Biosciences 128:109–120. Chan, M. D., and G. R. Parsons. 2000. Aspects of Brown Madtom, Noturus phaeus, life history in northern Mississippi. Copeia 2000:757–762. Chang, M.-M., and J. G. Maisey. 2003. Redescription of Ellima branneri and Diplomystus shengliensis, and relationships of some basal clupeomorphs. American Museum Novitates 3404:1–35. Chang, M.-M., D. S. Miao, Y. Y. Chen, J. J. Zhou, and P. F. Chen. 2001. Suckers (Fish, Catostomidae) from the Eocene of China account for the family’s current disjunct distributions. Science in China (Series D) 44:577–586. Chang, M.-M., J. Zhang, and D. Miao. 2006. A Lamprey from the Cretaceous Jehol biota of China. Nature 441:972–974. Chapman, G. B., and E. G. Johnson. 1997. An electron microscope study of intrusions into alarm substance cells of the Channel Catfish. Journal of Fish Biology 51:503–514.

Chapman, L. J., and C. A. Chapman. 1993. Fish populations in tropical floodplain pools: a re-evaluation of Holden’s data on the River Sokoto. Ecology of Freshwater Fish 2:23–30. Chaput, G., A. Locke, and D. Cairns. 1997. Status of American Eel (Anguilla rostrata) from the southern Gulf of St. Lawrence, p. 69–93. In The American Eel in Eastern Canada: Stock Status and Management Strategies. Proceedings of Eel Management Workshop. January 13–14. 1997. Quebec City, Quebec. R. H. Peterson (ed.). Canadian Technical Reports of Fisheries and Aquatic Science No. 2196. Chase, J. M. 2007. Drought mediates the importance of stochastic community assembly. Proceedings of the National Academy of Sciences 104:17430–17434. Chatto, D. A. 1979. Effects of salinity on hatching success of the Cui-ui. The Progressive Fish-Culturist 41:82–85. Chavalit, V. 2005. Aquatic alien species in Thailand (Part 1): biodiversity, p. 113–118. In International Mechanisms for the Control and Responsible Use of Alien Species in Aquatic Ecosystems. D. M. Bartley, R. Bhujel, S. Funge-Smith, P. G. Olin, and M. J. Phillips (eds.). Food and Agriculture Organization of the United Nations, Rome. Chebanov, N. A., N. V. Varnavskaya, and V. S. Vanavskiy. 1983. Effectiveness of spawning of male Sockeye Salmon, Oncorhynchus nerka (Salmonidae), of differing hierarchical rank by means of genetic-biochemical markers. Journal of Ichthyology 23:51–55. Chen, L. C., and R. L. Martinich. 1975. Pheromonal stimulation and metabolite inhibition of ovulation in Zebrafish, Brachydanio rerio. Fishery Bulletin 73:889–894. Chen, P. F., and H. H. Harvey. 1999. Spatial structuring of lengthat-age of the benthivorous White Sucker (Catostomus commersoni) in relation to environmental variables. Aquatic Living Resources 12:351–362. Chen, W., V. Lheknim, and R. L. Mayden. 2009. Molecular phylogeny of the Cobitoidea (Teleostei: Cypriniformes) revisited: position of enigmatic loach Ellopostoma resolved with six nuclear genes. Journal of Fish Biology 75:2197–2208. Chen, X.-Y. 1994. Morphology, phylogeny, biogeography and systematics of Phoxinus (Pisces: Cyprinidae). Unpubl. Ph.D. diss. The University of Kansas, Lawrence. Chen, X.-Y., and G. Arratia. 1994. Olfactory organ of Acipenseriformes and comparison with other actinnopterygians: patterns and diversity. Journal of Morphology 222:241–267. Chen, X.-Y., and G. Arratia. 1996. Breeding tubercles of Phoxinus (Teleostei: Cyprinidae): Morphology, distribution, and phylogenetic implications. Journal of Morphology 228:127–144. Cheong, T. S., M. L. Kavvas, and E. K. Anderson. 2006. Evaluation of adult White Sturgeon swimming capabilities and applications to fishway designs. Environmental Biology of Fishes 77:197–208. Cherr, G. N., and W. H. Clark, Jr. 1985. Gamete interaction in the White Sturgeon Acipenser transmontanus: a morphological and physiological review, p. 11–22. In North American Sturgeons: Biology and Aquaculture Potential. F. P. Binkowski and S. I. Doroshov (eds.). Dr. W. Junk Publishers, Dordrecht, The Netherlands. Chesney, E. J. 2008. Foraging behavior of Bay Anchovy larvae, Anchoa mitchilli. Journal of Experimental Marine Biology and Ecology 362:117–124. Chesney, E. J., and E. D. Houde. 1989. Laboratory studies of the effect of hypoxic waters on the survival of eggs and yolk-sac lar-

LITERATURE CITED

vae of the Bay Anchovy, Anchoa mitchilli, p. 184–191. In Population Biology of Bay Anchovy in mid-Chesapeake Bay. E. D. Houde, E. J. Chesney, T. A. Newberger, A. V. Vazquez, C. E. Zastrow, L. G. Morin, H. R. Harvey, and J. W. Gooch (eds.). Chesapeake Biological Laboratory, Solomons, Maryland. not seen Chesser, R. K., M. W. Smith, and M. H. Smith. 1984. Biochemical genetics of mosquitofish populations. II. Incidence and importance of multiple insemination. Genetica 64:77–81. Chew, R. L. 1974. Early life history of the Florida Largemouth Bass. Game and Fresh Water Fish Commission, Tallahassee, Florida, Fishery Bulletin No. 7. Chien, A. K. 1973. Reproductive behaviour of the angelfish Pterophyllum scalare (Pisces: Cichlidae) II. Influence of male stimuli upon the spawning rate of females. Animal Behaviour 21:457–463. Chipman, R. K. 1959. Studies of tolerance of certain freshwater fishes to brine water from oil wells. Ecology 40:299–302. Chipps, S. R., H. D. Symens, and H. Bollig. 2009. Influence of cladoceran composition and abundance on survival of age-0 Paddlefish, p. 411–422. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Chivers, D. P., G. E. Brown, and R. J. F. Smith. 1996. The evolution of chemical alarm signals: attracting predators benefits alarm signal senders. The American Naturalist 148:649–659. Chivers, D. P., R. S. Mirza, and J. G. Johnston. 2002. Learned recognition of heterospecific alarm cues enhances survival during encounters with predators. Behaviour 139:929–938. Chivers, D. P., M. H. Puttlitz, and A. R. Blaustein. 2000. Chemical alarm signaling by Reticulate Sculpins, Cottus perplexus. Environmental Biology of Fishes 57:347–352. Chivers, D. P., and R. J. F. Smith. 1993. The role of olfaction in chemosensory-based predator recognition in the Fathead Minnow, Pimephales promelas. Journal of Chemical Ecology 19:623–633. Chivers, D. P., and R. J. F. Smith. 1994a. Fathead Minnows, Pimephales promelas, acquire predator recognition when alarm substance is associated with the sight of an unfamiliar fish. Animal Behaviour 48:597–605. Chivers, D. P., and R. J. F. Smith. 1994b. The role of experience and chemical alarm signalling in predator recognition by Fathead Minnows (Pimephales promelas). Journal of Fish Biology 44:273–285. Chivers, D. P., and R. J. F. Smith. 1994c. Intra-and interspecific avoidance of areas marked with skin extract from Brook Sticklebacks (Culaea inconstans) in a natural habitat. Journal of Chemical Ecology 20:1517–1524. Chivers, D. P., and R. J. F. Smith. 1995a. Chemical recognition of risky habitats is culturally transmitted among Fathead Minnows, Pimephales promelas (Osteichthys, Cyprinidae). Ethology 99:286–296. Chivers, D. P., and R. J. F. Smith. 1995b. Free-living Fathead Minnows rapidly learn to recognize pike as predators. Journal of Fish Biology 46:949–954. Chivers, D. P., and R. J. F. Smith. 1995c. Fathead Minnows (Pimephales promelas) learn to recognize chemical stimuli from highrisk habitats by the presence of alarm substance. Behavioral Ecology 6:155–158.

521

Chivers, D. P., and R. J. F. Smith. 1998. Chemical alarm signalling in aquatic predator-prey systems: a review and prospectus. Ecoscience 5:338–352. Chivers, D. P., B. D. Wisenden, and R. J. F. Smith. 1995. The role of experience in the response of Fathead Minnows (Pimephales promelas) to skin extract of Iowa Darters (Etheostoma exile). Behaviour 132:665–674. Choe, K. P., and D. H. Evans. 2003. Compensation for hypercapnia by a euryhaline elasmobranch: effect of salinity and roles of gills and kidneys in fresh water. Journal of Experimental Zoology 297A:52–63. Choe, K. P., J. W. Verlander, C. S. Wingo, and D. H. Evans. 2004. A putative H+ -K+ -ATPase in the Atlantic Stingray, Dasytatis sabina: primary sequence and expression in gills. American Journal of Physiology: Regulatory, Integrative, and Comparative Physiology 287:R981–R991. Choudhury, A., and T. A. Dick. 1996. Diclybothrium atriatum n. sp. (Monogena: Diclybothriidae) from North American acipenserid fishes with observations on Diclybothrium armatum and Diclybothrium hamulatum. The Journal of Parasitology 82:965–976. Choudhury, A., and T. A. Dick. 1998. The historical biogeography of Sturgeons (Osteichthyes: Acipenseridae): a synthesis of phylogenetics, palaeontology and palaeogeography. Journal of Biogeography 25:623–640. Choudhury, A., and T. A. Dick. 2001. Sturgeons (Chondrostei: Acipenseridae) and their metazoan parasites: patterns and processes in historical biogeography. Journal of Biogeography 28: 1411–1439. Choudhury, A. and P. A. Nelson. 2000. Redescription of Crepidostomum opeongoensis Caira, 1985 (Trematoda: Allocreadiidae) from fish hosts Hiodon alosoides and Hiodon tergisus (Osteichthyes: Hiodontidae). Journal of Parasitology 86:1305–1312. Choudhury, A., T. L. Hoffnagle, and R. A. Cole. 2004. Parasites of native fishes of the Little Colorado River, Grand Canyon, Arizona. Journal of Parasitology 90:1042–1053. Christie, G. C., and C. I. Goddard. 2003. Sea Lamprey international symposium (SLIS II): advances in the integrated management of Sea Lamprey in the Great Lakes. Journal of Great Lakes Research 29(Suppl. 1):1–14. Chung-Davidson, Y-W., C. B. Rees, M. B. Bryan, and W. Li. 2008. Neurogenic and neuroendocrine effects of goldfish pheromones. Journal of Neuroscience 28:14492–14499. Chuno, A. J., J. S. McKinnon, and M. R. Servedio. 2007. Microhabitat variation and sexual selection can maintain male color polymorphisms. Evolution 61:2504–2515. Cicerello, R. R., and E. I. Landermilk. 1996. Nesting association of the cyprinid fishes Phoxinus cumberlandensis and Semotilus atromaculatus (Cyprinidae). Transactions of the Kentucky Academy of Science 57:47–48. Cimino, M. C. 1972a. Egg-production, polyploidization and evolution in a diploid all-female fish of the genus Poeciliopsis. Evolution 26:294–306. Cimino, M. C. 1972b. Meiosis in a triploid all-female fish (Poeciliopsis, Poeciliidae). Science 175:1484–1486. Clague, J. J., and T. S. James. 2002. History and isostatic effects of the last ice sheet in southern British Columbia. Quaternary Science Reviews 21:71–87. Clark, C. W., and D. A. Levy. 1988. Diel vertical migrations by juvenile sockeye salmon and the antipredation window. The American Naturalist 131:271–290.

522 LITERATURE CITED

Clark, D. W., and J. E. McInerney. 1974. Emigration of the Peamouth Chub, Mylocheilus caurinus, across a dilute seawater bridge: an experimental zoogeographic study. Canadian Journal of Zoology 52:457–469. Clark, K. E. 1978. Ecology and life history of the Speckled Madtom, Noturus leptacanthus (Ictaluridae). Unpubl. Master’s thesis, University of Southern Mississippi, Hattiesburg. Clark, P. U., J. M. Licciardi, D. R. MacAyeal, and J. W. Jenson. 1996. Numerical reconstruction of a soft-bedded Laurentide Ice Sheet during the last glacial maximum. Geology 24:679–682. Clark-Kolaks, S. J., J. R. Jackson, and S. E. Lochmann. 2009. Adult and juvenile Paddlefish in floodplain lakes along the lower White River, Arkansas. Wetlands 29:488–496. Clarkson, R. W., and M. R. Childs. 2000. Temperature effects of hypolimnial-release dams on early life history stages of Colorado River Basin big-river fishes. Copeia 2000:402–412. Clarkson, R. W., and W. L. Minckley. 1988. Morphology and foods of Arizona catostomid fishes: Catostomus insignis, Pantosteus clarki, and three putative hybrids. Copeia 1988:422–433. Clay, T. A., M. D. Suchy, W. Lorio, A. M. Ferrara, and Q. C. Fontenot. 2011. Early growth and survival of larval Alligator Gar Atractosteus spatula reared on artificial floating feed with or without a live Artemia spp. supplement. Journal of the World Aquaculture Society 42:412–416. Clay, W. C. 1975. The Fishes of Kentucky. Kentucky Department of Fish and Wildlife Resources, Frankfort, Kentucky. Cleland, C. E. 1966. The prehistoric animal ecology and ethnozoology of the Upper Great Lakes region. Anthropological Papers, Museum of Anthropology Univeristy of Michigan 29:1–294. Clements, M. D., H. L. Bart, and D. L. Hurley. 2004. Isolation and characterization of two distinct growth hormone cDNAs from the tetraploid Smallmouth Buffalofish (Ictiobus bubalus). General Comparative Endocrinology 136:411–418. Close, D. A., M. S. Fitzpatrick, and H. W. Li. 2002. The ecological and cultural importance of a species at risk of extinction, Pacific Lamprey. Fisheries 27:19–25. Cloutman, D. G., and W. A. Rogers. 2005. Determination of the Dactylogyrus banghami complex (Monogenea: Dactylogyridae) from North American Gulf of Mexico coastal drainages with descriptions of three new species. Comparative Parasitology 72:10–16. Coburn, M. M. 1982. Anatomy and relationships of Notropis atherinoides. Unpubl. Ph.D. diss., Ohio State University, Columbus. Coburn, M. M., and T. M. Cavender. 1992. Interrelationships of North American cyprinid fishes, p. 328–373. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Coburn, M. M., and L. M. Futey. 1996. The ontogeny of supraneurals and neural arches in the cypriniform Weberian apparatus (Teleostei: Ostariophysi). Zoological Journal of the Linnean Society 116:333–346. Cocherell, D. E., A. Kawabata, D. W. Kratville, S. A. Cocherell, R. C. Kaufman, E. K. Anderson, Z. Q. Chen, H. Bandeh, M. M. Rotondo, R. Padilla, R. Churchwell, M. L. Kavvas, and J. J. Cech, Jr. 2011. Passage performance and physiological stress response of adult White Sturgeon ascending a laboratory fishway. Journal of Applied Ichthyology 27:327–334. Cochnauer, T. G., J. R. Lukens, and F. E. Partridge. 1985. Status of White Sturgeon, Acipenser transmontanus, in Idaho, p. 127–133.

In North American Sturgeons: Biology and Aquaculture Potential. F. P. Binkowski and S. I. Doroshov (eds.). Dr. W. Junk Publishers, Dordrecht, The Netherlands. Cochran, P. A. 1986a. The daily timing of Lamprey attacks. Environmental Biology of Fishes 16:325–329. Cochran, P. A. 1986b. Attachment sites of parasitic Lampreys: comparisons among species. Environmental Biology of Fishes 17:71–79. Cochran, P. A. 2009. Predation on lampreys, p. 139–151. In Biology, Management, and Conservation of Lampreys in North America. L. R. Brown, S. D. Chase, M. G. Mesa, R. J. Beamish, and P. B. Moyle (eds.). American Fisheries Society, Symposium 72, Bethesda, Maryland. Cochran, P. A., and A. P. Gripentrog. 1992. Aggregation and spawning by Lampreys (genus Ichthyomyzon) beneath cover. Environmental Biology of Fishes 33:381–387. Cochran, P. A., and R. E. Jenkins. 1994. Small fishes as hosts for parasitic Lampreys. Copeia 1994:499–504. Cochran, P. A., and J. Lyons. 2004. Field and laboratory observations on the ecology and behavior of the Silver Lamprey (Ichthyomyzon unicuspis) in Wisconsin. Journal of Freshwater Ecology 19:245–253. Cochran, P. A., and J. Lyons. 2010. Attachments by parasitic Lampreys within the branchial cavities of their hosts. Environmental Biology of Fishes 88:343–348. Cochran, P. A., J. Lyons, and M. R. Gehl. 2003. Parasitic attachments by overwintering Silver Lampreys, Ichthyomyzon unicuspis, and Chestnut Lampreys, Ichthyomyzon castaneus. Environmental Biology of Fishes 68:65–71. Cochran, P. A., J. Lyons, and E. Merino-Nambo. 1996. Notes on the biology of the Mexican Lampreys Lampetra spadicea and L. geminis (Agnatha: Petromyzontidae). Ichthyological Exploration of Freshwaters 7:173–180. Cochran, P. A., and J. E. Marks. 1995. Biology of the Silver Lamprey, Ichthyomyzon unicuspis, in Green Bay and the lower Fox River, with a comparison to the Sea Lamprey, Petromyzon marinus. Copeia 1995:409–421. Cohen, A. 1997. Sturgeon poaching and black market caviar: a case study. Environmental Biology of Fishes 48:423–426. Cohen, D. M. 1970. How many recent fishes are there? Festschrift for George Sprague Myers. Proceedings of the California Academy of Science 38:341–346. Coker, R. E. 1923. Methuselah of the Mississippi. The Scientific Monthly 16:89–103. Coker, R. E. 1929. Keokuk Dam and the fisheries of the upper Mississippi River. Bulletin of the United States Bureau of Fisheries (Document 893) 45:87–139. Coker, R. E. 1930. Studies of common fishes of the Mississippi River at Keokuk. Bulletin of the United States Bureau of Fisheries 45:141–225. Coker, R. E., A. F. Shira, H. W. Clark, and A. D. Howard. 1921. Natural history and propagation of fresh-water mussels. Bulletin of the Bureau of Fisheries 37:77–181. Cole, K. S., and R. J. F. Smith. 1987. Release of chemicals by prostaglandin-treated female Fathead Minnows, Pimephales promelas, that stimulate male courtship. Hormones and Behavior 21:440–446. Cole, K. S., and R. J. F. Smith. 1992. Attraction of female Fathead Minnows, Pimephales promelas, to chemical stimuli from breeding males. Journal of Chemical Ecology 18:1269–1284.

LITERATURE CITED

Coleman, M. E., D. Winkelman, and H. L. Burge. 1987. Cui-ui egg and larvae temperature tolerance evaluations under controlled fluctuating temperature regimes. United States Fish and Wildlife Ser vice, Reno, Nevada. Coleman, R. M., and R. U. Fischer. 1991. Brood size, male fanning effort and the energetics of a nonshareable parental investment in Bluegill Sunfish, Lepomis macrochirus (Teleostei, Centrarchidae). Ethology 87:177–188. Coleman, R. M., M. R. Gross, and R. C. Sargent. 1985. Parental investment decision rules: a test in Bluegill Sunfish. Behavioral Ecology and Sociobiology 18:59–66. Coleman, S. W., and G. G. Rosenthal. 2006. Swordtail fry attend to chemical and visual cues in detecting predators and conspecifics. PLoSONE 1:1–4. Colgan, P. W., and M. R. Gross. 1977. Dynamics of aggression in male Pumpkinseed Sunfish (Lepomis gibbosus) over the reproductive phase. Zeitschrift für Tierpsychologie 43:139–151. Collette, B. B. 1965. Systematic significance of breeding tubercles in fishes of the family Percidae. Proceedings. United States National Museum 117:567–614. Collette, B. B. 1977. Epidermal breeding tubercles and bony contact organs in fishes. Symposia of the Zoological Society of London 39:225–268. Collette, B. B., and P. Banarescu. 1977. Systematics and zoogeography of the fishes of the family Percidae. Journal of the Fisheries Research Board of Canada 34:1450–1463. Collin, S. P. 2007. Ner vous and sensory systems. Fish Physiology 26:121–179. Collin, S. P., and H. B. Collin. 1993a. The visual system of the Florida garfish. Lepisosteus platyrhincus (Ginglymodi). I. Retina. Brain, Behavior and Evolution 42:77–97. Collin, S. P., and H. B. Collin. 1993b. The visual system of the Florida garfish. Lepisosteus platyrhincus (Ginglymodi). II. Cornea and lens. Brain, Behavior and Evolution 42:98–115. Collin, S. P., and R. G. Northcutt. 1993. The visual system of the Florida garfish. Lepisosteus platyrhincus (Ginglymodi). III. Retinal ganglion cells. Brain, Behavior and Evolution 42:295–320. Collin, S. P., and R. G. Northcutt. 1995. The visual system of the Florida garfish. Lepisosteus platyrhincus (Ginglymodi). IV. Bilateral projections and the binocular visual field. Brain, Behavior and Evolution 45:34–53. Collin, S. P., W. L. Davies, N. S. Hart, and D. M. Hunt. 2009. The evolution of early vertebrate photoreceptors. Philosophical Transactions of the Royal Society, Biological Sciences 364:2925–2940. Collinge, W. E. 1894. The sensory canal system of fishes. Part I.— Ganoidei. Quarterly Journal of Microscopical Science 36:499–537. Collins, M. R., S. G. Rogers, and T. I. J. Smith. 1996. Bycatch of Sturgeons along the southern Atlantic Coast of the USA. North American Journal of Fisheries Management 16:24–29. Collins, M. R., T. I. J. Smith, W. C. Post, and O. Pashuk. 2000a. Habitat utilization and biological characteristics of adult Atlantic Sturgeon in two South Carolina rivers. Transactions of the American Fisheries Society 129:982–988. Collins, M. R., S. G. Rogers, T. I. J. Smith, and M. L. Moser. 2000b. Primary factors affecting Sturgeon populations in the southeastern United States: fishing mortality and degradation of essential habitats. Bulletin of Marine Science 66:917–928.

523

Collins, M. R., W. C. Post, D. C. Russ, and T. I. J. Smith. 2002. Habitat use and movements of juvenile Shortnose Sturgeon in the Savannah River, Georgia-South Carolina. Transactions of the American Fisheries Society 131:975–979. Colombo, G., and R. Rossi. 1978. Environmental influences on growth and sex ratio in different Eel populations (Anguilla anguilla L.) of Adriatic coasts, p. 313–320. In Physiology and Behavior of Marine Organisms. D. S. McLusky and A. J. Berry (eds.). Pergamon Press, Oxford. Colombo, G., and G. Grandi. 1996. Histological study of the development and sex differentiation of the gonad in the European Eel. Journal of Fish Biology 48:493–512. Colombo, L., P. C. Belverdere, A. Marconato, and F. Bentivegna. 1982. Pheromones in teleost fish, p. 84–94. In Proceedings of the Second International Symposium on the Reproductive Physiology of Fish. C. J. J. Richter and H. J. th. Goos (eds.). Wageningen, The Netherlands. Colombo, R. E., J. E. Garvey, N. D. Jackson, R. Brooks, D. P. Herzog, R. A. Hrabik, and T. W. Spier. 2007a. Harvest of Mississippi River Sturgeon drives abundance and reproductive success: a harbinger of collapse? Journal of Applied Ichthyology 23:444–451. Colombo, R. E., J. E. Garvey, and P. S. Wills. 2007b. A guide to the embryonic development of the Shovelnose Sturgeon (Scaphirhynchus platorynchus), reared at a constant temperature. Journal of Applied Ichthyology 23:402–410. Colombo, R. E., J. E. Garvey, and P. S. Wills. 2007c. Gonadal development and sex-specific demographics of the Shovelnose Sturgeon in the middle Mississippi River. Journal of Applied Ichthyology 23:420–427. Colombo, R. E., P. S. Wills, and J. E. Garvey. 2004. Use of ultrasound imaging to determine sex of Shovelnose Sturgeon. North American Journal of Fisheries Management 24:322–326. Comabella, Y., A. Hurtado, and T. García-Galano. 2010. Ontogenetic changes in the morphology and morphometry of Cuban Gar (Atractosteus tristoechus). Zoological Science 27:931–938. Comabella, Y., R. Mendoza, C. Aguilera, O. Carrillo, A. Hurtado, and T. García-Galano. 2006. Digestive enzyme activity during early larval development of the Cuban Gar Atractosteus tristoechus. Fish Physiology and Biochemistry 32:147–157. Combs, D. L. 1982. Angler exploitation of Paddlefish in the Neosho River, Oklahoma. North American Journal of Fisheries Management 4:334–342. Combs, D. L. 1986. The role of regulations in managing Paddlefish populations, p. 68–76. In The Paddlefish: Status, Management and Propagation. J. G. Dillard, L. K. Graham, and T. R. Russell (eds.). American Fisheries Society Special Publication 7. Commens, A. M., and A. Mathis. 1999. Alarm pheromones of Rainbow Darters: responses to skin extracts of conspecifics and congeners. Journal of Fish Biology 55:1359–1362. Commens-Carson, A. M., and A. Mathis. 2007. Responses of three darter species (genus Etheostoma) to chemical alarm cues from conspecifics and congeners. Copeia 2007:838–843. Commonwealth of Pennsylvania. 2011. The Pennsylvania Code. Title 58, Part II, Subpart B, Chapter 75.3. Commonwealth of Pennsylvania, Harrisburg. Compagno, L. J. V. 1999a. Checklist of living elasmobranchs, p. 471–498. In Sharks, Skates, and Rays. The Biology of Elasmobranch Fishes. W. C. Hamlett (ed.). The Johns Hopkins University Press, Baltimore, Maryland.

524

LITERATURE CITED

Compagno, L. J. V. 1999b. Systematics and body form, p. 1–42. In Sharks, Skates, and Rays: the Biology of Elasmobranch Fishes. W. C. Hamlett (ed.). Johns Hopkins University Press, Baltimore, Maryland. Compagno, L. J. V. 2005. Checklist of Chondrichthyes: Sharks, Batoids, and Chimaeras. Science Publishers, Enfield, New Hampshire. Compagno, L. J. V., and T. R. Roberts. 1982. Freshwater Stingrays (Dasyatidae) of Southeast Asia and New Guinea, with description of a new species of Himantura and reports of unidentified species. Environmental Biology of Fishes 7:321–339. Compagno, L. J. V., and T. R. Roberts. 1984. Marine and freshwater Stingrays (Dasyatidae) of West Africa, with description of a new species. Proceedings of the California Academy of Science 43:283–300. Cone, D., D. J. Marcogliese, and R. Hussel. 2004. The myxozoan fauna of Spottail Shiner in the Great Lakes Basin: membership, richness, and geographical distribution. Journal of Parasitology 90:921–932. Congiu, L., F. Fontana, T. Patarnello, R. Rossi, and L. Zane. 2002. The use of AFLP in Sturgeon identification. Journal of Applied Ichthyology 18:286–289. Connell, J. H. 1983. On the prevalence and relative importance of interspecific competition: evidence from field experiments. The American Naturalist 122:661–696. Connell, J. H., and W. P. Sousa. 1983. On the evidence needed to judge ecological stability or persistence. The American Naturalist 121:789–824. Conner, J. V., and R. D. Suttkus. 1986. Zoogeography of freshwater fishes of the Western Gulf Slope, p. 413–456. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. Connor, E. F., and E. D. McCoy. 1979. The statistics and biology of the species-area relationship. The American Naturalist 113:791–833. Connor, E. F., and D. Simberloff. 1979. The assembly of species communities: chance or competition? Ecology 60:1132–1140. Connor, E. F., and D. Simberloff. 1986. Competition, scientific method, and null models in ecology. American Scientist 74:155–162. Conrow, R., A. V. Zale, and R. W. Gregory. 1990. Distributions and abundances of early life stages of fishes in a Florida lake dominated by aquatic macrophytes. Transactions of the American Fisheries Society 19:521–528. Constanz, G. D. 1979. Social dynamics and parental care in the Tessellated Darter (Pisces: Percidae). Proceedings of the Academy of Natural Sciences of Philadelphia 131:131–138. Constanz, G. D. 1984. Sperm competition in poeciliid fishes, p. 465–475. In Sperm Competition and the Evolution of Animal Mating Systems. R. L. Smith (ed.). Academic Press, Orlando, Florida. Constanz, G. D. 1985. Allopaternal care in the Tessellated Darter, Etheostoma olmstedi (Pisces: Percidae). Environmental Biology of Fishes 14:175–183. Contreras-MacBeath, T., and J. M. Rivas. 2007. Threatened fishes of the world: Notropis boucardi (Günther 1868) (Cyprinidae). Environmental Biology of Fishes 78:287–288. Conway, K. W., M. V. Hirt, L. Yang, R. L. Mayden, and A. M. Simons. 2010. Cypriniformes: systematics & paleontology, p. 295–316. In Origin and Phylogenetic Interrelationships of Teleosts. Fest-

schrift in honor of G. Arratia. H.-P. Schultze, J. S. Nelson, and M. V. H. Wilson (eds.). Verlag Dr. Friedrich Pfeil, München, Germany. Cook, A. G. 2001. A review of the comparative morphology and systematics of Utah Lake Suckers (Catostomidae). Journal of Zoology, London 254:293–308. Cook, D. A. 1994. Temporal patterns of food habits of the Atlantic Stingray, Dasyatis sabina (LeSueur, 1824) from the Banana River Lagoon, Florida. Unpubl. Master’s thesis, Florida Institute of Technology, Melbourne. Cook, S. J., and C. M. Bunt. 1999. Spawning and reproductive biology of the Greater Redhorse, Moxostoma valenciennesi, in the Grand River, Ontario. Canadian Field Naturalist 113:497–502. Cooke, D. W., S. D. Leach, and J. J. Isely. 2002. Behavior and lack of upstream passage of Shortnose Sturgeon at a hydroelectric facility and navigation lock complex, p. 101–110. In Biology, Management, and Protection of North American Sturgeon. W. Van Winkle, P. Anders, D. H. Secor, and D. Dixon (eds.). American Fisheries Society Symposium 28, Bethesda, Maryland. Cooke, S. J., C. M. Bunt, S. J. Hamilton, C. A. Jennings, M. P. Pearson, M. S. Cooperman, and D. F. Markle. 2005. Threats, conservation strategies, and prognosis for Suckers (Catostomidae) in North America: insights from regional case studies of a diverse family of non-game fishes. Biological Conservation 121:317–331. Coombs, S. 1994. Nearfield detection of dipole sources by the Goldfish (Carassius auratus) and the Mottled Sculpin (Cottus bairdi). Journal of Experimental Biology 190:109–129. Coombs, S., M. Hastings, and J. Finneran. 1996. Modeling and measuring lateral line excitation patterns to changing dipole source locations. Journal of Comparative Physiology 178:358–371. Cooper, G. P. 1935. Some results of forage fish investigations in Michigan. Transactions of the American Fisheries Society 66:242–266. Cooper, J. E. 1980. Egg, larval and juvenile development of Longnose Dace, Rhinichthys cataractae, and River Chub, Nocomis micropogon, with notes on their hybridization. Copeia 1980:469–478. Cooper, J. E. 1981. Development and replacement order of pharyngeal teeth in the Golden Shiner, Notemigonus crysoleucas. Ohio Journal of Science 81:14–18. Cooper, J. L. 1973. Vertical distribution patterns of Goldeye, Hiodon alosoides, in Fort Peck Reservoir, Montana. Fishery Bulletin 71:473–478. Cooperman, M. S., and D. F. Markle. 2003a. The Endangered Species Act and the National Research Council’s interim judgment in Klamath Basin. Fisheries 28:10–19. Cooperman, M. S., and D. F. Markle. 2003b. Rapid out-migration of Lost River and Shortnose Sucker larvae from in-river spawning beds to in-lake rearing grounds. Transactions of the American Fisheries Society 132:1138–1153. Coots, M. 1965. Occurrences of Lost River Sucker, Deltistes luxatus (Cope), and Shortnose Sucker, Chasmistes brevirostris (Cope), in northern California. California Fish and Game 51:68–73. Cope, E. D. 1870. On the fishes of a fresh-water Tertiary in Kiaho, discovered by Capt. Clarence King. Proceedings of the American Philosophical Society 18:219–231. Cope, E. D. 1872. On the Tertiary coal and fossils of Osino, Nevada. Proceedings of the American Philosophical Society 12:478–481.

LITERATURE CITED

Cope, E. D. 1874. Report on the vertebrate palaeontology of Colorado, p. 475–476. In 7th Annual Report U.S. Geological and Geographical Survey Territories, embracing Colorado:427–533. United States Government Printing Office, Washington, D.C. Cope, E. D. 1875. On the fishes of the Tertiary shales of the South Park. Bulletin of the United States Geological and Geographical Survey of Territories, 2nd series:3–5. Cope, E. D. 1879. The fishes of the Klamath Lake, Oregon. American Naturalist:784–785. Cope, E. D. 1884. The Vertebrata of the Tertiary formations of the West. Book 1. Report of the United States Geological Survey of Territories 3:1–1009. Cope, E. D. 1893. Fossil fishes from British Columbia. Proceedings of the Academy of Natural Sciences of Philadelphia 45:401–402. Cope, W. G., F. M. Holliman, T. J. Kwak, N. C. Oakley, P. R. Lazaro, D. Shea, T. Augspurger, J. M. Law, J. P. Henne, and K. M. Ware. 2011. Assessing water quality suitability for Shortnose Sturgeon in the Roanoke River, North Carolina, USA with an in situ bioassay approach. Journal of Applied Ichthyology 27:1–12. Copes, F. A. 1983. The Longnose Dace Rhinichthys cataractae Valenciennes in Wisconsin and Wyoming waters. Museum of Natural History 19:1–11. Copp, G. H., C. Vaughan, and A. Wheeler. 1993. First occurrence of the North American White Sucker Catostomus commersoni in Great Britain. Journal of Fish Biology 42:615–617. Corkum, L. D., B. Meunier, M. Moscicki, B. S. Zielinski, and A. P. Scott. 2008. Behavioural responses of female Round Gobies (Neogobius melanostomus) to putative steroidal pheromones. Behaviour 145:1347–1365. Cornell, H. V., and J. H. Lawton. 1992. Species interactions, local and regional processes, and limits to the richness of ecological communities: a theoretical perspective. Journal of Animal Ecology 61:1–12. COSEWIC (Committee on the Status of Endangered Wildlife in Canada). 2008. COSEWIC assessment and update status report on the Gravel Chub Erimystax x-punctatus in Canada. Committee on the Status of Endangered Wildlife in Canada, Ottawa, Canada. COSEWIC (Committee on the Status of Endangered Wildlife in Canada). 2010. Wildlife species search. Available from http:// www.cosewic.gc.ca/eng/sct1/index _e.cfm; as of February 2010. COSEWIC (Committee on the Status of Endangered Wildlife in Canada). 2011. Candidate wildlife species. Available from http://www.cosewic.gc.ca; as of June 2011 and October 2011. COSEWIC (Committee on the Status of Endangered Wildlife in Canada). 2012. COSEWIC update status report on the American Eel Anguilla rostrata in Canada. Committee on the Status of Endangered Wildlife in Canada. Ottawa. Cosson, J., O. Linhart, S. D. Mims, W. L. Shelton, and M. Rodina. 2000. Analysis of motility parameters from Paddlefish and Shovelnose Sturgeon spermatozoa. Journal of Fish Biology 56:1348–1367. Courtenay, W. R., Jr., D. A. Hensley, J. N. Taylor, and J. A. McCann. 1986. Distribution of exotic fishes in North America, p. 675–698. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. Coutant, C. C. 2004. A riparian habitat hypothesis for successful reproduction of White Sturgeon. Reviews in Fisheries Science 12:23–73.

525

Cowan, J. H., Jr., and E. D. Houde. 1990. Growth and survival of Bay Anchovy Anchoa mitchilli larvae in mesocosm enclosures. Marine Ecology Progress Series 68:47–57. Cowan, J. H., Jr., K. A. Rose, E. D. Houde, S. B. Wang, and J. Young. 1999. Modeling effects of increased larval mortality on Bay Anchovy population dynamics in the mesohaline Chesapeake Bay: evidence for compensatory reserve. Marine Ecology Progress Series 185:133–146. Cowley, D. E., and J. E. Sublette. 1987. Food habits of Moxostoma congestum (Baird and Girard) and Cycleptus elongatus (LeSueur) (Catostomidae: Cypriniformes) taken from Black River, Eddy County, New Mexico. Southwestern Naturalist 32:411–413. Cox, C. B., and P. D. Moore. 1993. Biogeography, an ecological and evolutionary approach. 5th edition. Blackwell Science, Oxford. Crabtree, C. B., and D. G. Buth. 1981. Gene duplication and diploidization in tetraploid catostomid fishes Catostomus fumeiventris and C. santaanae. Copeia 1981:705–708. Crabtree, C. B., and D. G. Buth. 1987. Biochemical systematics of the catostomid genus Catostomus: assessment of C. clarki, C. plebeius and C. discobolus including the Zuni Sucker, C. d. yarrowi. Copeia 1987:843–854. Cracraft, J. 1974. Continental drift and vertebrate distribution. Annual Review of Ecology and Systematics 5:215–261. Craig, J. F., K. Smiley, and J. A. Babaluk. 1989. Changes in the body composition with age of Goldeye, Hiodon alosoides. Canadian Journal of Fisheries and Aquatic Sciences 46:853–858. Craig, J. M., D. M. Papoulias, M. V. Thomas, M. L. Annis, and J. Boase. 2009. Sex assignment of Lake Sturgeon (Acipenser fulvescens) based on plasma sex hormone and vitellogenin levels. Journal of Applied Ichthyology 25 (Supplement 2):60–67. Craig, J. M., M. V. Thomas, and S. J. Nichols. 2005. Length-weight relationship and a relative condition factor equation for Lake Sturgeon (Acipenser fulvescens) from the St Clair system (Michigan, USA). Journal of Applied Ichthyology 21:81–85. Crapon de Caprona, M.-D., and M. J. Ryan. 1990. Conspecific mate recognition in swordtails, Xiphophorus nigrensis and X. pygmaeus: olfactory and visual cues. Animal Behaviour 39:290–296. Crawford, M. 1979. Reproductive modes of the Least Chub Iotichthys phlegethontis Cope. Unpubl. Master’s thesis, Utah State University, Logan. Crescitelli, F., M. McFall-Ngai, and J. Horwirz. 1985. The visual pigment sensitivity hypothesis: further evidence from fishes from varying habitats. Journal of Comparative Physiology A 157:323–333. Crisp, D. T., and P. A. Carling. 1989. Observations on siting, dimensions and structure of salmonid redds. Journal of Fish Biology 34:119–134. Crocker, C. E., and J. J. Cech, Jr. 1997. Effects of environmental hypoxia on oxygen consumption rates and swimming activity in juvenile White Sturgeon, Acipenser transmontanus, in relation to temperature and life intervals. Environmental Biology of Fishes 50:383–389. Cross, F. B. 1967. Handbook of Fishes of Kansas. University of Kansas, Museum of Natural History Miscellaneous Publications 45, Lawrence. Cross, F. B. 1970. Fishes as indicators of Pleistocene and Recent environments in the Central Plains, p. 241–258. In Pleistocene and Recent Environments of the Central Great Plains. W. Dort, Jr.,

526

LITERATURE CITED

and J. K. Jones, Jr. (eds.). University of Kansas, Special Publication No. 3. Cross, F. B., and W. L. Minkley. 1960. Five natural hybrid combinations in minnows (Cyprinidae). University of Kansas Publications, Museum of Natural History 13:1–18. Cross, F. B., and J. T. Collins. 1995. Fishes in Kansas. 2nd edition. University of Kansas Natural History Museum Public Education Series 14:1–315. Cross, F. B., R. L. Mayden, and J. D. Stewart. 1986. Fishes in the western Mississippi Basin (Missouri, Arkansas and Red Rivers), p. 363–412. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. Crossman, E. J., and D. E. McAllister. 1986. Zoogeography of freshwater fishes of the Hudson Bay drainage, Ungava Bay and the Arctic Archipelago, p. 53–104. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. Crow, K. D., P. F. Stadler, V. J. Lynch, C. Amemiya, and G. P. Wagner. 2006. The fish-specific hox cluster duplication is coincident with the origin of teleosts. Molecular Biology and Evolution 23:121–136. Crowl, T. A., and A. P. Covitch. 1990. Predator-induced life history shifts in a freshwater snail. Science 247:949–951. Crowley, B. E., P. L. Koch, and E. B. Davis. 2008. Stable isotope constraints on the elevation history of the Sierra Nevada Mountains, California. Geological Society of America Bulletin 120:588–598. Crumpton, J. 1971. Food habits of Longnose Gar (Lepisosteus osseus) and Florida Gar (Lepisosteus platyrhincus) collected from five central Florida lakes. Proceedings of the Southeastern Association of Game and Fish Commissioners. 24:419–424. Cruz, J., C. Aguilera, and R. Mendoza. 2010. Efectos de la contaminación en la fisiología del Catán, p. 164–187. In Biología, Ecología y Avances en el Cultivo de Catán Atractosteus spatula. R. C. Mendoza, C. Aguilera, and J. Montemayor (eds.). Universidad Autónoma de Nuevo León, Monterrey, Mexico. Cudmore-Vokey, B. and E. J. Crossman. 2000. Checklists of the fish fauna of the Laurentian Great Lakes and their connecting channels. Canadian Manuscript Report of Fisheries and Aquatic Science 2550:1–39. Cuerrier, J. P. 1951. The use of pectoral fin rays for determining age of Sturgeon and other species of fish. The Canadian Fish Culturist 11:1–9. Culp, J. M. 1989. Nocturnally constrained foraging of a lotic minnow (Rhinichthys cataractae). Canadian Journal of Zoology 67:2008–2012. Cumbie, P. M. 1975. Mercury levels in Georgia otter, mink and freshwater fish. Bulletin of Environmental Contamination and Toxicology 14:193–196. Cummings, M. E., F. J. García de León, D. M. Mollaghan, and M. J. Ryan. 2006. Is UV ornamentation an amplifier in swordtails? Zebrafish 3:91–100. Cummings, M. E., and R. Gelineau-Kattner. 2009. The energetic costs of alternative male reproductive strategies in Xiphophorus nigrensis. Journal of Comparative Physiology A 195:935–946. Cummings, M. E., G. G. Rosenthal, and M. J. Ryan. 2003. A private ultraviolet channel in visual communication. Proceedings of the Royal Society of London B 270:897–904.

Cunha, C., N. Mesquita, T. E. Dowling, A. Gilles, and M. M. Coelho. 2002. Phylogenetic relationships of Eurasian and American cyprinids using cytochrome b sequences. Journal of Fish Biology 61:929–944. Cunjak, R. A. 1988. Behaviour and microhabitat of young Atlantic Salmon (Salmo salar) during winter. Canadian Journal of Fisheries and Aquatic Sciences 45:2156–2160. Cunjak, R. A., and G. Power. 1986. Winter biology of the Blacknose Dace, Rhinichthys atratulus, in a southern Ontario stream. Environmental Biology of Fishes 17:53–60. Cunningham, M. E., D. F. Markle, V. G. Watral, M. L. Kent, and L. R. Curtis. 2005. Patterns of fish deformities and their association with trematode cysts in the Willamette River, Oregon. Environmental Biology of Fishes 73:9–19. Currie, C. R., B. Wong, A. E. Stuart, T. R. Schulltz, S. A. Rehner, U. G. Mueller, G. Sung, J. W. Spatafora, and N. A. Straus. 2003. Ancient tripartite coevolution in the attine ant-microbe symbiosis. Science 299:386–388. Curry, K. D., and A. Spacie. 1984. Differential use of stream habitat by spawning catostomids. American Midland Naturalist 111:267–279. Curtis, G. L., J. S. Ramsey, and D. L. Scarnecchia. 1997. Habitat use and movements of Shovelnose Sturgeon in Pool 13 of the upper Mississippi River during extreme low flow conditions. Environmental Biology of Fishes 50:175–182. Dadda, M., A. Pilastro, and A. Bisazza. 2005. Male sexual harassment and female schooling behaviour in the Eastern Mosquitofish. Animal Behaviour 70:463–471. Dadswell, M. J. 1979. Biology and population characteristics of the Shortnose Sturgeon, Acipenser brevirostrum (LeSueur, 1818) (Osteichthyes: Acipenseridae), in the Saint John River estuary, New Brunswick, Canada. Canadian Journal of Zoology 57:2186–2210. Dahlberg, M. D., and E. P. Odum. 1970. Annual cycles of species occurrence, abundance, and diversity in Georgia estuarine fish populations. American Midland Naturalist 83:382–392. Daly, R. J. 1970. Systematics of southern Florida anchovies (Pisces: Engraulidae). Bulletin of Marine Science 20:70–104. Daniels, R. A., and P. B. Moyle. 1983. Life history of Splittail (Cyprinidae: Pogonichthys macrolepidotus) in the Sacramento-San Joaquin Estuary. United States Fish and Wildlife Ser vice Fishery Bulletin 81:647–654. Danylchuk, A. J., and M. G. Fox. 1996. Size- and age-related variation in the seasonal timing of nesting activity, nest characteristics, and female choice of parental male Pumpkinseed Sunfish (Lepomis gibbosus). Canadian Journal of Zoology 74:1834–1840. Danylchuk, A. J., and W. M. Tonn. 2001. Effects of social structure on reproductive activity in male Fathead Minnows (Pimephales promelas). Behavioral Ecology 12:482–489. Darnell, R. M. 1958. Food habits of fishes and larger invertebrates of Lake Pontchartrain, Louisiana, an estuarine community. Publications of the Institute of Marine Science 5:353–416. Darnell, R. M. 1961. Trophic spectrum of an estuarine community, based on studies of Lake Pontchartrain, Louisiana. Ecology 42:553–568. Darnell, R. M. 1962. Fishes of the Rio Tamesi and related coastal lagoons in east central Mexico. Publications of the Institute of Marine Science University of Texas 8:299–365.

LITERATURE CITED

Dasgupta, S., R. J. Onders, D. T. Gunderson, and S. D. Mims. 2004. Methylmercury concentrations found in wild and farmraised Paddlefish. Journal of Food Science 69:122–125. Dauble, D. D. 1986. Life history and ecology of the Largescale Sucker (Catostomus macrocheilus) in the Columbia River. American Midland Naturalist 116:356–367. Dauble, D. D., and R. L. Buschbom. 1981. Estimates of hybridization between two species of catostomids in the Columbia River. Copeia 1981:802–810. Dauble, D. D., R. A. Moursund, and M. D. Bleich. 2006. Swimming behaviour of juvenile Pacific Lamprey, Lampetra tridentata. Environmental Biology of Fishes 75:167–171. Daugherty, D. J., T. D. Bacula, and T. M. Sutton. 2008. Reproductive biology of Blue Sucker in a large Midwestern river. Journal of Applied Ichthyology 2008:1–6. Davenport, D., and M. Warmuth. 1965. Notes on the relationship between the freshwater mussel Anodonta implicata Say and the Alewife, Pomolobus pseudoharengus (Wilson). Limnology and Oceanography 10 (supplement):R74–R78. David, S. 2009. Primitive fishes, living fossils and their ancestors. Available from http://www.primitivefishes.com/Gars.html; as of October 25, 2010. Davies, W. D. 1976. Lake Nicaragua fishery resources, p. 261–265. In Investigations of the Ichthyofauna of Nicaraguan Lakes. T. B. Thorson (ed.). University of Nebraska, Lincoln. Davis, B. J., and R. J. Miller. 1967. Brain patterns in minnows of the genus Hybopsis in relation to feeding habits and habitat. Copeia 1967:1–39. Davis, J. G. 2006. Reproductive biology, life history and population structure of a Bowfin Amia calva population in southeastern Louisiana. Unpubl. Master’s thesis. Nicholls State University, Thibodaux, Louisiana. Davis, J. T. 1993. Baitfish, p. 307–321. In Culture of Nonsalmonid Freshwater Fishes. R. R. Stickney (ed.). CRC Press, Boca Raton, Florida. Davis, K. B., and N. C. Parker. 1986. Plasma corticosteroid stress response of fourteen species of warmwater fish to transportation. Transactions of the American Fisheries Society 115:495–499. Davis, M. B. 1983. Quaternary history of deciduous forests of eastern North America and Europe. Annals of the Missouri Botanical Garden 70:550–563. Davis, M. C., N. H. Shubin, and A. Force. 2004. Pectoral fin and girdle development in the basal actinopterygians Polyodon spathula and Acipenser transmontanus. Journal of Morphology 262:608–628. Davis, R. M. 1967. Parasitism by newly-transformed anadromous Sea Lampreys on landlocked salmon and other fishes in a coastal Maine lake. Transactions of the American Fisheries Society 96:11–16. Dawley, R. M. 1987. Hybridization and polyploidy in a community of three Sunfish species (Pisces: Centrarchidae). Copeia 1987:326–335. Dawley, R. M., and K. A. Goddard. 1988. Diploid-triploid mosaics among unisexual hybrids of the minnows Phoxinus eos and Phoxinus neogaeus. Evolution 42:649–659. Dawley, R. M., R. J. Schultz, and K. A. Goddard. 1987. Clonal reproduction and polyploidy in unisexual hybrids of Phoxinus eos and Phoxinus neogaeus (Pisces; Cyprinidae). Copeia 1987:275–283.

527

Dawson, H. A., M. L. Jones, K. T. Scribner, and S. A. Gilmore. 2009. An assessment of age determination methods for Great Lakes larval Sea Lampreys. North American Journal of Fisheries Management 29:914–927. Daxobeck, C., D. K. Barnard, and D. J. Randall. 1981. Functional morphology of the gills of the Bowfin, Amia calva L., with special reference to their significance during air exposure. Respiration Physiology 43:349–364. Daye, P. G., and B. D. Glebe. 1984. Fertilization success and sperm motility of Atlantic Salmon (Salmo salar) in acidified water. Aquaculture 43:307–312. Deacon, J. E., and C. D. Williams. 1991. Ash Meadows and the legacy of the Devils Hole Pupfish, p. 69–87. In Battle Against Extinction: Native Fish Management in the American Southwest. W. L. Minckley and J. E. Deacon (eds.). University of Arizona Press, Tucson. Dean, B. 1895. The early development of garpike and Sturgeon. Journal of Morphology 11:1–62. Dean, B. 1898. On the dogfish (Amia calva), its habits and breeding. In Fourth Annual Report of the Commissioners of Fisheries, Game and Forests of the State of New York. Decker, E. A., A. D. Crum, S. D. Mims, and J. H. Tidwell. 1991. Processing yields and composition of Paddlefish (Polyodon spathula), a potential aquaculture species. Journal of Agricultural and Food Chemistry 39:686–688. de Chambrier, A., S. C. Coquille, J. Mariaux, and V. Tkach. 2009. Redescription of Testudotaenia testudo (Magath, 1924) (Eucestoda: Proteocephalidea), a parasite of Apalone spinifera (Le Sueur) (Reptilia: Trionychidae) and Amia calva L. (Pisces: Amiidae) in North America and erection of the Testudotaeniinae n. subfam. Systematic Parasitology 73:49–64. de Fraipont, M., G. J. FitzGerald, and H. Guderley. 1993. Agerelated differences in reproductive tactics in the three-spined Stickleback, Gasterosteus aculeatus. Animal Behaviour 46:961–968. de Gaudemar, B., and E. Beall. 1999. Reproductive behavioural sequences of single pairs of Atlantic Salmon in an experimental stream. Animal Behaviour 57:1207–1217. de Gaudemar, B., S. Schroder, and E. Beall. 2000. Nest placement and egg distribution in Atlantic Salmon redds. Environmental Biology of Fishes 57:37–47. DeHaan, P. W., S. V. Libants, R. F. Elliott, and K. T. Scribner. 2006. Genetic population structure of remnant Lake Sturgeon populations in the upper Great Lakes basin. Transactions of the American Fisheries Society 135:1478–1492. DeKay, J. E. 1842. Zoology of New-York, or the New-York fauna. Part IV. fishes. W. and A. White and J. Visscher, Albany, New York. de la Herrán, R., F. Fontana, M. Lanfredi, L. Congiu, M. Leis, R. Rossi, C. R. Rejón, M. R. Rejón, and M. A. Garrido-Ramos. 2001. Slow rates of evolution and sequence homogenation in an ancient satellite DNA family of Sturgeons. Molecular Biology and Evolution 18:432–436. DeLancey, L. B. 1989. Trophic relationship in the surf zone during the summer at Folly Beach, South Carolina. Journal of Coastal Research 5:477–488. Delany, M. F., and C. L. Abercrombie. 1986. American Alligator food habits in northcentral Florida. The Journal of Wildlife Management 50:348–353. Delarbre, C., H. Escriva, C. Gallut, V. Barriel, P. Kourilsky, P. Janvier, V. Laudet, and G. Gachelin. 2000. The complete nucleotide

528 LITERATURE CITED

sequence of the mitochondrial DNA of the agnathan Lampetra fluviatilis: bearings on the phylogeny of cyclostomes. Molecular Biology and Evolution 17:519–529. Delarbre, C., C. Gallut, V. Barriel, P. Janvier, and G. Gachelin. 2002. Complete mitochondrial DNA of the Hagfish, Eptatretus burgeri: the comparative analysis of mitochondrial DNA sequences strongly supports the cyclostome monophyly. Molecular Phylogenetics and Evolution 22:184–192. Delco, E. A. 1960. Sound discrimination by males of two cyprinid fishes. Texas Journal of Science 12:48–54. Deliagina, T. G., F. Ullén, M.-J. Gonzalez, H. Ehrsson, G. N. Orlovsky, and S. Grillner. 1995. Initiation of locomotion by lateral line photoreceptors in Lamprey: behavioural and neurophysiological studies. The Journal of Experimental Biology 198:2581–2591. DeLonay, A. J., D. M. Papoulias, M. L. Wildhaber, M. L. Annis, J. L. Bryan, S. A. Griffith, and S. H. Holan. 2007. Use of behavioral and physiological indicators to evaluate Scaphirhynchus Sturgeon spawning success. Journal of Applied Ichthyology 23:428–435. DeMarais, B. D., T. E. Dowling, M. E. Douglas, W. L. Minckley, and P. C. Marsh. 1992. Origin of Gila seminuda (Teleostei: Cyprinidae) through introgressive hybridization: implications for evolution and conservation. Proceedings of the National Academy of Sciences 89:2747–2751. DeMarais, B. D., T. E. Dowling, and W. L. Minckley. 1993. Postperturbation genetic changes in populations of endangered Virgin River chubs. Conservation Biology 7:334–341. DeMarais, B. D., and W. L. Minckley. 1992. Hybridization in native cyprinid fishes, Gila ditaenia and Gila sp., in northwestern Mexico. Copeia 1992:697–703. DeMartini, E. E. 1987. Paternal defense, cannibalism and polygamy: factors influencing the reproductive success of Painted Greenlings (Pisces, Hexagrammidae). Animal Behaviour 35:1145–1158. Demski, L. S., J. W. Gerald, and A. N. Popper. 1973. Central and peripheral mechanisms of teleost sound production. American Zoologist 13:1141–1167. Dence, W. A. 1933. Notes on a large Bowfin (Amia calva) living in a mudpuddle. Copeia 1933:35. Dence, W. A. 1948. Life history, ecology and habits of the Dwarf Sucker, Catostomus commersonii utawana Mather, at the Huntington Wildlife Station. Roosevelt Wildlife Bulletin 8:82–150. Deng, X., J. P. Van Eenennaam, and S. I. Doroshov. 2002. Comparison of early life stages and growth of Green and White Sturgeon, p. 237–248. In Biology, Management, and Protection of North American Sturgeon. W. Van Winkle, P. Anders, D. H. Secor, and D. Dixon (eds.). American Fisheries Society Symposium 28, Bethesda, Maryland. Denoncourt, R. F., T. B. Robbins, and R. Hesser. 1975. Recent introductions and reintroductions to the Pennsylvania fish fauna of the Susquehanna River drainage above Conowingo Dam. Proceedings of the Pennsylvania Academy of Science 49:57–58. Department of Fisheries and Oceans Canada. 2010. Status of American Eel and progress on achieving management goals. Canadian Science Advisory Secretariat, Science Advisory Report 2010/062. Depkin, F. C., M. C. Coulter, and A. L. Bryan, Jr. 1992. Food of nestling Wood Storks in east-central Georgia. Colonial Waterbirds 15:219–225. Derickson, W. K., and K. S. Price, Jr. 1973. The fishes of the shore zone of Rehoboth and Indian River bays, Delaware. Transactions of the American Fisheries Society 102:552–562.

De Roth, G. C. 1973. Effects of temperature and light on aerial breathing behavior of the Spotted Gar, Lepisosteus oculatus. The Ohio Journal of Science 73:35–41. DeSalle, R., and V. J. Birstein. 1996. PCR identification of black caviar. Nature 381:432–436. Detenbeck N. E., P. W. DeVore, G. J. Niemi, and A. Lima. 1992. Recovery of temperate-stream fish communities from disturbance: a review of case studies and synthesis of theory. Environmental Management 16:33–53. Dettlaff, T. A., A. S. Ginsburg, and O. J. Schmalhausen. 1993. Sturgeon Fishes. Developmental Biology and Aquaculture. Springer-Verlag, Berlin, Germany. Dettmers, J. M., S. Butreuter, D. H. Wahl, and D. A. Soluk. 2001. Patterns in abundance of fishes in the main channels of the upper Mississippi River system. Canadian Journal of Fisheries and Aquatic Sciences 58:933–942. Detwyler, R., and E. D. Houde. 1970. Food selection by laboratoryreared larvae of the Scaled Sardine Harengula pensacolae (Pisces, Clupeidae) and the Bay Anchovy Anchoa mitchilli (Pisces, Engraulidae). Marine Biology 7:214–222. Devitsina, G. V., and A. A. Kazhlaev. 1993. Chemoreceptor organs in early juvenile Paddlefish, Polyodon spathula. Journal of Ichthyology 33:143–149. De Vlaming, V. L. 1975. Effects of pinealectomy on gonadal activity in the cyprinid teleost, Notemigonus crysoleucas. General and Comparative Endocrinology 26:36–49. De Vlaming, V. L., and M. Sage. 1973. Osmoregulation in the euryhaline elasmobranch, Dasyatis sabina. Comparative Biochemistry and Physiology 45A:31–44. DeVore, J. D., B. W. James, C. A. Tracy, and D. A. Hale. 1995. Dynamics and potential production of White Sturgeon in the unimpounded lower Columbia River. Transactions of the American Fisheries Society 124:845–856. DeVries, D. R., G. M. Lein, and R. J. H. Hoxmeier. 2009. Paddlefish populations in the Alabama River drainage, p. 39–50. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. DeWoody, J. A., D. E. Fletcher, M. MacKiewicz, S. D. Wilkins, and J. C. Avise. 2000c. The genetic mating system of Spotted Sunfish (Lepomis punctatus): mate numbers and the influence of male reproductive parasites. Molecular Ecology 9:2119–2128. DeWoody, J. A., D. E. Fletcher, S. D. Wilkins, W. S. Nelson, and J. C. Avise. 1998. Molecular genetic dissection of spawning, parentage, and reproductive tactics in a population of Redbreast Sunfish, Lepomis auritus. Evolution 52:1802–1810. DeWoody, J. A., D. E. Fletcher, S. D. Wilkins, W. S. Nelson, and J. C. Avise. 2000a. Parentage and nest guarding in the Tessellated Darter (Etheostoma olmstedi) assayed by microsatellite markers (Perciformes: Percidae). Copeia 2000:740–747. DeWoody, J. A., D. E. Fletcher, S. D. Wilkins, W. S. Nelson, and J. C. Avise. 2000b. Genetic monogamy and biparental care in an externally fertilizing fish, the Largemouth Bass (Micropterus salmoides). Proceedings of the Royal Society of London B 267: 2431–2437. Diamond, J. N. 1975. Assembly of species communities, p. 342– 444. In Ecology and Evolution of Communities. M. L. Cody and J. M. Diamond (eds.). Belknap Press of Harvard University Press, Cambridge, Massachusetts.

LITERATURE CITED

Diamond, S. A., J. T. Oris, and S. I. Guttman. 1995. Adaptation to fluoranthene exposure in a laboratory population of Fathead Minnows. Environmental Toxicology and Chemistry 14:1393–1400. DiBenedetto, K. C. 2009. Life history characteristics of Alligator Gar, Atractosteus spatula in the Bayou DuLarge area of southcentral Louisiana. Unpubl. Master’s thesis, Louisiana State University, Baton Rouge. Dickerson, B. R., M. F. Willson, P. Bentzen, and T. P. Quinn. 2004. Size-assortative mating in salmonids: negative evidence for Pink Salmon in natural conditions. Animal Behaviour 68:381–385. Dickinson, W. R. 2004. Evolution of the North American Cordillera. Annual Review of Earth and Planetary Science 32:13–45. Di Dario, F. 2002. Evidence supporting a sister-group relationship between Clupeoidea and Engrauloidea (Clupeomorpha). Copeia 2002:496–503. Di Dario, F. 2009. Chirocentrids as engrauloids: evidence from suspensorium, branchial arches, and infraorbital bones (Clupeomorpha, Teleostei). Zoological Journal of the Linnean Society 156:363–383. Dieterman, D. J., M. S. Baird, and D. L. Galat. 2000, Mortality of Paddlefish in hoop nets in the lower Missouri River, Missouri. North American Journal of Fisheries Management 20:226–230. Dillard, J. G, L. K. Graham, and T. R. Russell (eds.). 1986. The Paddlefish: Status, Management and Propagation. American Fisheries Society Special Publication 7, Bethesda, Maryland. Dillman, C. B., R. M. Wood, B. R. Kuhajda, J. M. Ray, V. B. Salnikov, and R. L. Mayden. 2007. Molecular systematics of Scaphirhynchinae: an assessment of North American and central Asian freshwater Sturgeon species. Journal of Applied Ichthyology 23:290–296. Dimmick, W. W. 1987. Phylogenetic relationships of Notropis hubbsi, N. welaka, and N. emiliae (Cypriniformes: Cyprinidae). Copeia 1987:316–325. Dimmick, W. W. 1988. Ultrastructure of North American cyprinid barbels. Copeia 1988:72–80. Dimmick, W. W. 1993. A molecular perspective on the phylogenetic relationships of the barbeled minnows, historically assigned to the genus Hybopsis (Cyprinidae: Cypriniformes). Molecular Phylogenetics and Evolution 2:173–184. Dimmick, W. W., and B. M. Burr. 1999. Phylogenetic relationships of the Suckermouth Minnow, genus Phenacobius, inferred from parsimony analyses of nucleotide sequence, allozymic and morphological data (Cyprinidae: Cypriniformes). Biochemical Systematics and Ecology 27:469–485. Din, Z. B., and G. Gunter. 1986. The food and feeding habits of the common Bay Anchovy, Anchoa mitchilli (Valenciennes). Pertanika 9:99–108. Dingerkus, G., and W. M. Howell. 1976. Karyotypic analysis and evidence of tetraploidy in the North American Paddlefish, Polyodon spathula. Science 194:842–844. Di Santo, V., and W. A. Bennett. 2011. Is post-feeding thermotaxis advantageous in elasmobranch fishes? Journal of Fish Biology 78:195–207. Divers, S. J., S. S. Boone, J. J. Hoover, K. A. Boysen, K. J. Killgore, C. E. Murphy, S. G. George, and A. C. Camus. 2009. Field endoscopy for identifying gender, reproductive stage and gonadal anomalies in free-ranging Sturgeon (Scaphirhynchus) from the lower Mississippi River. Journal of Applied Ichthyology 25 (Supplement 2):68–74.

529

Dixon, D. A. (ed.). 2003. Biology, Management, and Protection of Catadromous Eels. American Fisheries Society Symposium 33. American Fisheries Society, Bethesda, Maryland. Doan, K. H. 1938. Observations on the dogfish (Amia calva) and their young. Copeia 1938:204. Dobzhansky, T. 1937. Genetics and the Origin of Species. 1st edition. Columbia University Press, New York. Docker, M. F. 2009. A review of the evolution of nonparasitism in Lampreys and an update of the paired species concept. American Fisheries Society Symposium 72:71–114. Docker, M. F., and F. W. H. Beamish. 1991. Growth, fecundity, and egg size of Least Brook Lamprey, Lampetra aepyptera. Environmental Biology of Fishes 31:219–227. Docker, M. F., and F. W. H. Beamish. 1994. Age, growth, and sex ratio among populations of Least Brook Lamprey, Lampetra aepyptera, larvae: an argument for environmental sex determination. Environmental Biology of Fishes 41:191–205. Docker, M. F., J. H. Youson, R. J. Beamish, and R. H. Devlin. 1999. Phylogeny of the Lamprey genus Lampetra inferred from mitochondrial cytochrome b and ND3 gene sequences. Canadian Journal of Fisheries and Aquatic Sciences 56:2340–2349. Doherty, C. A., R. A. Curry, and K. R. Munkittrick. 2010. Spatial and temporal movements of White Sucker: implications for use as a sentinel species. Transactions of the American Fisheries Society 139:1818–1827. Dolley, J. S. 1933. Preliminary notes on the biology of the St. Joseph River. American Midland Naturalist 14:193–227. Dominey, W. J. 1980. Female mimicry in male Bluegill Sunfish—a genetic polymorphism? Nature 284:546–548. Dominey, W. J. 1981. Anti-predator function of Bluegill Sunfish nesting colonies. Nature 290:586–588. Dominey, W. J. 1984. Alternative mating tactics and evolutionary stable strategies. American Zoologist 24:385–396. Donabauer, S. B., J. N. Stoeckel, and J. W. Quinn. 2009. Exploitation, survival, reproduction, and habitat use of gravid female Paddlefish in Ozark Lake, Arkansas River, Arkansas, p. 124–140. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Donahue, W. H. 1974. A karyotypic study of three species of Rajiformes (Chondrichthyes, Pisces). Canadian Journal of Genetics and Cytology 16:203–211. Donald, D. B. 1997. Relationship between year-class strength for Goldeyes and selected environmental variables during the first year of life. Transactions of the American Fisheries Society 126:361–368. Donald, D. B., and W. Aiken. 2005. Stock-yield model for a fish with variable annual recruitment. North American Journal of Fisheries Management 25:1226–1238. Donald, D. B., J. A. Babaluk, J. F. Craig, and W. A. Musker. 1992. Evaluation of the scale and operculum methods to determine age of adult Goldeyes with special reference to a dominant yearclass. Transactions of the American Fisheries Society 121:792–796. Donald, D. B., and A. H. Kooyman. 1977a. Migration and population dynamics of the Peace-Athabasca Delta Goldeye population. Occasional Paper, Canadian Wildlife Ser vice. 31:1–19. Donald, D. B., and A. H. Kooyman. 1977b. Food, feeding habits and growth of Goldeye, Hiodon alosoides (Rafinesque), in waters

530

LITERATURE CITED

of the Peace-Athabasca Delta. Canadian Journal of Zoology 55:1038–1047. Donald, D. B., and G. D. Sardella. 2010. Mercury and other metals in muscle and ovaries of Goldeye (Hiodon alosoides). Environmental Toxicology and Chemistry 29:373–379. Doosey, M., H. Bart, K. Saitoh, and M. Miya. 2009. Phylogenetic relationships of catostomid fishes (Actinopterygii: Cypriniformes) based on mitochondrial ND4/ND5 gene sequences. Molecular Phylogenetics and Evolution 54:1028–1034. Doroshov, S. I., G. P. Moberg, and J. P. Van Eenennaam. 1997. Observations on the reproductive cycle of cultured White Sturgeon, Acipenser transmontanus. Environmental Biology of Fishes 48:265–278. Dorsey, S. E., E. D. Houde, and J. C. Gamble. 1996. Cohort abundances and daily variability in mortality of eggs and yolk-sac larvae of Bay Anchovy, Anchoa mitchilli, in Chesapeake Bay. Fishery Bulletin 94:257–267. Douady, C. J., M. Dosay, M. S. Shiuji, and M. S. Stenhope. 2003. Molecular phylogenetic evidence refuting the hypotheses of Batoidea (Rays and Skates) as derived sharks. Molecular Phylogenetics and Evolution 26:215–221. Doucette, A. J., Jr. and J. M. Fitzsimons. 1988. Karyology of elopiform and clupeiform fishes. Copeia 1988:124–130. Douglas, M. E., and P. C. Marsh. 1998. Population and survival estimates of Catostomus latipinnis in northern Grand Canyon, with distribution and abundance of hybrids with Xyrauchen texanus. Copeia 1998:915–925. Douglas, M. E., P. C. Marsh, and W. L. Minckley. 1994. Indigenous fishes of western North America and the hypothesis of competitive displacement: Meda fulgida (Cyprinidae) as a case study. Copeia 1994:9–19. Douglas, M. E., M. R. Douglas, J. M. Lynch, and D. M. McElroy. 2001. Use of geometric morphometrics to differentiate Gila (Cyprinidae) within the upper Colorado River basin. Copeia 2001:389–400. Douglas, M. E., W. L. Minckley, and H. M. Tyus. 1989. Qualitiative characters, identification of Colorado River chubs (Cyprinidae: genus Gila) and the “art of seeing well”. Copeia 1989:653–662. Douglas, M. R., P. C. Brunner, and M. E. Douglas. 2003. Drought in an evolutionary context: molecular variability in Flannelmouth Sucker (Catostomus latipinnis) from the Colorado River basin of western North America. Freshwater Biology 48:1254–1273. Douglas, M. R., and M. E. Douglas. 2000. Late season reproduction by big-river Catostomidae in Grand Canyon (Arizona). Copeia 2000:238–244. Douglas, M. R., and M. E. Douglas. 2010. Molecular approaches to stream fish ecology, p. 157–195. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73, Bethesda, Maryland. Dovel, W. L. 1981. Ichthyoplankton of the lower Hudson Estuary, New York. New York Fish and Game Journal 28:21–39. Doving, K. B., E. H. Hamdani, E. Hogland, A. Kasumyan, and A. O. Tuvikene. 2005. Review of the chemical and physiological basis of alarm reactions in cyprinids, p. 133–163. In Fish Chemosenses. K. Reutter and B. G. Kapoor (eds.). Science Publishers, Inc., Enfield, New Hampshire. Dowling, T. E., R. E. Broughton, and B. D. DeMarais. 1997. Significant role for historical effects in the evolution of reproductive

isolation: evidence from patterns of introgression between the cyprinid fishes, Luxilus cornutus and Luxilus chrysocephalus. Evolution 51:1574–1583. Dowling, T. E., and B. D. DeMarais. 1993. Evolutionary significance of introgressive hybridization in cyprinid fishes. Nature 362:444–446. Dowling, T. E., and W. R. Hoeh. 1991. The extent of introgression outside the contact zone between Notropis cornutus and Notropis chrysocephalus (Teleostei: Cyprinidae). Evolution 45:944–956. Dowling, T. E., W. R. Hoeh, G. R. Smith, and W. M. Brown. 1992. Evolutionary relationships of shiners in the genus Luxilus (Cyprinidae), as determined by analysis of mitochondrial DNA. Copeia 1992:306–322. Dowling, T. E., P. C. Marsh, A. T. Kelsen, and C. A. Tibbets. 2005. Genetic monitoring of wild and repatriated populations of endangered Razorback Sucker (Xyrauchen texanus, Catostomidae, Teleostei) in Lake Mojave, Arizona-Nevada. Molecular Ecology 14:123–135. Dowling, T. E., and W. S. Moore. 1984. Level of reproductive isolation between two cyprinid fishes, Notropis cornutus and N. chrysocephalus. Copeia 1984:617–628. Dowling, T. E., and W. S. Moore. 1985a. Evidence for selection against hybrids in the family Cyprinidae (genus, Notropis). Evolution 39:152–158. Dowling, T. E., and W. S. Moore. 1985b. Genetic variation and divergence of the sibling pair of cyprinid fishes, Notropis cornutus and N. chrysocephalus. Biochemical Systematics and Ecology 13:471–476. Dowling, T. E., and G. J. P. Naylor. 1997. Evolutionary relationships of minnows in the genus Luxilus (Teleostei: Cyprinidae) as determined from cytochrome b sequences. Copeia 1997:758–765. Dowling, T. E., G. R. Smith, and W. M. Brown. 1989. Reproductive isolation and introgression between Notropis cornutus and Notropis chrysocephalus (family Cyprinidae): comparison of morphology, allozymes, and mitochondrial DNA. Evolution 43:620–634. Dowling, T. E., C. A. Tibbets, W. L. Minckley, and G. R. Smith. 2002. Evolutionary relationships of the plagopterins (Teleostei: Cyprinidae) from cytochrome b sequences. Copeia 2002:665–678. Downhower, J. F., and L. Brown. 1977. A sampling technique for benthic fish populations. Copeia 1977:403–406. Downhower, J. F., and L. Brown. 1979. Seasonal changes in the social structure of a Mottled Sculpin (Cottus bairdi) population. Animal Behaviour 27:451–458. Downhower, J. F., and L. Brown. 1980. Mate preferences of female Mottled Sculpins, Cottus bairdi. Animal Behaviour 28:728–734. Downhower, J. F., and L. Brown. 1981. The timing of reproduction and its behavioral consequences for Mottled Sculpins, Cottus bairdi, p. 78–95. In Natural Selection and Social Behavior. R. D. Alexander and D. W. Tinkle (eds.). Chiron Press, New York. Downhower, J. F., L. Brown, R. Pederson, and G. Staples. 1983. Sexual selection and sexual dimorphism in Mottled Sculpins. Evolution 37:96–103. Downhower, J. F., and D. B. Lank. 1994. Effect of previous experience on mate choice by female Mottled Sculpins. Animal Behaviour 47:369–372. Downhower, J. F., and R. Yost. 1977. The significance of male parental care in the Mottled Sculpin (Cottus bairdi). American Zoologist 17:936.

LITERATURE CITED

Doyon, C., R. Fortin, and P. A. Spear. 1999. Retinoic acid hydroxylation and teratogenesis in Lake Sturgeon (Acipenser fulvescens) from the St. Lawrence River and Abitibi region, Quebec. Canadian Journal of Fisheries and Aquatic Sciences 56:1428–1436. Drake, D. C., R. J. Naiman, and J. S. Bechtold. 2006. Fate of nitrogen in riparian forest soils and trees: an 15N tracer study simulating salmon decay. Ecology 87:1256–1266. Drake, J. A., T. E. Flum., G. J. Witteman, T. Voskuil, A. M. Hoylman, C. Creson, D. A. Kenny, G. R. Huxel, C. S. Larue, and J. R. Duncan. 1993. The construction and assembly of an ecological landscape. Journal of Animal Ecology 62:117–130. Draud, D. M., R. Macias-Ordonez, J. Verga, and M. Itzkowitz. 2004. Female and male Texas Cichlids (Herichthys cyanoguttatum) do not fight by the same rules. Behavioral Ecology 15:102–108. Drewry, G. E. 1962. Some observations of courtship behavior and sound production in five species of Fundulus. Unpubl. Master’s thesis, University of Texas, Austin. Drevnick, P. E., and M. B. Sandheinrich. 2003. Effects of dietary methylmercury on reproductive endocrinology of Fathead Minnows. Environmental Science & Technology 37:4390–4396. Dries, L. A. 2003. Peering through the looking glass at a sexual parasite: are Amazon Mollies red queens? Evolution 57:1387–1396. Driver, L. J., G. L. Adams, and S. R. Adams. 2009. Fish assemblage of a cypress wetland within an urban landscape. Southeastern Naturalist 8:527–536. Duarte-Bello, P. P., and R. J. Buesa. 1973. Catalago de peces cubanos (Primera revision). I. Indice Taxonomico, Ciencias, Serie 8 (Investigaciones Marinas no. 3). Dubois, N., D. J. Marcogliese, Magnan, P. 1996. Effects of the introduction of White Sucker, Catostomus commersoni, on the parasite fauna of Brook Trout, Salvelinus fontinalis. Canadian Journal of Zoology 74:1304–1312. Dudley, R. K., and S. P. Platania. 1999. Imitating the physical properties of drifting semiboyant fish (Cyprinidae) eggs with artificial eggs. Journal of Freshwater Ecology 14:423–430. Duff y, W. G. 1998. Population dynamics, production, and prey consumption of Fathead Minnows (Pimephales promelas) in prairie wetlands: a bioenergetics approach. Canadian Journal of Fisheries and Aquatic Sciences 55:15–27. Dugas, C. N., M. Konikoff, and M. F. Trahan. 1976. Stomach contents of Bowfin (Amia calva) and Spotted Gar (Lepisosteus oculatus) taken in Henderson Lake, Louisiana. Proceedings of the Louisiana Academy of Sciences 39:28–34. Dugatkin, L. A. 1992. Sexual selection and imitation: females copy the mate choice of others. American Naturalist 139:1384–1389. Dugatkin, L. A. 1996. The interface between culturally-based preferences and the genetic preferences: female mate choice in Poecilia reticulata. Proceedings of the National Academy of Sciences of the United States of America 93:2770–2773. Dugatkin, L. A. 1998. Genes, copying, and female mate choice; shifting thresholds. Behavioral Ecology 9:323–327. Dugatkin, L. A., and J. G. J. Godin. 1992. Reversal of female mate choice by copying in the Guppy (Poecilia reticulata). Proceedings of the Royal Society of London B 249:179–184. Dugatkin, L. A., and J. G. J. Godin. 1993. Female mate copying in the Guppy (Poecilia reticulata): age-dependent effects. Behavioral Ecology 4:289–282.

531

Dugatkin, L. A., and J. Höglund. 1995. Delayed breeding and the evolution of mate copying in lekking species. Journal of Theoretical Biology 174:261–267. Dugo, M. A., B. R. Kreiser, S. T. Ross, W. T. Slack, R. J. Heise, and B. R. Bowen. 2004. Conservation and management implications of fine-scale genetic structure of Gulf Sturgeon in the Pascagoula River, Mississippi. Journal of Applied Ichthyology 20:243–251. Dulka, J. G., N. E. Stacey, P. W. Sorensen, and G. J. van der Kraak. 1987. A steroid sex pheromone synchronizes male-female spawning readiness in Goldfish. Nature 325:251–253. Dumont, P., J. D’Amours, S. Thibodeau, N. Dubuc, R. Verdon, S. Garceau, P. Bilodeau, Y. Mailhot, and R. Fortin. 2011. Effects of the development of a newly created spawning ground in the Des Prairies River (Quebec, Canada) on the reproductive success of Lake Sturgeon (Acipenser fulvescens). Journal of Applied Ichthyology 27:394–404. Duncan, M. S., B. M. Wrege, F. M. Parauka, and J. J. Isely. 2011. Seasonal distribution of Gulf of Mexico Sturgeon in the Pensacola Bay system, Florida. Journal of Applied Ichthyology 27:316–321. Dunham, A. E., G. R. Smith, and J. N. Taylor. 1979. Evidence for ecological character displacement in western American catostomid fishes. Evolution 33:877–896. Dunstan, T. C., and J. F. Harper. 1975. Food habits of Bald Eagles in north-central Minnesota. The Journal of Wildlife Management 39:140–143. Dupuch, A., P. Magnan, and L. M. Dill. 2004. Sensitivity of Northern Redbelly Dace, Phoxinus eos, to chemical alarm cues. Canadian Journal of Zoology 82:407–415. Dupuis, H. M. C., and M. H. A. Keenleyside. 1988. Reproductive success of nesting male Longear Sunfish (Lepomis megalotis peltastes). I. Factors influencing spawning success. Behavioral Ecology and Sociobiology 23:109–116. Durham, B. W., and G. R. Wilde. 2005. Relationship between hatch date and first-summer growth of five species of prairie-stream cyprinids. Environmental Biology of Fishes 72:45–54. Durham, B. W., and G. R. Wilde. 2006. Influence of stream discharge on reproductive success of a prairie stream fish assemblage. Transactions of the American Fisheries Society 135:1644–1653. Durif, C., P. Elie, S. Dufour, J. Marchelidon, and B. Vidal. 2000. Analysis of morphological and physiological parameters during the silvering process of the European Eel (Anguilla anguilla) in the Lake of Grand-Lieu (France). Cybium 24:63–74. Dutil, J. -D., A. Giroux, A. Kemp, G. Lavoie, and J. P. Dallaire. 1988. Tidal influence on movements and on daily cycle of activity of American Eels. Transactions of the American Fisheries Society 117:488–494. Dutil, J. -D., B. Legare, and C. Desjardins. 1985. Discrimination d’un stock de poisson, l’anguille, Anguilla rostrata, basee sur le presence d’un produit chimique de synthese, le mirex. Canadian Journal of Fisheries and Aquatic Science 42:455–458. Dutil, J. -D., M. Michaud, and A. Giroux. 1989. Seasonal and diel patterns of stream invasion by American Eels (Anguilla rostrata) in the northern Gulf of St. Lawrence. Canadian Journal of Zoology 67:182–188. Dyer, B. S. 2000. Systematic review and biogeography of the freshwater fishes of Chile. Estudios Oceanológicos 19:77–98.

532 LITERATURE CITED

Dyke, A. S., J. T. Andrews, P. U. Clark, J. H. England, G. H. Miller, J. Shaw, and J. J. Veillette. 2002. The Laurentide and Innuitian ice sheets during the last glacial maximum. Quaternary Science Reviews 21:9–31. Dynesius, M., and C. Nilsson. 1994. Fragmentation and flow regulation of river systems in the northern third of the world. Science 266:753–762. Eales, J. G. 1968. The Eel fisheries of eastern Canada Fisheries Research Board of Canada Bulletin 166. Eastman, J. T. 1971. The pharyngeal bone musculature of the carp, Cyprinus carpio. Journal of Morphology 134:131–140. Eastman, J. T. 1977. The pharyngeal bones and teeth of catostomid fishes. American Midland Naturalist 97:68–88. Eastman, J. T. 1980. The caudal skeletons of catostomid fishes. American Midland Naturalist 103:133–148. Eaton, G. P. 2008. Epeirogeny in the southern Rocky Mountains region: evidence and origin. Geosphere 4:764–784. Eberhardt, L. L. and J. M. Thomas. 1991. Designing environmental field studies. Ecological Monographs 61:53–73. Echelle, A. A. 1967. The food of young-of-year Gars, Lepisosteus, in Lake Texoma with notes on spawning and development. Unpubl. Master’s thesis, University of Oklahoma, Norman. Echelle, A. A. 1968. Food habits of young-of-year Longnose Gar in Lake Texoma, Oklahoma. The Southwestern Naturalist 13:45–50. Echelle, A. A. 1973. Behavior of the pupfish, Cyprinodon rubrofluviatilis. Copeia 1973:68–76. Echelle, A. A. 1990. Nomenclature and non-Mendelian (“clonal”) vertebrates. Systematic Zoology 39:70–78. Echelle, A. A. 1991. Conservation genetics and genetic diversity in freshwater fishes of western North America, p. 141–153. In Battle Against Extinction: Native Fish Management in the American West. W. L. Minckley and J. E. Deacon (eds.). University of Arizona Press, Tucson. Echelle, A. A. 2008. The western Pupfish clade (Cyprinodontidae: Cyprinodon): mtDNA divergence times and drainage history. In Late Cenozoic Drainage History of the Southwestern Great Basin and lower Colorado River Region: Geologic and Biotic Perspectives. M. C. Reheis, R. Hershler, and D. M. Miller (eds.). Geological Society of America Special Paper 439:27–38. Echelle, A. A., T. E. Dowling, C. C. Moritz, and W. M. Brown. 1989b. Mitochondrial-DNA diversity and the origin of the Menidia clarkhubbsi complex of unisexual fishes (Atherinidae). Evolution 43:984–993. Echelle, A. A., and A. F. Echelle. 1992. Mode and pattern of speciation in the evolution of inland pupfishes in the Cyprinodon variegatus complex (Teleostei: Cyprinodontidae): an ancestordescendant hypothesis, p. 691–709. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Echelle, A. A., A. F. Echelle, and C. D. Crozier. 1983. Evolution of an all-female fish, Menidia clarkhubbsi (Atherinidae). Evolution 37:772–784. Echelle, A. A., A. F. Echelle, and D. W. Durham. 1988. Ploidy levels in silverside fishes (Atherinidae, Menidia) on the Texas coast: flow-cytometric analysis of the occurrence of allotriploidy. Journal of Fish Biology 32:835–844. Echelle, A. A., A. F. Echelle, L. E. DeBault, and D. P. Middaugh. 1989a. Evolutionary biology of the Menidia clarkhubbsi complex of unisexual fishes (Atherinidae): origins, clonal diversity, and

mode of reproduction, p. 144–152. In Evolution and Ecology of Unisexual Vertebrates. R. M. Dawley and J. P. Bogart (eds.). New York State Museum, Albany. Echelle, A. A., A. F. Echelle, and L. G. Hill. 1972. Interspecific interactions and limiting factors of abundance and distribution in the Red River Pupfish, Cyprinodon rubrofluviatilis. American Midland Naturalist 88:109–130. Echelle, A. A., and I. Kornfield. 1984. Evolution of Fish Species Flocks. University of Maine at Orono Press. Echelle, A. A., and C. D. Riggs. 1972. Aspects of the early life history of Gars (Lepisosteus) in Lake Texoma. Transactions of the American Fisheries Society 101:106–112. Eck, G. W., and L. Wells. 1987. Recent changes in Lake Michigan’s fish community and their probably causes with emphasis on the role of the Alewife (Alosa pseudoharengus). Canadian Journal of Fisheries and Aquatic Sciences 44:53–60. Eckman, S. 1953. Zoogeography of the Sea. Sidgwick and Jackson, London. Edds, D. R., W. J. Matthews, and F. P. Gelwick. 2002. Resource use by large Catfishes in a reservoir: is there evidence for interactive segregation and innate differences? Journal of Fish Biology 60:739–750. Eddy, S., and P. H. Simer. 1929. Notes on the food of the Paddlefish and the plankton of its habitat. Transactions of the Illinois Academy of Science 21:59–68. Eddy, S., and T. Surber. 1943. Northern Fishes with Special Reference to the Upper Mississippi Valley. University of Minnesota Press, Minneapolis. Eddy, S., and J. C. Underhill. 1974. Northern Fishes. North Central Publishing, St. Paul, Minnesota. Eder, S., and C. A. Carlson. 1977. Food habits of carp and White Suckers in the South Platte and St. Vrain Rivers and Goosequill Pond, Welsh County, Colorado. Transactions of the American Fisheries Society 106:339–346. Edwards, L. F. 1926. The protractile apparatus of the mouth of the catostomid fishes. Anatomical Records 33:257–270. Edwards, R. E., F. M. Parauka, and K. J. Sulak. 2007. New insights into marine migration and winter habitat of Gulf Sturgeon, p. 183–196. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Ege, V. 1939. A revision of the genus Anguilla Shaw. Dana Reports 16:8–256. Egusa, S. 1979. Notes on the culture of the European Eel (Anguilla anguilla L.) in Japanese Eel-farming ponds. Rapports et ProcèsVerbaux des Réunions du Conseil International pour l’Exploration de la Mer. 174:51–58. Ehlers, J. 1996. Quaternary and Glacial Geology. John Wiley and Sons, New York. Ehlinger, T. J., M. R. Gross, and D. P. Philipp. 1997. Morphological and growth rate differences between Bluegill males of alternative reproductive life histories. North American Journal of Fisheries Management 17:533–542. Ehrlich, P. R., and P. H. Raven. 1964. Butterflies and plants: a study in coevolution. Evolution 18:586–608. Eisenhour, D. J., and K. R. Piller. 1997. Two new intergeneric hybrids involving Semotilus atromaculatus and the genus Phoxinus with analysis of additional Semotilus atromaculatus-Phoxinus hybrids. Copeia 1997:204–209.

LITERATURE CITED

Eisenhour, D. J. 1999. Systematics of Macrhybopsis tetranema (Cypriniformes: Cyprinidae). Copeia 1999:969–980. Eisenhour, D. J. 2004. Systematics, variation, and speciation of the Macryhybopsis aetivalis complex west of the Mississippi River. Bulletin of the Alabama Museum of Natural History 23:9–48. Eizaguirre, C., S. Yeates, T. L. Lenz, M. Kalbe, and M. Milinski. 2009. MHC-based mate choice combines good genes and maintenance of MHC polymorphism. Molecular Ecology 18:3316–3329. Elder, J. F., Jr., and I. J. Schlosser. 1995. Extreme clonal uniformity of Phoxinus eos/neogaeus gynogens (Pisces: Cyprinidae) among variable habitats in northern Minnesota beaver ponds. Proceedings of the National Academy of Sciences 92:5001–5005. Electric Power Research Institute. 2001. Review and Documentation of Research and Technologies on Passage and Protection of Downstream Migrating Catadromous Eels at Hydroelectric Facilities. EPRI, Palo Alto, California. Ellerby, D. J., I. L. Y. Spierts, and J. D. Altringham. 2001. Slow muscle output of yellow-and silver-phase European Eels (Anguilla anguilla L.): changes in muscle performance prior to migration. Journal of Experimental Biology 204:1369–1379. Ellis, M. M., and G. C. Roe. 1917. Destruction of log perch eggs by Suckers. Copeia 47:69–71. Elofsson, H., B. G. Mcallister, D. E. Kime, I. Mayer, and B. Borg. 2003a. Long lasting Stickleback sperm; is ovarian fluid a key to success in fresh water? Journal of Fish Biology 63:240–253. Elofsson, H., K. van Look, B. Borg, and I. Mayer. 2003b. Influence of salinity and ovarian fluid on sperm motility in the fifteenspined Stickleback. Journal of Fish Biology 63:1429–1438. Elser, A. A. 1986. An overview of current management practices for Paddlefish fisheries, p. 62–67. In The Paddlefish: Status, Management and Propagation. J. G. Dillard, L. K. Graham, and T. R. Russell (eds.). American Fisheries Society Special Publication 7. Emanuel, M. E., and J. J. Dodson. 1979. Modification of the rheotropic behavior of male Rainbow Trout (Salmo gairdneri) by ovarian fluid. Journal of the Fisheries Research Board of Canada 36:63–68. Emery, A. R. 1978. The basis of fish community structure: marine and freshwater comparisons. Environmental Biology of Fishes 3:33–47. Endler, J. A. 1977. Geographic Variation, Speciation, and Clines. Princeton University Press, Princeton, New Jersey. Endler, J. A., and T. McLellan. 1988. The processes of evolution: towards a newer synthesis. Annual Review of Ecology and Systematics 19:395–421. Engeman, J. M., N. Aspinwall, and P. M. Mabee. 2009. Development of the pharyngeal arch skeleton in Catostomus commersonii (Teleostei: Cypriniformes). Journal of Morphology 270:291–305. Engström-öst, J., and U. Candolin. 2007. Human-induced water turbidity alters selection on sexual displays in Sticklebacks. Behavioral Ecology 18:393–398. Enquist, M., and O. Leimar. 1983. Evolution of fighting behaviordecision rules and assessment of relative strength. Journal of Theoretical Biology 102:387–410. Ensign, W. E., and E. E. Leonard. 2004. Diel habitat use by fishes in a Blue Ridge stream, Georgia. Journal of Freshwater Ecology 19:47–52. Epifanio, J. M., J. B. Koppelman, M. A. Nedbal, and D. P. Philipp. 1996. Geographic variation of Paddlefish allozymes and mito-

533

chondrial DNA. Transactions of the American Fisheries Society 125:546–561. Epifanio, J., and D. Philipp. 2001. Simulating the extinction of parental lineages from introgressive hybridization: the effects of fitness, initial proportions of parental taxa, and mate choice. Reviews in Fish Biology and Fisheries 10:339–354. Erbelding-Denk, C., J. H. Schröder, M. Schartl, I. Nanda, M. Schmid, and J. T. Epplen. 1994. Male polymorphism in Limia perugiae (Pisces: Poeciliidae). Behavior Genetics 24:95–101. Erickson, D. L., and J. E. Hightower. 2007. Oceanic distribution and behavior of Green Sturgeon, p. 197–211. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Erickson, D. L., J. A. North, J. E. Hightower, J. Weber, and L. Lauck. 2002. Movement and habitat use of Green Sturgeon Acipenser medirostris in the Rogue River, Oregon, USA. Journal of Applied Ichthyology 18:565–569. Erickson, D. L., and M. A. H. Webb. 2007. Spawning periodicity, spawning migration, and size at maturity of Green Sturgeon, Acipenser medirostris, in the Rogue River, Oregon. Environmental Biology of Fishes 79:255–268. Ernst, C. H., and E. M. Ernst. 2003. Snakes of the United States and Canada. Smithsonian Books, Washington, D.C. Eschmeyer, W. N. (ed.). 1998. Cata log of Fishes, Vol. 3. California Academy of Sciences, San Francisco. Eschmeyer, W. N., and R. Fricke (eds.). Catalog of Fishes Electronic Version. Available from http://research.calacademy.org/research /ichthyology/catalog/fishcatmain.asp; as of 30 September 2011. Eshelman, R. E. 1975. Geology and paleontology of the Early Pleistocene (Late Blancan) White Rock fauna from north-central Kansas. University of Michigan Papers on Paleontology 13:1–60. Eshenroder, R. L. 2009. Comment: mitochondrial DNA analysis indicates Sea Lampreys are indigenous to Lake Ontario. Transactions of the American Fisheries Society 138:1178–1189. Essex, H. E. 1929. The life-cycle of Haplobothrium globuliforme Cooper 1914. Science 69:677–678. Essington, T. E., and P. W. Sorenson. 1996. Overlapping sensitivities of Brook Trout and Brown Trout to putative hormonal pheromones. Journal of Fish Biology 48:1027–1029. Esteve, M. 2005. Observations of spawning behaviour in Salmoninae: Salmo, Oncorhynchus and Salvelinus. Reviews in Fisheries and Fish Biology 15:1–21. Etnier, D. A., and C. E. Skelton. 2003. Analysis of three Cisco forms (Coregonus, Salmonidae) from Lake Saganaga and adjacent lakes near the Minnesota/Ontario border. Copeia 2003: 739–749. Etnier, D. A, and W. C. Starnes. 1993. The Fishes of Tennessee. The University of Tennessee Press, Knoxville. Evans, A. N., B. J. Neilson, D. F. Markle, and S. A. Heppell. 2009. Threatened fishes of the world: Deltistes luxatus (Cope 1879) (Catostomidae). Environmental Biology of Fishes 86:401–402. Evans, A. N., J. M. Rimoldi, R. S. V. Gadepalli, and B. S. Nunez. 2010. Adaptation of a corticosterone ELISA to demonstrate sequence-specific effects of angiotensin II peptides and C-type natriuretic peptide on 1α-hydroxycorticosterone synthesis and steriodogenic mRNAs in the elasmobranch interrenal gland. Journal of Steroid Biochemistry and Molecular Biology 120: 149–154.

534

LITERATURE CITED

Evans, H. E., and J. Earl E. Duebler. 1955. Pharyngeal tooth replacement in Semotilus atromaculatus and Clinostomus elongatus, two species of cyprinid fishes. Copeia 1955:31–41. Evans, J. P., M. Pierotti and A. Pilastro. 2003. Male mating behavior and ejaculate expenditure under sperm competition risk in the Eastern Mosquitofish. Behavioral Ecology 14:268–273. Evans-White, M., W. K. Dodds, L. J. Gray, and K. M. Fritz. 2001. A comparison of the trophic ecology of the crayfishes (Orconectes nais (Faxon) and Orcontectes neglectus (Faxon)) and the Central Stoneroller minnow (Campostoma anomalum [Rafinesque]): omnivory in a tallgrass prairie stream. Hydrobiologia 462:131–144. Evernden, J. F., and G. T. James. 1964. Potassium-argon dates and the Tertiary floras of North America. American Journal of Science 262:945–974. Everett, S. R., D. L. Scarnecchia, G. J. Power, and C. J. Williams. 2003. Comparison of age and growth of Shovelnose Sturgeon in the Missouri and Yellowstone rivers. North American Journal of Fisheries Management 23:230–240. Ewers, L. A., and M. W. Boesel. 1935. The food of some Buckeye Lake fishes. Transactions of the American Fisheries Society 65:57–70. Eycleshymer, A. C., and J. M. Wilson. 1908. The adhesive organs of Amia. Biological Bulletin 14:134–148. Facey, D. E., and G. W. LaBar. 1981. Biology of American Eels in Lake Champlain, Vermont. Transactions of the American Fisheries Society 110:396–402. Facey, D. E., and M. J. Van den Avyle. 1987. Species profiles: life histories and environmental requirements of coastal fishes and invertebrates (North Atlantic)-American Eel. U.S. Fish and Wildlife Ser vice Biological Report No. 82(11.74), and U. S. Army Corps of Engineers Report No. TR EL-82-4, Washington, D.C. Fahay, M. P. 1978. Biological and fisheries data on American Eel, Anguilla rostrata (LeSueur). U.S. National Marine Fisheries Service Technical Series Report No. 17, Northeast Fisheries Center, Highlands, New Jersey. Falter, M. A., and J. J. Cech, Jr. 1991. Maximum pH tolerance of three Klamath Basin fishes. Copeia 1991:1109–1111. FAO (Food and Agricultural Organization). 2006. Fisheries global information system. Available from http://www.fao.org; as of September 2006. Fänge, R. 1986. Lymphoid organs in Sturgeons (Acipenseridae). Veterinary Immunology and Immunopathology 12:153–161. Fangue, N. A., and W. A. Bennett. 2003. Thermal tolerance responses of laboratory-acclimated and seasonally acclimatized Atlantic Stingray, Dasyatis sabina. Copeia 2003:315–325. Farabee, G. B. 1986. Fish species associated with revetted and natural main channel border habitats in pool 24 of the upper Mississippi River. North American Journal of Fisheries Management 6:504–508. Farmer, C. G. 1997. Did lungs and the intracardiac shunt evolve to oxygenate the heart in vertebrates? Paleobiology 23:358–372. Farmer, C. G., and D. C. Jackson. 1998. Air-breathing during activity in the fishes Amia calva and Lepisosteus oculatus. Journal of Experimental Biology 201:943–948. Farmer, G. J. 1980. Biology and physiology of feeding in adult Lampreys. Canadian Journal of Fisheries and Aquatic Sciences 37:1751–1761. Farmer, G. J., F. W. H. Beamish, and G. A. Robinson. 1975. Food consumption of the adult landlocked Sea Lamprey, Petromy-

zon marinus L. Comparative Biochemistry and Physiology 50A:753–757. Farmer, M. B., and S. H. Alonzo. 2008. Competition for territories does not explain allopaternal care in the Tessellated Darter. Environmental Biology of Fishes 83:391–395. Farr, J. A., and J. Travis. 1986. Fertility advertisement by female Sailfin Mollies, Poecilia latipinna (Pisces: Poeciliidae). Copeia 1986:467–472. Farr, J. A., and J. Travis. 1989. The effect of ontogenetic experience on variation in growth, maturation, and sexual-behavior in the Sailfin Molly, Poecilia latipinna (Pisces, Poeciliidae) Environmental Biology of Fishes 26:39–48. Farr, J. A., J. Travis, and J. C. Trexler. 1986. Behavioural allometry and interdemic variation in sexual behaviour of the Sailfin Molly, Poecilia latipinna (Pisces: Poeciliidae). Animal Behaviour 34:497–509. Farrell, B. C., and C. Mitter. 1993. Phylogenetic determinants of insect/plant community diversity, p. 253–266. In Species Diversity in Ecological Communities. R. E. Ricklefs and D. Schluter (eds.). University of Chicago Press, Chicago. Fast, A. W., L. H. Bottroff, and R. L. Miller. 1982. Largemouth Bass, Micropterus salmoides, and Bluegill, Lepomis macrochirus, growth rates associated with artificial destratification and Threadfin Shad, Dorosoma petenense, introductions at El Capitan Reservoir, California. California Fish and Game 68:4–20. Fausch, K. D. 1988. Tests of competition between native and introduced salmonids in streams: what have we learned? Canadian Journal of Fisheries and Aquatic Sciences 45:2238–2246. Fausch, K. D., and R. G. Bramblett. 1991. Disturbance and fish communities in intermittent tributaries of a western Great Plains river. Copeia 1991:659–674. Fausch, K. D., J. R. Karr, and P. R. Yant. 1984. Regional application of an index of biotic integrity based on stream fish communities. Transactions of the American Fisheries Society 113:39–55. Fausch, K. D., and R. J. White. 1986. Competition among juveniles of Coho Salmon, Brook Trout, and Brown Trout in a laboratory stream, and implications for Great Lakes tributaries. Transactions of the American Fisheries Society 115:363–381. Fava, J. A., and C. Tsai. 1974. The life history of the Pearl Dace, Semotilus margarita, in Maryland. Chesapeake Science 15:159–162. Felley, J. D. 1984. Multivariate identification of morphologicalenvironmental relationships within the Cyprinidae (Pisces). Copeia 1984:442–455. Felley, J. D., and L. G. Hill. 1983. Multivariate assessment of environmental preferences of cyprinid fishes of the Illinois River, Oklahoma. American Midland Naturalist 109:209–221. Felley, S. M., M. Vecchione, and S. G. F. Hare. 1987. Incidence of ectoparasitic copepods on ichthyoplankton. Copeia 1987:778–782. Ferguson, M. M., L. Bernatchez, M. Gatt, B. R. Konkle, S. Lee, M. L. Malott, and R. S. McKinley. 1993. Distribution of mitochondrial DNA variation in Lake Sturgeon (Acipenser fulvescens) from the Moose River basin, Ontario, Canada. Journal of Fish Biology 43 (Supplement A):91–101. Ferguson, M. M., and G. A. Duckworth. 1997. The status and distribution of Lake Sturgeon, Acipenser fulvescens, in the Canadian provinces of Manitoba, Ontario and Quebec: a genetic perspective. Environmental Biology of Fishes 48:299–309.

LITERATURE CITED

Fernet, D. A., and R. J. F. Smith. 1976. Agonistic behavior of captive Goldeye (Hiodon alosoides). Journal of the Fisheries Research Board of Canada 33:695–702. Ferrara, A. M. 2001. Life-history strategy of Lepisosteidae: implications for the conservation and management of Alligator Gar. Unpubl. Ph.D. diss., Auburn University, Auburn, Alabama. Ferrari, M. C. O., T. Capitania-Kwok, and D. P. Chivers. 2006. The role of learning in the acquisition of threat-sensitive responses to predator odours. Behavioral Ecology and Sociobiology 60:522–527. Ferrari, M. C. O., and D. P. Chivers. 2006. Learning threatsensitive predator avoidance: how do Fathead Minnows incorporate conflicting information? Animal Behaviour 71:19–26. Ferrari, M. C. O., J. J. Trowell, G. E. Brown, and D. P. Chivers. 2005. The role of learning in the development of threat-sensitive predator avoidance by Fathead Minnows. Animal Behaviour 70:777–784. Ferris, S. D. 1984. Tetraploidy and the evolution of the catostomid fishes, p. 55–93. In Evolutionary Genetics of Fishes. B. J. Turner (ed.). Plenum Press, New York. Ferris, S. D., D. G. Buth, and G. S. Whitt. 1982. Substantial genetic differentiation in populations of Catostomus plebeius. Copeia 1982:444–449. Ferris, S. D., and G. S. Whitt. 1977. Loss of duplicate gene expression after polyploidization. Nature 265:258–260. Ferris, S. D., and G. S. Whitt. 1978. Phylogeny of tetraploid catostomid fishes based on the loss of duplicate gene expression. Systematic Zoology 27:189–206. Ferris, S. D., and G. S. Whitt. 1979. Evolution of the differential regulation of duplicate genes after polyploidization. Journal of Molecular Evolution 12:267–317. Ferris, S. D., and G. S. Whitt. 1980. Genetic variability in species with extensive gene duplication: the tetraploid catostomid fishes. American Naturalist 115:650–666. Feyrer, F., B. Herbold, S. A. Matern, and P. B. Moyle. 2003. Dietary shifts in a stressed fish assemblage: consequences of a bivalve invasion in the San Francisco Estuary. Environmental Biology of Fishes 67:277–288. Ficke, A. D., C. A. Myrick, and L. J. Hansen. 2007. Potential impacts of global change on freshwater fisheries. Reviews in Fish Biology and Fisheries 17:581–613. Findeis, E. K. 1993. Skeletal anatomy of the North American Shovelnose Sturgeon Scaphirhynchus platorynchus (Rafinesque 1820) with comparisons to other Acipenseriformes. Unpubl. Ph.D. diss., University of Massachusetts, Amherst. Findeis, E. K. 1997. Osteology and interrelationships of Sturgeons (Acipenseridae). Environmental Biology of Fishes 48:73–126. Findlay, C. S., D. G. Bert, and L. Zheng. 2000. Effect of introduced piscivores on native minnow communities in Adirondack lakes. Canadian Journal of Fisheries and Aquatic Sciences 57:570–580. Fine, J. M., and P. W. Sorensen. 2010. Production and fate of the Sea Lamprey migratory pheromone. Fish Physiology and Biochemistry 36:1013–1020. Fine, J. M., L. A. Vrieze, and P. W. Sorensen. 2004. Evidence that petromyzontid Lampreys employ a common migratory pheromone that is partially comprised of bile acids. Journal of Chemical Ecology 30:2091–2110. Fine, M. L., H. E. Winn, and B. L. Olla. 1977. Communication in fishes, p. 472–518. In How Animals Communicate. T. A. Sebeok (ed.). Indiana University Press, Bloomington.

535

Fineran, B. A., and J. A. C. Nicol. 1976. Novel cones in the retina of the Anchovy (Anchoa). Journal of Ultrastructure Research 54:296–303. Fineran, B. A., and J. A. C. Nicol. 1977. Studies on the eyes of Anchovies, Anchoa mitchilli and A. hepsetus (Engraulidae) with particular reference to the pigment epithelium. Philosophical Transactions of the Royal Society of London, Series B, Biological Sciences 276:321–350. Fineran, B. A., and J. A. C. Nicol. 1978. Studies on the photoreceptors of Anchoa mitchilli and A. hepsetus (Engraulidae) with particular reference to the cones. Philosophical Transactions of the Royal Society of London, Series B, Biological Sciences 283:25–60. Fink, S. V., and W. L. Fink. 1981. Interrelationships of the ostariophysan fishes (Teleostei). Zoological Journal of the Linnean Society 72:297–353. Fink, W., and J. Humphries. 2010. Morphological description of the extinct North American Sucker Moxostoma lacerum (Ostariophysi, Catostomidae), based on high-resolution x-ray computed tomography. Copeia 2010:5–13. Finlay, J. C., J. M. Hood, M. P. Limm, M. E. Power, J. D. Schade, and J. R. Welter. 2011. Light-mediated thresholds in stream-water nutrient composition in a river network. Ecology 92:140–50. Firehammer, J. A., and D. L. Scarnecchia. 2006. Spring migratory movements by Paddlefish in natural and regulated river segments of the Missouri and Yellowstone rivers, North Dakota and Montana. Transactions of the American Fisheries Society 135:200–217. Firehammer, J. A., and D. L. Scarnecchia. 2007. The influence of discharge on duration, ascent distance, and fidelity of the spawning migration for Paddlefish of the YellowstoneSakakawea stock, Montana and North Dakota, USA. Environmental Biology of Fishes 78:23–36. Firehammer, J. A., D. L. Scarnecchia, and S. R. Fain. 2006. Modification of a passive gear to sample Paddlefish eggs in sandbed spawning reaches of the lower Yellowstone River. North American Journal of Fisheries Management 26:63–72. Fischer, C., and I. Schlupp. 2009. Differences in thermal tolerance in coexisting sexual and asexual mollies (Poecilia, Poeciliidae, Teleostei). Journal of Fish Biology 74:1662–1668. Fish, M. P. 1932. Contributions to the early life histories of sixtytwo species of fishes from Lake Erie and its tributary waters. Bulletin of the United States Bureau of Fisheries 47:293–398. Fish, M. P., and W. H. Mowbray. 1970. Sounds of Western North Atlantic Fishes. Johns Hopkins Press, Baltimore, Maryland. Fishbase. 2012. Fishbase: a global information system on fishes (ver. 02/2012). Available from http://www.fishbase.us/search. php; as of February 2012. Fisher, H. S., S. J. Mascuch, and G. G. Rosenthal. 2009. Multivariate male traits misalign with multivariate female preferences in the swordtail fish, Xiphophorus birchmanni. Animal Behaviour 78:265–269. Fisher, H. S., and G. G. Rosenthal. 2006. Female swordtails use chemical cues to select well-fed mates. Animal Behaviour 72:721–725. Fisher, H. S., and G. G. Rosenthal. 2010. Relative abundance of Xiphophorus fishes and its effect on sexual communication. Ethology 116:32–38. Fisher, H. S., B. B. M. Wong, and G. G. Rosenthal. 2006. Alteration of the chemical environment disrupts communication in a

536

LITERATURE CITED

freshwater fish. Proceedings of the Royal Society of London B 273:1187–1193. FOC (Fisheries and Oceans Canada) 2010. Fisheries and Oceans Canada. Available from http://www.dfo-mpo.gc.ca; as of 1 March 2010. Fitz, R. B. 1966. Unusual food of a Paddlefish (Polyodon spathula) in Tennessee. Copeia 1966:356. FitzGerald, G. J. 1991. The role of cannibalism in the reproductive ecology of the Threespine Stickleback. Ethology 89:177–194. FitzGerald, G. J. 1992. Filial cannibalism in fishes: why do parents eat their offspring? Trends in Ecology and Evolution 7:7–10. FitzPatrick, J. L., and N. R. Liley. 2008. Ejaculate expenditure and timing of gamete release in Rainbow Trout Oncorhynchus mykiss. Journal of Fish Biology 73:262–274. Fiumera, A. C., B. A. Porter, G. D. Grossman, and J. C. Avise. 2002. Intensive genetic assessment of the mating system and reproductive success in a semi-closed population of the Mottled Sculpin, Cottus bairdi. Molecular Ecology 11:2367–2377. Fives, J. M., S. M. Warlen, and D. E. Hoss. 1986. Aging and growth of larval Bay Anchovy, Anchoa mitchilli, from the Newport River Estuary, North Carolina. Estuaries 9:362–367. Flamarique, I. N., and F. I. Harosi. 1997. Photoreceptor morphology and visual pigment content in the retina of the common White Sucker (Catostomus commersoni). Biological Bulletin 193:209–210. Flamarique, I. N., and C. W. Hawryshyn. 1998. The common White Sucker (Catostomus commersoni): a fish with ultraviolet sensitivity that lacks polarization sensitivity. Journal of Comparative Physiology. A, Neuroethology, Sensory, Neural, and Behavioral Physiology 182:331–341. Flamarique, I. N., G. A. Mueller, C. L. Cheng, and C. R. Figiel. 2007. Communication using eye-roll reflective signalling. Proceedings of the Royal Society of London, Series B 274:877–882. Flecker, A. S., P. B. McIntyre, J. W. Moore, J. T. Anderson, B. W. Taylor, and R. O. Hall, Jr. 2010. Migratory fishes as material and process subsidies in riverine ecosystems, p. 559–592. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. Flemer, D. A., and W. S. Woolcott. 1966. Food habits and distribution of the fishes of Tuckahoe Creek, Virginia, with special emphasis on the Bluegill, Lepomis machrochirus Rafinesque. Chesapeake Science 7:75–89. Fleming, I. A. 1996. Reproductive strategies of Atlantic Salmon: ecology and evolution. Reviews in Fish Biology and Fisheries 6:379–416. Fleming, I. A. 1998. Pattern and variability in the breeding system of Atlantic Salmon (Salmo salar), with comparisons to other salmonids. Canadian Journal of Fisheries and Aquatic Sciences Suppl. 1 55:59–76. Fleming, I. A., and M. R. Gross. 1994. Breeding competition in a Pacific salmon (Coho: Oncorhynchus kisutch): measures of natural and sexual selection. Evolution 48:637–657. Fleming, I. A., B. Jonsson, M. R. Gross, and A. Lamberg. 1996. An experimental study of the reproductive behaviour and success of farmed and wild Atlantic Salmon (Salmo salar). Journal of Applied Ecology 33:893–905. Fleming, I. A., A. Lamberg, and B. Jonsson. 1997. Effects of early experience on the reproductive performance of Atlantic Salmon. Behavioral Ecology 8:470–480.

Fletcher, D. E. 1993. Nest association of Dusky Shiners (Notropis cummingsae) and Redbreast Sunfish (Lepomis auritus), a potentially parasitic relationship. Copeia 1993:159–167. Fletcher, D. E., and B. M. Burr. 1992. Reproductive biology, larval description, and the diet of the North American Bluehead Shiner, Pteronotropis hubbsi (Cypriniformes: Cyprinidae), with comments on conservation status. Ichthyological Exploration of Freshwaters 3:193–218. Fletcher, D. E., E. E. Dakin, B. A. Porter, and J. C. Avise. 2004. Spawning behavior and genetic parentage in the Pirate Perch (Aphredoderus sayanus), a fish with an enigmatic reproductive morphology. Copeia 2004:1–10. Flint, R. F. 1971. Glacial and Quaternary geology. John Wiley and Sons, Inc., New York. Flores-Coto, C., F. Barba-Torres, and J. Sánchez-Robles. 1983. Seasonal diversity, abundance, and distribution of ichthyoplankton in Tamiahua Lagoon, western Gulf of Mexico. Transactions of the American Fisheries Society 112:247–256. Font, W. F., and K. C. Corkum. 1975. Alloglossidium renale n. sp. (Digenea: Macroderoididae) from the fresh-water shrimp and A. progeneticum n. comb. Transactions of the American Microscopical Society 94:421–424. Fontaine, Y. A., M. Pisam, C. Le Moal, and A. Rambourg. 1995. Silvering and gill “mitochondria rich” cells in the Eel, Anguilla anguilla. Cell and Tissue Research 281:465–471. Fontana, F. 2002. A cytogenetic approach to the study of taxonomy and evolution in Sturgeons. Journal of Applied Ichthyology 18:226–233. Fontana, F., J. Tagliavini, and L. Congiu. 2001. Sturgeon genetics and cytogenetics: recent advancements and perspectives. Genetica 111:359–373. Fontana, F., R. M. Bruch, F. P. Binkowski, M. Lanfredi, M. Chicca, N. Beltrami, and L. Congiu. 2004. Karyotype characterization of the Lake Sturgeon, Acipenser fulvescens (Rafinesque 1817) by chromosome banding and fluorescent in situ hybridization. Genome 47:1–5. Foote, C. J. 1988. Male mate choice dependent on the male size in salmon. Behaviour 106:63–80. Foote, C. J. 1990. An experimental comparison of male and female spawning territoriality in a Pacific salmon. Behaviour 115: 283–314. Foote, C. J., G. S. Brown, and C. C. Wood. 1997. Spawning success of males using alternative mating tactics in Sockeye Salmon, Oncorhynchus nerka. Canadian Journal of Fisheries and Aquatic Sciences 54:1785–1795. Foote, C. J., J. W. Clayton, C. C. Lindsey, and R. A. Bodaly. 1992. Evolution of Lake Whitefish (Coregonus clupeaformis) in North America during the Pleistocene: evidence for a Nahanni glacial refuge race in the Northern Cordillera region. Canadian Journal of Fisheries and Aquatic Sciences 49:760–767. Foote, C. J., and P. A. Larkin. 1988. The role of male choice in the assortative mating of anadromous and non-anadromous salmon (Oncorhynchus nerka). Behaviour 106:43–62. Foran, C. M., and M. J. Ryan. 1994. Female-female competition in a unisexual/bisexual complex of Mollies. Copeia 1994:504–508. Forbes, S. A. 1878. The food of Illinois fishes. Bulletin of the Illinois State Laboratory of Natural History 1:71–89. Forbes, S. A. 1888. On the food relations of fresh-water fishes: a summary and discussion. Bulletin of the Illinois State Laboratory of Natural History 2:475–538.

LITERATURE CITED

Forbes, S. A., and R. E. Richardson. 1909. The Fishes of Illinois. 2nd edition. Illinois Natural History Survey, Urbana. Forbes, S. A., and R. E. Richardson. 1920. The Fishes of Illinois. 2nd edition. Illinois State Journal Co., State Printers, Springfield, Illinois. Ford, T. E., and E. Mercer. 1986. Density, size distribution and home range of American Eels, Anguilla rostrata in a Massachusetts salt marsh. Environmental Biology of Fishes 17:309–314. Forey, P. L. 1995. Agnathans recent and fossil, and the origin of jawed vertebrates. Reviews in Fish Biology and Fisheries 5:267–303. Forey, P. L., and P. Janvier. 1993. Agnathans and the origin of jawed vertebrates. Nature 361:129–134. Forey, P. L., D. T. Littlefoot, P. Ritchie, and A. Meyer. 1996. Interrelationships of elopomorph fishes, p. 175–191. In Interrelationships of Fishes. M. L. J. Stiassny, L. R. Parenti and G. D. Johnson (eds.). Academic Press, San Diego, California. Forlano, P. M., K. P. Maruska, S. A. Sower, J. A. King, and T. C. Tricas. 2000. Differential distribution of gonadotropin-releasing hormone-immunoreactive neurons in the Stingray brain: functional and evolutionary considerations. General and Comparative Endocrinology 118:226–248. Forrest, T. G., G. I. Miller, and J. R. Zagar. 1993. Sound propagation in shallow water: implications for acoustic communication by aquatic animals. Bioacoustics 4:259–270. Fortin, R., P. Dumont, and S. Guénette. 1996. Determinants of growth and body condition of Lake Sturgeon (Acipenser fulvescens). Canadian Journal of Fisheries and Aquatic Sciences 53:1150–1156. Foster, N. R. 1967. Comparative studies on the biology of Killifishes (Pisces, Cyprinodontidae). Unpubl. Ph.D. diss., Cornell University, Ithaca. Foster, S. 1990. Courting disaster in cannibal territory. Natural History Novitates 1990:52–61. Foster, S. A., J. A. Baker, and M. A. Bell. 2003. The case for conserving Threespine Stickleback populations. Fisheries 28:10–18. Fowler, H. W. 1912. Some features of ornamentation in freshwater fishes. The American Naturalist 46:470–476. Fowler, H. W. 1926. Fishes from Florida, Brazil, Bolivia, Argentina, and Chile. Proceedings of the Academy of Natural Sciences Philadelphia 78:249–285. Fowler, J. F., and C. A. Taber. 1985. Food habits and feeding periodicity in two sympatric stonerollers (Cyprinidae). American Midland Naturalist 113:217–224. Fox, B. J. 2001. The genesis and development of guild assembly rules, p. 23–57. In Ecological Assembly Rules. E. Weiher and P. Keddy (eds.). Cambridge University Press, Cambridge, United Kingdom. Fox, D. A., J. E. Hightower, and F. M. Parauka. 2000. Gulf Sturgeon spawning migration and habitat in the Choctawhatchee River system, Alabama-Florida. Transactions of the American Fisheries Society 129:811–826. Fox, D. A., J. E. Hightower, and F. M. Parauka. 2002. Estuarine and nearshore marine habitat use by Gulf Sturgeon from the Choctawhatchee River system, Florida, p. 111–126. In Biology, Management, and Protection of North American Sturgeon. W. Van Winkle, P. Anders, D. H. Secor, and D. Dixon (eds.). American Fisheries Society Symposium 28, Bethesda, Maryland. Fraker, M. E., J. W. Snodgrass, and F. Morgan. 2002. Differences in growth and maturation of Blacknose Dace (Rhinichthys

537

atratulus) across an urban-rural gradient. Copeia 2002: 1122–1127. Francis, D. R., D. J. Jude, and J. A. Barres. 1998. Mercury distribution in the biota of a Great Lakes estuary: Old Woman Creek, Ohio. Journal of Great Lakes Research 24:595–607. Fraser, D. F., and T. E. Sise. 1980. Observations on stream minnows in a patchy environment: a test of a theory of habitat distribution. Ecology 61:790–797. Fraser, D. F., and R. D. Cerri. 1982. Experimental evaluation of predator-prey relationships in a patchy environment: consequences for habitat use patterns in minnows. Ecology 63: 307–313. Fraser, D. F., D. A. DiMattia, and J. D. Duncan. 1987. Living among predators: the response of a stream minnow to the hazard of predation, p. 121–127. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Fraser, D. F., and E. E. Emmons. 1984. Behavioral response of Blacknose Dace (Rhinichthys atratulus) to varying densities of predatory Creek Chub (Semotilus atromaculatus). Canadian Journal of Fisheries and Aquatic Sciences 41:364–370. Fraser, G. A., and H. H. Harvey. 1984. Effects of environmental pH on the ionic composition of the White Sucker (Catostomus commersoni) and Pumpkinseed (Lepomis gibbosus). Canadian Journal of Zoology 62:249–259. Fredericks, J. P., and D. L. Scarnecchia. 1997. Use of surface visual counts for estimating relative abundance of age-0 Paddlefish in Lake Sakakawea. North American Journal of Fisheries Management 17:1014–1018. Freeman, M. C., M. K. Crawford, J. C. Barrett, D. E. Facey, M. G. Flood, J. Hill, D. J. Stouder, and G. D. Grossman. 1988. Fish assemblage stability in a southern Appalachian stream. Canadian Journal of Fisheries and Aquatic Sciences 45:1949–1958. Freeman, M. C., and G. D. Grossman. 1992. A field test for competitive interactions among foraging stream fishes. Copeia 1992:898–902. Freeman, M. C., and G. D. Grossman. 1993. Effects of habitat availability on dispersion of a stream cyprinid. Environmental Biology of Fishes 37:121–130. Freeman, M. C., and D. J. Stouder. 1989. Intraspecific interactions influence size specific depth distribution in Cottus bairdi. Environmental Biology of Fishes 24:231–236. French, W. E., B. D. S. Graeb, S. R. Chipps, K. N. Betrand, T. M. Selch, and R. A. Klumb. 2010. Vulnerability of age-0 Pallid Sturgeon Scaphirhynchus albus to fish predation. Journal of Applied Ichthyology 26:6–10. Fretwell, S. D. 1972. Populations in a seasonal environment. Princeton University Press, Princeton, New Jersey. Freund, J. A., L. Schimansky-Geier, B. Beisner, A. Neiman, D. F. Russell, T. Yakusheva, and F. Moss. 2002. Behavioral stochastic resonance: how the noise from a Daphnia swarm enhances individual prey capture by juvenile Paddlefish. Journal of Theoretical Biology 214:71–83. Frick, N. T., J. S. Bystriansky, and J. S. Ballantyne. 2007. The metabolic organization of a primitive air-breathing fish, the Florida Gar (Lepisosteus platyrhincus). Journal of Experimental Zoology 307A:7–17. Fricke, D. 1987. Reaction to alarm substance in cave populations of Astyanax fasciatus (Characidae, Pisces). Ethology 76:305–308.

538

LITERATURE CITED

Friedland, K. D., M. J. Miller, and B. Knights. 2007. Oceanic changes in the Sargasso Sea and declines in recruitment of the European Eel. ICES Journal of Marine Science 64:519–530. Friesen, R. G., and D. P. Chivers. 2006. Underwater video reveals strong avoidance of chemical cues by prey fishes. Ethology 112: 339–345. Frimpong, E. A., and P. L. Angermeier. 2010. Trait-based approaches to the analysis of stream fish communities, p. 109–136. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. Fritts, A. L., and T. N. Pearsons. 2006. Effects of predation by nonnative Smallmouth Bass on native salmonid prey: the role of predator and prey size. Transactions of the American Fisheries Society 135:853–860. Frommen, J. G., and T. C. M. Bakker. 2006. Inbreeding avoidance through non-random mating in Sticklebacks. Biology Letters 2:232–235. Frost, D. R., and J. W. Wright. 1988. The taxonomy of uniparental species, with special reference to parthenogenetic Cnemidophorus (Squamata: Teiidae). Systematic Zoology 37:200–209. Fry, B., P. L. Mumford, F. Tam, D. D. Fox, G. L. Warren, K. E. Havens, and A. D. Steinman. 1999. Trophic position and individual feeding histories of fish from Lake Okeechobee, Florida. Canadian Journal of Fisheries and Aquatic Sciences 56:590– 600. Frye, J. C., H. B. Willman, R. F. Black. 1965. Outline of glacial geology of Illinois and Wisconsin, p. 43–61. In The Quaternary of the United States. H. E. Wright, Jr., and D. G. Frey (eds.). Princeton University Press, Princeton, New Jersey. Fryer, G., and T. D. Iles. 1972. The Cichlid Fishes of the Great Lakes of Africa: Their Biology and Evolution. Oliver Boyd, Edinburgh. Fu, P., B. D. Neff, and M. R. Gross. 2001. Tactic-specific success in sperm competition. Proceedings of the Royal Society of London B 268:1105–1112. Fuiman, L. A. 1979. Descriptions and comparisons of catostomid fish larvae: northern Atlantic drainage species. Transactions of the American Fisheries Society 108:560–603. Fuiman, L. A. 1982. Family Catostomidae, Suckers. Identification of larval fishes of the Great Lakes basin with emphasis on the Lake Michigan drainage, p. 345–435. In Great Lakes Fishery Commision Special Publication, Ann Arbor, Michigan. Fuiman, L. A., J. V. Conner, B. F. Lathrop, G. R. Buynak, D. E. Snyder, J. J. Loos. 1983. State of the art of identification for cyprinid fish larvae from Eastern North America. Transactions of the American Fisheries Society 112:319–332. Fuiman, L. A., and D. C. Witman. 1979. Descriptions and comparisons of catostomid fish larvae: Catostomus catostomus and Moxostoma erythrurum. Transactions of the American Fisheries Society 108:604– 619. Fukayama, S., and H. Takahashi. 1983. Sex differentiation and development of the gonad in the Sand Lamprey, Lampetra reissneri. Bulletin of the Faculty of Fisheries, Hokkaido University 34:279–290. Fuller, P., and L. Nico. 2011. USGS Nonindigenous Aquatic Species Database. Gainesville, Florida, Revision Date: 8/23/2004. Available from http://nas.er.usgs.gov/queries/factsheet.aspx ?SpeciesID=310; as of February 2011.

Fuller, P. L., L. G. Nico, and J. D. Williams. 1999. Nonindigenous Fishes Introduced into Inland Waters of the United States. American Fisheries Society Special Publication 27, Bethesda, Maryland. Fuller, R. C. 2001. Patterns in male breeding behaviors in the Bluefin Killifish, Lucania goodei: a field study (Cyprinodontiformes: Fundulidae). Copeia 2001:823–828. Fuller, R. C. 2002. Lighting environment predicts the relative abundance of male colour morphs in Bluefin Killifish (Lucania goodei) populations. Proceedings of the Royal Society of London B 269:1457–1465. Fuller, R. C. 2003. Disentangling female mate choice and male competition in the Rainbow Darter, Etheostoma caeruleum. Copeia 2003:138–148. Fuller, R. C., K. L. Carleton, J. M. Fadool, T. C. Spady, and J. Travis. 2004. Population variation in opsin expression in the Bluefin Killifish, Lucania goodei: a real-time PCR study. Journal of Comparative Physiology A 190:147–154. Fuller, R. C., K. L. Carleton, J. M. Fadool, T. C. Spady, and J. Travis. 2005. Genetic and environmental variation in the visual properties of Bluefin Killifish, Lucania goodie. Journal of Evolutionary Biology 18:516–523. Fuller, R. C., L. J. Fleishman, M. Leal, J. Travis, and E. Loew. 2003. Intraspecific variation in ultraviolet cone production and visual sensitivity in the Bluefin Killifish, Lucania goodei. Journal of Comparative Physiology A 189:609–616. Fuller, R. C., and A. M. Johnson. 2009. A test for negative frequency-dependent mating success as a function of male colour pattern in the Bluefin Killifish. Biological Journal of the Linnean Society 98:489–500. Fuller, R. C., and J. Travis. 2001. A test for male parental care in a fundulid, the Bluefin Killifish, Lucania goodei. Environmental Biology of Fishes 6:419–426. Fuller, R. C., and J. Travis. 2004. Genetics, lighting environment, and heritable responses to lighting environment affect color morph expression in Bluefin Killifish, Lucania goodei. Evolution 58:1086–1098. Funderburg, J. B., and M. L. Gilbert. 1963. Observations on a probable new race of the Bowfin, Amia calva, from central Florida. Southeastern Biologists Bulletin 10:28. Funicelli, N. A. 1975. Taxonomy, feeding, limiting factors, and sex ratios of Dasyatis sabina, Dasyatis americana, Dasyatis sayi, and Narcine brasiliensis. Unpubl. Ph.D. diss., University of Southern Mississippi, Hattiesburg. Futuyma, D. J., and M. Slatkin (eds.). 1983. Coevolution. Sinauer Associates, Sunderland, Massachusetts. Gabor, C. R. 1999. Association patterns of Sailfin Mollies (Poecilia latipinna): alternative hypotheses. Behavioral Ecology and Sociobiology 46:333–340. Gabor, C. R., R. Gonzalez, M. Parmley, and A. S. Aspbury. 2010. Variation in male Sailfin Molly preference for female size: does sympatry with sexual parasites drive preference for smaller conspecifics? Behavioral Ecology and Sociobiology 64:783–792. Gabor, C. R., and M. J. Ryan. 2001. Character displacement in Sailfin Mollies, Poecilia latipinna: allozymes and behavior. Environmental Biology of Fishes 73:75–88. Gadomski, D. M., and C. A. Barfoot. 1998. Diel and distributional abundance patterns of fish embryos and larvae in the lower Columbia and Deschutes rivers. Environmental Biology of Fishes 51:353–368.

LITERATURE CITED

Gadomski, D. M., and M. J. Parsley. 2005a. Effects of turbidity, light level, and cover on predation of White Sturgeon larvae by Prickly Sculpins. Transactions of the American Fisheries Society 134:369–374. Gadomski, D. M., and M. J. Parsley. 2005b. Vulnerability of young White Sturgeon, Acipenser transmontanus, to predation in the presence of alternative prey. Environmental Biology of Fishes 74:389–396. Gage, M. J. G., P. Stockley, and G. A. Parker. 1995. Effects of alternative male mating strategies on characteristics of sperm production in the Atlantic Salmon (Salmo salar): theoretical and empirical investigations. Philosophical Transactions of the Royal Society of London B 350:391–399. Gale, W. F. 1983. Fecundity and spawning frequency of caged Bluntnose Minnows-fractional spawners. Transactions of the American Fisheries Society 112:398–402. Gale, W. F. 1986. Indeterminate fecundity and spawning behavior of captive Red Shiners-fractional, crevice spawners. Transactions of the American Fisheries Society 115:429–437. Gale, W. F., and C. A. Gale. 1977. Spawning habits of Spotfin Shiner (Notropis spilopterus) a fractional, crevice spawner. Transactions of the American Fisheries Society 106:170–177. Gammon, D. B., W. Li, A. P. Scott, B. S. Zielinski, and L. D. Corkum. 2005. Behavioural responses of female Neogobius melanostomus to odours of conspecifics. Journal of Fish Biology 67: 615–626. Gammon, J. R. 1973. The effect of thermal input on the populations of fish and macroinvertebrates in the Wabash River. Purdue University Water Resources Center, Technical Report 32. Ganzhorn, J., J. S. Rohovec, and J. L. Fryer. 1992. Dissemination of microbial pathogens through introductions and transfers of finfish, p. 175–192. In Dispersal of Living Organisms into Aquatic Ecosystems. A. Rosenfield and R. Mann (eds.). University of Maryland Sea Grant College Program, College Park, Maryland. Gao, Z., Y. Li, and W. Wang. 2008. Threatened fishes of the world: Myxocyprinus asiaticus Bleeker 1864 (Catostomidae). Environmental Biology of Fishes 83:345–346. Garant, D., J. J. Dodson, and L. Bernatchez. 2001. A genetic evaluation of mating system and determinants of individual reproductive success in Atlantic Salmon (Salmo salar L.). Journal of Heredity 92:137–145. García, C., E. Rolanalvarez, and L. Sanchez. 1992. Alarm reaction and alert state in Gambusia affinis (Pisces, Poeciliidae) in response to chemical stimuli from injured conspecifics. Journal of Ethology 10:41–46. Garcia de León, F. J., L. Gonzalez-Garcia, J. M. Herrera-Castillo, K. O. Winemiller, and A. Banda-Valdes. 2001. Ecology of the Alligator Gar, Atractosteus spatula, in the Vicente Guerrero Reservoir, Tamaulipas, Mexico. The Southwestern Naturalist 46:151–157. Gard, M. F. 2005. Ontogenetic microhabitat shifts in Sacramento Pikeminnow, Ptychocheilus grandis: reducing intraspecific predation. Aquatic Ecology 39:229–235. Gard, R., and G. A. Flittner. 1974. Distribution and abundance of fishes in Sagehen Creek, California. Journal of Wildlife Management 38:347–358. Gardiner, B. G. 1984. Sturgeons as living fossils, p. 148–152. In Living Fossils. N. Eldredge and S. M. Stanley (eds.). SpringerVerlag, New York.

539

Gardiner, B. G., D. T. J. Littlewood, and J. G. Maisey. 1996. Interrelationships of basal neopterygians, p. 117–146. In Interrelationships of Fishes. M. L. J. Stiassny, L. R. Parenti, and G. D. Johnson (eds.). Academic Press, San Diego, California. Garrigan, D., P. C. Marsh, and T. E. Dowling. 2002. Long-term effective population size of three endangered Colorado River fishes. Animal Conservation 5:95–102. Gasaway, C. R. 1970. Changes in the fish population in Lake Francis Case in South Dakota in the first 16 years of impoundment. Technical Papers of the Bureau of Sports Fisheries and Wildlife 56. Gatz, A. 1979. Ecological morphology of freshwater stream fishes. Tulane Studies in Zoology and Botany 21:91–124. Gazdewich, K. J., and D. P. Chivers. 2002. Acquired predator recognition by Fathead Minnows: influence of habitat characteristics on survival. Journal of Chemical Ecology 28:439–445. Geen, G. H., T. G. Northcote, G. F. Hartman, and C. C. Lindsey. 1966. Life histories of two species of catostomid fishes in Sixteenmile Lake, B.C., with particular reference to inlet stream spawning. Journal of the Fisheries Research Board of Canada 23:1761–1788. Geheber, A. D., and K. R. Piller. 2012. Spatio-temporal patterns of fish assemblage structure in a coastal plain stream: appropriate scales reveal historic tales. Ecology of Freshwater Fish 21:627–639. Gelsleichter, J., C. J. Walsh, N. J. Szabo, and L. E. L. Rasmussen. 2006. Organochlorine concentrations, reproductive physiology, and immune function in unique populations of freshwater Stringrays (Dasyatis sabina) from Florida’s St. Johns River. Chemosphere 63:1506–1522. Gelwick, F. P. 1990. Longitudinal and temporal comparisons of riffle and pool fish assemblages in a northeastern Oklahoma stream. Copeia 1990:1072–1082. Gelwick, F. P., and W. J. Matthews. 1992. Effects of an algivorous minnow on temperate stream ecosystem properties. Ecology 73:1630–1645. Gemballa, S., and P. Bartsch. 2002. Architecture of the integument in lower teleostomes: functional morphology and evolutionary implications. Journal of Morphology 253:290–309. Gende, S. M., A. E. Miller, and E. Hood. 2007. The effects of salmon carcasses on soil nitrogen pools in a riparian forest of southeastern Alaska. Canadian Journal of Forest Research 37:1194–1202. Gengerke, T. W. 1986. Distribution and abundance of Paddlefish in the United States, p. 22–35. In The Paddlefish: Status, Management and Propagation. J. G. Dillard, L. K. Graham, and T. R. Russell (eds.). American Fisheries Society Special Publication 7. George, A. L., B. R. Kuhajda, J. D. Williams, M. Cantrell, P. L. Rakes, and J. R. Shute. 2009. Guidelines for propagation and translocation for freshwater fish conservation. Fisheries 34: 529–545. Gerald, J. W. 1971. Sound production during courtship in six species of Sunfish (Centrarchidae). Evolution 25:75–87. Gerber, A. S., C. A. Tibbets, and T. E. Dowling. 2001. The role of introgressive hybridization in the evolution of the Gila robusta complex (Teleostei: Cyprinidae). Evolution 55:2028–2039. Gerken, J. E., and C. P. Paukert. 2009. Threats to Paddlefish habitat: implications for conservation, p. 173–183. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert

540

LITERATURE CITED

and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. German, D. P., B. C. Nagle, J. M. Villeda, A. M. Ruiz, A. W. Thomson, S. C. Balderas, and D. H. Evans. 2010. Evolution of herbivory in a carnivorous clade of minnows (Teleostei: Cyprinidae): effects on gut size and digestive physiology. Physiological and Biochemical Zoology 83:1–18. Gerrity, P. C., C. S. Guy, and W. M. Gardner. 2006. Juvenile Pallid Sturgeon are piscivorous: a call for conserving native cyprinids. Transactions of the American Fisheries Society 135:604–609. Gerrity, P. C., C. S. Guy, and W. M. Gardner. 2008. Habitat use of juvenile Pallid Sturgeon and Shovelnose Sturgeon with implications for water-level management in a downstream reservoir. North American Journal of Fisheries Management 28:832–843. Gershanovich, A. D. 1983. Factors determining variations in growth rate and size distribution in groups of young Paddlefish, Polyodon spathula (Polyodontidae). Journal of Ichthyology 23:56–61. Gess, R. W., M. I. Coates, and B. S. Rubidge. 2006. A Lamprey from the Devonian period of South Africa. Nature 443: 981–984. Gessner, J., S. Würtz, F. Kirschbaum, and M. Wirth. 2008. Biochemical composition of caviar as a tool to discriminate between aquaculture and wild origin. Journal of Applied Ichthyology 24(Supplement 1):52–56. Gibson, A. K., and A. Mathis. 2006. Opercular beat rate for Rainbow Darters, Etheostoma caeruleum exposed to chemical stimuli from conspecific and heterospecific fishes. Journal of Fish Biology 69:224–232. Gibson, J. R., and J. N. Fries. 2005. Culture studies of the Devils River Minnow. North American Journal of Aquaculture 67: 294–303. Gibson, R. M., and J. Höglund. 1992. Copying and sexual selection. Trends in Ecology and Evolution 7:229–232. Gido, K. B. 2001. Feeding ecology of three omnivorous fishes in Lake Texoma (Oklahoma-Texas). Southwestern Naturalist 46:23–33. Gido, K. B., K. N. Bertrand, J. N. Murdock, W. K. Dodds, and M. R. Whiles. 2010b. Disturbance-mediated effects of fishes on stream ecosystem processes: concepts and results from highly variable prairie streams, p. 593–617. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. Gido, K. B., and J. H. Brown. 1999. Invasion of North American drainages by alien fish species. Freshwater Biology 42:387–399. Gido, K. B., W. K. Dodds, and M. E. Eberle. 2010a. Retrospective analysis of fish community change during a half-century of landuse and streamflow changes. Journal of the North American Benthological Society 29:970–987. Gido, K. B., C. S. Guy, T. R. Strakosh, R. J. Bernot, K. J. Hase, and M. A. Shaw. 2002. Long-term changes in the fish assemblages of the Big Blue River Basin 40 years after the construction of Tuttle Creek Reservoir. Transactions of the Kansas Academy of Science 105:193–208. Gido, K. B., and W. J. Matthews. 2001. Ecosystem effects of water column minnows in experimental streams. Oecologia 126:247–253. Gido, K. B., W. J. Matthews, and W. C. Wolfinbarger. 2000. Longterm changes in a reservoir fish assemblage: stability in an unpredictable environment. Ecological Applications 10:1517–1529.

Gido, K. B., and D. L. Propst. 1999. Habitat use and association of native and nonnative fishes in the San Juan River, New Mexico and Utah. Copeia 1999:321–332. Gido, K. B., J. F. Schaefer, and J. Pigg. 2004. Patterns of fish invasions in the Great Plains of North America. Biological Conservation 118:121–131. Gilbert, C. H. 1898. The fishes of the Klamath River Basin. Bulletin of the United States Fisheries Commission 17:1–13. Gilbert, C. R. 1961. Hybridization versus intergradation: An inquiry into the relationship of two cyprinid fishes. Copeia 1961:181–192. Gilbert, C. R. 1964. The American cyprinid fishes of the subgenus Luxilus (genus Notropis). Bulletin of the Florida State Museum, Biological Sciences 8:95–194. Gilbert, C. R. 1976. Composition and derivation of the North American freshwater fish fauna. Florida Scientist 39:104–111. Gilbert, C. R. 1978. The nominal North American cyprinid fish, Notropis henryi interpreted as an intergeneric hybrid, Clinostomus funduloides × Nocomis leptocephalus. Copeia 1978:179–181. Gilbert, C. R. 1998. Type cata log of recent and fossil North American freshwater fishes: families Cyprinidae, Catostomidae, Ictaluridae, Centrarchidae and Elassomatidae. Florida Museum of Natural History, Special Publication No. 1, University of Florida, Gainesville. Gilbert, C. R., and R. M. Bailey. 1972. Systematics and zoogeography of the American cyprinid fish Notropis (Opsopoeodus) emiliae. Occasional Papers of the Museum of Zoology, University of Michigan 664. Gilbert, C. R., J. D. Williams, and National Audubon Society. 2002. National Audubon Society Field Guide to Fishes. North America. Alfred A. Knopf, New York. Gilbert, L. E., and P. E. Raven (eds.). 1980. Coevolution of Animals and Plants. Revised edition. Symposium V of the First International Congress of Systematic and Evolutionary Biology. University of Texas Press, Austin. Giles, L. W., and V. L. Childs. 1949. Alligator management of the Sabine National Wildlife Refuge. The Journal of Wildlife Management 13:16–28. Giles, N. 1987. Predation risk and reduced foraging activity in fish: experiments with parasitized and non-parasitized Three-spined Sticklebacks, Gasterosteus aculeatus L. Journal of Fish Biology 31:37–44. Gill, H. S., C. B. Renaud, F. Chapleau, R. L. Mayden, and I. C. Potter. 2003. Phylogeny of living parasitic Lampreys (Petromyzontiformes) based on morphological data. Copeia 2003:687–703. Gillette, D. P., A. M. Fortner, N. R. Franssen, S. Cartwright, C. M. Tobler, J. S. Wesner, P. C. Reneau, F. H. Reneau, I. Schlupp, E. C. Marsh-Matthews, W. J. Matthews, R. E. Broughton, and C. W. Lee. 2012. Patterns of change over time in darter (Teleostei: Percidae) assemblages of the Arkansas River basin, northeastern Oklahoma, USA. Ecography 35:855–864. Gilliam, J. F., and D. F. Fraser. 1987. Habitat selection under predation hazard: test of a model with foraging minnows. Ecology 68:1856–1862. Gillis, G. B. 1998. Environmental effects on undulatory locomotion in the American Eel Anguilla rostrata: kinematics in water and on land. Journal of Experimental Biology 201:949–961. Gilmour, K. M., and S. F. Perry. 2004. Branchial membraneassociated carbonic anhydrase activity maintains CO2 excretion in severely anemic dogfish. American Journal of

LITERATURE CITED

Physiology-Regulatory, Integrative, and Comparative Physiology 286:R1138–R1148. Gisbert, E., and S. I. Doroshov. 2006. Allometric growth in Green Sturgeon larvae. Journal of Applied Ichthyology 22(Supplement 1):202–207. Gisbert, E., and G. I. Ruban. 2003. Ontogenetic behavior of Siberian Sturgeon, Acipenser baerii: a synthesis between laboratory tests and field data. Environmental Biology of Fishes 67:311–319. Giudice, J. 1964. Production and comparative growth of three hybrids. Proceedings of the Southeastern Association of Game and Fish Commissioners 18:512–517. Gjerde, B. 1984. Variation in semen production of farmed Atlantic Salmon and Rainbow Trout. Aquaculture 40:109–114. Gleason, C. A., and T. M. Berra. 1993. Demonstration of reproductive isolation and observation of mismatings in Luxilus cornutus and L. chrysocephalus in sympatry. Copeia 1993:614–628. Glenn, C. L. 1975a. Annual growth-rates of Mooneye, Hiodon tergisus, in Assiniboine River. Journal of the Fisheries Research Board of Canada 32:407–410. Glenn, C. L. 1975b. Seasonal diets of Mooneye, Hiodon tergisus, in the Assiniboine River. Canadian Journal of Zoology 53:232–237. Glenn, C. L. 1976. Seasonal growth-rates of Mooneye (Hiodon tergisus) from Assiniboine River. Journal of the Fisheries Research Board of Canada 33:2078–2082. Glenn, C. L. 1978. Seasonal growth and diets of young-of-theyear Mooneye (Hiodon tergisus) from the Assiniboine River, Manitoba. Transactions of the American Fisheries Society 107:587–589. Glenn, C. L. 1980. Seasonal parasitic infections in Mooneye, Hiodon tergisus (LeSueur), from the Assiniboine River. Canadian Journal of Zoology 58:252–257. Glenn, C. L., and R. R. G. Williams. 1976. Fecundity of Mooneye, Hiodon tergisus, in the Assiniboine River. Canadian Journal of Zoology 54:156–161. Glover, J. B., M. E. Domino, K. C. Altman, J. W. Dillman, W. S. Castleberry, J. P. Eidson, and M. Mattocks. 2010. Mercury in South Carolina fishes, USA. Ecotoxicology 19:781–795. Goater, C. P., D. Bray, and D. B. Conn. 2005. Cellular aspects of early development of Ornithodiplostomum ptychocheilus metacercariae in the brain of Fathead Minnows, Pimephales promelas. Journal of Parasitology 91:814–821. Gobalet, K. W., P. D. Schulz, T. A. Wake, and N. Siefkin. 2004. Archaeological perspectives on Native American fisheries of California, with emphasis on steelhead and salmon. Transactions of the American Fisheries Society 133:801–833. Gobalet, K. W., and G. L. Fenenga. 1993. Terminal Pleistoceneearly Holocene fishes from Tulare Lake, San Joaquin Valley, California with comments on the evolution of Sacramento squawfish (Ptychocheilus grandis: Cyprinidae). Paleobios 15:1–8. Gobalet, K. W., and T. L. Jones. 1995. Prehistoric Native American fisheries of the central California coast. Transactions of the American Fisheries Society 124:813–823. Goddard, K. A., and R. M. Dawley. 1990. Clonal inheritance of a diploid nuclear genome by a hybrid freshwater minnow (Phoxinus eos-neogaeus, Pisces: Cyprinidae). Evolution 44:1052–1065. Goddard, K. A., R. M. Dawley, and T. E. Dowling. 1989. Origin and genetic relationships of diploid, triploid, and diploid-triploid

541

mosaic biotypes in the Phoxinus eos-neogaeus unisexual complex. New York State Museum Bulletin 466:268–280. Goddard, K. A., and R. J. Schultz. 1993. Aclonal reproduction by polyploidy members of the clonal hybrid species Phoxinus eosneogaeus (Cyprinidae). Copeia 1993:650–660. Godfrey, H. 1957. Feeding of Eels in four New Brunswick salmon streams. Fisheries Research Board of Canada Progress Report Atlantic Coast Station 67:19–22. Godin, J-G. 1995. Predation risk and alternative mating tactics in male Trinidadian guppies (Poecilia reticulata). Oecologia 103:224–229. Godin, J.-G. J., L. J. Classon and M. V. Abrahams. 1988. Group vigilance and shoal size in a small characin fish. Behavior 104:29–40. Goff, G. P. 1984. Brood care of Longnose Gar (Lepisosteus osseus) by Smallmouth Bass (Micropterus dolomieu). Copeia 1984:149–152. Gold, J. R., and C. T. Amemiya. 1986. Cytogenetic studies in North American minnows (Cyprinidae). XII. Patterns of chromosomal NOR variation among fourteen species. Canadian Journal of Zoology 65:1869–1877. Gold, J. R., and C. T. Amemiya. 1987. Genome size variation in North American minnows (Cyprinidae). II. Variation among 20 species. Genome 29:481–489. Gold, J. R., and H. J. Price. 1985. Genome size variation among North American minnows (Cyprinidae). I. Distribution of the variation in five species. Heredity 54:297–305. Gold, J. R., C. J. Ragland, and L. J. Schliesing. 1990. Genome size variation and evolution in North American cyprinid fishes. Genetics Selection Evolution 22:11–29. Gold, J. R., and P. K. Zoch. 1990. Intraspecific variation in chromosomal nucleolus organizer regions in Notropis chrysocephalus (Pisces: Cyprinidae). The Southwestern Naturalist 35:211–215. Gold, J. R., P. K. Zoch, and C. T. Amemiya. 1988. Cytogenetic studies in North American minnows (Cyprinidae). XIV Chromosomal NOR phenotypes of eight species from the genus Notropis. Cytobios 54:137–147. Golden, M. P. 1969. The Lost River Sucker, Catostomus luxatus (Cope). Oregon State Game Commission Report I-69. Goldman, C. R., and A. J. Horne. 1983. Limnology. McGraw-Hill, New York. Goldsborough, E. L., and H. W. Clark. 1908. Fishes of West Virginia. Bulletin of the Bureau of Fisheries Washington 27:29–39. Goldschmidt, T., and T. C. M. Bakker. 1990. Determinants of reproductive success of male Sticklebacks in the field and the laboratory. Netherlands Journal of Zoology 40:664–687. Goldschmidt, T., T. C. M. Bakker, and E. Feuth-De Bruijn. 1993. Selective copying in mate choice of female Sticklebacks. Animal Behaviour 45:541–547. Goldschmidt, T., S. A. Foster, and P. Sevenster. 1992. Inter-nest distance and sneaking in the Three-spined Stickleback. Animal Behaviour 44:793–795. Goldstein, R. M., and M. R. Meador. 2004. Comparisons of fish species traits from small streams to large rivers. Transactions of the American Fisheries Society 133:971–983. Goldthwait, R. P., A. Dreimanis, J. L. Forsyth, P. F. Karrow, and G. W. White. 1965. Pleistocene deposits of the Erie Lobe, p. 85–97. In The Quaternary of the United States. H. E. Wright, Jr., and D. G. Frey (eds.). Princeton University Press, Princeton, New Jersey.

542 LITERATURE CITED

Golub, J. L., and G. E. Brown. 2003. Are all signals the same? Ontogenetic change in the response to conspecific and heterospecific chemical alarm signals by juvenile Green Sunfish (Lepomis cyanellus). Behavioral Ecology and Sociobiology 54:113–118. Golub, J. L., V. Vermette, and G. E. Brown. 2005. Response to conspecific and heterospecific alarm cues by Pumpkinseeds in simple and complex habitats: field verification of an ontogenetic shift. Journal of Fish Biology 66:1073–1081. Gomulkiewicz, R., S. L. Nuismer, and J. N. Thompson. 2003. Coevolution in variable mutualisms. The American Naturalist 162:S80–S93. Gomulkiewicz, R., J. N. Thompson, R. D. Holt, S. L. Nuismer, and M. E. Hochberg. 2000. Hot spots, cold spots, and the geographic mosaic theory of coevolution. The American Naturalist 156:156–174. Goniakowska-Witalinska, L., G. Zaccone, S. Fasulo, A. Mauceri, A. Licata, and J. Youson. 1995. Neuroendocrine cells in the gills of the Bowfin Amia calva: an ultrastructural and immunocytochemical study. Folia Histochemica et Cytobiologica 33:171–177. Gonzalez, R. J., L. Milligan, A. Pagnotta, and D. G. McDonald. 2001. Effect of air breathing on acid-base and ion regulation after exhaustive exercise and during low pH exposure in the Bowfin, Amia calva. Physiological and Biochemical Zoology 74:502–509. Goode, G. B. 1884. The food fishes of the United States, p. 163– 682. In The Fisheries and Fishery Industries of the United States. Section I Part III. Washington, D.C. Goodfellow, W. L., Jr., C. H. Hocutt, R. P. Morgan, II, and J. R. Stauffer, Jr. 1984. Biochemical assessment of the taxonomic status of “Rhinichthys bowersi” (Pisces: Cyprinidae). Copeia 1984:652–659. Goodger, W. P., and T. A. Burns. 1980. The cardiotoxic effects of Alligator Gar (Lepisosteus spatula) roe on the isolated turtle heart. Toxicon 18:489–494. Goodrich, E. S. 1909. Vetebrata craniata: cyclostomes and fishes, p. 1–518. In A Treatise on Zoology, Part IX. R. Lankester (ed.). Adam and Charles Black, London, England. Goodrich, E. S. 1930. Studies on the Structure and Development of Vertebrates. Macmillan, London. Goodwin, K. P., and P. L. Angermeier. 2003. Demographic characteristics of American Eel in the Potomac River drainage, Virginia. Transactions of the American Fisheries Society 132:524–535. Goodyear, C. D., T. A. Edsall, D. M. Ormsby Dempsey, G. D. Moss, and P. E. Polanski. 1982. Atlas of the Spawning and Nursery Areas of Great Lakes Fishes. Vol. XIII. Reproductive Characteristics of Great Lakes Fishes. U.S. Fish and Wildlife Ser vice FWS/ OBS-82/52. Goodyear, C. P. 1966. Distribution of Gars on the Mississippi coast. Journal of the Mississippi Academy of Sciences 12:188–192. Goodyear, C. P. 1967. Feeding habits of three species of Gars, Lepisosteus, along the Mississippi Gulf Coast. Transactions of the American Fisheries Society 96:297–300. Gordon, B. 2009. Paddlefish harvest in Oklahoma, p. 223–233. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Gordon, M. E., and J. B. Layzer. 1993. Glochidial host of Alasmidonta atropurpurea (Bivalvia: Unionoidea, Unionidae). Transactions of the American Microscopical Society 112:145–150.

Gordon, M. S., and D. E. Rosen. 1962. A cavernicolous form of the poeciliid fish, Poecilia sphenops from Tabasco, Mexico. Copeia 2:360–368. Gorman, O. T. 1986. Assemblage organization of stream fishes: the effect of rivers on adventitious streams. The American Naturalist 128:611–616. Gorman, O. T. 1987. Habitat segregation in an assemblage of minnows in an Ozark stream, p. 33–41. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Gorman, O. T. 1988a. An experimental study of habitat use in an assemblage of Ozark minnows. Ecology 69:1239–1250. Gorman, O. T. 1988b. The dynamics of habitat use in a guild of Ozark minnows. Ecological Monographs 58:1–18. Gorman, O. T. 1992. Evolutionary ecology and historical ecology: assembly, structure, and organization of stream fish communities, p. 659–688. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Gorman, O. T., and J. R. Karr. 1978. Habitat structure and stream fish communities. Ecology 59:507–515. Gotelli, N. J., and G. R. Graves. 1996. Null models in ecology. Smithsonian Institution Press, Washington, D. C. Gotelli, N. J., and D. J. McCabe. 2002. Species co-occurrence: a meta-analysis of J. M. Diamond’s assembly rules model. Ecology 83:2091–2096. Gotelli, N. J., and C. M. Taylor. 1999. Testing macroecology models with stream-fish assemblages. Evolutionary Ecology Research 1:847–858. Goto, A. 1987. Polygyny in the River Sculpin, Cottus hangiongensis (Pisces: Cottidae), with special reference to male mating success. Copeia 1987:32–40. Goto, Y., S. Kubota, and S. Kohno. 1998. Highly repetitive DNA sequences that are restricted to the germline in the Hagfish, Eptatretus cirrhatus: a mosaic of eliminated elements. Chromosoma 107:17–32. Gould, J. L., S. L. Elliott, C. M. Masters, and J. Mukerji. 1999. Female preferences in a fish genus without female mate choice. Current Biology 9:497–500. Gould, S. J., and R. C Lewontin. 1979. The spandrels of San Marcos and the Panglossian paradigm. A critique of the adaptationist program. Proceedings of the Royal Society of London B 205: 581–598. Gould, S. J., and E. S. Vrba. 1982. Exaptation—a missing term in the science of form. Paleobiology 8:4–15. Govardovskii, V. I., P. Röhlich, Á. Szél, and L. V. Zueva. 1992. Immunocytochemical reactivity of rod and cone visual pigments in the Sturgeon retina. Visual Neuroscience 8:531–537. Govardovskii, V. I., and L. V. Zueva. 1987. Photoreceptors and visual pigments in Sturgeons. Journal of Evolutionary Biochemistry and Physiology 23:685–686. Gowanloch, J. N. 1933. Fishes and Fishing in Louisiana. Louisiana Department of Conservation, Baton Rouge. Gowanloch, J. N. 1933. Fishes and Fishing in Louisiana, Including Recipes for the Preparation of Seafoods. Bulletin No. 23, Department of Conservation, New Orleans, Louisiana. Grabowski, G. M., J. G. Blackburn, and E. R. Lacy. 1999. Morphology and epithelia ion transport of the alkaline gland in the Atlantic Stingray (Dasyatis sabina). Biology Bulletin 197:82–93.

LITERATURE CITED

Grady, J. M., and R. C. Cashner. 1988. Evidence of extensive intergeneric hybridization among the cyprinid fauna of Clark Creek, Wilkinson Co., Mississippi. The Southwestern Naturalist 33:137–146. Grady, J. M., and B. S. Elkington. 2009. Establishing and maintaining Paddlefish populations by stocking, p. 385–396. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Graham, J. B. 1997. Air-breathing Fishes: Evolution, Diversity, and Adaptation. Academic Press, New York. Graham, K. 1997. Contemporary status of the North American Paddlefish, Polyodon spathula. Environmental Biology of Fishes 48:279–289. Graham, L. K. 1986. Establishing and maintaining Paddlefish populations by stocking, p. 96–104. In The Paddlefish: Status, Management and Propagation. J. G. Dillard, L. K. Graham, and T. R. Russell (eds.). American Fisheries Society Special Publication 7. Graham, L. K., E. J. Hamilton, T. R. Russell, and C. E. Hicks. 1986. Culture of Paddlefish—a review of methods, p. 78–94. In The Paddlefish: Status, Management and Propagation. J. G. Dillard, L. K. Graham, and T. R. Russell (eds.). American Fisheries Society Special Publication 7. Graham, L. K., and J. L. Rasmussen. 1999. A MICRA perspective on closing Paddlefish and Sturgeon commercial fisheries, p. 130–142. In Proceedings of the Symposium on the Harvest and Conservation of North American Paddlefish and Sturgeon, May 7–8, 1998, Chattanooga, Tennessee. D. F. Williamson, G. W. Benz, and C. M. Hoover (eds.). TRAFFIC North America / World Wildlife Fund, Washington, D.C. Gramoni, R. and M. A. Ali. 1970. L’electroretinogramme et sa frequence de fusion chez Amia calva (Linne). Revue Canadienne de Biologie 29:353–363. Grande, L. 1979. Eohiodon falcatus, a new species of hiodontid (Pisces) from the late early Eocene Green River Formation of Wyoming. Journal of Paleontology 53:103–111. Grande, L. 1980. Paleontology of the Green River Formation, with a review of the fish fauna. Geological Survey of Wyoming Bulletin 63. 333 p. Grande, L. 1982. A revision of the fossil genus Knightia, with a description of a new genus from the Green River Formation (Teleostei, Clupeidae). American Museum Novitates 2731:1–22. Grande, L. 1984. Paleontology of the Green River formation, with a review of the fish fauna, 2nd ed. Geological Survey of Wyoming Bulletin 63:1–333. Grande, L. 1985. Recent and fossil clupeomorph fishes with materials for revision of the subgroups of clupeoids. Bulletin of the American Museum of Natural History 181:231–372. Grande, L. 1999. The first Esox (Esocidae: Teleostei) from the Eocene Green River formation, and a brief review of esocid fishes. Journal of Vertebrate Paleontology 19:271–292. Grande, L. 2001. An updated review of the fish faunas from the Green River Formation, the world’s most productive freshwater Lagerstatten, p. 1–38. In Eocene Biodiversity: Unusual Occurrences and Rarely Sampled Habitats. G. F. Gunnell (ed.). Kluwer Academic / Plenum Publishers, New York. Grande, L. 2005. Phylogenetic study of gars and closely related species, based mostly on skeletal morphology. The resurrection of Holostei, p. 119–122. In Fourth International Meeting on Me-

543

sozoic Fishes—Systematics, Homology and Nomenclature, Extended Abstracts. J. Poyato Ariza (ed.). UAM Ediciones, Madrid. Grande, L. 2010. An Empirical Synthetic Pattern Study of Gars (Lepisosteiformes) and Closely Related Species, Based Mostly on Skeletal Anatomy. The Resurrection of Holostei. American Society of Ichthyologists and Herpetologists Special Publication 6:1–871; Supplementary Issue of Copeia 10(2A). Grande, L. 2013. The Lost World of Fossil Lake: Snapshots from Deep Time. University of Chicago Press, Chicago, Illinois. Grande, L., and W. E. Bemis. 1991. Osteology and phylogenetic relationships of fossil and Recent Paddlefishes (Polyodontidae) with comments on the interrelationships of Acipenseriformes. Journal of Vertebrate Paleontology 11, Supplement 1:1–121. Grande, L., and W. E. Bemis. 1996. Interrelationships of Acipenseriformes, with comments on “Chondrostei,” p. 85–115. In Interrelationships of Fishes. M. Stiassny, L. Parenti, and G. D. Johnson (eds.). Academic Press, San Diego, California. Grande, L. and W. E. Bemis. 1998. A comprehensive phylogenetic study of amiid fishes (Amiidae) based on comparative skeletal anatomy. An empirical search for interconnected patterns of natural history. Society of Vertebrate Paleontology Memoir 4:1–690; Supplement to Journal of Vertebrate Paleontology 18(1). Grande, L., and W. E. Bemis. 1999. Historical biogeography and historical paleoecology of Amiidae and other halecomorph fishes, p. 413–424. In Systematics and Fossil Record. G. Arratia and H.-P. Schultze (eds.). Verlag Dr. Frierich Pfeil, Munchen Germany. Grande, L., J. T. Eastman, and T. M. Cavender. 1982. Amyzon gosiutensis, a new catostomid fish from the Green River Formation. Copeia 1982:523–532. Grande, L., and E. J. Hilton. 2006. An exquisitely preserved skeleton representing a primitive Sturgeon from the Upper Cretaceous Judith River formation of Montana (Acipenseriformes: Acipenseridae: n. gen. and sp.). Journal of Paleontology Memoir 65, volume 80, supplement to number 4:1–39. Grande, L., and E. J. Hilton. 2009. A replacement name for †Psammorhynchus Grande & Hilton, 2006 (Actinopterigii, Acipenseriformes, Acipenseridae). Journal of Paleontology 83:317–318. Grande, L., F. Jin, Y. Yabumoto, and W. E. Bemis. 2002. Protopsephurus liui, a well-preserved primitive Paddlefish (Acipenseriformes: Polyodontidae) from the lower Cretaceous of China. Journal of Vertebrate Paleontology 22:209–237. Grande, L., and J. G. Lundberg. 1988. Revision and redescription of the genus Astephus (Siluriformes: Ictaluridae) with a discussion of its phylogenetic relationships. Journal of Vertebrate Paleontology 8:139–171. Grande, L., and G. Nelson. 1985. Interrelationships of fossil and recent Anchovies (Teleostei: Engrauloidea) and description of a new species from the Miocene of Cyprus. American Museum Novitates 2826:1–16. Grant, B. F., P. M. Mehrle, and T. R. Russell. 1970. Serum characteristics of spawning Paddlefish (Polyodon spathula). Comparative Biochemistry and Physiology 37:321–330. Grant, D., P.-M. Fontaine, S. P. Good, J. J. Dodson, and L. Bernatchez. 2002a. The influence of male parental identity on growth and survival of offspring in Atlantic Salmon (Salmo salar). Evolutionary Ecology Research 4:537–549. Grant, G. C., B. Vondracek, and P. W. Sorensen. 2002b. Spawning interactions between sympatric Brown and Brook Trout may contribute to species replacement. Transactions of the American Fisheries Society 131:569–576.

544 LITERATURE CITED

Gray, R. W., and C. W. Andrews. 1971. Age and growth of the American Eel (Anguilla rostrata (LeSueur)) in Newfoundland waters. Canadian Journal of Zoology 49:121–128. Gray, S. M., L. M. Dill, F. Y. Tantu, E. R. Loew, F. Herder, and J. S. McKinnon. 2008. Environment-contingent sexual selection in a colour polymorphic fish. Proceedings of the Royal Society of London B 275:1785–1791. Greeley, J. R. 1927. Fishes of the Genesee region with annotated list, p. 47–66. In A biological survey of the Genesee River system. Supplemental to 16th Annual Report of the New York State Conservation Department, Albany. Green, O. L. 1966. Observations on the culture of the Bowfin. The Progressive Fish Culturist 28:179. Greenbank, J. 1956. Movement of fish under the ice. Copeia 1956:158–162. Greenberg, L. A. 1991. Habitat use and feeding behavior of thirteen species of benthic stream fishes. Environmental Biology of Fishes 31:389–401. Greene, J. M., and K. L. Brown. 1991. Demographic and genetic characteristics of multiple inseminated female mosquitofish (Gambusia affinis). Copeia 1991:434–444. Greenfield, D. W., S. T. Ross, and G. D. Deckert. 1970. Some aspects of the life history of the Santa Ana Sucker, Catostomus (Pantosteus) santaanae (Snyder). California Fish and Game 56:166–179. Greenfield, D. W., and J. E. Thomerson. 1997. Fishes of the continental waters of Belize. University Press of Florida, Gainesville. Greenwood, P. H. 1970. On the genus Lycoptera and its relationship with the family Hiodontidae (Pisces, Osteoglossomorpha). Bulletin of the British Museum (Natural History) Zoology 19:257–285. Greenwood, P. H. 1971. Hyoid and ventral gill arch musculature in osteoglossomorph fishes. Bulletin of the British Museum (Natural History) Zoology 22:1–55. Greenwood, P. H. 1973. Interrelationships of osteoglossomorphs, p. 307–332. In Interrelationships of Fishes. P. H. Greenwood, R. S. Miles and C. Patterson (eds.). Academic Press, London. Greenwood, P. H. 1984. African Cichlids and evolutionary theories, p. 141–154. In Evolution of Fish Species Flocks. A. A. Echelle and I. Kornfield (eds.). University of Maine at Orono Press. Gregor, P. D., and J. E. Deacon. 1987. Diel food utilization by Woundfin, Plagopterus argentissimus, in Virgin River, Arizona. Environmental Biology of Fishes 19:73–77. Gregor, P. D., and J. E. Deacon. 1988. Food partitioning among fishes of the Virgin River. Copeia 1988:314–323. Gregory, W. K. 1933. Fish skulls: a study of the evolution of natural mechanisms. Transactions of the American Philosophical Society 23:75–481. Griffith, R. E. 1953. Preliminary survey of the parasites of fish of the Palouse area. Transactions of the American Microscopal Society 72:51–57. Griffith, S. A., and D. L. Bechler. 1995. The distribution and abundance of the Bay Anchovy, Anchoa mitchilli, in a southeast Texas marsh lake system. Gulf Research Reports 9:117–122. Griffith, S. C., I. P. F. Owens, and K. A. Thuman. 2002. Extra pair paternity in birds: a review of interspecific variation and adaptive function. Molecular Ecology 11:2195–2212. Grimm, N. B. 1988. Feeding dynamics, nitrogen budgets, and ecosystem role of a desert stream omnivore, Agosia chrysogaster (Pisces: Cyprinidae). Environmental Biology of Fishes 21:143–152.

Grobstein, P., C. Comer, and S. Kostyk. 1980. The potential binocular field and its tectal representation in Rana pipiens. Journal of Comparative Neurology 190:175–185. Grohs, K. L., R. A. Klumb, S. R. Chipps, and G. A. Wanner. 2009. Ontogenetic patterns in prey use by Pallid Sturgeon in the Missouri River, South Dakota and Nebraska. Journal of Applied Ichthyology 25(Supplement 2):48–53. Grose, M. J., and E. O. Wiley. 2002. Phylogenetic relationships of the Hybopsis amblops species group (Teleostei: Cyprindae). Copeia 2004:1092–1097. Gross, E. L., P. J. Patchett, T. A. Dallegge, and J. E. Spencer. 2001. The Colorado River system and Neogene sedimentary formations along its course: apparent Sr isotopic connections. Journal of Geology 109:449–461. Gross, M. R. 1982. Sneaker, satellites and parentals: polymorphic mating strategies in North American Sunfishes. Zeitschrift für Tierpsychologie 60:1–26. Gross, M. R. 1984. Sunfish, salmon, and the evolution of alternative reproductive strategies and tactics in fishes, p. 57–75. In Fish Reproduction: Strategies and Tactics. G. W. Potts and R. J. Wootton (eds.). Academic Press, London, United Kingdom. Gross, M. R. 1985. Disruptive selection for alternative life histories in salmon. Nature 313:47–48. Gross, M. R. 1991a. Evolution of alternative reproductive strategies: frequency-dependent sexual selection in male Bluegill Sunfish. Philosophical Transactions of the Royal Society of London B 332:59–66. Gross, M. R. 1991b. Salmon breeding behavior and life history evolution in changing environments. Ecology 72:1180–1186. Gross, M. R. 1996. Alternative reproductive strategies and tactics: diversity within sexes. Trends in Ecology and Evolution 11:92–98. Gross, M. R., and E. L. Charnov. 1980. Alternative male life histories in Bluegill Sunfish. Proceedings of the National Academy of Sciences of the United States of America 77:6937–6940. Gross, M. R., and A. M. McMillan. 1981. Predation and the evolution of colonial nesting in Bluegill Sunfish (Lepomis macrochirus). Behavioral Ecology and Sociobiology 8:163–174. Grosslein, M. D., and L. L. Smith, Jr. 1959. The Goldeye, Amphiodon alosoides (Rafinesque), in the commercial fishery of the Red Lakes Minnesota. Fishery Bulletin 60:33–41. Grossman, G. D., and V. Boulé. 1991. Effects of Rosyside Dace (Clinostomus funduloides) on microhabitat use of Rainbow Trout (Oncorhynchus mykiss). Canadian Journal of Fisheries and Aquatic Sciences 48:1235–1243. Grossman, G. G., and J. L. Sabo. 2010. Incorporating environmental variation into models of community stability: examples from stream fish, p. 407–426. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. Grossman, G. D., A. de Sostoa, M. C. Freeman, and J. LobonCervia. 1987a. Microhabitat use in a Mediterranean riverine fish assemblage: fishes of the lower Matarrana. Oecologia (Berlin) 73:490–500. Grossman, G. D, A. de Sostoa, M. C. Freeman, and J. LobonCervia. 1987b. Microhabitat use in a Mediterranean riverine fish assemblage: fishes of the upper Matarrana. Oecologia (Berlin) 73:501–512.

LITERATURE CITED

Grossman, G. D., J. F. Dowd, and M. Crawford. 1990. Assemblage stability in stream fishes: a review. Environmental Management 14:661–671. Grossman, G. D., and M. C. Freeman. 1987. Microhabitat use in a stream fish assemblage. Journal of Zoology, London 212:151–176. Grossman, G. D., P. B. Moyle, and J. O. Whittaker, Jr. 1982. Stochasticity in structural and functional characteristics of an Indiana stream fish assemblage: a test of community theory. American Naturalist 120:423–454. Grossman, G. D., and R. E. Ratajczak, Jr. 1998. Long-term patterns of microhabitat use by fish in a southern Appalachian stream from 1983 to 1992: effects of hydrologic period, season and fish length. Ecology of Freshwater Fish 7:108–131. Grossman, G. D., R. E. Ratajczak, Jr., M. Crawford, and M. C. Freeman. 1998. Assemblage organization in stream fishes: effects of environmental variation and interspecific interactions. Ecological Monographs 68:395–420. Grossman, G. D., P. A. Rincon, M. D. Farr, and R. E. Ratajczak, Jr. 2002. A new optimal foraging model predicts habitat use by drift-feeding stream minnows. Ecology of Freshwater Fish 11:2–10. Grunina, A. S., and A. V. Recoubratsky. 2005. Induced androgenesis in fish: obtaining viable nucleocytoplasmic hybrids. Russian Journal of Developmental Biology 36:208–217. Grunina, A. S., A. V. Recoubratsky, V. A. Barmintsev, E. D. Vasil’eva, and M. S. Chebanov. 2009. Dispermic androgenesis as a method for recovery of endangered Sturgeon species, p. 187–204. In Biology, Conservation and Sustainable Development of Sturgeons. R. Carmona, A. Domezain, M. García-Gallego, J. A. Hernando, F. Rodríguez, and M. Ruiz-Rejón (eds.). Springer Science + Business Media B. V. Fish & Fisheries Series 29. Grunina, A. S., A. V. Recoubratsky, L. I. Tsvetkova, and V. A. Barmintsev. 2006. Investigation on dispermic androgenesis in Sturgeon fish. The first successful production of androgenetic Sturgeons with cryopreserved sperm. International Journal of Refrigeration 29:379–386. Grunwald, C., L. Maceda, J. Waldman, J. Stabile, and I. Wirgin. 2008. Conservation of Atlantic Sturgeon Acipenser oxyrinchus oxyrinchus: delineation of stock structure and distinct population segments. Conservation Genetics 9:1111–1124. Grunwald, C., J. Stabile, J. R. Waldman, R. Gross, and I. Wirgin. 2002. Population genetics of Shortnose Sturgeon Acipenser brevirostrum based on mitochondrial DNA control region sequences. Molecular Ecology 11:1885–1898. Guénette, S., R. Fortin, and R. Rassart. 1993. Mitochondrial DNA in Lake Sturgeon (Acipenser fulvescens) from the St. Lawrence River and James Bay drainage in Quebec, Canada. Canadian Journal of Fisheries and Aquatic Sciences 50:659–664. Guénette, S., E. Rassart, and R. Fortin. 1992. Morphological differentiation of Lake Sturgeon (Acipenser fulvescens) from the St. Lawrence River and Lac des Deux Montagnes (Quebec, Canada). Canadian Journal of Fisheries and Aquatic Sciences 49:1959–1965. Guilbard, F., J. Munro, P. Dumont, D. Hatin, and R. Fortin. 2007. Feeding ecology of Atlantic Sturgeon and Lake Sturgeon cooccurring in the St. Lawrence estuarine transition zone, p. 85–104. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland.

545

Guilford, T., and M. S. Dawkins. 1991a. Receiver psychology and the evolution of animal signals. Animal Behaviour 42:1–14. Guilford, T., and M. S. Dawkins. 1991b. Receiver psychology and the design of animal signals. Trends in Neurosciences 16:430–436. Guilford, T., and M. S. Dawkins. 1992. Understanding signal design—a reply to Blumberg and Alberts. Animal Behaviour 44:384–385. Gumm, J. M. and C. R. Gabor. 2005. Asexuals looking for sex: conflict between species and mate-quality recognition in Sailfin Mollies (Poecilia latipinna). Behavioral Ecology and Sociobiology 58:558–565. Gundersen, D. T., M. D. Krahling, J. J. Donosky, R. G. Cable, and S. D. Mims. 1998. Polychlorinated biphenyls and chlordane in the gonads of Paddlefish, Polyodon spathula, from the Ohio River. Bulletin of Environmental Contamination and Toxicology 61:650–657. Gundersen, D. T., R. Miller, A. Mischler, K. Elpers, S. D. Mims, J. G. Millar, and V. Blazer. 2000. Biomarker response and health of polychlorinated biphenyl- and chlordane-contaminated Paddlefish from the Ohio River basin, USA. Environmental Toxicology and Chemistry 19:2275–2285. Gundersen, D. T., and W. D. Pearson. 1992. Partitioning of PCBs in the muscle and reproductive tissues of Paddlefish, Polyodon spathula, at the Falls of the Ohio River. Bulletin of Environmental Contamination and Toxicology 49:455–462. Gunning, G. E., and C. R. Shoop. 1962. Restricted movements of the American Eel, Anguilla rostrata (LeSueur) in freshwater streams, with comments on growth rate. Tulane Studies in Zoology 9:265–272. Gunning, G. E., and R. D. Suttkus. 1991. Species dominance in the fish populations of the Pearl River at two study areas in Mississippi and Louisiana: 1966–1988. Southeastern Fishes Council Proceedings 23:7–15. Gunter, G. 1938. Notes on invasion of fresh-water by fishes of the Gulf of Mexico, with special reference to the MississippiAtchafalaya River system. Copeia 1938:69–72. Gunter, G. 1942. A list of fishes of the mainland of North and Middle America recorded from both freshwater and sea water. American Midland Naturalist 28:305–326. Gunter, G. 1945. Studies on marine fishes of Texas. Institute of Marine Science Publications 1:1–190. Gunter, G., and G. E. Hall. 1963. Biological investigations of the St. Lucie estuary (Florida) in connection with Lake Okeechobee discharges through the St. Lucie Canal. Gulf Research Report 1:189–307. Gunter, G., and G. E. Hall. 1965. A biological investigation of the Caloosahatchee estuary of Florida. Gulf Research Report 2:1–71. Gurgens, C., D. F. Russell, and L. A. Wilkens. 2000. Electrosensory avoidance of metal obstacles by the Paddlefish. Journal of Fish Biology 57:277–290. Guthrie, S. 1990. The physiology of the teleostean optic tectum, p. 279–344. In The Visual System of Fish. R. Douglas and M. Djamgoz (eds.). Chapman and Hall, London. Gutreuter, S., J. M. Dettmers, and D. H. Wahl. 2003. Estimating mortality rates of adult fish from entrainment through the propellers of river towboats. Transactions of the American Fisheries Society 132:646–661. Gutreuter, S., J. M. Vallazza, and B. C. Knights. 2006. Persistent disturbance by commercial navigation alters the relative abundance

546

LITERATURE CITED

of channel-dwelling fishes in a large river. Canadian Journal of Fisheries and Aquatic Sciences 63:2418–2433. Gutreuter, S., J. M. Vallazza, and B. C. Knights. 2010. Lateral distribution of fishes in the main-channel trough of a large floodplain river: implications for restoration. River Research and Applications 26:619–635. Haag, W. R., A. M. Commens-Carson, and M. L. Warren, Jr. 2007. Life history variation in the Yazoo Shiner (Notropis rafinesquei) in three Mississippi streams. American Midland Naturalist 158:306–320. Haag, W. R., and M. L. Warren, Jr. 1997. Host fishes and reproductive biology of six freshwater mussel species from the Mobile Basin, USA. Journal of the North American Benthological Society 16:576–585. Haag, W. R., and M. L. Warren, Jr. 2003. Host fishes and infection strategies of freshwater mussels in large Mobile Basin streams, USA. Journal of the North American Benthological Society 22:78–91. Haase, B. L. 1969. An ecological life history of the Longnose Gar, Lepisosteus osseus (Linnaeus), in Lake Mendota and in several other lakes of southern Wisconsin. Unpubl. Ph.D. diss., University of Wisconsin, Madison. Häberli, M. A., and P. B. Aeschlimann. 2004. Male traits influence odour-based mate choice in the three-spined Stickleback. Journal of Fish Biology 64:702–710. Hackney, P. A., W. M. Tatum, and S. L. Spencer. 1968. Life history of the River Redhorse, Moxostoma carinatum (Cope) in the Cahaba River, Alabama, with notes on the management of the species as a sport fish. Proceedings of the Southeastern Association of Game and Fish Commissioners 21:324–342. Haddad, V., Jr., N. D. Garrone, N. J. B. de Paula, F. P. L. Marques, K. C. Barbaro. 2004. Freshwater Stingrays: study of epidemiologic, clinic and therapeutic aspects based on 84 envenomings in humans and some enzymatic activities of the venom. Toxicon 43:287–294. Hagan, S. M., and K. W. Able. 2003. Seasonal changes of the pelagic fish assemblage in a temperate estuary. Estuarine, Coastal and Shelf Science 56:15–29. Hageman, J. R., D. C. Timpe, and R. D. Hoyt. 1986. The biology of the Paddlefish in Lake Cumberland, Kentucky. Proceedings of the Annual Conference of the Southeastern Association of Fish and Wildlife Agencies 40:237–248. Haines, S. E., and J. L. Gould. 1994. Female platys prefer long tails. Nature 370:512. Haire, R. N., A. L. Miracle, J. P. Rast, and G. W. Litman. 2000. Members of the Ikaros gene family are present in early representative vertebrates. The Journal of Immunology 165:306–312. Hale, M. C., J. R. Jackson, and J. A. DeWoody. 2010. Discovery and evaluation of candidate sex-determining genes and xenobiotics in the gonads of Lake Sturgeon (Acipenser fulvescens). Genetica 138:745–756. Hall, A. T., and J. T. Oris. 1991. Anthracene reduces reproductive potential and is maternally transferred during long-term exposure in Fathead Minnows. Aquatic Toxicology 19:249–264. Hall, D. J., and E. E. Werner. 1977. Seasonal distribution and abundance of fishes in the littoral zone of a Michigan lake. Transactions of the American Fisheries Society 106:545–555. Halliday, R. G. 1991. Marine distribution of the Sea Lamprey (Petromyzon marinus) in the northwest Atlantic. Canadian Journal of Fisheries and Aquatic Sciences 48:832–842.

Halstead, B. W., R. R. Ocampo, and F. R. Modglin. 1955. A study on the comparative anatomy of the venom apparatus of certain North American Stingrays. Journal of Morphology 97:1–21. Hamdanji, E. H., and K. B. Døving. 2007. The functional organization of the fish olfactory system. Progress in Neurobiology 82:80–86. Hamilton, W. D. 1971. Geometry of the selfish herd. Journal of Theoretical Biology 31:295–311. Hamilton, S. J. 2004. Review of selenium toxicity in aquatic food chains. Science of the Total Environment 326:1–31. Hamman, R. L. 1982. Induced spawning and culture of Bonytail Chub. Progressive Fish-Culturist 44:201–203. Haney, D. C., J. C. Vokoun, and D. B. Noltie. 2001. Alarm pheromone recognition in a Missouri darter assemblage. Journal of Fish Biology 59:810–817. Hankins, R. L., II. 1995. Status and distribution of selected populations of Relict Dace (Relictus solitarius) in White Pine County. Supplemental Report, Nevada Department of Conservation and Natural Resources. Hankison, S. J., and M. R. Morris. 2002. Sexual selection and species recognition in the Pygmy Swordtail, Xiphophorus pygmaeus: conflicting preferences. Behavioral Ecology and Sociobiology 51:140–145. Hankison, S. J., and M. R. Morris. 2003. Avoiding a compromise between sexual selection and species recognition: female swordtail fish assess multiple species-specific cues. Behavioral Ecology 14:282–287. Hann, H. W. 1927. The history of the germ of Cottus bairdi Girard. Journal of Morphology and Physiology 43:427–497. Hansen, K. A., and C. P. Paukert. 2009. Current management of Paddlefish sport fisheries, p. 277–290. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society, Symposium 66, Bethesda, Maryland. Hansen, M. J., and C. W. Ramm. 1994. Persistence and stability of fish community structure in a southwest New York stream. American Midland Naturalist 132:52–67. Hanson, A. J., and H. D. Smith. 1967. Mate selection in a population of Sockeye Salmon (Oncorhynchus nerka) of mixed age group. Journal of the Fisheries Research Board of Canada 24:1955–1977. Hanson, K. C., and S. J. Cooke. 2009. Why does size matter? A test of the benefits of female mate choice in a teleost fish based on morphological and physiological indicators of male quality. Physiological and Biochemical Zoology 82:617–624. Hanson, M. A., and M. R. Riggs. 1995. Potential effects of fish predation on wetland invertebrates: a comparison of wetlands with and without Fathead Minnows. Wetlands 15:167–175. Hanson, R. C., and W. R. Fleming. 1979. Serum cortisol levels of juvenile Bowfin, Amia calva: effects of hypophysectomy, hormone replacement and environmental salinity. Comparative Biochemistry and Physiology 63A:499–502. Hanson, R. C., D. Duff, J. Brehe, and W. R. Fleming. 1976. The effect of various salinities, hypophysectomy, and hormone treatments on the survival and sodium and potassium content of juvenile Bowfin, Amia calva. Physiological Zoology 49:376–385. Hara, T. J. 1986. Role of olfaction in fish behavior, p. 152–176. In The Behavior of Teleost Fishes. T. J. Pitcher (ed.). The Johns Hopkins University Press, Baltimore, Maryland.

LITERATURE CITED

Hara, T. J., and C. Zhang. 1998. Topographic bulbar projections and dual neural pathways of the primary olfactory neurons in salmonid fishes. Neuroscience 82:301–313. Hardisty, M. W. 1961. Studies on an isolated population of the brook Lamprey (Lampetra planeri). The Journal of Animal Ecology 30:339–355. Hardisty, M. W. 1964. The fecundity of Lampreys. Archiv für Hydrobiologie 60:340–357. Hardisty, M. W. 1965. Sex differentiation and gonadogenesis in Lampreys. II. The ammocoete gonads of the landlocked Sea Lamprey, Petromyzon marinus. Journal of Zoology, London 146:346–387. Hardisty, M. W. 1971. Gonadogenesis, sex differentiation and gametogenesis, p. 295–359. In The Biology of Lampreys. Vol. 1. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Hardisty, M. W. 1979. Biology of the cyclostomes. Chapman and Hall, London. Hardisty, M. W. 1982. Lampreys and Hagfishes: analysis of cyclostome relationships, p. 165–260. In The Biology of Lampreys. Vol. 4B. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Hardisty, M. W. 2006. Lampreys. Life without jaws. Forrest Text, Tresaith, United Kingdom. Hardisty, M. W., and I. C. Potter. 1971a. The behaviour, ecology and growth of larval Lampreys, p. 85–125. In The Biology of Lampreys. Vol. 1. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Hardisty, M. W., and I. C. Potter. 1971b. The general biology of adult Lampreys, p. 127–206. In The Biology of Lampreys. Vol. 1. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Hardisty, M. W., and I. C. Potter. 1971c. Paired species, p. 249–277. In The Biology of Lampreys. Vol. 1. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Hardisty, M. W., I. C. Potter, and R. W. Hilliard. 1989. Physiological adaptations of the living agnathans. Transactions of the Royal Society of Edinburgh: Earth Sciences 80:241–254. Hardisty, M. W., I. C. Potter, and R. Sturge. 1970. A comparison of the metamorphosing and macrophthalmia stages of the Lampreys, Lampetra fluviatilis and L. planeri. Journal of Zoology, London 162:383–400. Hardy, R. S., and M. K. Litvak. 2004. Effects of temperature on the early development, growth, and survival of Shortnose Sturgeon, Acipenser brevirostrum, and Atlantic Sturgeon, Acipenser oxyrhynchus, yolk-sac larvae. Environmental Biology of Fishes 70:145–154. Haro, A. J. 1991. Thermal preferenda and behavior of Atlantic Eels (genus Anguilla) in relation to their spawning migration. Environmental Biology of Fishes 31:171–184. Haro, A. 2003. Downstream migration of silver phase anguillid Eels, p. 215–222. In Eel Biology. Aida, K., K. Tsukamoto, and K. Yamauchi (eds.). Springer, Tokyo. Haro, A. J., and W. H. Krueger, 1988. Pigmentation, size, and migration of elvers (Anguilla rostrata (LeSueur)) in a coastal Rhode Island stream. Canadian Journal of Zoology 66:2528–2533. Haro, A. J., and W. H. Krueger. 1991. Pigmentation, otolith rings, and upstream migration of juvenile American Eels (Anguilla rostrata) in a coastal Rhode Island stream. Canadian Journal of Zoology 69:812–814. Haro, A., W. Richkus, K. Whalen, A. Hoar, W.-D. Busch, S. Lary, T. Brush, and D. Dixon. 2000. Population decline of the Ameri-

547

can Eel: implications for research and management. Fisheries 25:7–16 Harrell, R. M., and H. A. Loyacano. 1982. Age, growth and sex ratio of the American Eel in the Cooper River, South Carolina. Proceedings of the Annual Conference of the Southeastern Association of Fish and Wildlife Agencies 34:349–359. Harrington, R. W. J. 1955. The osteocranium of the American cyprinid fish, Notropis bifrenatus, with an annotated synonymy of teleost skull bones. Copeia 1955:267–290. Harris, P. M., and R. L. Mayden. 2001. Phylogenetic relationships of major clades of Catostomidae (Teleostei: Cypriniformes) as inferred from mitochondrial SSU and LSU rDNA sequences. Molecular Phylogenetics and Evolution 20:225–237. Harris, P. M., R. L. Mayden, H. E. Perez, and F. Garcia de Leon. 2002. Phylogenetic relationships of redhorse (Moxostoma) and jumprock (Scartomyzon) Suckers (Cypriniformes: Catostomidae) based on mitochondrial cytochrome b sequence data. Journal of Fish Biology 61:1433–1452. Hart, J. L. 1968. Respiration of the Winnipeg Goldeye (Hiodon alosoides). Journal of the Fisheries Research Board of Canada 25:2603–2608. Hart, M. 1985. The influence of fertilization on the creeping through cycle of the three-spined Stickleback. Behaviour 93:194–202. Hartman, E. J., and M. V. Abrahams. 2000. Sensory compensation and the detection of predators: the interaction between chemical and visual information. Proceedings of the Royal Society of London B 267:571–575. Hartman, K. J., and S. B. Brandt. 1995. Trophic resource partitioning, diets, and growth of sympatric estuarine predators. Transactions of the American Fisheries Society 124:520–537. Hartman, K. J., J. Howell, and J. A. Sweka. 2004. Diet and daily ration of Bay Anchovy in the Hudson River, New York. Transactions of the American Fisheries Society 133:762–771. Harvey, B. C. 1987. Susceptibility of young-of-the year fishes to downstream displacement by flooding. Transactions of the American Fisheries Society 116:851–855. Harvey, B. C. 1991a. Interactions among stream fishes: predatorinduced habitat shifts and larval survival. Oecologia 87:29–36. Harvey, B. C. 1991b. Interaction of abiotic and biotic factors influences larval fish survival in an Oklahoma stream. Canadian Journal of Fisheries and Aquatic Sciences 48:1476–1480. Harvey, B. C., R. C. Cashner, and W. J. Matthews. 1988. Differential effects of Largemouth and Smallmouth Bass on habitat use by stoneroller minnows in stream pools. Journal of Fish Biology 33:481–487. Harvey, B. C., and R. J. Nakamoto. 1999. Diel and seasonal movements by adult Sacramento Pikeminnow (Ptychocheilus grandis) in the Eel River, northwestern California. Ecology of Freshwater Fish 8:209–215. Harvey, B. C., and A. J. Stewart. 1991. Fish size and habitat depth relationships in headwater streams. Oecologia 87:336–342. Harvey, B. C., J. L. White, and R. J. Nakamoto. 2004. An emergent multiple predator effect may enhance biotic resistance in a stream fish assemblage. Ecology 85:127–133. Harvey, M. C., and G. E. Brown. 2004. Dine or dash? Ontogenetic shift in the response of Yellow Perch to conspecific alarm cues. Environmental Biology of Fishes 70:345–352. Haslouer, S. G., M. E. Eberle, D. R. Edds, K. B. Gido, C. S. Mammoliti, J. R. Triplett, J. T. Collins, D. A. Distler, D. G. Huggins,

548 LITERATURE CITED

and W. J. Stark. 2005. Current status of native fish species in Kansas. Transactions of the Kansas Academy of Science 108:32–46. Hassan, E. S. 1989. Hydrodynamic imaging of the surroundings by the lateral line of the blind cave fish, Anoptichthys jordani, p. 217–228. In The Mechanosensory Lateral Line Neurobiology and Evolution. S. Coombs, P. Gorner, and H. Munz (eds.). Springer, Berlin. Hatin, D., R. Fortin, and F. Caron. 2002. Movements and aggregation areas of adult Atlantic Sturgeon (Acipenser oxyrinchus) in the St. Lawrence River estuary, Québec, Canada. Journal of Applied Ichthyology 18:586–594. Hatin, D., S. Lachance, and D. Fournier. 2007a. Effect of dredged sediment deposition on use by Atlantic Sturgeon and Lake Sturgeon at an open-water disposal site in the St. Lawrence estuarine transition zone, p. 235–255. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Hatin, D., J. Munro, F. Caron, and R. D. Simons. 2007b. Movements, home range size, and habitat use and selection of early juvenile Atlantic Sturgeon in the St. Lawrence estuarine transition zone, p. 129–155. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Hauser, W. J. 1969. Life history of the Mountain Sucker, Catostomus platyrhynchus, in Montana. Transactions of the American Fisheries Society 98:209–215. Hawkins, A. D. 1993. Underwater sound and fish behaviour, p. 129–169. In The Behavior of Teleost Fishes. T. J. Pitcher (ed.). Chapman and Hall, London. Hawkins, D. K., and C. J. Foote. 1998. Early survival and development of coastal Cutthroat Trout (Oncorhynchus clarki clarki), steelhead (Oncorhynchus mykiss), and reciprocal hybrids. Canadian Journal of Fisheries and Aquatic Sciences 55:2097–2104. Hay, M. E., J. D. Parker, D. E. Burkepile, C. C. Caudill, A. E. Wilson, Z. P. Hallinan, and A. D. Chequer. 2004. Mutualisms and aquatic community structure: the enemy of my enemy is my friend. Annual Review of Ecology and Systematics 35:175–197. Hayes, D. B., W. W. Taylor, and J. C. Schneider. 1992. Response of Yellow Perch and the benthic invertebrate community to a reduction in the abundance of White Suckers. Transactions of the American Fisheries Society 121:36–53. Hazel, P.-P., J. Lamoureux, E. Magnin, and R. Nault. 1983. Croissance de six especes de poissoins vivant pres de leur limite de repartition en latitude et en altitude sur le territoire de la Baie James. Cybium 7:57–69. Hazelton, P. D., and G. D. Grossman. 2009. The effects of turbidity and an invasive species on foraging success of Rosyside Dace (Clinostomus funduloides). Freshwater Biology 54:1977–1989. Healey, M. C., R. Lake, and S. G. Hinch. 2003. Energy expenditures during reproduction by Sockeye Salmon (Oncorhynchus nerka). Behaviour 140:161–182. Heard, W. R. 1958. Studies in the genus Ictiobus (buffalofishes). Unpubl. Master’s thesis, Oklahoma State University, Stillwater. Heatherly, T. H., M. R. Whiles, D. Knuth, and J. E. Garvey. 2005. Diversity and community structure of littoral zone macroinver-

tebrates in southern Illinois reclaimed surface mine lakes. American Midland Naturalist 154:67–77. Heß, M., R. R. Melzer, R. Eser, and U. Smola. 2006. The structure of Anchovy outer retinae (Engraulididae, Clupeiformes): a comparative light- and electron-microscopic study using museumstored material. Journal of Morphology 267:1356–1380. Heckmann, R. A., J. E. Deacon, and P. D. Greger. 1986. Parasites of the Woundfin minnow, Plagopterus argentissimus and other endemic fishes from the Virgin River, Utah. Great Basin Naturalist 46:662–678. Heckmann, R. A., P. D. Greger, and R. C. Furtek. 1993. The Asian fish tapeworm, Bothriocephalus acheilognathi, in fishes from Nevada. Journal of the Helminthological Society of Washington 60:127–128. Heczko, E. J., and B. H. Seghers. 1981. Effects of alarm substance on schooling in the Common Shiner (Notropis cornutus, Cyprinidae). Environmental Biology of Fishes 6:25–29. Hedrick, M. S., M. L. Burleson, D. R. Jones, and W. K. Milsom. 1991. An examination of central chemosensitivity in an airbreathing fish (Amia calva). Journal of Experimental Biology 155:165–174. Hedrick, M. S., and D. R. Jones. 1993. The effects of altered aquatic and aerial respiratory gas concentrations on airbreathing patterns in a primitive fish (Amia calva). Journal of Experimental Biology 181:81–94. Hedrick, M. S., and D. R. Jones. 1999. Control of gill ventilation and air-breathing in the Bowfin Amia calva. The Journal of Experimental Biology 202:87–94. Hedrick, M. S., S. L. Katz, and D. R. Jones. 1994. Periodic airbreathing behavior in a primitive fish revealed by spectral analysis. Journal of Experimental Biology 197:429–436. Heff ron, J. K., and T. S. Moerland. Parvalbumin characterization from the euryhaline Stingray Dasyatis sabina. Comparative Biochemistry and Physiology, Part A 150:339–346. Heimberg, A. M., R. Cowper-Sal-lari, M. Sémon, P. C. J. Donoghue, and K. J. Peterson. 2010. microRNAs reveal the interrelationships of Hagfish, Lampreys, and gnathostomes and the nature of the ancestral vertebrate. Proceedings of the National Academy of Sciences 107:19379–19383. Heins, D. C., and J. A. Baker. 1987. Analysis of factors associated with intraspecific variation in propagule size of a streamdwelling fish, p. 223–231. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Heins, D. C., and J. A. Baker. 1992. Historical and recent influences on reproduction in North American stream fishes, p. 573–599. In Systematics, Historical Ecology and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Heins, D. C., and J. A. Baker. 1993. Clutch production in the darter Etheostoma lynceum and its implications for life-history study. Journal of Fish Biology 42(6):819–829. Heins, D. C., and F. G. Rabito, Jr. 1986. Spawning performance in North American minnows: direct evidence of the occurrence of multiple clutches in the genus Notropis. Journal of Fish Biology 28:343–357. Heins, D. C., and F. G. Rabito, Jr. 1988. Reproductive traits in populations of the Weed Shiner, Notropis texanus, from the Gulf coastal plain. The Southwestern Naturalist 33:147–156.

LITERATURE CITED

Heise, R. J., R. B. Bringolf, R. Patterson, W. G. Cope, and S. T. Ross. 2009. Plasma vitellogenin and estradiol concentrations in adult Gulf Sturgeon from the Pascagoula River drainage, Mississippi. Transactions of the American Fisheries Society 138:1028–1035. Heise, R. J., W. T. Slack, S. T. Ross, and M. A. Dugo. 2005. Gulf Sturgeon summer habitat use and fall migration in the Pascagoula River, Mississippi, USA. Journal of Applied Ichthyology 21:461–468. Heist, E. J., and A. Mustapha. 2008. Rangewide genetic structure in Paddlefish inferred from DNA microsatellite loci. Transactions of the American Fisheries Society 137:909–915. Helfman, G. S. 1979. Twilight activities of Yellow Perch, Perca flavescens. Journal of the Fisheries Research Board of Canada 36:173–179. Helfman, G. S. 1981. Twilight activities and temporal structure in a freshwater fish community. Canadian Journal of Fisheries and Aquatic Sciences 38:1405–1420. Helfman, G. S., and J. B. Clark. 1986. Rotational feeding: overcoming gape limited foraging in anguillid Eels. Copeia 1986:679–685. Helfman, G. S., B. B. Collette, D. E. Facey, and B. W. Bowen. 2009. The Diversity of Fishes. 2nd edition. Wiley-Blackwell, Chichester, West Sussex, United Kingdom. Helfman, G. S., D. E. Facey, L. S. Hales, and E. L. Bozeman. 1987. Reproductive ecology of the American Eel. American Fisheries Society Symposium 1:42–56. Helfman, G. S., D. L. Stoneburner, E. L. Bozeman, P. A. Christian, and R. Whalen. 1983. Ultrasonic telemetry of American Eel movements in a tidal creek. Transactions of the American Fisheries Society 112:105–110. Henley, D., L. Frankland, S. Hale, C. O’Bara, and T. Stefanavage. 2001. The case for multi-jurisdictional management of Ohio River Paddlefish. Proceedings of the Southeastern Association of Fish and Wildlife Agencies 55:243–256. Hennig, W. 1966. Phylogenetic Systematics. University of Illinois Press, Urbana. Henson, S. A., and R. R. Warner. 1997. Male and female alternative reproductive behaviors in fishes: a new approach using intersexual dynamics. Annual Review of Ecology and Systematics 28:571–592. Hermoso, V., M. Clavero, F. Blanco-Garrido, and J. Prenda. 2011. Invasive species and habitat degradation in Iberian streams: an analysis of their role in freshwater fish diversity loss. Ecological Applications 21:175–88. Hernandez, L. P., N. C. Bird, and K. L. Staab. 2007. Using Zebrafish to investigate cypriniform evolutionary novelties: functional development and evolutionary diversification of the kinethmoid. Journal of Experimental Zoology (Mol. Dev. Evol.) 308B:625–641. Herrington, S. J., K. N. Hettiger, E. J. Heist, and D. B. Keeney. 2008. Hybridization between Longnose and Alligator Gars in captivity, with comments on possible Gar hybridization in nature. Transactions of the American Fisheries Society 137:158–164. Herrington, S. J., and K. J. Popp. 2004. Observations on the reproductive behavior of nonindigenous Rough Shiner, Notropis baileyi, in the Chattahoochee River system. Southeastern Naturalist 3:267–276. Herting, G. E., and A. Witt, Jr. 1967. The role of physical fitness of forage fishes in relation to their vulnerability to predation by

549

Bowfin (Amia calva). Transactions of the American Fisheries Society 96:427–430. Herting, G. E., and A. Witt, Jr. 1968. Rate of digestion in the Bowfin. The Progressive Fish Culturist 30:26–28. Hesse, L. W., Q. P. Bliss, and G. J. Zuerlein. 1982. Some aspects of the ecology of adult fishes in the channelized Missouri River with special reference to the effects of two nuclear power generating plants, p. 225–276. In The Missouri River. L. W. Hesse, G. L. Hergenrader, H. S. Lewis, S. D. Reetz, and A. B. Schlesinger (eds.). The Missouri River Study Group, Norfolk, Nebraska. Hesse, L. W., and G. E. Mestl. 1993. The status of Nebraska fishes in the Missouri River. 1. Paddlefish (Polyodontidae: Polyodon spathula). Transactions of the Nebraska Academy of Sciences 20:53–65. Hesse, L. W., and B. A. Newcomb. 1982. On estimating the abundance of fish in the upper channelized Missouri River. North American Journal of Fisheries Management 2:80–83. Hesse, L. W., J. C. Schmulback, J. M. Carr, K. D. Keenlyne, D. G. Unkenholz, J. W. Robinson, and G. E. Mestl. 1989. Missouri River fishery resources in relation to past, present, and future stresses, p. 352–371. In Proceedings of the International Large River Symposium. D. P. Dodge (ed.). Canadian Special Publication of Fisheries and Aquatic Sciences 106. Heubel, K. U., K. Hornhardt, T. Ollmann, J. Parzefall, M. J. Ryan, and I. Schlupp. 2008. Geographic variation in female matecopying in the species complex of a unisexual fish, Poecilia formosa. Behaviour 145:1041–1064. Heubel, K. U., and I. Schlupp. 2006. Turbidity affects association behaviour of male Poecilia latipinna. Journal of Fish Biology 68:513–520. Heufelder, G. R. 1982. Lepisosteidae, p. 45–55. In Identification of Larval Fishes of the Great Lakes Basin with Emphasis on the Lake Michigan Drainage. N. A. Auer (ed.). Great Lakes Fishery Commission, Special Publication 82–3, Ann Arbor, Michigan. Heuschele, J., and U. Candolin. 2007. An increase in pH boosts olfactory communication in Sticklebacks. Biology Letters 3:411–413. Heuschele, J., M. Mannerla, P. Gienapp, and U. Candolin. 2009. Environment-dependent use of mate choice cues in Sticklebacks. Behavioral Ecology 20:1223–1227. Hewitt, G. M. 1996. Some genetic consequences of ice ages, and their role in divergence and speciation. Biological Journal of the Linnean Society 58:247–276. Higgs, D. M. 2004. Neuroethology and sensory ecology of teleost ultrasound detection, p. 173–188. In The Senses of Fishes: Adaptations for the Reception of Natural Stimuli. G. von der Emde, J. Mogdans, and B. G. Kapoor (eds.). Kluwer Academic Publishers, Boston, Massachusetts. Higgs, D. M., and L. A. Fuiman. 1998. Associations between sensory development and ecology in three species of clupeoid fish. Copeia 1998:133–144. Hightower, J. E., K. P. Zehfuss, D. A. Fox, and F. M. Parauka. 2002. Summer habitat use by Gulf Sturgeon in the Choctawhatchee River, Florida. Journal of Applied Ichthyology 18:595–600. Hildebrand, S. F. 1943. A review of the American Anchovies (Family Engraulidae). Bulletin of the Bingham Oceanographic Collection 8:1–165. Hildebrand, S. F. 1963. Family Engraulidae, p. 152–249. In Fishes of the Western North Atlantic Part Three: Soft-rayed Bony

550

LITERATURE CITED

Fishes. H. B. Bigelow, C. M. Cohen, G. W. Mead, D. Merriman, Y. H. Olsen, W. C. Schroeder, L. P. Schultz, and J. Tee-Van (eds.). Sears Foundation for Marine Research, New Haven, Connecticut. Hildebrand, S. F., and L. E. Cable. 1931. Development and life history of fourteen teleostean fishes at Beaufort, N.C. Bulletin of the United States Bureau of Fisheries 46:383–488. Hildebrand, S. F., and W. C. Schroeder. 1928. Fishes of Chesapeake Bay. Bulletin of the United States Bureau of Fisheries 43:1–366. Hildemann, W. H., and E. D. Wagner. 1954. Intraspecific sperm competition in Lebistes. American Naturalist 88:87–91. Hill, B. J., and I. C. Potter. 1970. Oxygen consumption in ammocoetes of the Lamprey, Ichthyomyzon hubbsi Raney. The Journal of Experimental Biology 53:47–57. Hill, L. G. 1972. Social aspects of aerial respiration of young Gar (Lepisosteus). The Southwestern Naturalist 16:239–247. Hill, L. G., L. Renfro, and R. Reynolds. 1972. Effects of dissolved oxygen tensions upon the rate of aerial respiration of young Spotted Gar, Lepisosteus oculatus (Lepisosteidae). The Southwestern Naturalist 17:273–278. Hill, L. G., G. D. Schnell, and A. A. Echelle. 1973. Effect of dissolved oxygen concentration on locomotory reactions of the Spotted Gar, Lepisosteus oculatus (Pisces: Lepisosteidae). Copeia 1973:119–124. Hill, S. E., and M. J. Ryan. 2006. The role of model female quality in the mate copying behaviour of Sailfin Mollies. Biology Letters 2:203–205. Hilliard, R. W., D. J. Bird, and I. C. Potter. 1983. Metamorphic changes in the intestine of three species of Lampreys. Journal of Morphology 176:181–196. Hilton, B. L., and D. J. Grosse. 1981. A reproductive pheromone in the Mexican poeciliid fish Poecilia chica. Copeia 1981:219–223. Hilton, E. J. 2001. Tongue bite apparatus of osteoglossomorph fishes: variation of a character complex. Copeia 2:372–381. Hilton, E. J. 2002. Osteology of the extant North American fishes of the genus Hiodon (LeSueur), 1818 (Teleostei: Osteoglossomorpha: Hiodontiformes). Fieldiana (Zoology) new series 100:1–142. Hilton, E. J. 2003. Comparative osteology and phylogenetic systematics of fossil and living bony-tongue fishes (Actinopterygii, Teleostei, Osteoglossomorpha). Zoological Journal of the Linnean Society 137:1–100. Hilton, E. J. 2004. The caudal skeleton of Acipenseriformes (Actinopterygii: Chondrostei): recent advances and new observations, p. 599–617. In Recent Advances in the Origin and Early Radiation of Vertebrates. G. Arratia, M. V. H. Wilson, and R. Cloutier (eds.). Verlag Dr. Friedrich Pfeil, München, Germany. Hilton, E. J. 2005. Observations on the skulls of Sturgeons (Acipenseridae): shared similarities of Pseudoscaphirhynchus kaufmanni and juvenile specimens of Acipenser stellatus. Environmental Biology of Fishes 72:135–144. Hilton, E. J., and W. E. Bemis. 1999. Skeletal variation in Shortnose Sturgeon (Acipenser brevirostrum) from the Connecticut River: implications for comparative osteological studies of fossil and living fishes, p. 69–94. In Mesozoic Fishes 2—Systematics and Fossil Record. G. Arratia and H. P. Schultze (eds.). Verlag Dr. Friedrich Pfeil, München, Germany. Hilton, E. J., and P. L. Forey. 2009. Redescription of †Chondrosteus acipenseroides Egerton, 1858 (Acipenseriformes, †Chondroste-

idae) from the Lower Lias of Lyme Regis (Dorset, England), with comments on the early evolution of Sturgeons and Paddlefishes. Journal of Systematic Palaeontology 7:427–453. Hilton, E. J., and L. Grande. 2006. Review of the fossil record of Sturgeons, family Acipenseridae (Actinopterygii: Acipenseriformes), from North America. Journal of Paleontology 80:672–683. Hilton, E. J., and L. Grande. 2008. Fossil Mooneyes (Teleostei, Hiodontiformes, Hiodontidae) from the Eocene of western North America, with a reassessment of their taxonomy, p. 221–251. In Fishes and the Break-Up of Pangea. L. Cavin, A. Longbottom, and M. Richter (eds.). Geological Society of London, Special Publication 295. Hilton, E. J., L. Grande, and W. E. Bemis. 2011. Skeletal anatomy of the Shortnose Sturgeon, Acipenser brevirostrum Lesueur, 1818, and the systematics of Sturgeons (Acipenseriformes, Acipenseridae. Fieldiana Life and Earth Sciences 3:1–168. Hines, A. H., R. B. Whitlatch, S. F. Thrush, J. E. Hewitt, V. J. Cummings, P. K. Dayton, and P. Legendre. 1997. Nonlinear foraging response of a large marine predator to benthic prey: Eagle Rays, pits, and bivalves in a New Zealand sandflat. Journal of Experimental Marine Biology and Ecology 216:191–201. Hinton, D. E. 1998. Multiple stressors in the Sacramento River watershed, p. 303–317. In Fish Ecotoxicology. T. Braunbeck, D. E. Hinton, and B. Streit (eds.). Birkhäuser Verlag, Basel, Switzerland. Hitt, N. P., and P. L. Angermeier. 2011. Fish community and bioassessment responses to stream network position. Journal of the North American Benthological Society 30:296–309. Hlohowskyj, C. P., M. M. Coburn, and T. M. Cavender. 1989. Comparison of a pharyngeal filtering apparatus in seven species of the herbivorous cyprinid genus, Hybognathus (Pisces: Cyprinidae). Copeia 1989:172–183. Hobe, H., and B. R. McMahon. 1988. Mechanisms of acid-base and ionoregulation in White Suckers (Catostomus commersoni) in natural soft water II. Exposure to a fluctuating ambient pH regime. Journal of Comparative Physiology B 158:67–79. Hobe, H., C. M. Wood, and B. R. McMahon. 1984. Mechanisms of acid-base and ionoregulation in White Suckers (Catostomus commersoni) in natural soft water I. Acute exposure to low ambient pH. Journal of Comparative Physiology B 154:35–46. Hochachka, P. W., and T. P. Mommsen. 1983. Protons and anaerobiosis. Science 219:1391–1397. Hochleithner, M., and J. Gessner. 1999. The Sturgeons and Paddlefishes of the World, Biology and Aquaculture. Aqua Tech Publications, Kitzbuehel, Austria. Hocutt, C. H., and E. O. Wiley (eds.). 1986. The Zoogeography of North American Freshwater Fishes. John Wiley & Sons, New York. Hocutt, C. H. 1987. Evolution of the Indian Ocean and the drift of India: a vicariant event. Hydrobiologia 150:203–223. Hodgens, L. S., S. C. Blumenshine, and J. C. Bednarz. 2004. Great Blue Heron predation on stocked Rainbow Trout in an Arkansas tailwater fishery. North American Journal of Fisheries Management 24:63–75. Hoeinghaus, D. J., K. O. Winemiller, and J. S. Birnbaum. 2007. Local and regional determinants of stream fish assemblage structure: inferences based on taxonomic vs. functional groups. Journal of Biogeography 34:324–338.

LITERATURE CITED

Hoetker, G. M., and K. W. Gobalet. 1999. Fossil Razorback Sucker (Pisces: Catostomidae, Xyrauchen texanus) from southeastern California. Copeia 1999:755–759. Hofer, R. 1991. Digestion, p. 413–425. In Cyprinid Fishes: Systematics, Biology and Exploitation. I. J. Winfield and J. S. Nelson (eds.). Chapman and Hall, London. Hoff, M. H., and C. R. Bronte. 1999. Structure and stability of the midsummer fish communities in Chequamegon Bay, Lake Superior, 1973–1996. Transactions of the American Fisheries Society 128:362–373. Hoff man, G. L. 1999. Parasites of North American Freshwater Fishes. 2nd edition. Cornell University Press, Ithaca, New York. Hoff man, G. L., and G. Schubert. 1984. Some parasites of exotic fishes, p. 233–261. In Distribution, Biology, and Management of Exotic Fishes. W. R. Courtenay, Jr. and J. R. Stauffer, Jr. (eds.). The Johns Hopkins University Press, Baltimore, Maryland. Hoff man, R. D., and R. D. Curnow. 1979. Mercury in herons, egrets, and their foods. The Journal of Wildlife Management 43(1):85–93. Hoff nagle, T. L., and T. J. Timmons. 1989. Age, growth, and catch analysis of the commercially exploited Paddlefish population in Kentucky Lake, Kentucky-Tennessee. North American Journal of Fisheries Management 9:316–326. Hofmann, M. H., W. Wojtenek, and L. A. Wilkens. 2002. Central organization of the electrosensory systems in the Paddlefish (Polyodon spathula). The Journal of Comparative Neurology 446:25–36. Hogue, C. C., D. R. Sutherland, and B. M. Christensen. 1993. Ecology of metazoan parasites infecting Catostomus spp. (Catostomidae) from southwestern Lake Superior. Canadian Journal of Zoology 71:1646–1652. Hogue, J. J., and J. P. Buchanan. 1977. Larval development of Spotted Sucker, Minytrema melanops. Transactions of the American Fisheries Society 102:778–785. Hogue, J. J., Jr., J. V. Conner, and V. R. Kranz. 1981. Descriptions and methods for identifying larval Blue Sucker, Cycleptus elongatus (LeSueur). Rapports et. Proces, Verbaux des Reunions, Conseil Permanent Pour Lí Exporation de la Mer 178:585–587. Hohler, D. B. 1981. A dwarfed population of Catostomus rimiculus (Catostomidae: Pisces) in Jenny Creek, Jackson County, Oregon. Unpubl. Master’s thesis, Oregon State University, Corvallis. Holčik, J. 2006. Is the naturalization of the Paddlefish in the Danube River Basin possible? Journal of Applied Ichthyology 22 (Supplement 1):40–43. Holčík, J., and V. Šorić. 2004. Redescription of Eudontomyzon stankokaramani (Petromyzontes, Petromyzontidae)—a little known Lamprey from the Drin River drainage, Adriatic Sea basin. Folia Zoologica 53:399–410. Holden, P. B. 1991. Ghosts of the Green River: impacts of Green River poisoning on management of native fishes, p. 43–54. In Battle Against Extinction: Native Fish Management in the American West. W. L. Minckley and J. E. Deacon (eds.). The University of Arizona Press, Tuscon. Holden, P. B., and E. J. Wick. 1982. Life history and prospects for recovery of Colorado squawfish, p. 98–108. In Fishes of the Upper Colorado River System: Present and Future. W. H. Miller, H. M. Tyus, and C. A. Carlson (eds.). American Fisheries Society, Bethesda, Maryland. Holder, D. R. 1970. A study of fish movements from the Okefenokee Swamp into the Suwannee River. Proceedings of the South-

551

eastern Association of Game and Fish Commissioners 24: 591–608. Holey, M., B. Hollender, M. Imhof, R. Jesien, R. Konopacky, M. Toneys, and D. Coble. 1979. Never give a Sucker an even break. Fisheries 4:2–6. Holland, H. T. 1964. Ecology of the Bowfin (Amia calva Linnaeus) in southeastern Missouri. Unpubl. Master’s thesis, University of Missouri, Columbia. Holland, P. W. H. 1999. Gene duplication: past, present and future. Cell and Developmental Biology 10:541–547. Holland, P. W. H. 2003. More genes in vertebrates? Journal of Structural and Functional Genomics 3:75–84. Hollander, E. E., and J. W. Avault, Jr. 1975. Effects of salinity on survival of buffalo fish eggs through yearlings. The Progressive Fish-Culturist 37:47–51. Holldobler, B. 1995. The chemistry of social regulation: multicomponent signals in ant societies. Proceedings of the National Academy of Sciences of the United States of America 92:19–22. Hollingsworth, Jr., P. R., and C. D. Hulsey. 2011. Reconciling gene trees of eastern North American minnows. Molecular Phylogenetics and Evolution 61:149–156. Holloway, A. D. 1954. Notes on the life history and management of the Shortnose and Longnose Gars in Florida Waters. Journal of Wildlife Management 18:438–339. Holmes, J. A., and P. Lin. 1994. Thermal niche of larval Sea Lamprey, Petromyzon marinus. Canadian Journal of Fisheries and Aquatic Sciences 51:253–262. Holmes, J. A., H. Chu, S. A. Khanam, R. G. Manzon, and J. H. Youson. 1999. Spontaneous and induced metamorphosis in the American Brook Lamprey, Lampetra appendix. Canadian Journal of Zoology 77:959–971. Holtby, L. B., and M. C. Healey. 1986. Selection for adult size in female Coho Salmon (Oncorhynchus kisutch). Canadian Journal of Fisheries and Aquatic Sciences 43:1946–1959. Holzkamm, T. E., and L. G. Waisberg. 2004. Native American utilization of Sturgeon, p. 22–39. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. H. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. Honda, H. 1980. Female sex-pheromone of Rainbow Trout, Salmo gairdneri, involved in courtship behavior. Bulletin. Japanese Society of Scientific Fisheries 46:1109–1112. Hoopes, D. T. 1960. Utilization of mayflies and caddis flies by some Mississippi River fishes. Transactions of the American Fisheries Society 89:32–34. Hoover, C. 1999. Import and export of Sturgeon and Paddlefish in the United States, p. 162–170. In Proceedings of the Symposium on the Harvest and Conservation of North American Paddlefish and Sturgeon, May 7–8, 1998, Chattanooga, Tennessee. D. F. Williamson, G. W. Benz, and C. M. Hoover (eds.). TRAFFIC North America / World Wildlife Fund, Washington, D.C. Hoover, J. J. 2004. Rare treasure from river bottom: juvenile Goldeye (Hiodon alosoides) for medium-sized aquaria. American Currents 30(3):1–6. Hoover, J. J., K. A. Boysen, J. A. Beard, and H. Smith. 2011a. Assessing the risk of entrainment by cutterhead dredges to juvenile Lake Sturgeon (Acipenser fulvescens) and juvenile Pallid Sturgeon (Scaphirhynchus albus). Journal of Applied Ichthyology 27:369–375.

552 LITERATURE CITED

Hoover, J. J., K. A. Boysen, C. E. Murphy, and S. G. George. 2009a. Morphological variation in juvenile Paddlefish, p. 157–171. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Hoover, J. J., J. Collins, K. A. Boysen, A. W. Katzenmeyer, and K. J. Killgore. 2011b. Critical swimming speeds of adult Shovelnose Sturgeon in rectilinear and boundary-layer flow. Journal of Applied Ichthyology 27:226–230. Hoover, J. J., S. G. George, and K. J. Killgore. 2000. Rostrum size of Paddlefish (Polyodon spathula) (Acipenseriformes: Polyodontidae) from the Mississippi Delta. Copeia 2000:288–290. Hoover, J. J., S. G. George, and K. J. Killgore. 2007. Diet of Shovelnose Sturgeon and Pallid Sturgeon in the free-flowing Mississippi River. Journal of Applied Ichthyology 23:494–499. Hoover, J. J., A. Turnage, and K. J. Killgore. 2009b. Swimming performance of juvenile Paddlefish: quantifying risk of entrainment, p. 141–155. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Hopkins, G. S. 1890. Structure of the stomach of Amia calva. Proceedings of the American Society of Microscopists 12:165–169. Horn, M. H. 1972. The amount of space available for marine and freshwater fishes. Fishery Bulletin, U.S. 70:1295–1297. Horn, M. H., and C. D. Riggs. 1973. Effects of temperature and light on the rate of air breathing of the Bowfin, Amia calva. Copeia 1973:653–657. Horton, R. E. 1945. Erosional development of streams and their drainage basins. Bulletin of the Geological Society of America 56:275–370. Houde, A. E. 1987. Mate choice based on naturally occurring colour-pattern in a guppy population. Evolution 41:1–10. Houde, A. E. 1988. The effects of female choice and male-male competition on the mating success of male guppies. Animal Behaviour 36:888–896. Houde, E. D. 1974. Effects of temperature and delayed feeding on growth and survival of larvae of three species of subtropical marine fishes. Marine Biology 26:271–285. Houde, E. D. 1977. Food concentration and stocking density effects on survival and growth of laboratory-reared larvae of Bay Anchovy, Anchoa mitchilli and Lined Sole, Achirus lineatus. Marine Biology 43:333–341. Houde, E. D., and J. A. Lovdal. 1984. Seasonality of occurrence, foods and food preferences of ichthyoplankton in Biscayne Bay, Florida. Estuarine, Coastal and Shelf Science 18:403–419. Houde, E. D., and J. D. A. Lovdal. 1985. Patterns of variability in ichthyoplankton occurrence and abundance in Biscayne Bay, Florida. Estuarine, Coastal and Shelf Science 20:79–103. Houde, E. D., and R. C. Schekter. 1983. Oxygen uptake and comparative energetics among eggs and larvae of three subtropical marine fishes. Marine Biology 72:283–293. Houde, E. D., and C. E. Zastrow. 1991. Bay Anchovy, p. 8.1–8.14. In Habitat Requirements for Chesapeake Bay Living Resources. 2nd edition. S. L. Funderburk, J. A. Mihursky, S. J. Jordan, and D. Riley (eds.). Chesapeake Bay Program, Annapolis, Maryland. not seen

Houser, A. 1965. Growth of Paddlefish in Fort Gibson Reservoir, Oklahoma. Transactions of the American Fisheries Society 94:91–93. Houser, A., and M. G. Bross. 1959. Observations on growth and reproduction of the Paddlefish. Transactions of the American Fisheries Society 88:50–52. Hove, M. C., R. A. Engelking, M. Peteler, and L. Sovell. 1994. Life history research on Ligumia recta and Lasmigona costata. Triannual Unionid Report 4:1. Howard, A. D. 1914a. Some cases of narrowly restricted parasitism among commercial species of fresh water mussels. Transactions of the American Fisheries Society 44:41–44. Howard, A. D. 1914b. Experiments in propagation of fresh-water mussels of the Quadrula group. Report of the U.S. Commissioner of Fisheries for 1913. Appendix 4:1–52. Howell, W. M., and T. E. Denton. 1969. Chromosomes of ammocoetes of the Ohio brook Lamprey Lampetra aepyptera. Copeia 1969:393–395. Howell, W. M., and C. R. Duckett. 1971. Somatic chromosomes of the Lamprey, Ichthyomyzon gagei (Agnatha: Petromyzonidae). Experientia 27:222–223. Howes, G. 1984. Phyletics and biogeography of the aspinine cyprinid fishes. Bulletin of the British Museum of Natural History (Zoology) 47:283–303. Hoxmeier, R. J. H., and D. R. DeVries. 1996. Status of Paddlefish in the Alabama waters of the Tennessee River. North American Journal of Fisheries Management 16:935–938. Hoxmeier, R. J. H., and D. R. DeVries. 1997. Habitat use, diet, and population structure of adult and juvenile Paddlefish in the lower Alabama River. Transactions of the American Fisheries Society 126:288–301. Hoysak, D. J., and N. R. Liley. 2001. Fertilization dynamics in Sockeye Salmon and a comparison of sperm from alternative male phenotypes. Journal of Fish Biology 58:1286–1300. Hoysak, D. J., and N. E. Stacey. 2008. Large and persistent effect of a female steroid pheromone on ejaculate size in Goldfish, Carassius auratus. Journal of Fish Biology 73:1573–1584. Hoyt, R. D. 1972. Anatomy and osteology of the cephalic lateralline system of the Silverjaw Minnow, Ericymba buccata (Pisces: Cyprinidae). Copeia 1972:812–816. Hoyt, R. D. 1984. Notes on various growth features of the Paddlefish in the Ohio River. Transactions of the Kentucky Academy of Science 45:75–76. Hoyt, R. D., G. J. Overmann, and G. A. Kindschi. 1979. Observations on the larval ecology of the Smallmouth Buffalo, p. 1–16. In Proceedings of the Third Symposium on Larval Fish. R. D. Hoyt (ed.), Western Kentucky University, Bowling Green. Hoyt, R. D., A. T. Waite, and B. M. DiPasqualie. 1976. Population dynamics and catch susceptibility of Smallmouth Buffalo in Rough River Reservoir. Kentucky Department of Fish and Wildlife Resources Fisheries Bulletin No. 62, Frankfort. Hrabik, R. A., D. P. Herzog, D. E. Ostendorf, and M. D. Petersen. 2007. Larvae provide first evidence of successful reproduction by Pallid Sturgeon, Scaphirhynchus albus, in the Mississippi River. Journal of Applied Ichthyology 23:436–443. Hrbek, T., and A. Larson. 1999. The evolution of diapause in the killifish family Rivulidae (Atherinomorpha, Cyprinodontiformes): a molecular phylogenetic and biogeographic perspective. Evolution 54:1200–1216.

LITERATURE CITED

Hubbs, C. 1957. Duration of sperm viability as a factor in frequency of fish hybridization. Texas Journal of Science 9:472–474. Hubbs, C. 1964. Interaction between a bisexual fish species and its gynogenetic sexual parasite. Bulletin of the Texas Memorial Museum 8:1–72. Hubbs, C. 1984. Changes in fish abundance with time of day and among years at a station in Lake Texoma. Annual Proceedings of the Texas Chapter of the American Fisheries Society 6:42–57. Hubbs, C., and K. Strawn. 1956. Interfertility between two sympatric fishes, Notropis lutrensis and Notropis venustus. Evolution 10:341–344. Hubbs, C. L. 1921. An ecological study of the fresh-water atherine fish, Labidesthes sicculus. Ecology 2:267–276. Hubbs, C. L. 1930. Materials for a revision of the catostomid fishes of eastern North America. Miscellaneous Publications Museum Zoology, University of Michigan, No. 20. Hubbs, C. L. 1943. Terminology of early stages of fishes. Copeia 1943:260. Hubbs, C. L. 1951. The American cyprinid fish, Notropis germanus Hay interpreted as an intergeneric hybrid. The American Midland Naturalist 45:446–454. Hubbs, C. L. 1955. Hybridization between fish species in nature. Systematic Zoology 4:1–20. Hubbs, C. L., and R. M. Bailey. 1952. Identification of Oxygeneum pulverulentum Forbes, from Illinois, as a hybrid cyprinid fish. Papers of the Michigan Academy of Science, Arts and Letters 37:143–153. Hubbs, C. L., and G. P. Cooper. 1936. Minnows of Michigan. Cranbrook Institute of Science Bulletin 8:1–95. Hubbs, C. L., and A. A. Echelle. 1972. Endangered non-game fishes of the upper Rio Grande basin, p. 147–167. In Symposium on Endangered Vertebrates of Southwestern United States. New Mexico Game and Fish Department, Santa Fe. Hubbs, C. L., and L. C. Hubbs. 1932. Apparent parthenogenesis in nature in a form of fish of hybrid origin. Science 76:628–630. Hubbs, C. L., and L. C. Hubbs. 1947. Natural hybrids between two species of catostomid fishes. Papers of the Michigan Academy of Sciences, Arts, and Letters 31:147–167. Hubbs, C. L., L. C. Hubbs, and R. E. Johnson. 1943a. Hybridization in nature between species of catostomid fishes. Contributions of the University of Michigan Laboratory of Vertebrate Biology 22:1–76. Hubbs, C. L., and K. F. Lagler. 1947. Fishes of the Great Lakes Region. Cranbrook Institute of Science, Bloomfield Hills, Michigan. Hubbs, C. L., and K. F. Lagler 1958. Fishes of the Great Lakes Region. Cranbrook Institute of Science, Bloomfield Hills, Michigan. Hubbs, C. L., and K. F. Lagler. 1964. Fishes of the Great Lakes Region. The University of Michigan Press, Ann Arbor, Michigan. Hubbs, C. L., and R. R. Miller. 1943. Mass hybridization between two genera of cyprinid fishes in the Mohave Desert, California. Paper of the Michigan Academy of Science, Arts, and Letters 28:343–378. Hubbs, C. L., and R. R. Miller. 1953. Hybridization in nature between the fish genera Catostomus and Xyrauchen. Papers of the Michigan Academy of Sciences, Arts, and Letters 38:207–233.

553

Hubbs, C. L., and R. R. Miller. 1948. Two new, relict genera of cyprinid fishes from Nevada. Occasional Papers of the Museum of Zoology, University of Michigan 507:1–30. Hubbs, C. L., and R. R. Miller. 1977. Six distinctive cyprinid fish species referred to Dionda inhabiting segments of the Tampico embayment drainage of Mexico. Transactions of the San Diego Society of Natural History 18:267–336. Hubbs, C. L., R. R. Miller, and L. C. Hubbs. 1974. Hydrographic history and relict fishes of the North-Central Great Basin. Memoirs of the California Academy of Sciences 7:1–259. Hubbs, C. L., and I. C. Potter. 1971. Distribution, phylogeny and taxonomy, p. 1–65. In The Biology of Lampreys. Vol. 1. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Hubbs, C. L. and E. C. Raney. 1948. Subspecies of Notropis altipinnis, a cyprinid fish of the eastern United States. Occasional Papers of the Museum of Natural History, University of Michigan 506:1–20. Hubbs, C. L., and M. B. Trautman. 1937. A revision of the Lamprey genus Ichthyomyzon. Miscellaneous Publications of the Museum of Zoology, University of Michigan 35:7–109. Hubbs, C. L., B. W. Walker, and R. E. Johnson. 1943b. Hybridization in nature between species of American cyprinodont fishes. Contributions of the University of Michigan Laboratory of Vertebrate Biology 23:1–21. Hubenova, T., A. Zaikov, and P. Vasileva. 2007. Management of Paddlefish fry and juveniles in Bulgarian conditions. Aquaculture International 15:249–253. Hubert, W. A., D. D. Harris, and T. A. Wesche. 1994. Diurnal shifts in use of summer habitat by age-0 Brown Trout in a regulated mountain stream. Hydrobiologia 284:147–156. Hubley, R. C., Jr. 1961. Incidence of Lamprey scarring on fish in the upper Mississippi River, 1956–58. Transactions of the American Fisheries Society 90:83–85. Hughes, A. L. 1985a. Male size, mating success, and mating strategy in the mosquitofish, Gambusia affinis (Poeciliidae). Behavioral Ecology and Sociobiology 17:271–278. Hughes, A. L. 1985b. Seasonal trends in body size of adult male mosquitofish, Gambusia affinis, with evidence for their social control. Environmental Biology of Fishes 14:251–258. Hughes, R. L., and I. C. Potter. 1969. Studies on gametogenesis and fecundity in the Lampreys, Mordacia praecox and M. mordax (Petromyzonidae). Australian Journal of Zoology 17:447– 464. Hugueny, B., T. Oberdorff, and P. A. Tedesco. 2010. Community ecology of river fishes: a large-scale perspective, p. 29–62. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. Humphries, J. M., F. L. Bookstein, B. Chernoff, G. R. Smith, J. F. Elder, Jr., and S. G. Poss. 1981. Multivariate discrimination by shape in relation to size. Systematic Zoology 30:291–308. Humphries, J. M., and R. C. Cashner. 1994. Notropis suttkusi, a new cyprinid from the Ouachita uplands of Oklahoma and Arkansas, with comments on the status of Ozarkian populations of N. rubellus. Copeia 1994:82–90. Hunt, B. P. 1954. Food relationships between Florida Spotted Gar and other organisms in the Tamiami Canal, Dade County, Florida. Transactions of the American Fisheries Society 82:13–33.

554 LITERATURE CITED

Hunt, B. P. 1960. Digestion rate and food consumption of Florida Gar, Warmouth and Largemouth Bass. Transactions of the American Fisheries Society 89:206–211. Hunter, G. W., III. 1932. A new trematodes (Plesiocreadium parvum, sp. nov.) from fresh water fish. Transactions of the American Microscopical Society 51:16–21. Hunter, J. R., and W. J. Wisby. 1961. Utilization of the nests of Green Sunfish (Lepomis cyanellus) by the Redfin Shiner (Notropis umbratilus cyanocephalus). Copeia 1961:113–115. Hunter, J. R. 1963. The reproductive behavior of the Green Sunfish, Lepomis cyanellus. Zoologica 48:13–24. Hunter, J. R., and A. D. Hasler. 1965. Spawning association of the Redfin Shiner, Notropis umbratilis, and the Green Sunfish, Lepomis cyanellus. Copeia 1965:265–281. Huntingford, F. A. 1976. A comparison of the reaction of Sticklebacks in different reproductive conditions towards conspecifics and predators. Animal Behaviour 24:694–697. Huntingford, F. A., N. B. Metcalfe, J. E. Thorpe, W. D. Graham, and C. E. Adams. 1990. Social-dominance and body size in Atlantic Salmon parr, Salmo salar L. Journal of Fish Biology 36:877–881. Huntingford, F. A., and A. Turner. 1987. Animal Conflict. Chapman and Hall, London, United Kingdom. Huntsman, G. R. 1967. Nuptial tubercles in carpsuckers (Carpoides). Copeia 1967:457–458. Hurlbert, S. H. 1984. Pseudoreplication and the design of ecological field experiments. Ecological Monographs 54:187–211. Hurley, I. A., R. L. Mueller, K. A. Dunn, E. J. Schmidt, M. Friedman, R. K. Hol, V. E. Prince, Y. Yang, M. G. Thomas, and M. I. Coates. 2007. A new time-scale for Ray-Finned Fish evolution. Proceedings of the Royal Society. B 274:489–498. Hurley, K. L., R. J. Sheehan, and R. C. Heidinger. 2004a. Accuracy and precision of age estimates for Pallid Sturgeon from pectoral fin rays. North American Journal of Fisheries Management 24:715–718. Hurley, K. L., R. J. Sheehan, R. C. Heidinger, P. S. Wills, and B. Clevenstine. 2004b. Habitat use by middle Mississippi River Pallid Sturgeon. Transactions of the American Fisheries Society 133:1033–1041. Hussakof, L. 1911. The spoonbill fishery of the lower Mississippi. Transactions of the American Fisheries Society 40:245–248. Hutchings, J. A., and R. A. Myers. 1988. Mating success of alternative maturation phenotypes in male Atlantic Salmon, Salmo salar. Oecologia 75:169–174. Hutton, R. F. 1964. A second list of parasites from marine and coastal animals of Florida. Transactions of the American Microscopical Society 83:439–447. Huxley, T. H. 1861. Preliminary essay upon the systematic arrangement of the fishes of the Devonian epoch. Memoirs of the Geological Survey of the United Kingdom, Decade 10:1–40. Hyde, D. A., T. W. Moon, and S. F. Perry. 1987. Physiological consequences of prolonged aerial exposure in the American Eel, Anguilla rostrata: blood respiratory and acid-base status. Journal of Comparative Physiology B: Biochemical, Systemic, and Environmental Physiology 157:635–642. IGFA (International Game Fish Association). 2006. 2006 World record game fishes: freshwater, saltwater, and flyfishing. International Game Fish Association, Dania Beach, Florida. IGFA (International Game Fish Association). 2011. All-tackle world records database. Available from http://www.igfa.org/

Records/Fish-Records.aspx?LC=ATR&Fish=Bowfin;as of 29 July 2011. Ijiri, S., C. Berard, and J. M. Trant. 2000. Characterization of gonadal and extra-gonadal forms of the cDNA endcoding the Atlantic Stingray (Dasyatis sabina) cytochrome P450 aromatase (CY19). Molecular and Cellular Endocrinology 164:169–181. Illick, H. J. 1956. A comparative study of the cephalis lateral-line system of North American Cyprinidae. American Midland Naturalist 56:204–223. Imms, A. D. 1904. Notes on the gill-rakers of the spoonbill sturgeon, Polyodon spathula. Proceedings of the Zoological Society of London 2:22–35. Inebnit T. E., III. 2009. Aspects of the reproductive and juvenile ecology of Alligator Gar in the Fourche LaFave River, Arkansas. Unpubl. Master’s thesis, University of Central Arkansas, Conway. Infante, D. M., and J. D. Allan. 2010. Response of stream fish assemblages to local-scale habitat as influenced by landscape: a mechanistic investigation of stream fish assemblages, p. 371– 397. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. Inoue, J. G., M. Miya, K. Tsukamoto, and M. Nishida. 2003. Basal actinopterygian relationships: a mitogenomic perspective on the phylogeny of the ‘ancient fish’. Molecular Phylogenetics and Evolution 26:110–120. Inouye, B. 2001. Relationships between ecological interaction modifications and diff use coevolution; similarities, differences, and causal links. Oikos 95:353–360. Ireland, S. C., R. C. P. Beamesderfer, V. L. Paragamian, V. D. Wakkinen, and J. T. Siple. 2002. Success of hatchery-reared juvenile White Sturgeon (Acipenser transmontanus) following release in the Kootenai River, Idaho, USA. Journal of Applied Ichthyology 18:642–650. Irving, P. W., and A. E. Magurran. 1997. Context-dependent fright reactions in captive European minnows: the importance of naturalness in laboratory experiments. Animal Behaviour 53:1193–1201. Israel, J. A., J. F. Cordes, M. A. Blumberg, and B. May. 2004. Geographic patterns of genetic differentiation among collections of Green Sturgeon. North American Journal of Fisheries Management 24:922–931. Itzkowitz, M. 1971. Preliminary study of the social behavior of male Gambusia affinis (Baird and Girard) (Pisces: Poeciliidae) in aquaria. Chesapeake Science 12:219–224. Itzkowitz, M. 1974. The effects of other fish on the reproductive behavior of the male Cyprinodon variegatus (Pisces: Cyprinodontidae). Behaviour 48:1–22. IUCN (International Union for Conservation of Nature). 2010. IUCN Red List of Threatened Species. Version 2009.2. Available from http://www.iucnredlist.org/; as of February 2010. IUCN (International Union for Conservation of Nature). 2011. IUCN Red List of Threatened Species. Version 2011.2 Available from http://www.iucnredlist.org/; accessed December 2011, July 2011. Jablonski, D., and J. J. Sepkoski, Jr. 1996. Paleobiology, community ecology, and scales of ecological pattern. Ecology 77: 1367–1378.

LITERATURE CITED

Jackson, D. A., and H. H. Harvey. 1989. Biogeographic associations in fish assemblages: local vs. regional processes. Ecology 70:1472–1484. Jackson, D. A., P. R. Peres-Neto, and J. D. Olden. 2001. What controls who is where in freshwater fish communities—the roles of biotic, abiotic, and spatial factors. Canadian Journal of Fisheries and Aquatic Sciences 58:157–170. Jackson, D. A., K. M. Somers, and H. H. Harvey. 1992. Null models and fish communities: evidence of nonrandom patterns. The American Naturalist 139:930–951. Jackson, H. E., and S. L. Scott. 2003. Patterns of elite faunal utilization at Moundville, Alabama. American Antiquity 68:552–572. Jackson, N. D., J. E. Garvey, and R. E. Colombo. 2007. Comparing aging precision of calcified structures in Shovelnose Sturgeon. Journal of Applied Ichthyology 23:525–528. Jackson, J. R., A. J. VanDeValk, T. E. Brooking, O. A. vanKeeken, and L. G. Rudstam. 2002. Growth and feeding dynamics of Lake Sturgeon, Acipenser fulvescens, in Oneida Lake, New York: results from the first five years of a restoration program. Journal of Applied Ichthyology 18:439–443. Jackson, S. W. 1957. Comparison of the age and growth of four fishes from Lower and Upper Spavinaw Lakes, Oklahoma. Proceedings of the Southeastern Association Game and Fish Commissioners 11:232–249. Jacob, A., S. Nusslé, A. Britschgi, G. Evanno, R. Müller, and C. Wedekind. 2007. Male dominance linked to size and age, but not to ‘good genes’ in Brown Trout (Salmo trutta). BMC Evolutionary Biology 7:207 Jacobs, R. P., W. A. Hyatt, N. T. Hagstrom, E. B. O’Donnell, E. C. Schluntz, P. Howell, and D. R. Molnar. 2003. Trends in abundance, distribution and growth of freshwater fishes from the Connecticut River in Connecticut (1988–2002). Connecticut Department of Environmental Protection, Bureau of Natural Resources, Inland Fisheries Division and Marine Fisheries Division, Hartford. Jaensson, A., and K. H. Olsén. 2010. Effects of copper on olfactorymediated endocrine responses and reproductive behaviour in mature male Brown Trout, Salmo trutta parr to conspecific females. Journal of Fish Biology 76:800–817. Jaensson, A., A. P. Scott, A. Moore, H. Kylin, and K. H. Olsén. 2007. Effects of a pyrethroid pesticide on endocrine responses to female odours and reproductive behaviour in male parr of Brown Trout (Salmo trutta L.). Aquatic Toxicology 81:1–9. Jager, H. I., M. S. Bevelhimer, K. B. Lepla, J. A. Chandler, and W. Van Winkle. 2007. Evaluation of reconnection options for White Sturgeon in the Snake River using a population viability model, p. 319–335. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Jager, H. I., J. A. Chandler, K. B. Lepla, and W. Van Winkle. 2001. A theoretical study of river fragmentation by dams and its effects on White Sturgeon population. Environmental Biology of Fishes 60:347–361. Jamieson, B. G. M. 1991. Fish Evolution and Systematics: Evidence from Spermatozoa. Cambridge University Press, Cambridge, England. Jamieson, I. 1995. Do female fish prefer to spawn in nests with eggs for reasons of mate choice copying or egg survival? American Naturalist 145:824–832.

555

Jamieson, I. G. 1994. Mate choice in three-spined Sticklebacks: a reply to Goldschmidt et al. Animal Behaviour 47:991–993. Jamieson, I. G., and P. W. Colgan. 1989. Eggs in the nests of males and their effects on mate choice in the three-spined Stickleback. Animal Behaviour 38:859–865. Jamieson, I. G., and P. W. Colgan. 1992. Sneak spawning and egg stealing by male Threespine Sticklebacks. Canadian Journal of Zoology 70:963–967. Janech, M. G., W. R. Fitzgibbon, R. Chen, M. W. Nowak, D. H. Miller, R. V. Paul, and D. W. Ploth. 2003. Molecular and functional characterization of a urea transporter from the kidney of the Atlantic Stingray. American Journal of Physiology: Renal Physiology 284:F996–F1005. Janech, M. G., W. R. Fitzgibbon, M. Nowak, D. H. Miller, and D. W. Ploth. 2006a. Cloning and functional characterization of a second urea transporter isoform (strUT-2) from the kidney of the Atlantic Stingray. American Journal of Physiology: Regulatory, Integrative, and Comparative Physiology 291:R844–R853. Janech, M. G.,W. R. Fitzgibbon, D. W. Ploth, E. R. Lacy, and D. H. Miller. 2006b. Effect of low environmental salinity on plasma composition and renal function of the Atlantic Stingray, a euryhaline elasmobranch. American Journal of Physiology: Renal Physiology 291:F770–F780. Janech, M. G., and P. M. Piermarini. 2002. Renal water and solute excretion in the Atlantic Stingray in fresh water. Journal of Fish Biology 61:1053–1057. Janvier, P. 1981. The phylogeny of the Craniata, with particular reference to the significance of fossil “agnathans”. Journal of Vertebrate Paleontology 1:121–159. Janvier, P. 2009. Les premiers vertébrés et les premières étapes de l’évolution du crâne. Comptes Rendus Palevol 8:209–219. Janvier, P. 2010. microRNAs revive old views about jawless vertebrate divergence and evolution. Proceedings of the National Academy of Sciences 107:19137–19138. Janvier, P., and R. Lund. 1983. Hardistiella montanensis n. gen. et sp. (Petromyzontida) from the Lower Carboniferous of Montana, with remarks on the affinities of the Lampreys. Journal of Vertebrate Paleontology 2:407–413. Janvier, P., R. Lund, and E. D. Grogan. 2004. Further consideration of the earliest known Lamprey, Hardistiella montanensis Janvier and Lund, 1983, from the Carboniferous of Bear Gulch, Montana, U.S.A. Journal of Vertebrate Paleontology 24:742–743. Jaroszewska, M., and K. Dabrowski. 2008. Morphological analysis of the functional design of the connection between the alimentary tract and the gas bladder in air-breathing lepisosteid fish. Annals of Anatomy 190:383–390. Jaroszewska, M., and K. Dabrowski. 2009: Early ontogeny of Semionotiformes and Amiiformes (Neopterygii: Actinopterygii), p. 230–274. In Development of Non-Teleost Fish. Y. W. Kunz, C. A. Luer, and B. G. Kapoor (eds.). Science Publishers Inc., Enfield, New Hampshire. Jaroszewska, M., K. Dabrowski, and G. Rodríguez. 2010. Development of testis and digestive tract in Longnose Gar (Lepisosteus osseus) at the onset of exogenous feeding of larvae and in juveniles. Aquaculture Research 41:1486–1497. Jarrett, R. D., and H. E. Malde. 1987. Palaeodischarge of the late Pleistocene Bonneville flood, Snake River, Idaho, computed from new evidence. The Geological Society of America Bulletin 99:126–134.

556

LITERATURE CITED

Järvi, T. 1990. The effects of male dominance, secondary sexual characteristics and female mate choice on the mating success of male Atlantic Salmon, Salmo salar. Ethology 84:123–132. Jarvik, E. 1959. De tidiga fossila ryggradsdjuren. Paleoanatomiska arbetsmetoder och resultat. [The early fossil vertebrates. Paleoanatomical methods and results]. Svensk Naturvetenskap 1959:5–80. Jarvik, E. 1980. Basic Structure and Evolution of Vertebrates. Vol. 1. Academic Press, New York. Jeff ries, H. P. 1960. Winter occurrences of Anguilla rostrata elvers in New England and Middle Atlantic estuaries. Limnology and Oceanography 5:338–340. Jelks, H. L., S. J. Walsh, N. M. Burkhead, S. Contreras-Balderas, E. Díaz-Pardo, D. A. Hendrickson, J. Lyons, N. E. Mandrak, F. McCormick, J. S. Nelson, S. P. Platania, B. A. Porter, C. B. Renaud, J. J. Schmitter-Soto, E. B. Taylor, and M. L. Warren, Jr. 2008. Conservation status of imperiled North American freshwater and diadromous fishes. Fisheries 33:372–407. Jellyman, D. and K. Tsukamoto. 2005. Swimming depths of offshore migrating longfineels Anguilla dieffenbachia. Marine Ecology Progress Series 286:261–267. Jenkins, R. E. 1970. Systematic studies of the catostomid fish tribe Moxostomatini. Unpubl. Ph.D. diss., Cornell University, Ithaca, New York. Jenkins, R. E. 1994. Harelip Sucker Moxostoma lacerum (Jordan and Brayton), p. 519–523. In Freshwater Fishes of Virgina. R. E. Jenkins and N. M. Burkhead (eds.). American Fisheries Society, Bethesda, Maryland. Jenkins, R. E., and N. M. Burkhead. 1994. Freshwater Fishes of Virginia. American Fisheries Society, Bethesda, Maryland. Jenkins, R. E., and D. J. Jenkins. 1980. Reproductive behavior of the Greater Redhorse, Moxostoma valenciennesi, in the Thousand Islands region. Canadian Field Naturalist 94:426–430. Jenkins, R. E., and E. A. Lachner. 1971. Criteria for analysis and interpretation of the American fish genera Nocomis Girard, and Hybopsis Agassiz. Smithsonian Contributions to Zoology 90. Jennings, C. A., and S. J. Zigler. 2000. Ecology and biology of Paddlefish in North America: historical perspectives, management approaches, and research priorities. Reviews in Fish Biology and Fisheries 10:167–181. Jennings, C. A., and S. J. Zigler. 2009. Biology and life history of Paddlefish in North America: an update, p. 1–22. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society, Symposium 66, Bethesda, Maryland. Jennings, M. J., and D. P. Philipp. 1992. Female choice and male competition in Longear Sunfish. Behavioral Ecology 3:84–94. Jennings, M. J., and D. P. Philipp. 2002. Alternative mating tactics in Sunfishes (Centrarchidae): a mechanism for hybridization? Copeia 2002:1102–1105. Jensen, C., and F. J. Schwartz. 1994. Atlantic Ocean occurrences of the Sea Lamprey, Petromyzon marinus (Petromyzontiformes: Petromyzontidae), parasitizing Sandbar, Carcharhinus plumbeus, and Dusky, C. obscurus (Carcharhiniformes: Carcharhinidae), Sharks off North and South Carolina. Brimleyana 21:69–72. Jeppesen, E., M. Meerhoff, K. Holmgren, I. González-Bergonzoni, F. Teixeira-de Mello, S. A. J. Declerck, L. Meester, et al. 2010. Impacts of climate warming on lake fish community structure

and potential effects on ecosystem function. Hydrobiologia 646:73–90. Jessop, B. 2010. Geographic effects on American Eel (Anguilla rostrata) life history characteristics and strategies. Canadian Journal of Fisheries and Aquatic Science 67:326–346. Jessop, B. M. 1987. Migrating American Eels in Nova Scotia. Transactions of the American Fisheries Society 116:161–170. Jessop, B. M. 2000. Estimates of population size and instream mortality rate of American Eel elvers in a Nova Scotia river. Transactions of the American Fisheries Society 129:514–526. Jessop, B. M. 2003. Annual variability in the effects of water temperature, discharge, and tidal stage on the migration of American Eel elvers from estuary to river. American Fisheries Society Symposium 33:3–16. Jessop, B. M., and C. J. Harvie. 2003. A CUSUM analysis of discharge patterns by a hydroelectric dam and discussion of potential effects on the upstream migration of American Eel elvers. Canadian Technical Report of Fisheries and Aquatic Sciences No. 2454. Jessop B. M., J. C. Shiao, Y. Iizuka, and W. N. Tzeng. 2006. Migration between freshwater and estuary of juvenile American Eels Anguilla rostrata as revealed by otolith microchemistry. Marine Ecology Progress Series 310:219–233. Jester, D. B. 1972. Life history, ecology and management of the River Carpsucker, Carpoides carpio (Rafinesque), with reference to Elephant Butte Lake. New Mexico State University Agricultural Experiment Station Research Report 243, Las Cruces. Jester, D. B. 1973. Life history, ecology, and management of the Smallmouth Buffalo, Ictiobus bubalus (Rafinesque), with reference to Elephant Butte Lake. New Mexico State University Agricultural Experiment Station Research Report 261, Las Cruces. Jin, F. 1999. Middle and Late Mesozoic Acipenseriformes from northern Hebei and western Liaoning, China, p. 188–260. In Palaeoworld. No. 11. P. J. Chen and F. Jin (eds.). Jehol Biota. Press of University of Science and Technology of China, Hefei, China. Jobling, M. 1995. Environmental Biology of Fishes. Chapman & Hall, London, United Kingdom. Johansen, K., D. Hanson, and C. Lenfant. 1970. Respiration in a primitive air breather, Amia calva. Respiration Physiology 9:162–174. Johnson, S. K., L. T. Fries, J. Williams, and D. G. Huff man. 1995. Presence of the parasitic swim bladder nematode, Anguillicola crassus, in Texas aquaculture. World Aquaculture 26:35–36. Johnson, B. L., and D. B. Noltie. 1996. Migratory dynamics of stream-spawning Longnose Gar (Lepisosteus osseus). Ecology of Freshwater Fish 5:97–107. Johnson, B. L., and D. B. Noltie. 1997. Demography, growth, and reproductive allocation in stream-spawning Longnose Gar. Transactions of the American Fisheries Society 126:438–466. Johnson, D. L. 2000. Sound production in Cyprinodon bifasciatus (Cyprinodontiformes). Environmental Biology of Fishes 59:341–346. Johnson, D. W., and W. L. Minckley. 1969. Natural hybridization in buffalofishes, genus Ictiobus. Copeia 1969:198–200. Johnson, J. B. 2002. Evolution after the flood: phylogeography of the desert fish Utah Chub. Evolution 56:948–960. Johnson, J. B., and M. C. Belk. 1999. Effects of predation on lifehistory evolution in Utah Chub (Gila atraria). Copeia 1999: 948–957.

LITERATURE CITED

Johnson, J. B., and S. Jordan. 2000. Phylogenetic divergence in Leatherside Chub (Gila copei) inferred from mitochondrial cytochrome b sequences. Molecular Ecology 9:1029–1035. Johnson, J. B., T. E. Dowling, and M. C. Belk. 2004. Neglected taxonomy of rare desert fishes: congruent evidence for two species of Leatherside Chub. Systematic Biology 53:841–855. Johnson, J. H. 1981. The summer food habits of the Cutlips Minnow, Exoglossum maxillingua, in a central New York stream. Copeia 1981:484–487. Johnson, J. H., and E. Z. Johnson. 1982. Observations on the eyepicking behavior of the Cutlips Minnow, Exoglossum maxillingua. Copeia 1982:711–712. Johnson, J. H., S. R. LaPan, R. M. Klindt, and A. Schiavone. 2006b. Lake Sturgeon spawning on artificial habitat in the St Lawrence River. Journal of Applied Ichthyology 22:465–470. Johnson, L. 1994. Long-term experiments on the stability of two fish populations in previously unexploited arctic lakes. Journal of Fisheries and Aquatic Science 51:209–225. Johnson, M. R., and F. F. Snelson, Jr. 1996. Reproductive life history of the Atlantic Stingray, Dasyatis sabina (Pisces, Dasyatidae), in the freshwater St. Johns River, Florida. Bulletin of Marine Science 59:74–88. Johnson, N. S., M. A. Luehring, M. J. Siefkes, and W. M. Li. 2006a. Mating pheromone reception and induced behavior in ovulating female Sea Lampreys. North American Journal of Fisheries Management 26:67–72. Johnson, N. S., M. J. Siefkes, and W. M. Li. 2005. Capture of ovulating female Sea Lampreys in traps baited with spermiating male Sea Lampreys. North American Journal of Fisheries Management 25:67–72. Johnson, N. S., S.-S. Yun, H. T. Thompson, C. O. Brant, and W. Li. 2009. A synthesized pheromone induces upstream movement in female Sea Lamprey and summons them into traps. Proceedings of the National Academy of Sciences 106:1021–1026. Johnson, R. P. 1963. Studies on the life history and ecology of the Bigmouth Buffalo, Ictiobus cyprinellas (Valenciennes). Journal of the Fisheries Research Board of Canada 20:1397–1430. Johnson, W. S., D. M. Allen, M. V. Ogburn, and S. E. Stancyk. 1990. Short-term predation responses of adult Bay Anchovies, Anchoa mitchilli to estuarine zooplankton availability. Marine Ecology Progress Series 64:55–68. Johnston, C. E. 1989. Male minnows build spawning nests. The Illinois Natural History Survey Reports 291:1–2. Johnston, C. E. 1991. Spawning activities of Notropis chlorocephalus, Notropis chiliticus, and Hybopsis hypsinotus, nest associates of Nocomis leptocephalus in the southeastern United States, with comments on nest association (Cypriniformes: Cyprinidae). Brimleyana 17:77–78. Johnston, C. E. 1994a. The benefit to some minnows of spawning in the nests of other species. Environmental Biology of Fishes 40:213–218. Johnston, C. E. 1994b. Nest association in fishes: evidence for mutualism. Behavioral Ecology and Sociobiology 35:379–383. Johnston, C. E. 1999. The relationship of spawning mode to conservation of North American minnows (Cyprinidae). Environmental Biology of Fishes 55:21–30. Johnston, C. E. 2000a. Allopaternal care in the Pygmy Sculpin (Cottus pygmaeus). Copeia 2000:262–264.

557

Johnston, C. E. 2000b. Movement patterns of imperiled Blue Shiners (Pisces: Cyprinidae) among habitat patches. Ecology of Freshwater Fish 9:170–176. Johnston, C. E., and W. S. Birkhead. 1988. Spawning in the Bandfin Shiner, Notropis zonistius. Journal of the Alabama Academy of Science 59:30–33. Johnston, C. E., and H. M. Buchanan. 2007. Learned or innate production of acoustic signals in fishes: a test using a cyprinid. Environmental Biology of Fishes 78:183–187. Johnston, C. E., and D. L. Johnson. 2000a. Sound production during the spawning season in cavity-nesting darters of the subgenus Catonotus (Percidae: Etheostoma). Copeia 2000:475–481. Johnston, C. E., and D. L. Johnson. 2000b. Sound production in Pimephales notatus (Rafinesque) (Cyprinidae). Copeia 2000:567–571. Johnston, C. E., and K. J. Kleiner. 1994. Reproductive behavior of the Rainbow Shiner (Notropis chrosomus) and the Rough Shiner (Notropis baileyi), nest associates of the Bluehead Chub (Nocomis leptocephalus) (Pisces: Cyprinidae) in the Alabama River drainage. The Journal of the Alabama Academy of Science 65:230–240. Johnston, C. E., and C. L. Knight. 1999. Life-history traits of the Bluenose Shiner, Pteronotropis welaka (Cypriniformes: Cyprinidae). Copeia 1999:200–205. Johnston, C. E., and M. J. Maceina. 2009. Fish assemblage shifts and species declines in Alabama, USA streams. Ecology of Freshwater Fish 18:33–40. Johnston, C. E., and L. M. Page. 1992. The evolution of complex reproductive strategies in North American minnows (Cyprinidae), p. 600–621. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Johnston, C. E., and C. T. Phillips. 2003. Sound production in Sturgeon, Scaphirhynchus albus and S. platorynchus (Acipenseridae). Environmental Biology of Fishes 68:59–64. Johnston, C. E., and S. P. Vives. 2003. Sound production in Codoma ornata (Girard) (Cyprinidae). Environmental Biology of Fishes 68:81–85. Johnstone, R. A. 1995. Honest advertisement of multiple qualities using multiple signals. Journal of Theoretical Biology 177:87–94. Johnstone, R. A. 1996. Multiple displays in animal communication: ‘backup signals’ and ‘multiple messages’. Philosophical Transactions of the Royal Society of London B 351:329–338. Jollie, M. 1984. Development of cranial and pectoral girdle bones of Lepisosteus, with a note on scales. Copeia 1984:476–502. Jones, H. M., and C. A. Paszkowski. 1997. Effects of exposure to predatory cues on territorial behaviour of Fathead Minnows. Environmental Biology of Fishes 49:97–109. Jones, J. C., and J. D. Reynolds. 1997. Effects of pollution on reproductive behaviour of fishes. Reviews in Fish Biology and Fisheries 7:463–491. Jones, J. W. 1959. The Salmon. New Naturalist, London. Jones, J. W., and G. M. King. 1952. The spawning of the male salmon parr (Salmo salar Linn. Juv.). Proceedings of the Zoological Society of London 122:615–619. Jones, P. W., F. D. Martin, and J. D. Hardy, Jr. 1978. Development of Fishes of the Mid-Atlantic Bight: An Atlas of Egg, Larval and Juvenile Stages. Volume 1. Acipenseridae through Ictaluridae.

558

LITERATURE CITED

Biological Ser vices Program, Fish and Wildlife Ser vice, U.S. Department of the Interior, FWS/OBS-78/12. Jones, S. R. M., and P. T. K. Woo. 1992. Vector specificity of Trypanosoma catostomi and its infectivity to freshwater fishes. The Journal of Parasitology 78:87–92. Jonsson, B., N. Jonsson, K. Hindar, T. G. Northcote, and S. Engen. 2008. Asymmetric competition drives lake use of coexisting salmonids. Oecologia 157:553–560. Jordan, D. S. 1876. Concerning the fishes of the Ichthyologia Ohiensis. Bulletin of the Buffalo Society of Natural Sciences 3:91–97. Jordan, D. S. 1878. Manual of the Vertebrates of the Northern United States, including the District East of the Mississippi River and North of North Carolina and Tennessee, Exclusive of Marine Species. 2nd edition. Jansen, McClurg & Co., Chicago, Illinois. Jordan, D. S. 1917. Changes in names of American fishes. Copeia 1917:85–89. Jordan, D. S. 1924. Concerning the genus Hybopsis of Agassiz. Copeia 1924:51–52. Jordan, D. S., and A. W. Brayton. 1877. On Lagochila, a new genus of catostomoid fishes. Proceedings of the Academy of Natural Sciences of Philadelphia 29:280–283. Jordan, D. S., and B. W. Evermann. 1896. The fishes of North and Middle America: a descriptive cata logue of the species of fishlike vertebrates found in the waters of North America, north of the Isthmus of Panama. Bulletin of the United States National Museum 47:1–1240. Jordan, D. S., and B. W. Evermann. 1902. American Food and Game Fishes. Doubleday, Page and Company, Garden City, New York. Jordan, D. S., and W. F. Thompson. 1910. Note on the gold-eye, Amphiodon alosoides Rafinesque, or Elattonistius chrysopsis (Richardson). Proceedings of the U.S. National Museum 38:353–357. Jordan, F., and D. A. Arrington. 2001. Weak trophic interactions between large predatory fishes and herpetofauna in the channelized Kissimmee River, Florida, USA. Wetlands 21:155–159. Jordan, G. R., R. A. Klumb, G. A. Wanner, and W. J. Stancill. 2006. Poststocking movements and habitat use of hatcheryreared juvenile Pallid Sturgeon in the Missouri River below Fort Randall Dam, South Dakota and Nebraska. Transactions of the American Fisheries Society 135:1499–1511. Jordan, R. C., A. M. Gospodarek, E. T. Schultz, R. K. Cowen, and K. Lwiza. 2000. Spatial and temporal growth rate variation of Bay Anchovy (Anchoa mitchilli) larvae in the mid Hudson River Estuary. Estuaries 23:683–689. Jordan, W. C., and A. F. Youngson. 1992. The use of genetic marking to assess the reproductive success of mature male Atlantic Salmon parr (Salmo salar, L.) under natural spawning conditions. Journal of Fish Biology 41:613–618. Jørgensen, J. M., Å. Flock, and J. Wersäll. 1972. The Lorenzinian ampullae of Polyodon spathula. Zeitschrift für Zellforschung und Mikroskopische Anatomie 130:362–377. Joy, J. E. 2008. Intestinal parasites of Bowfin, Amia calva L., from the Green Bottom Wildlife Management Area, West Virginia, USA. Comparative Parasitology 75:138–140. Joy, J. E., W. E. Triest, and E. M. Walker. 2009. Adaptation of Haplobothrium globuliforme (Cestoda: Pseudophyllidea) to the intestinal architecture of the Bowfin (Amia calva L). Journal of Parasitology 95:69–74.

Juanes, F., R. E. Marks, K. A. McKown, and D. O. Conover. 1993. Predation by age-0 Bluefish on age-0 anadromous fishes in the Hudson River Estuary. Transactions of the American Fisheries Society 122:348–356. Jumper, G. Y., and R. C. Baird. 1991. Location by olfaction: a model and application to the mating problem in the deep-sea hatchetfish, Argyropelecus hemigymnus. American Naturalist 138:141–1458. Juneau, C. L. 1975. An inventory and study of Vermillion Bay and Atchafalaya Bay complex. Louisiana Wildlife and Fisheries Commission. Oyster, Water Bottom Seafoods Division Technical Bulletin 13:2–76. Kaeding, L. R., and D. B. Osmundson. 1988. Interaction of slow growth and increased early-life mortality: an hypothesis on the decline of Colorado squawfish in the upstream regions of its historic range. Environmental Biology of Fishes 22:287–298. Kahnle, A. W., K. A. Hattala, and K. A. McKown. 2007. Status of Atlantic Sturgeon of the Hudson River estuary, New York, USA, p. 347–363. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.), American Fisheries Society Symposium 56, Bethesda, Maryland. Kajiura, S. M., A. P. Sebastian, and T. C. Tricas. 2000. Dermal bite wounds as indicators of reproductive seasonality and behavior in the Atlantic Stingray, Dasyatis sabina. Environmental Biology of Fishes 58:23–31. Kajiura, S. M., and T. C. Tricas. 1996. Seasonal dynamics of dental sexual dimorphism in the Atlantic Stingray, Dasyatis sabina. Journal of Experimental Zoology 276:219–232. Kalbe, M., C. Eizaguirre, I. Dankert, T. B. H. Reusch, R. D. Sommerfeld, K. M. Wegner, and M. Milinski. 2009. Lifetime reproductive success is maximized with optimal major histocompatibility complex diversity. Proceedings of the Royal Society of London B 276:925–934. Kallemeyn, L. 1983. Status of the Pallid Sturgeon. Fisheries 8:3–9. Kallman, K. D. 1984. A new look at sex determination in poeciliid fishes, p. 95–171. In Evolutionary Genetics of Fishes. B. J. Turner (ed.). Plenum Publishing Co., New York. Kallman, K. D. 1989. Genetic control of size at maturity in Xiphophorus, p. 163–184. In Ecology and Evolution of Livebearing Fishes (Poeciliidae). G. K. Meffe and F. F. S. Snelson (eds.). Prentice Hall, New Jersey. Kalmijn, A. J. 1982. Electric and magnetic field detection in elasmobranch fishes. Science 218:916–918. Kalmijn, A. J. 1989. Functional evolution of lateral line and inner ear sensory systems, p. 187–215. In The Mechanosensory Lateral Line-Neurobiology and Evolution. S. Coombs, P. Gorner, and H. Munz (eds.). Springer-Verlag, New York. Kammerer, C. F., L. Grande, and M. W. Westneat. 2006. Comparative and developmental functional morphology of the jaws of living and fossil Gars (Actinopterygii: Lepisosteidae). Journal of Morphology 267:1017–1031. Kaneko, T., S. Hasegawa, and S. Sasai. 2003. Chloride cells in the Japanese Eel during their early life stages and downstream migration, p. 457–468. In Eel Biology. K. Aida, K. Tsukamoto, and K. Yamauchi (eds.). Springer, Tokyo. Karlsson, L., G. Ekbohm, and G. Seinholtz. 1984. Comments on a study of the thermal behaviour of the American Eel (Anguilla rostrata) and some statistical suggestions for temperature preference studies. Hydrobiologia 109:75–78.

LITERATURE CITED

Karplus, I., and D. Algom. 1996. Polymorphism and pair formation in the mosquitofish Gambusia holbrooki (Pisces: Poeciliidae). Environmental Biology of Fishes 45:169–176. Kasumyan, A. O. 1999. Olfaction and taste senses in Sturgeon behaviour. Journal of Applied Ichthyology 15:228–232. Kasumyan, A. O. 2002. Sturgeon food searching behaviour evoked by chemical stimuli: a reliable sensory mechanism. Journal of Applied Ichthyology 18:685–690. Kasumyan, A. O. 2007. Paddlefish, Polyodon spathula juveniles food searching behaviour evoked by natural food odour. Journal of Applied Ichthyology 23:636–639. Katechis, C. T., P. C. Sakaris, and E. R. Irwin. 2007. Population demographics of Hiodon tergisus (Mooneye) in the Lower Tallapoosa River. Southeastern Naturalist 6:461–470. Katula, R. S., and L. M. Page. 1998. Nest association between a large predator, the Bowfin (Amia calva), and its prey, the Golden Shiner (Notemigonus crysoleucas). Copeia 1998:220–221. Katz, S. L. 1996. Ventilatory control in a primitive fish: signal conditioning via non-linear O2 affinity. Respiration Physiology 103:165–175. Kaufman, R. C., A. G. Houck, and J. J. Cech, Jr. 2007. Effects of temperature and carbon dioxide on Green Sturgeon bloodoxygen equilibria. Environmental Biology of Fishes 79:201–210. Kavanau, J. L. 1990. Conservative behavioural evolution, the neural substrate. Animal Behaviour 39:758–767. Kay, L. K., R. Wallus, and B. L. Yeager. 1994. Reproductive Biology and Early Life History of Fishes in the Ohio River Drainage. Volume 2: Catostomidae. Tennessee Valley Authority, Chattanooga. Kazakov, R. V. 1981. Peculiarities of sperm production by anadromous and parr Atlantic Salmon (Salmo salar L.) and fish cultural characteristics of such sperm. Journal of Fish Biology 18:1–8. Keast, A., and M. G. Fox. 1992. Space use and feeding patterns of an offshore fish assemblage in a shallow mesotrophic lake. Environmental Biology of Fishes 34:159–170. Keast, A., J. Harker, and D. Turnbull. 1978. Nearshore fish habitat utilization and species association in Lake Opinicon (Ontario, Canada). Environmental Biology of Fishes 3:173–184. Keck, B. P., and T. J. Near. 2009. Patterns of natural hybridization in darters (Percidae: Etheostomatinae). Copeia 2009:758–773. Keddy, P., and E. Weiher. 2001. Introduction: the scope and goals of research on assembly rules, p. 1–20. In Ecological Assembly Rules. E. Weiher and P. Keddy (eds.). Cambridge University Press, Cambridge, United Kingdom. Keegan-Rogers, V. 1984. Unfamiliar-female mating advantage among clones of unisexual fish (Poeciliopsis: Poeciliidae). Copeia 1984:169–174. Keene, D. A. 2004. Reevaluating late prehistoric coastal subsistence and settlement strategies: new data from Grove’s Creek site, Skidaway Island, Georgia. American Antiquity 69:671–688. Keenleyside, M. H. 1967. Behavior of male Sunfishes (genus Lepomis) towards females of three species. Evolution 21:688–695. Keenleyside, M. H. A. 1972. Intraspecific intrusions into nests of spawning Longear Sunfish (Pisces: Centrarchidae). Copeia 1972:272–278. Keenleyside, M. H. A., and M. H. C. Dupuis. 1988. Courtship and spawning competition in Pink Salmon (Oncorhynchus gorbuscha). Canadian Journal of Zoology 66:262–265. Keenlyne, K. D. 1997. Life history and status of the Shovelnose Sturgeon, Scaphirhynchus platorynchus. Environmental Biology of Fishes 48:291–298.

559

Keenlyne, K. D., E. M. Grossman, and L. G. Jenkins. 1992. Fecundity of the Pallid Sturgeon. Transactions of the American Fisheries Society 121:139–140. Keenlyne, K. D., C. J. Henry, A. Tews, and P. Clancey. 1994. Morphometric comparisons of upper Missouri River Sturgeons. Transactions of the American Fisheries Society 123:779–785. Keenlyne, K. D., and L. G. Jenkins. 1993. Age at sexual maturity of the Pallid Sturgeon. Transactions of the American Fisheries Society 122:393–396. Keevin, T. M., S. G. George, J. J. Hoover, B. R. Kuhajda, and R. L. Mayden. 2007. Food habits of the endangered Alabama Sturgeon, Scaphirhynchus suttkusi Williams and Clemmer, 1991 (Acipenseridae). Journal of Applied Ichthyology 23:500–505. Keister, J. E., E. D. Houde, and D. L. Breitburg. 2000. Effects of bottom-layer hypoxia on abundances and depth distributions of organisms in Patuxent River, Chesapeake Bay. Marine Ecology Progress Series 205:43–59. Keister, J. E., E. D. Houde, and D. L. Breitburg. 2000. Effects of bottom-layer hypoxia on abundances and depth distributions of organisms in Patuxent River, Chesapeake Bay. Marine Ecology Progress Series 205:43–59. Kells, V., and K. Carpenter. 2011. A Field Guide to Coastal Fishes from Maine to Texas. John Hopkins University Press, Baltimore, Maryland. Kelly, H. A. 1924. Amia calva guarding its young. Copeia 133:73–74. Kelly, J. T., A. P. Klimley, and C. E. Crocker. 2007. Movements of Green Sturgeon, Acipenser medirostris, in the San Francisco Bay estuary, California. Environmental Biology of Fishes 79:281–295. Kemp, P. S., T. Tsuzaki, and M. L. Moser. 2009. Linking behaviour and performance: intermittent locomotion in a climbing fish. Journal of Zoology 277:171–178. Kempinger, J. J. 1988. Spawning and early life history of Lake Sturgeon in the Lake Winnebago system, Wisconsin, p. 110– 122. In 11th Annual Larval Fish Conference. R. D. Hoyt (ed.). American Fisheries Society Symposium 5, Bethesda, Maryland. Kempinger, J. J., K. J. Otis, and J. R. Ball. 1998. Fish kills in the Fox River, Wisconsin, attributable to carbon monoxide from marine engines. Transactions of the American Fisheries Society 127:669–672. Kendle, E. R. 1970. Sexual discrimination by olfaction in the Black Bullhead, Ictalurus melas. State of Nebraska Game and Parks Commission Job Progress Reports, March 1, 1969–February 28, 1970:21–31. Kennedy, A. J., and T. M. Sutton. 2007. Effects of harvest and length limits on Shovelnose Sturgeon in the upper Wabash River, Indiana. Journal of Applied Ichthyology 23:465–475. Kennedy, A. J., T. M. Sutton, and B. E. Fisher. 2006. Reproductive biology of female Shovelnose Sturgeon in the upper Wabash River, Indiana. Journal of Applied Ichthyology 22:177–182. Kennedy, T. B., and G. L. Vinyard. 1997. Drift ecology of western catostomid larvae with emphasis on Warner Suckers, Catostomus warnerensis (Teleostei). Environmental Biology of Fishes 49:187–195. Kennedy, T. B., and G. L. Vinyard. 2006. Ecology of young stream-resident Warner Sucker (Catostomus warnerensis) in Warner Basin, Oregon. American Midland Naturalist 156:403– 407. Kennedy, W. A., and W. M. Sprules. 1967. Goldeye in Canada. Fisheries Research Board of Canada 161:1–45.

560

LITERATURE CITED

Kennen, J. G., S. J. Wisniewski, N. H. Ringler, and H. M. Hawkins. 1994. Application and modification of an auger trap to quantify emigrating fishes in Lake Ontario tributaries. North American Journal of Fisheries Management 14:828–836. Kent, B. 1999. Sharks from the Fisher/Sullivan Site, p. 11–37. In Early Eocene Vertebrates and Plants from the Fisher/Sullivan Site (Nanjemoy Formation) Stafford County, Virginia. R. Weems and G. Grimsley (eds.). Virginia Division of Mineral Resources, Publication 152. Kenyon, T. N., F. Ladich, and H. Y. Yan. 1998. A comparative study of hearing ability in fishes: the auditory brainstem response approach. Journal of Comparative Physiology A 182:307–318. Kereliuk, M. R., Y. Katare, K. Tierney, A. Laframboise, A. P. Scott, S. J. Loeb, and B. Zielinski. 2009. Attraction of female Round Gobies to steroids released by males. Association for Chemorecpetion Science, 31st Annual Meeting, Sarasota, Florida. Kerns, J. A., P. W. Bettoli, and G. D. Scholten. 2009. Mortality and movements of Paddlefish released as bycatch in a commercial fishery in Kentucky Lake, Tennessee, p. 329–343. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Kettratad, J., and D. F. Markle. 2010. Redescription of the Tyee Sucker, Catostomus tsiltcoosensis (Catostomidae). Western North American Naturalist 70:273–287. Khidir, K. T., and C. B. Renaud. 2003. Oral fimbriae and papillae in parasitic Lampreys (Petromyzontiformes). Environmental Biology of Fishes 66:271–278. Kieffer, M. C., and B. Kynard. 1993. Annual movements of Shortnose and Atlantic Sturgeons in the Merrimack River, Massachusetts. Transactions of the American Fisheries Society 122:1088–1103. Kihslinger, R. L., and A. P. Klimley. 2002. Species identity and the temporal characteristics of fish acoustic signals. Journal of Comparative Physiology 116:210–214. Kikugawa, K., K. Katoh, S. Kuraku, H. Sakurai, O. Ishida, N. Iwabe, and T. Miyata. 2004. Basal jawed vertebrate phylogeny inferred from multiple nuclear DNA-coded genes. BMC Biology 2:1–11. Killgore, K. J, and J. J. Hoover. 2001. Effects of hypoxia on fish assemblages in a vegetated waterbody. Journal of Aquatic Plant Management 39:40–44. Killgore, K. J., J. J. Hoover, J. P. Kirk, S. G. George, B. R. Lewis, and C. E. Murphy. 2007. Age and growth of Pallid Sturgeon in the free-flowing Mississippi River. Journal of Applied Ichthyology 23:452–456. Killgore, K. J., S. T. Maynord, M. D. Chan, and R. P. Morgan, II. 2001. Evaluation of propeller-induced mortality on early life stages of selected fish species. North American Journal of Fisheries Management 21:947–955. Killgore, K. J., L. E. Miranda, C. E. Murphy, D. M. Wolff, J. J. Hoover, T. M. Keevin, S. T. Maynord, and M. A. Cornish. 2011. Fish entrainment rates through towboat propellers in the upper Mississippi and Illinois rivers. Transactions of the American Fisheries Society 140:570–581. Killgore, K. J., R. P. Morgan II, and N. B. Rybicki. 1989. Distribution and abundance of fishes associated with submersed aquatic plants in the Potomac River. North American Journal of Fisheries Management 9:101–111.

Kim, J. B., B. A. Barton, and J. M. Conlon. 2001. Characterization of two molecular forms of vasoactive intestinal polypeptide (VIP) from the Pallid Sturgeon, Scaphirhynchus albus (Acipenseriformes). Fish Physiology and Biochemistry 25:231–238. Kim, J.-W., G. E. Brown, I. J. Dolinsek, N. N. Brodeur, A. O. H. C. Leduc, and J. W. A. Grant. 2009. Combined effects of chemical and visual information eliciting antipredator behaviour in juvenile Atlantic Salmon, Salmo salar. Journal of Fish biology 74:1280–1290. Kimmel, P. G. 1975. Fishes of the Miocene-Pliocene Deer Butte Formation, southeast Oregon. University of Michigan Museum of Paleontology, Claude W. Hibbard Memorial Volume 5, Papers on Paleontology 14:69–87. Kimura, R., D. H. Secor, E. D. Houde, and P. M. Piccoli. 2000. Upestuary dispersal of young-of-the-year Bay Anchovy, Anchoa mitchilli in the Chesapeake Bay: inferences from microprobe analysis of strontium in otoliths. Marine Ecology Progress Series 208:217–227. Kimura, S., and Y. Tao. 1937. Notes on the nuptial coloration and pearl organs in Chinese fresh-water fishes. Shanghai Sizenkagaku Kenkyusyo Iho 6:277–318. Kinch, J. C. 1979. Trophic habits of the juvenile fishes within artificial waterways: Marco Island, Florida. Contributions in Marine Science 22:77–90. King, E. L., Jr. 1980. Classification of Sea Lamprey (Petromyzon marinus) attack marks on Great Lakes Lake Trout (Salvelinus namaycush). Canadian Journal of Fisheries and Aquatic Sciences 37:1989–2006. King, T. L., B. A. Lubinski, and A. P. Spidle. 2001. Microsatellite DNA variation in Atlantic Sturgeon (Acipenser oxyrinchus oxirynchus) and cross-species amplification in the Acipenseridae. Conservation Genetics 2:103–119. King, T. L., E. G. Zimmerman, and T. L. Beitinger. 1985. Concordant variation in thermal tolerance and allozymes of the Red Shiner, Notropis lutrensis, inhabiting tailwater sections of the Brazos River, Texas. Environmental Biology of Fishes 13:49–57. Kingsford, R. T. 2011. Conservation management of rivers and wetlands under climate change—a synthesis. Marine and Freshwater Research 62:217–222. Kirsch, P. H. 1889. Notes on a collection of fishes obtained in the Gila River, at Fort Thomas, Arizona, by Lieut. W. L. Carpenter, U.S. Army. Proceedings of the National Museum 11:555–558. Kistler, H. D. 1906. The primitive pores of Polyodon spathula. The Journal of Comparative Neurology and Psychology 16:294–298. Kitamura, S., H. Ogata, and F. Takashima. 1994. Olfactory responses of several species of teleost to f-prostaglandins. Comparative Biochemistry and Physiology A 107:463–467. Kitchell, J. F., M. G. Johnson, C. K. Minns, K. H. Loftus, L. Greig, and C. H. Olver. 1977. Percid habitat: the river analogy. Journal of the Fisheries Research Board of Canada 34:1936–1940. Klaassen, H. E., and K. L. Morgan. 1974. Age and growth of Longnose Gar in Tuttle Creek Reservoir, Kansas. Transactions of the American Fisheries Society 103:402–405. Kleckner, R. C., and W. H. Krueger. 1981. Changes in the swim bladder retial morphology in Anguilla rostrata during premigration metamorphosis. Journal of Fish Biology 18:569–577. Kleckner, R. C., and J. D. McCleave. 1982. Entry of migrating American Eel leptocephali into the Gulf Stream system. Helgoländer Meeresuntersuchungen 35:329–339.

LITERATURE CITED

Kleckner, R. C., and J. D. McCleave. 1985. Spatial and temporal distribution of American Eel larvae in relation to North Atlantic Ocean current systems. Dana 4:67–92. Kleckner, R. C., and J. D. McCleave. 1988. The northern limit of spawning by Atlantic Eels (Anguilla spp.) in the Sargasso Sea in relation to thermal fronts and surface water masses. Journal of Marine Research 46:647–667. Kleckner, R. C., J. D. McCleave, and G. S. Wippelhauser. 1983. Spawning of American Eel, Anguilla rostrata, relative to thermal fronts in the Sargasso Sea. Environmental Biology of Fishes 9:289–293. Kleerekoper, H. 1972. The sense organs, p. 373–404. In The Biology of Lampreys. Vol. 2. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Knapp, R. A., and R. C. Sargent. 1989. Egg-mimicry as a mating strategy in the Fantail Darter, Etheostoma flabellare: females prefer males with eggs. Behavioral Ecology and Sociobiology 25:321–326. Knights, B. C., J. M. Vallazza, S. J. Zigler, and M. R. Dewey. 2002. Habitat and movement of Lake Sturgeon in the upper Mississippi River system, USA. Transactions of the American Fisheries Society 131:507–522. Knotek, W. L., and D. J. Orth. 1998. Survival for specific life intervals of Smallmouth Bass, Micropterus dolomieu, during parental care. Environmental Biology of Fishes 51:285–296. Koch, B. T., J. E. Garvey, J. You, and M. J. Lydy. 2006. Elevated organochlorines in the brain-hypothalamic-pituitary complex of intersexual Shovelnose Sturgeon. Environmental Toxicology and Chemistry 25:1689–1697. Koch, D. L. 1973. Reproductive characteristics of the Cui-ui lakesucker (Chasmistes cujus Cope) and its spawning behavior in Pyramid Lake, Nevada. Transactions of the American Fisheries Society 102:145–149. Koch, D. L. 1976. Life history information on the Cui-ui lakesuckers (Chasmistes cujus Cope, 1883) in Pyramid Lake, Nevada. Occasional Papers of the Biological Society of Nevada 40:1–12. Koch, J. D., and M. C. Quist. 2010. Current status and trends in Shovelnose Sturgeon (Scaphirhynchus platorynchus) management and conservation. Journal of Applied Ichthyology 26:491–498. Koch, J. D., M. C. Quist, and K. A. Hansen. 2009a. Precision of hard structures used to estimate age of Bowfin in the upper Mississippi River. North American Journal of Fisheries Management 29:506–511. Koch, J. D., M. C. Quist, K. A. Hansen, and G. A. Jones. 2009b. Population dynamics and potential management of Bowfin (Amia calva) in the upper Mississippi River. Journal of Applied Ichthyology 25:545–550. Koch, J. D., M. C. Quist, C. L. Pierce, K. A. Hansen, and M. J. Steuck. 2009c. Effects of commercial harvest on Shovelnose Sturgeon populations in the upper Mississippi River. North American Journal of Fisheries Management 29:84–100. Kodric-Brown, A. 1977. Reproductive success and the evolution of breeding territories in Pupfish (Cyprinodon). Evolution 31:750–766. Kodric-Brown, A. 1983. Determinants of male reproductive success in Pupfish (Cyprinodon pecosensis). Animal Behaviour 31:128–137. Kodric-Brown, A. 1985. Female preference and sexual selection for male coloration in the Guppy (Poecilia reticulata). Behavioral Ecology and Sociobiology 17:199–205.

561

Kodric-Brown, A. 1988. Effect of population density, size of habitat and oviposition substrate on the breeding system of Pupfish (Cyprinodon pecosensis). Ethology 77:28–43. Kodric-Brown, A. 1995. Does past reproductive history predict competitive interactions and male mating success in Pupfish? Animal Behaviour 50:1433–1440. Kodric-Brown, A. 1996. Role of male-male competition and female choice in the development of breeding coloration in Pupfish (Cyprinodon pecosensis). Behavioral Ecology 7:431–437. Kodric-Brown, A., and P. F. Nicoletto. 1993. The relationship between physical condition and social status in Pupfish, Cyprinodon pecosensis. Animal Behaviour 46:1234–1236. Kodric-Brown, A., and U. Strecker. 2001. Responses of Cyprinodon maya and C. labiosus females to visual and olfactory cues of conspecific and heterospecific males. Biological Journal of the Linnean Society 74:541–548. Koehler, A., P. Romans, S. S. Desser, and M. Ringuette. 2004. Encapsulation of Myxobolus pendula (Myxosporidia) by epithelioid cells of its cyprinid host Semotilus atromaculatus. Journal of Parasitology 90:1401–1405. Kofoid, C. A. 1900. Notes on the natural history of Polyodon, p. 252 (Abstract). In The meeting of naturalists in Chicago. C. V. Davenport. Science (New Series) 11:246–253. Kohlhorst, D. W., L. W. Botsford, J. S. Brennan, and G. M. Cailliet. 1991. Aspects of the structure and dynamics of an exploited central California population of White Sturgeon (Acipenser transmontanus), p. 277–293. In Acipenser: Actes du Premier Colloque International sur l’Esturgeon, Bordeaux, 3–6 octobre 1989. P. Williot. (ed.). CEMAGREF-DICOVA, Bordeaux, France. Kohn, M. J., and T. J. Fremd. 2008. Miocene tectonics and climate forcing of biodiversity, western United States. Geology 36:783–786. Kojima, N. F., K. J. Kojima, S. Kobayakawa, N. Higashide, C. Hamanaka, A. Nitta, I. Koeda, T. Yamaguchi, M. Shichiri, S.-i. Kohno, and S. Kubota. 2010. Whole chromosome elimination and chromosome terminus elimination both contribute to somatic differentiation in Taiwanese Hagfish Paramyxine sheni. Chromosome Research 18:383–400. Kolok, A. S., R. M. Spooner, and A. P. Farrell. 1993. The effect of exercise on the cardiac output and blood flow distribution of the Largescale Sucker, Catostomus macrocheilus. Journal of Experimental Biology 183:301–321. Kolok, A. S., E. B. Peake, L. L. Tierney, S. A. Roark, R. B. Noble, K. See, and S. I. Guttman. 2004. Copper tolerance in Fathead Minnow: I. The role of genetic and nongenetic factors. Environmental Toxicology and Chemistry. 23:200–207. Koppelman, J. B. 1994. Hybridization between Smallmouth Bass, Micropterus dolomieu, and Spotted Bass, M. punctulatus, in the Missouri River system, Missouri. Copeia 1994:204–210. Korf, H. W., C. Schomerus, and J. H. Stehle. 1998. The pineal organ, its hormone melatonin, and the photoneuroendocrine system. Advances in Anatomy, Embryology and Cell Biology 146:1–100. Körner, K. E., O. Lutjens, J. Parzefall, and I. Schlupp. 1999. The role of experience in mating preferences of the unisexual Amazon Molly. Behaviour 136:257–268. Koseki, Y., and K. Maekawa. 2000. Sexual selection on mature parr of Masu Salmon (Oncorhynchus masou): does sneaking behavior favor small body size and less-developed sexual characters? Behavioral Ecology and Sociobiology 48:211–217.

562

LITERATURE CITED

Kottelat, M. and J. Freyhof. 2007. Handbook of European Freshwater Fishes. Kottelat, Cornol, Switzerland and Freyhof, Berlin, Germany. Kowalski, K. T., J. P. Schubauer, C. L. Scott, and J. R. Spotila. 1978. Interspecific and seasonal differences in the temperature tolerance of streamfish. Journal of Thermal Biology 3:105–108. Kozfkay, J. R., and D. L. Scarnecchia. 2002. Year-class strength and feeding ecology of age-0 and age-1 Paddlefish (Polyodon spathula) in Fort Peck Lake, Montana, USA. Journal of Applied Ichthyology 18:601–607. Kraak, S. B. M. 1996. “Copying mate choice”: which phenomena deserve this term? Behavioural Processes 36:99–102. Kraak, S. B. M., and T. G. G. Groothuis. 1994. Female preference for nests is based on the presence of the eggs themselves. Behaviour 131:189–206. Kramer, D. L., and J. B. Graham. 1976. Synchronous air breathing, a social component of respiration in fishes. Copeia 1976:689–697. Kramer, R. H., and L. L. Smith, Jr. 1960. Utilization of nests of large-mouth bass, (Micropterus salmoides) by Golden Shiners (Notemigonus crysoleucas). Copeia 1960:73–74. Krayushkina, L. S. 2006. Considerations on evolutionary mechanisms of osmotic and ionic regulation in Acipenseridae: an overview. Journal of Applied Ichthyology 22 (Supplement 1):70–76. Krebs, C. J. 1994. Ecology, the experimental analysis of distribution and abundance. 4th edition. Harper Collins College Publishers, New York. Krieger, J., and P. A. Fuerst. 2002a. Evidence of multiple alleles of nuclear 18S ribosomal RNA gene in Sturgeon (family: Acipenseridae). Journal of Applied Ichthyology 18:290–297. Krieger, J., and P. A. Fuerst. 2002b. Evidence for a slowed rate of molecular evolution in the order Acipenseriformes. Molecular Biology and Evolution 19:891–897. Krieger, J., and P. A. Fuerst. 2009. Molecular markers and the study of phylogeny and genetic diversity in North American Sturgeons and Paddlefish, p. 63–83. In Biology, Conservation and Sustainable Development of Sturgeons. R. Carmona, A. Domezain, M. García-Gallego, J. A. Hernando, F. Rodríguez, and M. Ruiz-Rejón (eds.). Springer Science + Business Media B. V. Fish & Fisheries Series 29. Krieger, J., P. A. Fuerst, and T. M. Cavender. 2000. Phylogenetic relationships of the North American Sturgeons (order Acipenseriformes) based on mitochondrial DNA sequences. Molecular Phylogenetics and Evolution 16:64–72. Krieger, J., A. K. Hett, P. A. Fuerst, E. Artyukhin, and A. Ludwig. 2008. The molecular phylogeny of the order Acipenseriformes revisited. Journal of Applied Ichthyology 24 (Supplement 1):36–45. Krieger, J., A. K. Hett, P. A. Fuerst, V. J. Birstein, and A. Ludwig. 2006. Unusual intraindividual variation of the nuclear 18S rRNA gene is widespread within the Acipenseridae. Journal of Heredity 97:218–225. Kristensen, J. 1980. Large Flathead Chub (Platygobio gracilis) from the Peace-Athabasca Delta, Alberta, including a Canadian record. Canadian Field-Naturalist 94:342. Kroll, K. J., J. P. Van Eenennaam, S. I. Doroshov, J. E. Hamilton, and T. R. Russell. 1992. Effect of water temperature and formulated diets on growth and survival of larval Paddlefish. Transactions of the American Fisheries Society 121:538–543.

Kroll, K. J., J. P. Van Eenennaam, S. I. Doroshov, J. Linares, E. J. Hamilton, and T. R. Russell. 1994. Growth and survival of Paddlefish fry raised in the laboratory on natural and artificial diets. The Progressive Fish-Culturist 56:169–174. Krout, R. T., and W. A. Dunson. 1985. Stimulation of sodium efflux in air-breathing fish exposed to low pH. Comparative Biochemistry and Physiology. C, Comparative Pharmacology and Toxicology 82:49–53. Krueger, W. H., and K. Oliveira. 1997. Sex, size, and gonad morphology of silver American Eels Anguilla rostrata. Copeia 1997:415–420. Krueger, W. H., and K. Oliveira. 1999. Evidence for environmental sex determination in the American Eel, Anguilla rostrata. Environmental Biology of Fishes 55:381–389. Kruesi, K., and G. Alcaraz. 2007. Does a sexually selected trait represent a burden in locomotion? Journal of Fish Biology 70:1161–1170. Kruse, G. O., and D. L. Scarnecchia. 2002. Assessment of bioaccumulated metal and organochlorine compounds in relation to physiological biomarkers in Kootenai White Sturgeon. Journal of Applied Ichthyology 18:430–438. Kubach, K. M., M. C. Scott, and J. S. Bulak. 2011. Recovery of a temperate riverine fish assemblage from a major diesel oil spill. Freshwater Biology 56:503–518. Kubota, S., J.-i. Takano, R. Tsuneishi, S. Kobayakawa, N. Fujikawa, M. Nabeyama, and S.-i. Kohno. 2001. Highly repetitive DNA families restricted to germ cells in a Japanese Hagfish (Eptatretus burgeri): a hierarchical and mosaic structure in eliminated chromosomes. Genetica 111:319–328. Kucheryavyi, A. V., K. A. Savvaitova, M. A. Gruzdeva, and D. S. Pavlov. 2007. Sexual dimorphism and some special traits of spawning behavior of the Arctic Lamprey, Lethenteron camtschaticum. Journal of Ichthyology 47:481–485. Kuehne, R. A. 1962. A classification of streams, illustrated by fish distribution in an eastern Kentucky creek. Ecology 43:608–614. Kuhajda, B. R., R. L. Mayden, and R. M. Wood. 2007. Morphologic comparisons of hatchery-reared specimens of Scaphirhynchus albus, Scaphirhynchus platorynchus, and S. albus × S. platorynchus hybrids (Acipenseriformes: Acipenseridae). Journal of Applied Ichthyology 23:324–347. Kullander, S. O., and C. J. Ferraris, Jr. 2003. Family Engraulidae, p. 39–42. In Checklist of the Freshwater Fishes of South and Central America. R. E. Reis, S. O. Kullander, and C. J. Ferraris, Jr. (eds.). EDIPUCRS, Porto Alegre, Brasil. Kuntz, A. 1915. The embryology and larval development of Bairdiella chrysura and Anchovia mitchilli. Bulletin of the United States Bureau of Fisheries 33:3–19. Kuraku, S., D. Hoshiyama, K. Katoh, H. Suga, and T. Miyata. 1999. Monophyly of Lampreys and Hagfishes supported by nuclear DNA-coded genes. Journal of Molecular Evolution 49:729–735. Kuraku, S., and S. Kuratani. 2006. Time scale for cyclostome evolution inferred with a phylogenetic diagnosis of Hagfish and Lamprey cDNA sequences. Zoological Science 23:1053–1064. Kuraku, S., A. Meyer, and S. Kuratani. 2008. Timing of genome duplications relative to the origin of the vertebrates: did cyclostomes diverge before or after? Molecular Biology and Evolution 26:47–59. Kurobe, T., G. O. Kelley, T. B. Waltzek, and R. P. Hedrick. 2008. Revised phylogenetic relationships among herpesviruses iso-

LITERATURE CITED

lated from Sturgeons. Journal of Aquatic Animal Health 20: 96–102. Kurtz, J., M. Kalbe, P. B. Aeschlimann, M. A. Häberli, K. M. Wegner, T. B. H. Reusch, and M. Milinski. 2004. Major histocompatibility complex diversity influences parasite resistance and innate immunity in Sticklebacks. Proceedings of the Royal Society of London B 271:197–204. Kyle, A. L., P. W. Sorensen, N. E. Stacey, and J. G. Dulka. 1987. Medial olfactory tract pathways controlling sexual reflexes and behavior in teleosts. Annals of the New York Academy of Sciences 519:97–107. Kynard, B. 1997. Life history, latitudinal patterns, and status of the Shortnose Sturgeon, Acipenser brevirostrum. Environmental Biology of Fishes 48:319–334. Kynard, B., and M. Horgan. 2001. Guidance of yearling Shortnose and Pallid Sturgeon using vertical bar rack and louver arrays. North American Journal of Fisheries Management 21:561–570. Kynard, B., and M. Horgan. 2002a. Ontogenetic behavior and migration of Atlantic Sturgeon, Acipenser oxyrinchus oxyrinchus, and Shortnose Sturgeon, A. brevirostrum, with notes on social behavior. Environmental Biology of Fishes 63:137–150. Kynard, B., and M. Horgan. 2002b. Attraction of prespawning male Shortnose Sturgeon Acipenser brevirostrum to the odor of prespawning females. Journal of Ichthyology 42:205–209. Kynard, B., and E. Parker. 2004. Ontogenetic behavior and migration of Gulf of Mexico Sturgeon, Acipenser oxyrinchus desotoi, with notes on body color and development. Environmental Biology of Fishes 70:43–55. Kynard, B., and E. Parker. 2005. Ontogenetic behavior and dispersal of Sacramento River White Sturgeon, Acipenser transmontanus, with a note on body color. Environmental Biology of Fishes 74:19–30. Kynard, B., E. Henyey, and M. Horgan. 2002a. Ontogenetic behavior, migration, and social behavior of Pallid Sturgeon, Scaphirhynchus albus, and Shovelnose Sturgeon, S. platorynchus, with notes on the adaptive significance of body color. Environmental Biology of Fishes 63:389–403. Kynard, B., M. Horgan, and M. Kieffer. 2000. Habitats used by Shortnose Sturgeon in two Massachusetts rivers, with notes on estuarine Atlantic Sturgeon: a hierarchical approach. Transactions of the American Fisheries Society 129:487–503. Kynard, B., E. Parker, and T. Parker. 2005. Behavior of early life intervals of Klamath River Green Sturgeon, Acipenser medirostris, with a note on body color. Environmental Biology of Fishes 72:85–97. Kynard, B., E. Parker, D. Pugh, and T. Parker. 2007. Use of laboratory studies to develop a dispersal model for Missouri River Pallid Sturgeon early life intervals. Journal of Applied Ichthyology 23:356–374. Kynard, B., P. Zhuang, L. Zhang, T. Zhang, and Z. Zhang. 2002b. Ontogenetic behavior and migration of Volga River Russian Sturgeon, Acipenser gueldenstaedtii, with a note on adaptive significance of body color. Environmental Biology of Fishes 65:411–421. Kynard, B, P. Zhuang, T. Zhang, and L. Zhang. 2003. Ontogenetic behavior and migration of Darby’s Sturgeon, Acipenser dabryanus, from the Yangtze River, with notes on body color and development rate. Environmental Biology of Fishes 66:27–36.

563

LaBar, G. W., and D. E. Facey. 1983. Local movements and inshore population sizes of American Eels in Lake Champlain, Vermont. Transactions of the American Fisheries Society 112:111–116. Laberge, F., and T. J. Hara. 2001. Neurobiology of fish olfaction: a review. Brain Research Reviews 36:46–59. Laberge, F., and T. J. Hara. 2003a. Non-oscillatory discharges of an F-prostaglandin responsive neuron population in the olfactory bulb-telencephalon transition area in Lake Whitefish. Neuroscience 116:1089–1095. Laberge, F., and T. J. Hara. 2003b. Behavioural and electrophysiological responses to F-prostaglandins, putative spawning pheromones, in three salmonid fishes. Journal of Fish Biology 62:206–221. Lacepède, B. G. E. 1797. Mémoire sur la Polyodon feuille. Bulletin de la Société Philomathique de Paris 1:49. Lachner, E. A. 1952. Studies of the biology of the cyprinid fishes of the chub genus Nocomis of Northeastern United States. The American Midland Naturalist 48:433–457. Lachner, E. A., and R. E. Jenkins. 1967. Systematics, distribution, and evolution of the chub genus Nocomis (Cyprinidae) in the southwestern Ohio River basin, with the description of a new species. Copeia 1967:557–580. Lagler, K. F. 1943. Food habits and economic relations of the turtles of Michigan with special reference to fish management. American Midland Naturalist 29:257–312. Lagler, K. F., and V. C. Applegate. 1942. Further studies of the food of the Bowfin (Amia calva) in southern Michigan, with notes on the inadvisability of using trapped fish in food analyses. Copeia 1942:190–191. Lagler, K. F., and F. V. Hubbs. 1940. Food of the long-nosed Gar (Lepisosteus osseus oxyurus) and the Bowfin (Amia calva) in southern Michigan. Copeia 1940:239–241. Lahnsteiner, F., B. Berger, T. Weismann, and R. A. Patzner. 1995. Fine structure and motility of the spermatozoa and composition of the seminal plasma in the perch, Perca fluviatilis. Journal of Fish Biology 47:492–508. Lahnsteiner, F., B. Berger, T. Weismann and R. A. Patzner. 1996. Motility of spermatozoa of Alburnus alburnus and its relationship to seminal plasma composition and sperm metabolism. Fish Physiology and Biochemistry 15:167–179. Lalancette, L. M. 1977. Feeding in White Suckers (Catostomus commersoni) from Gamelin Lake, Quebec, over a twelve month period. Canadian Field Naturalist 104:369–376. Lamb, T. D., S. P. Collin, and E. N. Pugh, Jr. 2007. Evolution of the vertebrate eye: opsins, photoreceptors, retina and eye cup. Nature Reviews / Neuroscience 8:960–975. Lambe, L. M. 1906. On Amyzon brevipinne Cope, from the Amyzon beds of the Southern Interior of British Columbia. Royal Society of Canada, Transactions 12:151–156. Lambert, J. G. D., and J. W. Resink. 1991. Steroid glucuronides as male pheromones in the reproduction of the African Catfish, Clarias gariepinus: a brief review. Journal of Steroid Biochemistry and Molecular Biology 40:549–556. Lambert, J. G. D., R. van den Hurk, W. G. E. J. Schoonen, J. W. Resink, and P. G. W. J. van Oordt. 1986. Gonadal steroidiogenesis and the possible role of steroid glucuronides as sex pheromones in two species of teleosts. Fish Physiology and Biochemistry 2:101–107. Lamberti, G. A., S. V. Gregory, L. R. Ashkenas, R. C. Wildman, and K. M. S. Moore. 1991. Stream ecosystem recovery following

564

LITERATURE CITED

a catastrophic debris flow. Canadian Journal of Fisheries and Aquatic Sciences 48:196–208. Lambou, V. W. 1961. Efficiency and selectivity of flag gillnets fished in Lake Bistineau, Louisiana. Proceedings of the Southeastern Association of Game and Fish Commissioners 15:319–359. Lamouroux, N., N. L. Poff, and P. L. Angermeier. 2002. Intercontinental convergence of stream fish community traits along geomorphic and hydraulic gradients. Ecology 83:1792–1807. Lamson, H. M., J. -C. Shiao, Y. Iizuka, W. N. Tzeng, and D. K. Cairns. 2006. Movement patterns of American Eels (Anguilla rostrata) between salt and fresh water in a small coastal watershed, based on otolith microchemistry. Marine Biology 149:1567–1576. Landini, W., and E. Menesini. 1978. L’ittiofauna plio-pleistocenica della sezione della Vrica (Crotone-Calabria). Bollettino della Societa Paleontologica Italiana 17:143–175. Landolt, J. C., and L. G. Hill. 1975. Observations of the gross structure and dimensions of the gills of three species of Gars (Lepisosteidae). Copeia 1975:470–475. Landry, C., D. Garant, P. Duchesne, and L. Bernatchez. 2001. Good genes as heterozygosity: the major histocompatibility complex and mate choice in Atlantic Salmon (Salmo salar). Proceedings of the Royal Society of London B 268:1279–1285. Laney, R. W., J. E. Hightower, B. R. Versak, M. F. Mangold, W. W. Cole, Jr., and S. E. Winslow. 2007. Distribution, habitat use, and size of Atlantic Sturgeon captured during Cooperative Winter Tagging Cruises, 1988–2006, p. 167–182. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Lang, N. J., K. J. Roe, C. B. Renaud, H. S. Gill, I. C. Potter, J. Freyhof, A. M. Naseka, P. Cochran, H. Espinosa Pérez, E. M. Habit, B. R. Kuhajda, D. A. Neely, Y. S. Reshetnikov, V. B. Salnikov, M. T. Stoumboudi, and R. L. Mayden. 2009. Novel relationships among Lampreys (Petromyzontiformes) revealed by a taxonomically comprehensive molecular dataset. American Fisheries Society Symposium 72:41–55. Langecker, T. G., H. Wilkens, and J. Parzefall. 1996. Studies on the trophic structure of an energy-rich Mexican cave (Cueva de las Sardinas) containing sulfurous water. Memoirs Biospeologie 23:121–125. Lankford, S. E., T. E. Adams, and J. J. Cech, Jr. 2003. Time of day and water temperature modify the physiological stress response in Green Sturgeon, Acipenser medirostris. Comparative Biochemistry and Physiology—Part A 135:291–302. Lanzing, W. J. R. 1958. Structure and function of the suction apparatus of the Lamprey. Proceedings. Koninklijke Nederlandse akademie van Wetenschappen 61:300–307. Lapolla, A. E. 2001a. Bay Anchovy, Anchoa mitchilli in Narragansett Bay, Rhode Island I: population structure, growth and mortality. Marine Ecology Progress Series 217:93–102. Lapolla, A. E. 2001b. Bay Anchovy, Anchoa mitchilli in Narragansett Bay, Rhode Island II: spawning season, hatch-date distribution and young-of-the-year growth. Marine Ecology Progress Series 217:103–109. Largiadér, C. R., V. Fries, and T. C. M. Bakker. 2001. Genetic analysis of sneaking and egg-thievery in a natural population of three-spined Stickleback (Gasterosteus aculeatus). Heredity 86:459–468.

Larimore, R. W. 1949. Changes in the cranial nerves of the Paddlefish, Polyodon spathula, accompanying development of the rostrum. Copeia 1949:204–212. Larimore, R. W. 1950. Gametogenesis of Polyodon spathula (Walbaum): a basis for regulation of the fishery. Copeia 1950:116–124. La Rivers, I. 1962. Fishes and Fisheries of Nevada. Nevada State Fish and Game Commission, Carson City. Lauder, G. V., Jr. 1979. Feeding mechanics in primitive teleosts and in the helecomorph fish Amia calva. J. Zool. London 187:543–578. Lauder, G. V. 1980. Evolution of the feeding mechanism in primitive actinopterygian fishes: a functional anatomical analyses of Polypterus, Lepisosteus and Amia. Journal of Morphology 163:283–317. Lauder, G. V., and K. F. Liem. 1983. The evolution and interrelationships of the actinopterygian fishes. Bulletin of the Museum of Comparative Zoology 150:1–197. Lauder, G. V., and S. F. Norton. 1980. Asymmetrical muscle activity during feeding in the Gar, Lepisosteus oculatus. Journal of Experimental Biology 84:17–32. Lauder, G. V., and P. C. Wainwright. 1992. Function and history: the pharyngeal jaw apparatus in primitive Ray-Finned Fishes, p. 456–491. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, California. Laurent, G. 1999. A systems perspective on early olfactory coding. Science 286:723–728. Lavoué, S., M. Miya, and M. Nishida. 2010. Mitochondrial phylogenomics of Anchovies (family Engraulidae) and recurrent origins of pronounced miniaturization in the order Clupeiformes. Molecular Phylogenetics and Evolution 56:480–485. Lavoué, S., M. Miya, K. Saitoh, N. B. Ishiguro, and M. Nishida. 2007. Phylogenetic relationships among Anchovies, sardines, Herrings and their relatives (Clupeiformes), inferred from whole mitogenome sequences. Molecular Phylogenetics and Evolution 43:1096–1105. Lavoué, S., and J. P. Sullivan. 2004. Simultaneous analysis of five molecular markers provides a well-supported phylogenetic analysis for the living bony-tongue fishes (Osteoglossomorpha: Teleostei). Molecular Phylogenetics and Evolution 2004:171–183. Lawlor, L. R. 1980. Structure and stability in natural and randomly constructed competitive communities. American Naturalist 116:394–408. Lawrence, B. J., and R. J. F. Smith. 1989. Behavioral responses of solitary Fathead Minnow, Pimephales promelas, to alarm substances. Journal of Chemical Ecology 15:209–219. Lê, H. L., V. G. Lecointre, and R. Perasso. 1993. A 28S rRNA-based phylogeny of the gnathostomes: first steps in the analysis of conflict and congruence with morphologically based cladograms. Molecular Phylogenetics and Evolution 2:31–51. Leach, B., and R. Montgomerie. 2000. Sperm characteristics associated with different male reproductive tactics in Bluegill Sunfish (Lepomis macrochirus). Behavioral Ecology and Sociobiology 49:31–37. Leak, J. C., and E. D. Houde. 1987. Cohort growth and survival of Bay Anchovy, Anchoa mitchilli larvae in Biscayne Bay, Florida. Marine Ecology Progress Series 37:109–122. LeBreton, G. T. O., and F. W. H. Beamish. 1998. The influence of salinity on ionic concentrations and osmolarity of blood serum

LITERATURE CITED

in Lake Sturgeon, Acipenser fulvescens. Environmental Biology of Fishes 52:477–482. LeBreton, G. T. O., and F. W. H. Beamish. 2004. Growth, bioenergetics and age, p. 195–216. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. H. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. LeBreton, G. T. O., F. W. Beamish, and R. S. McKinley (eds.). 2004. Sturgeons and Paddlefish of North America. Kluwer Academic Publishers, Dordrecht, The Netherlands. Lecomte-Finiger, R. 1994. The early life of the European Eel. Nature 370:424. LDWF (Louisiana Department of Wildlife and Fisheries). 2011a. Louisiana Fishing Regulations 2001-Recreational. Louisiana Department of Wildlife and Fisheries, Baton Rouge. LDWF (Louisiana Department of Wildlife and Fisheries). 2011b. Louisiana Commercial Fishing Regulations 2011. Louisiana Department of Wildlife and Fisheries, Baton Rouge. Leduc, A. O. H. C., J. M. Kelly, and G. E. Brown. 2004. Detection of conspecific alarm cue by juvenile salmonids under neutral and weakly acidic conditions: laboratory and field tests. Oecologia 139:318–324. Leduc, A. O. H., C. F. C. Lamaze, L. McGraw, and G. E. Brown. 2008. Response to chemical alarm cues under weakly acidic conditions: a graded loss of antipredator behaviour in juvenile Rainbow Trout. Water, Air and Soil Pollution 189:179–187. Leduc, A. O. H., C. M. K. Noseworthy, J. C. Adrian, Jr., and G. E. Brown. 2003. Detection of conspecific and heterospecific alarm signals by juvenile Pumpkinseed under weak acidic conditions. Journal of Fish Biology 63:1331–1336. Leduc, A. O. H., C. E. Roh, C. Breau, and G. E. Brown. 2007a. Learned recognition of a novel odour by wild juvenile Atlantic Salmon, Salmo salar, under fully natural conditions. Animal Behaviour 73:471–477. Leduc, A. O. H., C. E. Roh, C. Breau, and G. E. Brown. 2007b. Effects of ambient acidity on chemosensory learning: an example of an environmental constraint on acquired predator recognition in wild juvenile Atlantic Salmon (Salmo salar). Ecology of Freshwater Fish 16:385–394. Leduc, A. O. H., C. E. Roh, and G. E. Brown. 2009. Effects of acid rainfall on juvenile Atlantic Salmon (Salmo salar) antipredator behaviour: loss of chemical alarm function and potential survival consequences during predation. Marine and Freshwater Research 60:1223–1230. Leduc, A. O. H., C. E. Roh, M. C. Harvey, and G. E. Brown. 2006. Impaired detection of chemical alarm cues by juvenile wild Atlantic Salmon (Salmo salar) in a weakly acidic environment. Canadian Journal of Fisheries and Aquatic Sciences 63:2356–2363. Lee, D. S., C. R. Gilbert, C. H. Hocutt, R. E. Jenkins, D. E. McAllister, and J. R. Stauffer, Jr. 1980. Atlas of North American Freshwater Fishes. North Carolina State Museum of Natural History, Raleigh. Leidy, P. 1886. On Amia and its probable Taenia. Proceedings of the Academy of Natural Sciences of Philadelphia 38:62–63. Lein, G. M., and D. R. DeVries. 1997. Boat electrofishing as a technique for sampling Paddlefish. Transactions of the American Fisheries Society 126:334–337. Lein, G. M., and D. R. DeVries. 1998. Paddlefish in the Alabama River drainage: population characteristics and the adult spawn-

565

ing migration. Transactions of the American Fisheries Society 127:441–454. Lellis, W. A. 2002. Freshwater mussel survey of the Delaware Water Gap National Recreation Area: qualitative survey 2001. Report to the National Park Ser vice. Lenhardt, M., G. Markovic, A. Hegedis, S. Maletin, M. Cirkovic, and Z. Markovic. 2011. Non-native and translocated fish species in Serbia and their impact on the native ichthyofauna. Reviews of Fish Biology and Fisheries 21:407–421. Lennon, R. E. 1954. Feeding mechanism of the Sea Lamprey and its effect on host fishes. Fisheries Bulletin of the Fish and Wildlife Ser vice 56:247–293. Lenon, N., K. Stave, T. Burke, and J. E. Deacon. 2002. Bonytail (Gila elegans) may enhance survival of Razorback Suckers (Xyrauchen texanus) in rearing ponds by preying on exotic crayfish. Journal of the Arizona-Nevada Academy of Sciences 34:46–52. Lenz, T. L., C. Eizaguirre, J. P. Scharsack, M. Kalbe, and M. Milinski. 2009. Disentangling the role of MHC-dependent ‘good genes’ and ‘compatible genes’ in mate-choice decisions of threespined Sticklebacks, Gasterosteus aculeatus under semi-natural conditions. Journal of Fish Biology 75:2122–2142. Léonard, N. J., W. W. Taylor, and C. Goddard. 2004. Multijurisdictional management of Lake Sturgeon in the Great Lakes and St. Lawrence River, p. 231–251. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. Lesueur, C. A. 1817a. A new genus of fishes, of the order Abdominales, proposed, under the name of Catostomus; and the characters of this genus, with those of its species, indicated. Journal of the Academy of Natural Sciences, Philadelphia 1(pt 1, nos 5/6):88–96, 102–111. Lesueur, C. A. 1817b. A short description of five (supposed) new species of the genus Muraena, discovered by Mr. Le Sueur, in the year 1816. Journal of the Academy of Natural Sciences, Philadelphia v. 1 (pt 1, no. 5):81–83. [P. 88–96 published in Sept., p. 102–111 in Oct.] Lethbridge, R. C., and I. C. Potter. 1979. The oral fimbriae of the Lamprey, Geotria australis. Journal of Zoology, London 188:267–277. Levesley, P. B., and A. E. Magurran. 1988. Population differences in the reaction of minnows to alarm substance. Journal of Fish Biology 32:699–706. Levine, J. S., P. S. Lobel, and E. F. MacNichol. 1980. Visual communication in fishes, p. 447–475. In Environmental Physiology of Fishes. M. A. Ali (ed.). Plenum Press, New York. Leviton, A. E., and R. H. Gibbs, Jr. 1988. Standards in herpetology and ichthyology. Standard symbolic codes for institutional resource collections in herpetology and ichthyology. Supplement No. 1: additions and corrections. Copeia 1988:280–282. Leviton, A. E., R. H. Gibbs, Jr., E. Heal, and C. E. Dawson. 1985. Standards in herpetology and ichthyology: part I. Standard symbolic codes for institutional resource collections in herpetology and ichthyology. Copeia 1985:802–832. Lewis, S. V., and I. C. Potter. 1976a. Gill morphometrics of the Lampreys, Lampetra fluviatilis (L.) and Lampetra planeri (Bloch). Acta Zoologica 57:103–112. Lewis, S. V., and I. C. Potter. 1976b. A scanning electron microscope study of the gills of the Lamprey, Lampetra fluviatilis (L.). Micron 7:205–211.

566

LITERATURE CITED

Lewis, S. V., and I. C. Potter. 1977. Oxygen consumption during the metamorphosis of the parasitic Lamprey, Lampetra fluviatilis (L.) and its non-parasitic derivative, Lampetra planeri (Bloch). The Journal of Experimental Biology 69:187–198. Lewis, S. V., and I. C. Potter. 1982. A light and electron microscope study of the gills of larval Lampreys (Geotria australis) with particular reference to the water-blood pathway. Journal of Zoology, London 198:157–176. Lewis, T. C. 1982. The reproductive anatomy, seasonal cycles, and development of the Atlantic Stingray, Dasyatis sabina (LeSueur) (Pisces, Dasyatidae), from the Northeastern Gulf of Mexico. Unpubl. Ph.D. diss., Florida State University, Tallahassee. Lewontin, R. C., and L. C. Birch. 1966. Hybridization as a source of variation for adaptation to new environments. Evolution 20:315–336. Li, C., and G. Ortí. 2007. Molecular phylogeny of Clupeiformes (Actinopterygii) inferred from nuclear and mitochondrial DNA sequences. Molecular Phylogenetics and Evolution 44:386–398. Li, G.-Q., and M. V. H. Wilson. 1994. An Eocene species of Hiodon from Montana, its phylogenetic relationships, and the evolution of the postcranial skeleton in the Hiodontidae (Teleostei). Journal of Vertebrate Paleontology 14:153–167. Li, G.-Q., and M. V. H. Wilson. 1996. Phylogeny of Osteoglossomorpha, p. 163–174. In Interrelationships of Fishes. M. J. L. Stiassny, L. R. Parenti, and G. D. Johnson (eds.). Academic Press, San Diego, California. Li, G.-Q., and M. V. H. Wilson. 1999. Early divergence of Hiodontiformes sensu stricto in East Asia and phylogeny of some Late Mesozoic teleosts from China, p. 369–384. In Mesozoic Fishes 2—Systematics and Fossil Record. G. Arratia and H.-P. Schultze (eds.). Verlag Dr Friedrich Pfeil, München, Germany. Li, H. W., C. B. Schreck, C. E. Bond, and E. Rexstad. 1987. Factors influencing changes in fish assemblages of Pacific Northwest streams, p. 193–202. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Li, S. K., and D. H. Owings. 1978a. Sexual selection in the threespined Stickleback: I. Normative observations. Zeitschrift für Tierpsychologie 46:359–371. Li, S. K., and D. H. Owings. 1978b. Sexual selection in the threespined Stickleback: II. Nest raiding during the courtship phase. Behaviour 64:298–304. Li, W. M. 2005. Potential multiple functions of a male Sea Lamprey pheromone. Chemical Senses 30:I307–I308. Li, W., A. P. Scott, M. J. Siefkes, H. Yan, Q. Liu, S.-S. Yun, and D. A. Gage. 2002. Bile acid secreted by male Sea Lamprey that acts as a sex pheromone. Science 296:138–141. Li, W., A. P. Scott, M. J. Siefkes, S.-S. Yun, and B. Zielinski. 2003b. A male pheromone in the Sea Lamprey (Petromyzon marinus): an overview. Fish Physiology and Biochemistry 28:259–262. Li, W. M., M. J. Siefkes, A. P. Scott, and J. H. Teeter. 2003a. Sex pheromone communication in the Sea Lamprey: implications for integrated management. Journal of Great Lakes Research 29:85–94. Li, C. H., G. Q. Lu, and G. Orti. 2008. Optimal data partitioning and a test case for Ray-Finned Fishes (Actinopterygii) based on ten nuclear loci. Systematic Biology 57:519–539. Light, J. E., A. C. Fiumera, and B. A. Porter. 2005. Egg-feeding in the freshwater piscicolid leech, Cystobranchus virginicus (Annelida, Hirudinea). Invertebrate Biology 124:50–56.

Liley, N. R. 1966. Ethological isolating mechanisms in four sympatric species of poeciliid fishes. Behaviour Supplement 13: 1–197. Liley, N. R. 1982. Chemical communication in fish. Canadian Journal of Fisheries and Aquatic Sciences 39:22–35. Liley, N. R., K. H. Olsén, C. J. Foote, and G. J. Van Der Kraak. 1993. Endocrine changes associated with spawning behavior in male Kokanee Salmon (Oncorhynchus nerka) and the effects of anosmia. Hormones and Behavior 27:470–487. Liley, N. R., P. Tamkee, R. Tsai, and D. J. Hoysak. 2002. Fertilization dynamics in Rainbow Trout (Oncorhynchus mykiss): effect of male age, social experience, and sperm concentration and motility on in vitro fertilization. Canadian Journal of Fisheries and Aquatic Sciences 59:144–152. Lim, S. T., and G. S. Bailey. 1977. Gene duplication in the salmonid fishes: evidence for duplicated but catalytically equivalent A4 lactate dehydrogenases. Biochemical Genetics 15:707–721. Lima, N. R. W., C. J. Kobak, and R. C. Vrijenhoek. 1996. Evolution of sexual mimicry in sperm-dependent all-female forms of Poeciliopsis (Pisces: Poeciliidae). Journal of Evolutionary Biology 9:185–203. Lima, N. R. W., and R. C. Vrijenhoek. 1996. Avoidance of filial cannibalism by sexual and clonal forms of Poeciliopsis (Pisces: Poeciliidae). Animal Behaviour 51:293–301. Lima, S. L., and L. M. Dill. 1990. Behavioral decisions made under the risk of predation: a review and prospectus. Canadian Journal of Zoology 68:619–640. Linares-Casenave, J., J. P. Van Eenennaam, and S. I. Doroshov. 2002. Ultrastructural and histological observations on temperature-induced follicular ovarian atresia in the White Sturgeon. Journal of Applied Ichthyology 18:382–390. Lincicome, D. R., and H. J. Van Cleave. 1949. Distribution of Leptorhynchoides thecatus, a common acanthocephalan parasite in fishes. American Midland Naturalist 41:421–431. Lindley, S. T., M. L. Moser, D. L. Erickson, M. Belchik, D. W. Welch, E. L. Rechisky, J. T. Kelly, J. Heublein, and A. P. Klimley. 2008. Marine migration of North American Green Sturgeon. Transactions of the American Fisheries Society 137:182–194. Lindquist, D. G., J. R. Shute, and P. W. Shute. 1981. Spawning and nesting behavior of the Waccamaw Darter, Etheostoma perlongum. Environmental Biology of Fishes 6:177–191. Lindström, K., and N. Kangas. 1996. Egg presence, egg loss, and female mate preferences in the Sand Goby (Pomatoschistus minutus). Behavioral Ecology 7:213–217. Lindström, K., and R. C. Sargent. 1997. Food access, brood size and filial cannibalism in the Fantail Darter, Etheostoma flabellare. Behavioral Ecology and Sociobiology 40:107–110. Linhart, O., J. Cosson, S. D. Mims, W. L. Shelton, and M. Rodina. 2002. Effects of ions on the motility of fresh and demembranated Paddlefish (Polyodon spathula) spermatozoa. Reproduction 124:713–719. Linhart, O., and S. Kudo. 1997. Surface ultrastructure of Paddlefish eggs before and after fertilization. Journal of Fish Biology 51:573–582. Linnaeus, C. 1758. Systema Naturae, 10th Edition. Laurentii Salvii, Holmiae, 824 p. Lintlop, S. P., and J. H. Youson. 1983. Concentration of triiodothyronine in the sera of the Sea Lamprey, Petromyzon marinus, and the brook Lamprey, Lampetra lamottenii, at various phases of

LITERATURE CITED

their life cycle. General and Comparative Endocrinology 49: 187–194. Linton, E. D., B. Jónsson, and D. L. G. Noakes. 2007. Effects of water temperature on the swimming and climbing behavior of glass Eels, Anguilla spp. Environmental Biology of Fishes 78:189–192. Linville, R. G., S. N. Luoma, L. Cutter, and G. A. Cutter. 2002. Increased selenium threat as a result of invasion of the exotic bivalve Potamocorbula amurensis into the San Francisco Bay-Delta Aquatic Toxicology 57:51–64. Lippe, C., Dumont, P., Bernatchez, L. 2004. Isolation and identification of 21 microsatellite loci in the Copper Redhorse (Moxostoma hubbsi; Catostomidae) and their variability in other catostomids. Molecular Ecology Notes 4:638–641. Lippe, C., P. Dumont, and L. Bernatchez. 2006. High genetic diversity and no inbreeding in the endangered Copper Redhorse, Moxostoma hubbsi (Catostomidae, Pisces): the positive sides of a long generation time. Molecular Ecology 15:1769–1780. Litvak, M. K., and N. E. Mandrak. 1993. Ecology of freshwater baitfish use in Canada and the United States. Fisheries 18:6–13. Liu, C., and Y. Zeng. 1988. Notes on the Chinese Paddlefish, Psephurus gladius (Martens). Copeia 1988:482–484. Liu, J., and M.-m. Chang. 2009. A new Eocene catostomid (Teleostei: Cypriniformes) from northeastern China and early divergence of Catostomidae. Science In China Series D-Earth Sciences 52:189–202. Livingston, C. A., and R. B. Leonard. 1990. Locomotion evoked by stimulation of the brain stem in the Atlantic Stingray, Dasyatis sabina. Journal of Neuroscience 10:194–204. Livingston, R. J. 1982. Trophic organization of fishes in a coastal seagrass system. Marine Ecology Progress Series 7:1–12. Livingston, R. J., G. J. Kobylinski, F. G. Lewis III, and P. F. Sheridan. 1976. Long-term fluctuations of epibenthic fish and invertebrate populations in Apalachicola Bay, Florida. Fishery Bulletin 74:311–321. Lobchenko, V., A. Vedrasco, and R. Billard. 2002. Rearing Paddlefish, Polyodon spathula to maturity in ponds in the Republic of Moldavia. International Review of Hydrobiology 87:553–559. Locatello, L., M. B. Rasotto, B. Adiaenssens, and A. Pilastro. 2008. Ejaculate traits in relation to male body size in the Eastern Mosquitofish, Gambusia holbrooki. Journal of Fish Biology 73:1600–1611. Lodge, D. M. 1993. Biological invasions: lessons for ecology. Trends in Ecology and Evolution 8:133–137. Loew, E. R., and J. N. Lythgoe. 1978. The ecology of cone pigments in teleost fish. Vision Research 16:851–856. Loew, E. R., and A. J. Sillman. 1993. Age-related changes in the visual pigments of the White Sturgeon (Acipenser transmontanus). Canadian Journal of Zoology 71:1552–1557. Loftus, K. H. (ed.). 1982. Proceedings of the 1980 North American Eel Conference. Ontario Fisheries Technical Report Series No. 4. Loftus, W. F., and J. A. Kushlan. 1987. Freshwater fishes of southern Florida. Bulletin of the Florida State Museum, Biological Sciences 31:147–344. Logiudice, F. T., and R. J. Laird. 1994. Morphology and density distribution of cone photoreceptor in the retina of the Atlantic Stingray, Dasyatis sabina. Journal of Morphology 221:277–289.

567

Lohr, S. C., and K. D. Fausch. 1997. Multiscale analysis of natural variability in stream fish assemblages of a western Great Plains watershed. Copeia 1997:706–724. Loiselle, P. V. 1982. Male spawning-partner preference in an arena-breeding teleost Cyprinodon macularius californiensis Girard (Atherinomorpha: Cyprinodontidae). American Naturalist 120:721–732. Long, J. H., M. E. Hale, M. J. McHenry, and M. W. Westneat. 1996. Functions of fish skin: flexural stiff ness and steady swimming of Longnose Gar, Lepisosteus osseus. Journal of Experimental Biology 199:2139–2151. Long, W. L., and W. W. Ballard. 1976. Normal embryonic stages of the White Sucker, Catostomus commersonii. Copeia 1976:342–351. Long, W. L., and W. W. Ballard. 2001. Normal embryonic stages of the Longnose Gar, Lepisosteus osseus. BMC Developmental Biology 1:6. Available from http://www.biomedcentral.com/1471– 213X/1/6; as of February 2012. Lonzarich, D. G., M. L. Warren, Jr., and M. R. Lonzarich. 1998. Effects of habitat isolation on the recovery of fish assemblages in experimentally defaunated stream pools in Arkansas. Canadian Journal of Fisheries and Aquatic Sciences 55:2141–2149. Lonzarich, D. G., M. R. Lonzarich, and M. L. Warren, Jr. 2000. Effects of riffle length on the short-term movement of fishes among stream pools. Canadian Journal of Fisheries and Aquatic Sciences 57:1508–1514. Lookabaugh, P. S., and P. L. Angermeier. 1992. Diet patterns of American Eel, Anguilla rostrata, in the James River drainage, Virginia. Journal of Freshwater Ecology 7:425–431. Lorenz, J., and J. Serafy. 2006. Subtroprical wetland fish assemblages and changing salinity regimes: implications for everglades restoration. Hydrobiologia 569:401–422. Lorion, C. M., D. F. Markle, S. B. Reid, and M. F. Docker. 2000. Redescription of the presumed-extinct Miller Lake Lamprey, Lampetra minima. Copeia 2000:1019–1028. Losey, G. S., Jr., F. G. Stanton, T. M. Telecky, and W. A. Tyler III. 1986. The Zoology 691 graduate seminar class, 1986. Copying others, an evolutionary stable strategy for mate choice: a model. American Naturalist 128:653–664. Losey, G. S., T. W. Cronin, T. H. Goldsmith, D. Hyde, N. J. Marshall, and W. N. McFarland. 1999. The UV visual world of fishes: a review. Journal of Fish Biology 54:921–943. Love, J. W. 2002. Sexual dimorphism in Spotted Gar, Lepisosteus oculatus from southeastern Louisiana. The American Midland Naturalist 147:393–399. Love, J. W. 2004. Age, growth, and reproduction of Spotted Gar, Lepisosteus oculatus (Lepisosteidae), from the Lake Pontchartrain estuary, Louisiana. The Southwestern Naturalist 49:18–23. Lovejoy, N. 1996. Systematics of myliobatoid elasmobranchs: with emphasis on the phylogeny and historical biogeography of neotropical freshwater Stingrays (Pomatotrygonidae: Rajiformes). Zoological Journal of the Linnaean Society 117:207–257. Lovell, J. M., M. M. Findlay, G. M. Harper, and R. M. Moate. 2006. The inner ear ultrastructure from the Paddlefish (Polyodon spathula) using transmission electron microscopy. Journal of Microscopy 222:36–41. Lovell, J. M., M. M. Findlay, R. M. Moate, J. R. Nedwell, and M. A. Pegg. 2005. The inner ear morphology and hearing abilities of the Paddlefish (Polyodon spathula) and the Lake Sturgeon

568 LITERATURE CITED

(Acipenser fulvescens). Comparative Biochemistry and Physiology, Part A 142:286–296. Lovis, W. A. 1985. Seasonal settlement dynamics and the role of the Fletcher site in the woodland adaptations of the Saginaw drainage basin. Arctic Anthropology 22:153–170. Lowe, D. R., F. W. H. Beamish, and I. C. Potter. 1973. Changes in the proximate body composition of the landlocked Sea Lamprey, Petromyzon marinus (L.) during larval life and metamorphosis. Journal of Fish Biology 5:673–682. Lowe, D. W., J. R. Matthews, C. J. Moseley, and W. Beacham. 1990. The Official World Wildlife Fund Guide to Endangered Species of North America. Beacham Publications, Washington D.C. Lowe, J. J., and M. J. C. Walker. 1997. Reconstructing Quaternary Environments. 2nd edition. Addison Wesley Longman, Ltd., Essex, England, United Kingdom. Lucas, M. C., and E. Baras. 2001. Migration of Freshwater Fishes. Blackwell Science, Malden, Massachusetts. Luczkovich, J. J., M. W. Sprague, S. E. Johnson, and R. C. Pullinger. 1999. Delimiting spawning areas of Weakfish, Cynoscion regalis (family Sciaenidae), in Pamlico Sound, North Carolina using passive hydroacoustic surveys. Bioacoustics 10:143–160. Ludwig, A. 2006. A Sturgeon view of conservation genetics. European Journal of Wildlife Research 52:3–8. Ludwig, A. 2008. Identification of Acipenseriformes species in trade. Journal of Applied Ichthyology 24(Supplement 1):2–19. Ludwig, A., N. M. Belfiore, C. Pitra, V. Svirsky, and I. Jenneckens. 2001. Genome duplication events and functional reduction of ploidy levels in Sturgeon (Acipenser, Huso, and Scaphirhynchus). Genetics 158:1203–1215. Ludwig, A., S. Lippold, L. Debus, and R. Reinartz. 2009. First evidence of hybridization between endangered Sterlets (Acipenser ruthenus) and exotic Siberian Sturgeons (Acipenser baerii) in the Danube River. Biological Invasions 11:753–760. Ludwig, A., B. May, L. Debus, and I. Jenneckens. 2000. Heteroplasmy in the mtDNA control region of Sturgeon (Acipenser, Huso, and Scaphirhynchus). Genetics 156:1933–1947. Luer, C. A., C. J. Walsh, A. B. Bodine, and J. T. Wyffels. 2007. Normal embryonic development in the Clearnose Skate, Raja eglanteria, with experimental observations on artificial insemination. Environmental Biology of Fishes 80:239–255. Lund, R., and E. D. Grogan. 1997. Relationships of the Chimaeriformes and the basal radiation of the Chondrichthyes. Reviews in Fish Biology and Fisheries 7:65–123. Lundberg, J. G. 1967. Pleistocene fishes of the Good Creek Formation, Texas. Copeia 1967:453–455. Lundberg, J. G., M. Kottelat, G. R. Smith, M. L. J. Stiassny, and A. C. Gill. 2000. So many fishes, so little time: an overview of recent ichthyological discovery in continental waters. Annals of the Missouri Botanical Garden 87:26–62. Lundberg, J. G., and E. Marsh. 1976. Evolution and functional anatomy of the pectoral fin rays in cyprinoid fishes, with emphasis on the Suckers (Family Catostomidae). American Midland Naturalist 96:332–349. Luo, J. 1993. Tidal transport of the Bay Anchovy, Anchoa mitchilli, in darkness. Journal of Fish Biology 42:531–539. Luo, J., and S. B. Brandt. 1993. Bay Anchovy, Anchoa mitchilli production and consumption in mid-Chesapeake Bay based on a bioenergetics model and acoustic measures of fish abundance. Marine Ecology Progress Series 98:223–236.

Luo, J., and J. A. Musick. 1991. Reproductive biology of the Bay Anchovy in Chesapeake Bay. Transactions of the American Fisheries Society 120:701–710. Lutnesky, M. M. F., and J. W. Adkins. 2003. Putative chemical inhibition of development by conspecifics in mosquitofish, Gambusia affinis. Environmental Biology of Fishes 66:181–186. Lutterschmidt, W. I., and V. H. Hutchison. 1997. The critical thermal maximum: history and critique. Canadian Journal of Zoology 75:1561–1574. Lutz, P. 1975. Adaptive and evolutionary aspects of the ionic content of fishes. Copeia 1975:369–373. Lyons, J. 1989. Changes in the abundance of small littoral-zone fishes in Lake Mendota, Wisconsin. Canadian Journal of Zoology 67:2910–2916. Lyons, J. 1993. Status and biology of Paddlefish (Polyodon spathula) in the lower Wisconsin River. Transactions of the Wisconsin Academy of Science, Arts and Letters 81:123–135. Lyons, J., and J. J. Kempinger. 1992. Movements of adult Lake Sturgeon in the Lake Winnebago system. Wisconsin Department of Natural Resources Research Report 156. Lythgoe, J. N. 1980. Vision in fishes: ecological adaptations, p. 431–445. In Environmental Physiology of Fishes. M. A. Ali (ed.). Plenum Press, New York. Ma, X., X. Bangxi, W. Yindong, and W. Mingxue. 2003. Intentionally introduced and transferred fishes in China’s inland waters. Asian Fisheries Science 16:279–290. Mabee, P. M., and M. Noordsy. 2004. Development of paired fins in the Paddlefish, Polyodon spathula. Journal of Morphology 261:334–344. MacAlpin, A. 1947. Paleopsephurus wilsoni, a new polyodontid fish from the Upper Cretaceous of Montana, with a discussion of allied fishes, living and fossil. Contributions from the Museum of Paleontology, University of Michigan 6:167–234. MacArthur, R. H., and E. O. Wilson. 1967. The Theory of Island Biogeography. Princeton University Press, Princeton, New Jersey. MacAvoy, S. E., S. A. Macko, S. P. McIninch, and G. C. Garman. 2000. Marine nutrient contributions to freshwater apex predators. Oecologia 122:568–573. MacConnell, E., R. P. Hedrick, C. Hudson, and C. A. Speer. 2001. Identification of an iridovirus in cultured Pallid (Scaphirhynchus albus) and Shovelnose Sturgeon (S. platorynchus). Fish Health Newsletter 29:1–3. MacDonald, D. D., M. G. Ikonomou, A. Rantalaine, I. H. Rogers, D. Sutherland, and J. Van Oostdam. 1997. Contaminants in White Sturgeon (Acipenser transmontanus) from the upper Fraser River, British Columbia, Canada. Environmental Toxicology & Chemistry 16:479–490. Macesic, L. J., and S. M. Kajiura. 2010. Comparative punting kinematics and pelvic fin musculature of benthic batoids. Journal of Morphology 271:1219–1228. Macey, D. J., and I. C. Potter. 1978. Lethal temperatures of ammocoetes of the Southern Hemisphere Lamprey, Geotria australis Gray. Environmental Biology of Fishes 3:241–243. Macey, D. J., and I. C. Potter. 1982. The effect of temperature on the oxygen dissociation curves of whole blood of larval and adult Lampreys (Geotria australis). The Journal of Experimental Biology 97:253–261. MacGregor, J. M., and E. D. Houde. 1996. Onshore-offshore pattern and variability in distribution and abundance of Bay An-

LITERATURE CITED

chovy, Anchoa mitchilli eggs and larvae in Chesapeake Bay. Marine Ecology Progress Series 138:15–25. MacGregor, R., J. Casselman, L. Greig, W. A. Allen, L. McDermott, and T. Haxton. 2010. DRAFT Recovery Strategy for the American Eel (Anguilla rostrata) in Ontario. Ontario Recovery Strategy Series. Prepared for Ontario Ministry of Natural Resources, Peterborough, Ontario. MacGregor, R. B., J. M. Casselman, W. A. Allen, T. Haxton, J. M. Dettmers, A. Mathers, S. LaPan, T. C. Pratt, P. Thompson, M. Stanfield, L. Marcogliese, and J.-D. Dutil. 2009. Natural heritage, anthropogenic impacts and bio-political issues related to the status and sustainable management of American Eel: a retrospective analysis and management perspective at the population level, p. 713–739. In Challenges for Diadromous Fishes in a Dynamic Global Environment. A. J. Haro, K. L. Smith, R. A. Rulifson, C. M. Moffitt, R. J. Klauda, M. J. Dadswell, R. A. Cunjak, J. E. Cooper, K. L. Beal, and T. S. Avery (eds.). American Fisheries Society, Symposium 69, Bethesda, Maryland. MacGregor, R. B., A. Mathers, P. Thompson, J. M. Casselman, J. M. Dettmers, S. LaPan, T. C. Pratt, and W. A. Allen. 2008. Declines of American Eel in North America: complexities associated with bi-national management, p. 357–381. In International Governance of Fisheries Ecosystems: Learning from the Past, Finding Solutions for the Future. M. G. Schechter, W. W. Taylor, and N. J. Leonard (eds.). American Fisheries Society, Bethesda, Maryland. Machado, M. D., D. C. Heins, and H. L. Bart. 2002. Microgeographical variation in ovum size of the Blacktail Shiner, Cyprinella venusta Girard, in relation to stream flow. Ecology of Freshwater Fish. 11:11–19. Macías Garcia, C. 1991. Sexual behaviour and trade-offs in the viviparous fish Girardinichthys multiradiatus. Unpubl. Ph.D. diss., University of East Anglia, Norwich, United Kingdom. Macías Garcia, C., and T. B. de Perera. 2002. Ultraviolet-based female preferences in a viviparous fish. Behavioral Ecology and Sociobiology 52:1–6. Macías Garcia, C., G. Jimenez, and B. Contreras. 1994. Correlational evidence of a sexually-selected handicap. Behavioral Ecology and Sociobiology 35:253–259. Macías Garcia, C., E. Saborío, and C. Berea. 1998. Does malebiased predation lead to male scarcity in viviparous fish? Journal of Fish Biology 53 (Suppl. A):104–117. MacKay, H. H. 1963. Fishes of Ontario. Ontario Department of Lands and Forests, Toronto. MacKiewicz, M., D. E. Fletcher, S. D. Wilkins, J. A. DeWoody, and J. C. Avise. 2002. A genetic assessment of parentage in a natural population of Dollar Sunfish (Lepomis marginatus) based on microsatellite markers. Molecular Ecology 11:1877–1883. MacLaren, R. D. 2006. The effects of male proximity, apparent size, and absolute size on female preference in the Sailfin Molly, Poecilia latipinna. Behaviour 143:1457–1472. MacLaren, R. D., and W. J. Rowland. 2006. Female preference for male lateral projection area in the Shortfin Molly, Poecilia mexicana: evidence for a pre-existing bias in sexual selection. Ethology 112:678–690. MacLaren, R. D., W. J. Rowland, and N. Morgan. 2004. Female preferences for sailfin and body size in the Sailfin Molly, Poecilia latipinna. Ethology 110:363–379. MacLean, J., and J. J. Manguson. 1977. Species interactions in percid communities. Journal of the Fisheries Research Board of Canada 34:1941–1951.

569

Maclean, N., and R. D. Jurd. 1972. The control of haemoglobin synthesis. Biological Reviews 47:393–437. Madenjian, C. P., T. J. Desorcie, J. R. McClain, A. P. Woldt, J. D. Holuszko, and C. A. Bowen II. 2004. Status of Lake Trout rehabilitation on Six Fathom Bank and Yankee Reef in Lake Huron. North American Journal of Fisheries Management 24:1003–1016. Madsen, M. L. 1971. The presence of nuptial tubercles on female Quillback (Carpiodes cyprinus). Transactions of the American Fisheries Society 100:132–134. Maekawa, K. 1983. Streaking behavior of mature male parrs of the Miyabe Charr, Salvelinus malma miyabei, during spawning. Japanese Journal of Ichthyology 30:227–234. Maekawa, K., and T. Hino. 1986. Spawning behavior of Dolly Varden in southeastern Alaska, with special reference to the mature male parr. Japanese Journal of Ichthyology 32:454–458. Maekawa, K., and H. Onozato. 1986. Reproductive tactics and fertilization success of mature male Miyabe Char, Salvelinus malma miyabei. Environmental Biology of Fishes 15:119–129. Magnan, P. 1988. Interactions between brook charr, Salvelinus fontinalis, and nonsalmonid species: ecological shift, morphological shift, and their impact on zooplankton communities. Canadian Journal of Fisheries and Aquatic Sciences 45:999–1009. Magnhagen, C. 1994. Sneak or challenge: alternative spawning tactics in non-territorial male Common Gobies. Animal Behaviour 47:1212–1215. Magnuson, J. J., and R. C. Lathrop. 1992. Historical changes in the fish community, p. 193–231. In Food Web Management: A Case Study of Lake Mendota, Wisconsin. J. F. Kitchell (ed.). Springer Verlag, New York. Magnuson, J. J., B. J. Benson, and A. S. McLain. 1994. Insights on species richness and turnover from long-term ecological research: fishes in north temperate lakes. American Zoologist 34:437–451. Magurran, A. E., and A. Higham. 1988. Information transfer across fish shoals under predator threat. Ethology 78:153–158. Magurran, A. E., P. W. Irving, and P. A. Henderson. 1996. Is there a fish alarm pheromone? A wild study and critique. Proceedings of the Royal Society of London B 263:1551–1556. Magurran, A. E., and B. H. Seghers. 1994. A cost of sexual harassment in the Guppy, Poecilia reticulata. Proceedings of the Royal Society of London B 258:89–92. Mahon, R., and C. B. Portt. 1985. Local size related segregation of fishes in streams. Archives of Hydrobiology 103:267–271. Mahy, G. 1975a. Ostéologie comparée et phylogénie des poissons cyprinoïdes I. Ostéologie crâneinne du goujon à fines écailles, Chrosomus neogaeus (Cope). Le Naturaliste Canadien 102:1–31. Mahy, G. 1975b. Ostéologie comparée et phylogénie des poissons cyprinioïdes, II. L’appareil de Wéber, le squelette axial et les ceintures du goujon à fines écailles Chrosomus neogaeus (Cope). Le Naturaliste Canadien. 102:165–180. Mahy, G. 1975c. Ostéologie comparée et phylogénie des poissons cyprinoïdes III. Ostéologie comparée de C. erythrogaster Rafinesque, C. eos Cope, C. oreas Cope, C. neogaeus (Cope), et P. phoxinus (Linné) et phylogénie du genre Chrosomus. Le Naturaliste Canadien. 102:617–642. Mailhot, Y., P. Dumont, and N. Vachon. 2011. Management of the Lake Sturgeon Acipenser fulvescens population in the lower St Lawrence River (Québec, Canada) from the 1910s to the present. Journal of Applied Ichthyology 27:405–410.

570

LITERATURE CITED

Maisey, J. G., G. J. P. Naylor, and D. J. Ward. 2004. Mezoic elasmobranchs, neoselachian phylogeny, and the rise of modern neoselachian diversity, p. 17–56. In Mesozoic Fishes 3—Systematics, Paleoenvironments, and Biodiversity. G. Arratia and A. Tintori (eds.). Verlag Dr. Friedrich Pfeil, Munchen, Germany. Makeyeav, A. P. 1980. Early ontogenetic characteristics of the Bigmouth Buffalo, Ictiobus cyprinellus (Catostomidae). Journal of Ichthyology 20:73–89. Mallatt, J., and J. Sullivan. 1998. 28S and 18S rDNA sequences support the monophyly of Lampreys and Hagfishes. Molecular Biology and Evolution 15:1706–1718. Mandrak, N. E. 1995. Biogeographic patterns of fish species richness in Ontario lakes in relation to historical and environmental factors. Canadian Journal of Fisheries and Aquatic Sciences 52:1462–1474. Mandrak, N. E., and E. J. Crossman. 1992. Postglacial dispersal of freshwater fishes into Ontario. Canadian Journal of Zoology 70:2247–2259. Manion, P. J., and B. R. Smith. 1978. Biology of larval and metamorphosing Sea Lampreys, Petromyzon marinus, of the 1960 year class in the Big Garlic River, Michigan, Part II, 1966–1972. Great Lakes Fisheries Commission Technical Report No. 30:1–35. Manley, G. A., and J. A. Clack. 2004. An outline of the evolution of vertebrate hearing organs, p. 1–26. In Evolution of the Vertebrate Auditory System. G. A. Manley, A. N. Popper, and R. R. Fay (eds.). Springer-Verlag, New York. Mann, D. A., P. A. Cott, B. W. Hanna, and A. N. Popper. 2007. Hearing in eight species of northern Canadian freshwater fishes. Journal of Fish Biology 70:109–120. Mann, D. A., D. M. Higgs, W. N. Tavolga, M. J. Souza, and A. N. Popper. 2001. Ultrasound detection by clupeiform fishes. Journal of the Acoustical Society of America 109:3048–3054. Manny, B. A., and G. W. Kennedy. 2002. Known Lake Sturgeon (Acipenser fulvescens) spawning habitat in the channel between lakes Huron and Erie in the Laurentian Great Lakes. Journal of Applied Ichthyology 18:486–490. Mansueti, A. J., and J. D. Hardy, Jr. 1967. Development of Fishes of the Chesapeake Bay Region: An Atlas of Egg, Larval, and Juvenile Stages, Part 1. Natural Resources Institute University of Maryland, College Park. Manwell, C. 1963. The blood proteins of cyclostomes. A study in phylogenetic and ontogenetic biochemistry, p. 372–455. In The Biology of Myxine. A. Brodal and R. Fänge (eds.). Universitetsforlaget, Oslo. Marcogliese, D. J. 1991. Seasonal occurrence of Lernaea cyprinacea on fishes in Belews Lake, North Carolina. Journal of Parasitology 77:326–327. Marcogliese, D. J., and D. K. Cone. 1993. What metazoan parasites tell us about the evolution of American and European Eels. Evolution 47:1632–1635. Marcogliese, D. J., and D. Cone. 2001. Myxozoan communities parasitizing Notropis hudsonius (Cyprinidae) at selected localities on the St. Lawrence River, Quebec: possible effects of urban effluents. Journal of Parasitology 88:467–473. Marcoux, R. G. 1966. Occurrence of a melanistic Paddlefish (Polyodon spathula) in Montana. Copeia 1966:876. Marcus, J. P, and G. E. Brown. 2003. Response of Pumpkinseed Sunfish to conspecific alarm cues: an interaction between ontogeny and stimulus concentration. Canadian Journal of Zoology 81:1671–1677.

Marcy, B. C., Jr., D. E. Fletcher, F. D. Martin, M. H. Paller, and M. J. M. Reichert. 2005. Fishes of the Middle Savannah River Basin. The University Press of Georgia, Athens. Marentette, J. R., and L. D. Corkum. 2008. Does the reproductive status of male Round Gobies (Neogobius melanostomus) influence their response to conspecific odours? Environmental Biology of Fish 81:447–455. Margolis, L., and J. R. Arthur. 1979. Synopsis of the parasites of fishes of Canada. Bulletin of the Fisheries Research Board of Canada. Marin, B. A., and M. K. Saiki. 1999. Effects of ambient water quality on the endangered Lost River Sucker in Upper Klamath Lake. Transactions of the American Fisheries Society 128:953–961. Mark, E. L. 1890. Studies on Lepisosteus. Part 1. Bulletin of the Museum of Comparative Zoology 19:1–128. Markle, D. F. 1997. Audubon’s hoax: Ohio River fishes described by Rafinesque. Archives of Natural History 24:439–447. Markle, D. F., M. R. Cavalluzzi, and D. C. Simon. 2005. Morphology and taxonomy of Klamath Basin Suckers (Catostomidae). Western North American Naturalist 65:473–489. Markle, D. F., and K. Clauson. 2009. Ontogenetic and habitatrelated changes in diet of late larval and juvenile Suckers (Catostomidae) in Upper Klamath Lake, Oregon. Western North American Naturalist 66:492–501. Markle, D. F., and M. S. Cooperman. 2001. Relationships between Lost River and Shortnose Sucker biology and management of Upper Klamath Lake, p. 93–118. In Water Allocation in the Klamath Reclamation Project, 2001: An Assessment of Natural Resource, Economic, Social, and Institutional Issues with a Focus on the Upper Klamath Basin. Oregon State University Extension Publication, Corvallis. Markle, D. F., and L. K. Dunsmoor. 2007. Effects of habitat volume and Fathead Minnow introduction on larval survival of two endangered Sucker species in upper Klamath Lake, Oregon. Transactions of the American Fisheries Society 136:567–579. Markle, D. F., T. N. Pearsons, and D. T. Bills. 1991. Natural history of Oregonichthys (Pisces: Cyprinidae), with a description of a new species from the Umpqua River of Oregon. Copeia 1991:277–293. Markus, H. C. 1934. Life history of the blackhead minnow (Pimephales promelas). Copeia 1934:116–112. Marler, C. A., and M. J. Ryan. 1997. Origin and maintenance of a female mating preference. Evolution 51:1244–1248. Marley, R. D. 1983. Spatial distribution patterns of planktonic fish eggs in lower Mobile Bay, Alabama. Transactions of the American Fisheries Society 112:257–266. Marrin, D. L. 1983. Ontogenetic changes and intraspecific resource partitioning in the Tahoe Sucker, Catostomus tahoensis. Environmental Biology of Fishes 8:39–47. Marsh, P. C. 1985. Effect of incubation temperature on survival of embryos of native Colorado River fishes. The Southwestern Naturalist 30:129–140. Marsh, P. C. 1987. Digestive tract contents of adult Razorback Suckers in Lake Mohave, Arizona-Nevada. Transactions of the American Fisheries Society 116:117–119. Marsh, P. C. 1996. Threatened fishes of the world: Xyrauchen texanus (Abbott, 1860) (Catostomidae). Environmental Biology of Fishes 45:258.

LITERATURE CITED

Marsh, P. C., and J. E. Brooks. 1989. Predation by introduced ictalurid Catfishes as a deterrent to re-establishment of hatcheryreared Razorback Suckers. Southwestern Naturalist 34: 188–195. Marsh, P. C., C. A. Pacey, and B. R. Kesner. 2003. Decline of the Razorback Sucker in Lake Mohave, Colorado River, Arizona and Nevada. Transactions of the American Fisheries Society 132:1251–1256. Marsh-Matthews, E., and W. J. Matthews. 2000. Geographic, terrestrial and aquatic factors: which most influence the structure of stream fish assemblages in the midwestern United States? Ecology of Freshwater Fish 9:9–21. Marsh-Matthews, E., and W. J. Matthews. 2002. Temporal stability of minnow species co-occurrence in streams of the central United States. Transactions of the Kansas Academy of Science 105:162–177. Marsh-Matthews, E., and W. J. Matthews. 2010. Proximate and residual effects of exposure to simulated drought on prairie stream fishes, p. 461–486. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society, Symposium 73, Bethesda, Maryland. Marsh-Matthews E., W. J. Matthews, and N. R. Franssen. 2011. Can a highly invasive species re-invade its native habitat? The paradox of the Red Shiner. Biological Invasions 13:2911–2924. Marsh-Matthews, E., W. J. Matthews, K. B. Gido, and R. L. Marsh. 2002. Reproduction by young-of-year Red Shiner (Cyprinella lutrensis) and its implications for invasion success. Southwestern Naturalist 47:605–610. Martin, B. A., and M. K. Saiki. 1999. Effects of ambient water quality on the endangered Lost River Sucker in Upper Klamath Lake, Oregon. Transactions of the American Fisheries Society 128:953–961. Martin, R. G. 1975. Sexual and aggressive behavior, density, and social structure in a natural population of mosquitofish, Gambusia affinis holbrooki. Copeia 1975:445–454. Martin, R. G. 1977. Density dependent aggressive advantage in melanistic male mosquitofish, Gambusia affinis holbrooki (Girard). Florida Science 40:393–400. Martino, E. J., and K. W. Able. 2003. Fish assemblages across the marine to low salinity transition zone of a temperate estuary. Estuarine, Coastal, and Shelf Science 56:969–987. Maruska, K. P. 2001. Morphology of the mechanosensory lateral line system in elasmobranch fishes: ecological and behavioral considerations. Environmental Biology of Fishes 60:47–75. Maruska, K. P., E. G. Cowie, and T. C. Tricas. 1996. Periodic gonadal activity and protracted mating in elasmobranch fishes. Journal of Experimental Zoology 276:219–332. Maruska, K. P., and T. C. Tricas. 1998. Morphology of the mechanosensory lateral line system in the Atlantic Stingray, Dasyatis sabina: the mechanotactile hypothesis. Journal of Morphology 238:1–22. Marvit, P., and J. D. Crawford. 2000. Auditory discrimination in a sound-producing electric fish (Pollimyrus): tone frequency and click-rate difference detection. Journal of the Acoustical Society of America 108:1819–1825. Mason, W. T., Jr., and J. P. Clugston. 1993. Foods of the Gulf Sturgeon in the Suwannee River, Florida. Transactions of the American Fisheries Society 122:378–385.

571

Massman, W. H. 1954. Marine fishes in fresh and brackish waters of Virginia rivers. Ecology 35:75–78. Mateos, M., O. I. Sanjur, and R. C. Vrijenhoek. 2002. Historical biogeography of the livebearing fish genus Poeciliopsis (Poeciliidae: Cyprinodontiformes). Evolution 56:972–984. Matheney, M. P., and C. F. Rabeni. 1995. Patterns of movement and habitat use by Northern Hog Suckers in an Ozark Stream. Transactions of the American Fisheries Society 124:886–897. Mathis, A., D. P. Chivers, and J. F. Smith. 1993. Population differences in responses of Fathead Minnows (Pimephales promelas) to visual and chemical stimuli from predators. Ethology 93: 31–40. Mathis, A., D. P. Chivers, and J. F. Smith. 1996. Cultural transmission of predator recognition in fishes: intraspecific and interspecific learning. Animal Behaviour 51:185–201. Mathis, A., and R. J. F. Smith. 1992. Avoidance of areas marked with a chemical alarm substance by Fathead Minnows (Pimephales promelas) in a natural habitat. Canadian Journal of Zoology 70:1473–1476. Mathis, A., and R. J. F. Smith. 1993a. Intraspecific and crosssuperorder responses to chemical alarm signals by Brook Stickleback. Ecology 74:2395–2404. Mathis, A., and R. J. F. Smith. 1993b. Fathead Minnows, Pimephales promelas, learn to recognize Northern Pike, Esox lucius, as predators on the basis of chemical stimuli from minnows in the pike’s diet. Animal Behaviour 46:645–656. Mathis, A., and R. J. F. Smith. 1993c. Chemical alarm signals increase the survival time of Fathead Minnows (Pimephales promelas) during encounters with Northern Pike (Esox lucius). Behavioral Ecology 4:260–265. Mathis, A., and R. J. F. Smith. 1993d. Avoidance of areas marked with a chemical alarm substance by Fathead Minnows (Pimephales promelas) in a natural habitat. Canadian Journal of Zoology 70:1473–1476. Mathis, A., and R. J. F. Smith. 1993e. Chemical labelling of Northern Pike, Esox lucius, by the alarm pheromone of Fathead Minnows, Pimephales promelas. Journal of Chemical Ecology 19:1967–1979. Mattei, X. 1991. Spermatozoon ultrastructure and its systematic implications in fishes. Canadian Journal of Zoology 69:3038–3055. Matthews, W. J. 1977. Influence of physico-chemical factors on habitat selection by Red Shiners, Notropis lutrensis (Pisces: Cyprinidae). Unpubl. Ph.D. diss., University of Oklahoma, Norman. Matthews, W. J. 1982. Small fish community structure in Ozark streams: structured assembly patterns or random abundance of species? American Midland Naturalist 107:42–54. Matthews, W. J. 1985. Critical current speeds and microhabitats of the benthic fishes Percina roanoka and Etheostoma flabellare. Environmental Biology of Fishes 2:303–308. Matthews, W. J. 1986a. Fish faunal breaks and stream order in the eastern and central United States. Environmental Biology of Fishes 17:81–92. Matthews, W. J. 1986b. Diel differences in gill net and seine catches of fish in winter in a cove of Lake Texoma, OklahomaTexas. Texas Journal of Science 38:153–158. Matthews, W. J. 1986c. Fish faunal structure in an Ozark stream: stability, persistence and a catastrophic flood. Copeia 1986: 388–397.

572 LITERATURE CITED

Matthews, W. J. 1986d. Geographic variation in thermal tolerance of a widespread minnow Notropis lutrensis of the North American mid-west. Journal of Fish Biology 28:407–417. Matthews, W. J. 1987. Geographic variation in Cyprinella lutrensis (Pisces: Cyprinidae) in the United States, with notes on Cyprinella lepida. Copeia 1987:616–637. Matthews, W. J. 1998. Patterns in Freshwater Fish Ecology. Chapman and Hall, New York. Matthews, W. J., J. R. Bek, and E. Surat. 1982a. Comparative ecology of the darters, Etheostoma podostemone, E. flabellare, and Percina roanoka in the Upper Roanoke River drainage Virginia. Copeia 1982:805–814. Matthews, W. J., R. C. Cashner, and F. P. Gelwick. 1988. Stability and persistence of fish faunas and assemblages in three Midwestern streams. Copeia 1988:945–955. Matthews, W. J., K. B. Gido, and E. Marsh-Matthews. 2001. Density-dependent overwinter survival and growth of Red Shiners from a southwestern river. Transactions of the American Fisheries Society 130:478–488. Matthews, W. J., B. C. Harvey, and M. E. Power. 1994. Spatial and temporal patterns in the fish assemblages of individual pools in a midwestern stream (USA). Environmental Biology of Fishes 39:381–397. Matthews, W. J., and D. C. Heins. 1987. Community and Evolutionary Ecology of North American Stream Fishes. University of Oklahoma Press, Norman. Matthews, W. J., and L. G. Hill. 1977. Tolerance of the Red Shiner, Notropis lutrensis (Cyprinidae) to environmental parameters. The Southwestern Naturalist 22:89–98. Matthews, W. J., and L. G. Hill. 1979. Influence of physicochemical factors on habitat selection by Red Shiners, Notropis lutrensis (Pisces: Cyprinidae). Copeia 1979:70–81. Matthews, W. J., and L. G. Hill. 1980. Habitat partitioning in the fish community of a southwestern river. Southwestern Naturalist 25:51–66. Matthews, W. J., R. E. Jenkins, and J. T. Styron, Jr. 1982b. Systematics of two forms of Blacknose Dace, Rhinichthys atratulus (Pisces: Cyprinidae) in a zone of syntopy, with a review of the species group. Copeia 1982:902–920. Matthews, W. J., and J. D. Maness. 1979. Critical thermal maxima, oxygen tolerances and success of cyprinid fishes in a southwestern river. American Midland Naturalist 102:374–377. Matthews, W. J., and E. Marsh-Matthews. 2006a. Temporal changes in replicated experimental stream fish assemblages: predictable or not? Freshwater Biology 51:1605–1622. Matthews, W. J., and E. Marsh-Matthews. 2006b. Persistence of fish species associations in pools of a small stream of the southern Great Plains. Copeia 2006:696–710. Matthews, W. J., and E. Marsh-Matthews. 2007. Extirpation of Red Shiner in direct tributaries of Lake Texoma (OklahomaTexas): a cautionary case history from a fragmented riverreservoir system. Transactions of the American Fisheries Society 136:1041–1062. Matthews, W. J., M. E. Power, and A. J. Stewart. 1986. Depth distribution of Campostoma grazing scars in an Ozark stream. Environmental Biology of Fishes 17:291–297. Matthews, W. J., and H. W. Robison. 1998. Influence of drainage connectivity, drainage area, and regional species richness on fishes of the Interior Highlands in Arkansas. American Midland Naturalist 139:1–19.

Matthews, W. J., W. Shelton, and E. Marsh-Matthews. 2012. First year growth of Longnose Gar (Lepisosteus osseus) from zygote to recruited juvenile. The Southwestern Naturalist 57:335–337. Matthews, W. J., A. J. Stewart, and M. E. Power. 1987. Grazing fishes as components of North American stream ecosystems: effects of Campostoma anomalum, p. 128–135. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Matthews, W. J., and J. T. Styron, Jr. 1981. Tolerance of headwater vs. mainstream fishes for abrupt physicochemical changes. American Midland Naturalist 105:149–158. Matuszek, J. E., J. Goodier, and D. L. Wales. 1990. The occurrence of Cyprinidae and other small fish species in relation to pH in Ontario lakes. Transactions of the American Fisheries Society 119:850–861. Maurakis, E. G. 1998. Breeding behaviors in Nocomis platyrhynchus and Nocomis raneyi (Actinopterygii: Cyprinidae). Virginia Journal of Science 49:227–236. Maurakis, E. G., and W. S. Woolcott. 1993. Spawning behaviors in Luxilus albeolus and Luxilus cerasinus (Cyprinidae). Virginia Journal of Science 44:275–278. Maurakis, E. G., W. S. Woolcott, and J. T. Magee. 1990. Pebblenests of four Semotilus species. Southeastern Fishes Council Proceedings 22:7–13. Maurakis, E. G., W. S. Woolcott, and W. R. McGuire. 1995. Nocturnal reproductive behavior in Semotilus atromaculatus (Pisces, Cyprinidae). Proceedings of the Southeastern Fishes Council 31:1–3. Maurakis, E. G., W. S. Woolcott, and M. H. Sabaj. 1991a. Reproductive-behavioral phylogenetics of Nocomis speciesgroups. American Midland Naturalist 126:103–110. Maurakis, E. G., W. S. Woolcott, and M. H. Sabaj. 1991b. Reproductive behavior of Exoglossum species. Bulletin of the Alabama Museum of Natural History 10:11–16. May, E. B., and A. A. Echelle. 1968. Young-of-year Alligator Gar in Lake Texoma, Oklahoma. Copeia 1968:629–630. Mayden, R. L. 1985a. Biogeography of Ouachita Highland fishes. Southwestern Naturalist 30:195–211. Mayden, R. L. 1985b. Nuptial structures in the subgenus Catonotus, genus Etheostoma (Percidae). Copeia 1985:580–583. Mayden, R. L. 1987a. Historical ecology and North American highland fishes: a research program in community ecology, p. 210–222. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Mayden, R. L. 1987b. Pleistocene glaciation and historical biogeography of North American central-highland fishes, p. 141–151. In Quaternary Environments of Kansas. W. C. Johnson (ed.). Kansas Geological Survey, Guidebook Series 5. Mayden, R. L. 1988. Vicariance biogeography, parsimony, and evolution in North American freshwater fishes. Systematic Zoology 37:329–355. Mayden, R. L. 1989. Phylogenetic studies of North American minnows, with emphasis on the genus Cyprinella (Teleostei: Cypriniformes). Miscellaneous Publications of the Museum of Natural History, University of Kansas:1–189. Mayden, R. L. 2002. Phylogenetic relationships of the enigmatic Ornate Shiner, Cyprinella ornata, a species endemic to Mexico

LITERATURE CITED

(Teleostei: Cyprindae). Reviews in Fish Biology and Fisheries 12:339–347. Mayden, R. L., and B. R. Kuhajda. 1996. Systematics, taxonomy, and conservation status of the endangered Alabama Sturgeon, Scaphirhynchus suttkusi Williams and Clemmer (Actinopterygii, Acipenseridae). Copeia 1996:241–273. Mayden, R. L., and B. R. Kuhajda. 1997a. Threatened fishes of the world: Scaphirhynchus suttkusi Williams and Clemmer, 1991 (Acipenseridae). Environmental Biology of Fishes 48:418– 419. Mayden, R. L., and B. R. Kuhajda. 1997b. Threatened fishes of the world: Scaphirhynchus albus (Forbes & Richardson, 1905) (Acipenseridae). Environmental Biology of Fishes 48:420–421. Mayden, R. L., R. H. Matson, and D. M. Hillis. 1992. Speciation in the North American genus Dionda (Teleostei: Cypriniformes), p. 710–746. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Mayden, R. L., W. J. Rainboth, and D. G. Buth. 1991. Phylogenetic systematics of the cyprinid genera Mylopharodon and Ptychocheilus: comparative morphometry. Copeia 1991:819–834. Mayden, R. L., and A. M. Simons. 2002. Crevice spawning behavior in Dionda dichroma, with comments on the evolution of spawning modes in North American shiners (Teleostei: Cyprinidae). Reviews in Fish Biology and Fisheries 12:327–337. Mayden, R. L., A. M. Simons, R. M. Wood, P. M. Harris, and B. R. Kuhajda. 2006. Molecular systematics and classification of North American notropin shiners and minnows (Cypriniformes: Cyprinidae) p. 72–101. In Studies of North American Desert Fishes in Honor of E. P. (Phil) Pister, Conservationist. M. De Lourdes Lozano-Vilano and A. J. Contreras-Balderas (eds.). Dirección de Puclicaciones. Universidad Autónoma de Nuevo Leon, Monterrey, México. Mayden, R. L., W.-J. Chen, H. L. Bart, M. H. Doosey, A. M. Simons, K. L. Tang, R. M. Wood, M. K. Agnew, L. Yang, M. V. Hirt, M. D. Clements, K. Saitoh, T. Sado, M. Miya, and M. Nishida. 2009. Reconstructing the phylogenetic relationships of the earth’s most diverse clade of freshwater fishes—order Cypriniformes (Actinopterygii: Ostariophysi): a case study using multiple nuclear loci and the mitochondrial genome. Molecular Phylogenetics and Evolution 51:500–514. Mayden, R. L., K. L. Tang, R. M. Wood, W.-J. Chen, M. K. Agnew, K. W. Conway, L. Yang, A. M. Simons, H. L. Bart, P. M. Harris, J. Li, X. Wang, K. Saitoh, S. He, H. Liu, Y. Chen, M. Nishida, and M. Miya. 2008. Inferring the tree of life of the order Cypriniformes, the earth’s most diverse clade of freshwater fishes: implications of varied taxon and character sampling. Journal of Systematics and Evolution 46:424–438. Mayer, T. D., and R. D. Congdon. 2008. Evaluating climate variability and pumping effects in statistical analyses. Ground Water 46:212–227. Mayhew, D. A. 1983. A new hybrid cross, Notropis atherinoides × Notropis volucellus (Pisces: Cyprinidae), from the lower Monongahela River, Western Pennsylvania. Copeia 1983:1077–1082. Mayhew, R. L. 1924. The skull of Lepidosteus platostomus. Journal of Morphology 38:315–346. Maynard Smith, J. 1982. Evolution and the Theory of Games. Cambridge University Press, Cambridge. Mayr, E. 1953. Principles of Systematic Zoology. McGraw-Hill, New York.

573

McAda, C. W., and R. S. Wydoski. 1983. Maturity and fecundity of the Bluehead Sucker, Catostomus discobolus (Catostomidae), in the upper Colorado River Basin. Southwestern Naturalist 28: 120–123. McAlister, W. H. 1958. The correlation of coloration with social rank in Gambusia hurtadoi. Ecology 39:477–482. McAllister, D. E., S. P. Platania, F. W. Schueler, M. E. Baldwin, and D. S. Lee. 1986. Ichthyofaunal patterns on a geographic grid, p. 17–51. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. McAuliffe, J. R., and D. H. Bennett. 1981. Observations on the spawning habits of the yellowfin Shiner, Notropis lutipinnis. The journal of the Elisha Mitchell Scientific Society 97: 200–203. McCabe, D. J., M. A. Beekey, A. Mazloff, and J. E. Marsden. 2006. Negative effects of Zebra Mussels on foraging and habitat use by Lake Sturgeon (Acipenser fulvescens). Aquatic Conservation: Marine and Freshwater Ecosystems 16:493–500. McCart, P., and N. Aspinwall. 1970. Spawning habits of the Largescale Sucker, Catostomus macrocheilus, at Slave Lake, British Columbia. Journal of the Fisheries Research Board of Canada 27:1154–1158. McCarthy, M. S., and W. L. Minckley. 1987. Age estimation for Razorback Sucker (Pisces: Catostomidae) from Lake Mohave, Arizona and Nevada. Journal of the Arizona-Nevada Academy of Sciences 21:87–91. McClanahan, L. L., C. R. Feldmeth, J. Jones, and D. L. Soltz. 1986. Energetics, salinity and temperature tolerance in the Mohave Tui Chub, Gila bicolor mohavensis. Copeia 1986:45–52. McCleave, J. D. 1980. Swimming performance of European Eel [Anguilla anguilla (L.)] elvers. Journal of Fish Biology 16:445–452. McCleave, J. D. 2003. Spawning areas of Atlantic Eels, p. 141–155. In Eel Biology. K. Aida, K. Tsukamoto, and K. Yamauchi (eds.). Springer, Tokyo. McCleave, J. D., and R. C. Kleckner. 1985. Oceanic migrations of Atlantic Eels (Anguilla spp.): adults and their offspring. Contributions in Marine Science 27:316–337. McCleave, J. D., R. C. Kleckner, and M. Castonguay. 1987. Reproductive sympatry of American and European Eel and implications for migration and taxonomy. American Fisheries Society Symposium 1:268–297. McComb D. M., and S. M. Kajiura. 2008. Visual fields of four batoid fishes: a comparative study. The Journal of Experimental Biology 211:482–490. McComish, T. S. 1964. Food habits of Bigmouth and Smallmouth Buff alo in Lewis and Clark Lake and the Missouri River. Unpubl. Master’s thesis, South Dakota State College, Brookings. McComish, T. S. 1967. Food habits of Bigmouth and Smallmouth Buffalo in Lewis and Clark Lake and the Missouri River. Transactions of the American Fisheries Society 96:70–73. McCord, J. W., M. R. Collins, W. C. Post, and T. I. J. Smith. 2007. Attempts to develop an index of abundance for age-1 Atlantic Sturgeon in South Carolina, USA, p. 397–403. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland.

574

LITERATURE CITED

McCormack, B. 1967. Aerial respiration in the Florida Spotted Gar. Quarterly Journal of the Florida Academy of Science 30:68–72. McCormick, C. A. 1981. Central projections of the lateral line and eighth nerves in the Bowfin, Amia calva. Journal of Comparative Neurology 197:1–15. McCormick, C. A. 1982. The organization of the octavolateralis area in actinopterygian fishes: a new interpretation. Journal of Morphology 171:159–181. McCune, A. R., K. S. Thomson, and P. E. Olsen. 1984. Semionotid fishes from the Mesozoic great lakes of North America, p. 27–44. In Evolution of Species Flocks. A. A. Echelle and I. Kornfield (eds.). University of Maine at Orono Press. McCusker, M. R., E. Parkinson, and E. B. Taylor. 2000. Mitochondrial DNA variation in Rainbow Trout (Oncorhynchus mykiss) across its native range: testing biogeographical hypotheses and their relevance to conservation. Molecular Ecology 9:2089–2108. McDonald, C. G., and C. W. Hawryshyn. 1995. Intraspecific variation of spectral sensitivity in Threespine Stickleback (Gasterosteus aculeatus) from different photic regimes. Journal of Comparative Physiology A 176:255–260. McDonald, C. G., T. E. Reimchen, and C. W. Hawryshyn. 1995. Nuptial colour loss and signal masking in Gasterosteus: an analysis using video imaging. Behaviour 132:963–977. McEachran, J. D., and N. Aschliman. 2004. Phylogeny of the Batoidea, p. 79–113. In Biology of Sharks and Their Relatives. J. C. Carrier, J. A. Musick, and M. R. Heithaus (eds.). CRC Press, Boca Raton, Florida. McEachran, J. D., and M. R. de Carvalho. 2003. Dasyatidae, p. 562–571. In The Living Marine Resources of the Western Central Atlantic. K. E. Carpenter (ed.). FAO Species Identification Guide for Fishery Purposes. Vol. 1. Food and Agriculture Organization, Rome. McEachran, J., K. Dunn, and T. Miyake. 1996. Interrelationships of the batoid fishes (Chondrichthyes: Batoidea), p 63–84. In Interrelationships of Fishes. M. Stiassny, L. Parenti, and G. D. Johnson (eds.). Academic Press, San Diego, California. McElroy, D. M., and M. E. Douglas. 1995. Patterns of morphological variation among endangered populations of Gila robusta and Gila cypha (Teleostei: Cyprinidae) in the upper Colorado River basin. Copeia 1995:636–649. McEnroe, M., and J. J. Cech, Jr. 1985. Osmoregulation in juvenile and adult White Sturgeon, Acipenser transmontanus, p. 23–30. In North American Sturgeons: Biology and Aquaculture Potential. F. P. Binkowski and S. I. Doroshov (eds.). Dr. W. Junk Publishers, Dordrecht, The Netherlands. McEwan, L. C., and D. H. Hirth. 1980. Food habits of the Bald Eagle in north-central Florida. The Condor 82:229–231. McGarvey, D. J., and R. M. Hughes. 2008. Longitudinal zonation of Pacific Northwest (U.S.A.) fish assemblages and the speciesdischarge relationship. Copeia 2008:311–321. McGeachin, R. B. 1993. Carp and buffalo, p. 117–143. In Culture of Nonsalmonid Freshwater Fishes. R. R. Stickney (ed.). CRC Press, Boca Raton, Florida. McGhee, K. E., R. C. Fuller, and J. Travis. 2007. Male competition and female choice interact to determine mating success in the Bluefin Killifish. Behavioral Ecology 18:822–830. McGowan D. W., and S. M. Kajiura. 2009. Electroreception in the euryhaline Stingray, Dasyatis sabina. The Journal of Experimental Biology 212:1544–1552.

McGrath, K., J. Bernier, S. Ault, J. -D. Dutil, and K. Reid. 2003. Differentiating downstream migrating Eels Anguilla rostrata from resident Eels in the St. Lawrence River. American Fisheries Society Symposium 33:315–327. McGrath, P. E. 2010. The life history of Longnose Gar, Lepisosteus osseus, an apex predator in the tidal waters of Virginia. Unpubl. Ph.D. diss., The College of William and Mary, Williamsburg, Virginia. McGree, M., T. A. Whitesel, and J. Stone. 2008. Larval metamorphosis of individual Pacific Lampreys reared in captivity. Transactions of the American Fisheries Society 137:1866–1878. McInerney, J. E. 1969. Reproductive behaviour of the Blackspotted Stickleback, Gasterosteus wheatlandi. Journal of the Fisheries Research Board of Canada 26:2061–2075. McIntyre, P. B., and A. S. Flecker. 2010. Ecological stoichiometry as an integrative framework in stream fish ecology, p. 539–558. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. McKaye, F. E. 1971. Behavioral aspects of population dynamics in unisexual-bisexual Poeciliopsis (Pisces: Poeciliidae). Ecology 52:778–790. McKaye, K. R. 1981. Natural Selection and the evolution of interspecific brood care in fishes, p. 173–183. In Natural Selection and Social Behavior. R. D. Alexander and D. W. Tinkle (eds.). Chiron Press, New York. McKaye, K. R., and N. M. McKaye. 1977. Communal care and the kidnapping of young by parental cichlids. Evolution 31:674–681. McKenzie, D. J., S. Aota, and D. J. Randall. 1991a. Ventilatory and cardiovascular responses to blood pH, plasma PCO2, blood O2 content, and catecholamines in an air-breathing fish, the Bowfin (Amia calva). Physiological Zoology 64:432– 450. McKenzie, D. J., M. L. Burleson, and D. J. Randall. 1991b. The effects of branchial denervation and pseudobranch ablation on cardioventilatory control in an air-breathing fish. Journal of Experimental Biology 161:347–365. McKenzie D. J., M. Piccolella, A. Z. Dalla Valle, E. W. Taylor, C. L. Bolis, and J. F. Steffensen. 2003. Tolerance of chronic hypercapnia by the European Eel Anguilla anguilla. Journal of Experimental Biology 206:1717–1726. McKenzie, D. J., and D. J. Randall. 1990. Does Amia calva aestivate? Fish Physiology and Biochemistry 8:147–158. McKenzie, J. A., and M. H. A. Keenleyside. 1970. Reproductive behavior of Ninespine Sticklebacks (Pungitius pungitius (L.)) in South Bay, Manitoulin Island, Ontario. Canadian Journal of Zoology 48:55— 61. McKinley, D. 1984. History of a neglected account of the Paddlefish, Polyodon spathula. Copeia 1984:201–204. McKinley, R. S., T. D. Singer, J. S. Ballantyne, and G. Power. 1993. Seasonal variation in plasma nonesterified fatty acids of Lake Sturgeon (Acipenser fulvescens) in the vicinity of hydroelectric facilities. Canadian Journal of Fisheries and Aquatic Sciences 50:2440–2447. McKinley, S., G. Van Der Kraak, and G. Power. 1998. Seasonal migrations and reproductive patterns in the Lake Sturgeon, Acipenser fulvescens, in the vicinity of hydroelectric stations in northern Ontario. Environmental Biology of Fishes 51:245–256.

LITERATURE CITED

McKinney, T., W. R. Persons, and R. S. Rogers. 1999. Ecology of Flannelmouth Sucker in the Lee’s Ferry tailwater, Colorado River, Arizona. Great Basin Naturalist 59:259–265. McLane, W. M. 1955. The fishes of the St. Johns River System. Unpubl. Ph.D. diss., University of Florida, Gainesville. McLaughlin, R. L., J. E. Marsden, and D. B. Hayes. 2003. Achieving the benefits of Sea Lamprey control while minimizing effects on nontarget species: Conceptual synthesis and proposed policy. Journal of Great Lakes Research 29(Suppl. 1):755–765. McLennan, D. A. 1993. Temporal changes in the structure of the male nuptial signal in the Brook Stickleback, Culaea inconstans (Kirtland). Canadian Journal of Zoology 71:1111–1119. McLennan, D. A. 1994. Changes in female colour across the ovulatory cycle in Brook Sticklebacks, Culaea inconstans (Kirtland). Canadian Journal of Zoology 72:144–153. McLennan, D. A. 1995. Male mate choice based on female nuptial coloration in Brook Sticklebacks, Culaea inconstans. Animal Behaviour 50:213–221. McLennan, D. A. 2003. The importance of olfactory signals in the gasterosteid mating system: sticklebacks go multimodal. Biological Journal of the Linnean Society 80:555–572. McLennan, D. A. 2004. Male Brook Sticklebacks’ (Culaea inconstans) response to olfactory cues. Behaviour 141:1411–1424. McLennan, D. A. 2005. Changes in response to olfactory cues across the ovulatory cycle in Brook Sticklebacks. Animal Behaviour 69:181–188. McLennan, D. A. 2006. The umwelt of the Threespine Stickleback, p. 179–224. In The Biology of the Threespine Stickleback. S. Östlund-Nilsson, I. Mayer and F. Huntingford (eds.). CRC Press. McLennan, D. A., and McPhail, J. D. 1989. Experimental investigations of the evolutionary significance of sexually dimorphic nuptial colouration in Gasterosteus aculeatus (L.): temporal changes in the structure of the male mosaic signal. Canadian Journal of Zoology 67:1767–1777. McLennan, D. A., and McPhail, J. D. 1990. Experimental investigations of the evolutionary significance of sexually dimorphic nuptial colouration in Gasterosteus aculeatus (L.): the relationships between male colour and female behaviour. Canadian Journal of Zoology 68:482–492. McLennan, D. A., and M. J. Ryan. 1997. Responses to conspecific and heterospecific olfactory cues in the swordtail, Xiphophorus cortezi. Animal Behaviour 54:1077–1088. McLennan, D. A., and M. J. Ryan. 1999. Interspecific recognition and discrimination based upon olfactory cues in northern swordtails. Evolution 53:880–888. McLennan, D. A., and M. J. Ryan. 2008. Female swordtails, Xiphophorus continens, prefer the scent of heterospecific males. Animal Behaviour 75:1731–1737. McLeod, C., L. Hildebrand, and D. Radford. 1999. A synopsis of Lake Sturgeon management in Alberta, Canada. Journal of Applied Ichthyology 15:173–179. McMahon, T. E., and J. C. Tash. 1979. The use of chemosenses by Threadfin Shad, Dorosoma petenense to detect conspecifics, predators and food. Journal of Fish Biology 14:289–296. McMillan, M., and D. Wilcove. 1994. Gone but not forgotten: why have species protected by the Endangered Species Act become extinct? Endangered Species Update 11:5–6. McNaughton, S. J. 1984. Grazing lawns: animals in herds, plant form, and coevolution. The American Naturalist 124:863–886.

575

McNeely, D. I., and C. E. Wade. 2003. Relative abundance of the gynogen Poecilia formosa and its sexual host Poecilia latipinna (Teleostei: Poeciliidae) in some southern Texas habitats. Southwestern Naturalist 48:451–453. McNeely, D. L. 1987. Niche relations within an Ozark stream cyprinid assemblage. Environmental Biology of Fishes 18:195–208. McPeek, M. A. 1992. Mechanisms of sexual selection operating on body size in the mosquitofish (Gambusia holbrooki). Behavioral Ecology 3:1–12. McPhail, J. D., and C. C. Lindsey. 1970. Freshwater fishes of northwestern Canada and Alaska. Bulletin 173. Fisheries Research Board of Canada, Ottawa. McPhail, J. D., and C. C. Lindsey. 1986. Zoogeography of the freshwater fishes of Cascadia (the Columbia System and Rivers North to the Stikine), p. 615–637. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. McPhail, J. D. 2007. The Freshwater fishes of British Columbia. The University of Alberta Press, Edmonton. McPhee, M. V., M. J. Osborne, and T. F. Turner. 2008. Genetic diversity, population structure, and demographic history of the Rio Grande Sucker, Catostomus (Pantosteus) plebeius, in New Mexico. Copeia 2008:191–199. McPherson, T. D., R. S. Mirza, and G. G. Pyle. 2004. Responses of wild fishes to alarm chemicals in pristine and metalcontaminated lakes. Canadian Journal of Zoology 82:694–700. McQuown, E., G. A. E. Gall, and B. May. 2002. Characterization and inheritance of six microsatellite loci in Lake Sturgeon. Transactions of the American Fisheries Society 131:299–307. McQuown, E., C. C. Krueger, H. L. Kincaid, G. A. E. Gall, and B. May. 2003. Genetic comparison of Lake Sturgeon populations: differentiation based on allelic frequencies at seven microsatellite loci. Journal of Great Lakes Research 29:3–13. McSwain, L. E., and R. M. Gennings. 1972. Spawning behavior of the Spotted Sucker Minytrema melanops (Rafinesque). Transactions of the American Fisheries Society 101:738–740. Meador, M. R., and D. M. Carlisle. 2007. Quantifying tolerance indicator values for common stream fish species of the United States. Ecological Indicators 7:329–338. Mecklenburg, C. W., T. A. Mecklenburg, and L. K. Thorsteinson. 2002. Fishes of Alaska. American Fisheries Society, Bethesda, Maryland. Medland, T. E., and F. W. H. Beamish. 1987. Age validation for the Mountain Brook Lamprey, Ichthyomyzon greeleyi. Canadian Journal of Fisheries and Aquatic Sciences 44:901–904. Meesters, A. 1940. Über die Organisation des Gesichtsfeldes der Fische. Zeitschrift für Tierpsychologie 4:84–149. Meek, S. E., and S. F. Hildebrand. 1923. The marine fishes of Panama. Publication Field Museum, Natural History, Zoology (Series 215) 15:1–330. Meffe, G. K. 1985. Predation and species replacement in American Southwestern fishes: a case study. Southwestern Naturalist 30:173–187. Meffe, G. K. 1986. Conservation genetics and the management of endangered fishes. Fisheries 11:14–23. Meffe, G. K. 1990. Offspring size variation in Eastern Mosquitofish (Gambusia holbrooki: Poeciliidae) from contrasting thermal environments. Copeia 1990:10–18. Meffe, G. K., and T. M. Berra. 1988. Temporal characteristics of fish assemblage structure in an Ohio stream. Copeia 1988:684–690.

576

LITERATURE CITED

Meffe, G. K., D. A. Hendrickson, W. L. Minckley, and J. N. Rinne. 1983. Factors resulting in decline of the endangered Sonoran Topminnow (Atheriniformes: Poeciliidae) in the United States. Biological Conservation 25:135–159. Meffe, G. K., and W. L. Minckley. 1987. Persistence and stability of fish and invertebrate assemblages in a repeatedly disturbed Sonoran Desert stream. American Midland Naturalist 117:177–191. Meffe, G. K., and R. C. Vrijenhoek. 1988. Conservation genetics in the management of desert fishes. Conservation Biology 2:149–155. Mehlis, M., T. C. M. Bakker, and J. G. Frommen. 2008. Smells like sib spirit: recognition in three-spined Sticklebacks (Gasterosteus aculeatus) is mediated by olfactory cues. Animal Cognition 11:643–650. Mehlis, M., T. C. M. Bakker, K. Langen, and J. G. Frommen. 2009. Cain and Abel reloaded? Kin recognition and male-male aggression in three-spined Sticklebacks, Gasterosteus aculeatus L. Journal of Fish Biology 75:2154–2162. Meinkoth, N. A. 1947. Notes on the life cycle and taxonomic position of Haplobothrium globuliforme Cooper, a tapeworm of Amia calva L. Transactions of the American Microscopical Society 66:256–261. Mendelson, J. 1975. Feeding relationships among species of Notropis (Pisces: Cyprinidae) in a Wisconsin stream. Ecological Monographs 45:199–230. Mendoza, R. C. Aguilera, L. Carreon, J. Montemayor, and M. Gonzalez. 2008a. Weaning of Alligator Gar (Atractosteus spatula) larvae to artificial diets. Aquaculture Nutrition 14:223–231. Mendoza, R., C. Aguilera, and A. M. Ferrara. 2008b. Gar biology and culture: status and prospects. Aquaculture Research 39:748–763. Mendoza, R., C. Aguilera, and J. Montemayor. 2010a. Ecología, de los lepisostéidos, p. 21–41. In Biología, Ecología y Avances en el Cultivo de Catán Atractosteus spatula. R. C. Mendoza, C. Aguilera, and J. Montemayor (eds.). Universidad Autónoma de Nuevo León, Monterrey, Mexico. Mendoza, R., C. Aguilera, and J. Montemayor (eds.). 2010b. Biología, ecología y avances en el cultivo de catán Atractosteus spatula. Universidad Autónoma de Nuevo León, Monterrey, Mexico. Mendoza, R., C. Aguilera, G. Rodriquez, M. Gonzalez, and R. Castro. 2002. Morphophysiological studies on Alligator Gar (Atractosteus spatula) larval development as a basis for their culture and repopulation of their natural habitats. Reviews in Fish Biology and Fisheries 12:133–142. Meng, L., and P. B. Moyle. 1995. Status of Splittail in the Sacramento-San Joaquin estuary. Transactions of the American Fisheries Society 124:538–549. Meredith, T. L., and S. M. Kajiura. 2010. Olfactory morphology and physiology of elasmobranchs. The Journal of Experimental Biology 213:3449–3456. Mero, S. W., D. W. Willis, and G. J. Power. 1994. Walleye and Sauger predation on Paddlefish in Lake Sakakawea, North Dakota. North American Journal of Fisheries Management 14:226–227. Meronek, T. G., F. A. Copes, and D. W. Coble. 1997a. The bait industry in Illinois, Michigan, Minnesota, Ohio, South Dakota, and Wisconsin. Technical Bulletin Series, North Central Regional Aquaculture Center, Iowa State University 105:1–8.

Meronek, T. G., F. A. Copes, and D. W. Coble. 1997b. A survey of the bait industry in the North-Central region of the United States. North American Journal of Fisheries Management 17:703–711. Merritt, R. B., J. F. Rogers, and B. J. Kurz. 1978. Genic variability in the Longnose Dace, Rhinichthys cataractae. Evolution 32:116–124. Merz, J. E., and P. B. Moyle. 2006. Salmon, wildlife, and wine: marine-derived nutrients in human-dominated ecosystems of central California. Ecological Applications 16:999–1009. Mesa, M. G., J. M. Bayer, and J. G. Seelye. 2003. Swimming performance and physiological responses to exhaustive exercise in radio-tagged and untagged Pacific Lampreys. Transactions of the American Fisheries Society 132:483–492. Mestl, G., and J. Sorensen. 2009. Joint management of an interjurisdictional Paddlefish snag fishery in the Missouri River below Gavins Point Dam, South Dakota and Nebraska, p. 235–259. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Mettee, M. F. 2000. Blue Sucker research. River Crossings 9:9. Mettee, M. F., P. E. O’Neil, and J. M. Pierson. 1996. Fishes of Alabama and the Mobile Basin. Oxmoor House, Inc., Birmingham, Alabama. Mettee, M. F., P. E. O’Neil, and S. J. Rider. 2009. Paddlefish movement in the lower Mobile River basin, Alabama, p. 63–81. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Mettee, M. F., T. E. Shepard, P. E. O’Neil, S. W. McGregor, and W. P. Henderson 2003. Biology of a spawning population of Cycleptus meridionalis in the Alabama River. Proceedings of the Annual Conference of the Southeastern Association of Fish and Wildlife Agencies 2003:59–67. Meyer, A., and C. Lydeard. 1993. The evolution of copulatory organs, internal fertilization, placentae and viviparity in killifishes (Cyprinodontiformes) inferred from a DNA phylogeny of the tyrosine kinase gene X-src. Proceedings of the Royal Society of London, B 254:153–162. Meyer, A., J. Morrissey, and M. Schartl. 1994. Recurrent origin of a sexually selected trait in Xiphophorus fishes inferred from a molecular phylogeny. Nature 368:539–542. Meyer, F. P., and J. H. Stevenson. 1962. Studies on the artificial propagation of the Paddlefish. The Progressive Fish-Culturist 24:65–67. Meyer, W. H. 1962. Life history of three species of Redhorse (Moxostoma) in the Des Moines River, Iowa. Transactions of the American Fisheries Society 91:412–419. Michaletz, P. H., C. F. Rabeni, W. W. Taylor, and T. R. Russell. 1982. Feeding ecology and growth of young-of-the-year Paddlefish in hatchery ponds. Transactions of the American Fisheries Society 111:700–709. Milinski, M. 2003. The function of mate choice in Sticklebacks: optimizing Mhc genes. Journal of Fish Biology 63 (Supplement A):1–16. Milinsky, M., S. Griffiths, K. M. Wegner, T. B. H. Reusch, A. HaasAssenbaum, and T. Boehm. 2005. Mate choice decisions of Stickleback females predictably modified by MHC peptide

LITERATURE CITED

ligands. Proceedings of the National Academy of Sciences of the United States of America 102:4414–4418. Miller, A. I., and L. G. Beckman. 1996. First record of predation on White Sturgeon eggs by sympatric fishes. Transactions of the American Fisheries Society 125:338–340. Miller, D. L., and R. J. Behnke. 1985. Two new intergeneric cyprinid hybrids from the Bonneville Basin, Utah. Copeia 1985:509–515. Miller, G. L., and W. R. Nelson. 1974. Goldeye, Hiodon alosoides, in Lake Oahe: abundance, age, growth, maturity, food, and the fishery, 1963–69. United States Fish and Wildlife Ser vice, Technical Paper 79, Washington, D. C. Miller, H. C. 1963. The behavior of the Pumpkinseed Sunfish, Lepomis gibbosus (Linnaeus), with notes on the behavior of other species of Lepomis and the pygmy Sunfish, Elassoma evergladei. Behaviour 22:88–151. Miller, M. J. 2004. The ecology and functional morphology of feeding of North American Sturgeon and Paddlefish, p. 87–102. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. H. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. Miller, M. J. 2009. Ecology of anguilliform leptocephali: remarkable transparent fish larvae of the ocean surface layer. AquaBioScience Monographs 2:1–94. Miller, R. J. 1962. Reproductive behavior of the stoneroller minnow, Campostoma anomalum pullum. Copeia 1962:407–417. Miller, R. J. 1964. Behavior and ecology of some North American cyprinid fishes. American Midland Naturalist 72:313–357. Miller, R. J., and H. E. Evans. 1965. External morphology of the brain and lips in catostomid fishes. Copeia 1965:467–487. Miller, R. R. 1945. Snyderichthys, a new generic name for the Leatherside Chub of the Bonneville and Upper Snake drainages in western United States. Journal of the Washington Academy of Sciences 35:28. Miller, R. R. 1948. The cyprinodont fishes of the Death Valley system of eastern California and southwestern Nevada. Miscellaneous Publications, Museum of Zoology, University of Michigan No. 68. Miller, R. R. 1959. Origin and affinities of the freshwater fish fauna of western North America, p. 187–222. In Zoogeography. C. L. Hubbs (ed.). American Association for the Advancement of Science, Publication 51. Miller, R. R. 1960. Four new species of viviparous fishes, Poeciliopsis, from northwestern Mexico. Occasional Papers. Museum of Zoology. University of Michigan 433:1–9. Miller, R. R. 1961. Man and the changing fish fauna of the American southwest. Papers of the Michigan Academy of Sciences, Arts, and Letters 46:365–404. Miller, R. R. 1981. Coevolution of deserts and Pupfishes (genus Cyprinodon) in the American Southwest, p. 39–94. In Fishes in North American Deserts. R. J. Naiman and D. L. Soltz (eds.). John Wiley and Sons, New York. Miller, R. R. 1986. Composition and derivation of the freshwater fish fauna of Mexico. Anales de la Escuela Nacional de Ciencias Biologicas 30:121–153. Miller, R. R., and C. L. Hubbs. 1960. The spiny-rayed cyprinid fishes (Plagopterini) of the Colorado River systems. Miscellaneous Publications of the Museum of Zoology, University of Michigan 115:1–39.

577

Miller, R. R., W. L. Minckley, and S. M. Norris. 2005. Freshwater Fishes of Mexico. University of Chicago Press, Chicago, Illinois. Miller, R. R., and G. R. Smith. 1967. New fossil fishes from PlioPleistocene Lake Idaho. Occasional Papers of the Museum of Zoology University of Michigan 654:1–24. Miller, R. R., and G. R. Smith. 1981. Distribution and evolution of Chasmistes (Pisces: Catostomidae) in western North America. Occasional Papers, Museum of Zoology, University of Michigan 696:1–46. Miller, R. R., and M. L. Smith. 1986. Origin and geography of the fishes of central Mexico, p. 487–517. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. Miller, R. R., J. D. Williams, and J. E. Williams. 1989. Extinctions of North American fishes during the past century. Fisheries 14:22–38. Miller, S. A., and T. A. Crowl. 2006. Effects of Common Carp (Cyprinus carpio) on macrophytes and invertebrate communities in a shallow lake. Freshwater Biology 51:85–94. Miller, S. E., and D. L. Scarnecchia. 2008. Adult Paddlefish migrations in relation to spring river conditions of the Yellowstone and Missouri rivers, Montana and South Dakota. Journal of Applied Ichthyology 24:221–228. Mills, E. L., D. L. Strayer, M. D. Scheuerell, and J. T. Carlton. 1996. Exotic species in the Hudson River basin: a history of invasions and introductions. Estuaries 19:814–823. Mills, K. H., and D. W. Schindler. 1986. Biological indicators of lake acidification. Water, Air, & Soil Pollution 30:779–789. Milsom, W. K., and D. R. Jones. 1985. Characteristics of mechanoreceptros in the air-breathing organ of the holostean fish, Amia calva. Journal of Experimental Biology 117:389–399. Mims, S. D. 2001. Aquaculture of Paddlefish in the United States. Aquatic Living Resources 14:391–398. Mims, S. D., R. J. Onders, and W. L. Shelton. 2009. Propagation and culture of Paddlefish, p. 357–383. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society, Symposium 66. Bethesda, Maryland. Mims, S. D., and H. R. Schmittou. 1989. Influence of Daphnia density on survival and growth of Paddlefish larvae at two temperatures. Proceedings of the Annual Conference of the Southeastern Association of Fish and Wildlife Agencies 43:112–118. Mims, S. D., and W. L. Shelton. 1998. Induced meiotic gynogenesis in Shovelnose Sturgeon. Aquaculture International 6:323–329. Mims, S. D., and W. L. Shelton. 1999. Monosex culture of Paddlefish and Shovelnose Sturgeon, p. 67–76. In Proceedings of the Symposium on the Harvest and Conservation of North American Paddlefish and Sturgeon, May 7–8, 1998, Chattanooga, Tennessee. D. F. Williamson, G. W. Benz, and C. M. Hoover (eds.). TRAFFIC North America / World Wildlife Fund, Washington, D.C. Mims, S. D., W. L. Shelton, O. Linhart, and C. Wang. 1997. Induced meiotic gynogenesis of Paddlefish, Polyodon spathula. Journal of the World Aquaculture Society 28:334–343. Minckley, W. L. 1971. Keys to native and introduced fishes of Arizona. Journal of the Arizona Academy of Science 6:183–188.

578 LITERATURE CITED

Minckley, W. L. 1973. Fishes of Arizona. Arizona Game and Fish Department, Phoenix. Minckley, W. L. 1983. Status of the razorback sucker, Xyrauchen texanus (Abbott), in the Lower Colorado River basin. The Southwestern Naturalist 28:165–187. Minckley, W. L. 1991a. Native fishes of arid lands: a dwindling resource of the desert southwest. United States Forest Ser vice General Technical Report RM 206:1–45. Minckley, W. L. 1991b. Native fishes of the Grand Canyon Region: an obituary? p. 124–177. In Committee on Glen Canyon Environmental Studies. Colorado River Ecology and Dam Management. Proceedings of a symposium, May 24–25, 1990, Santa Fe, New Mexico. National Academy Press, Washington. Minckley, W. L., and W. E. Barber. 1971. Some aspects of biology of the Longfin Dace, a cyprinid fish characteristic of streams in the Sonoran desert. The Southwestern Naturalist 15:459–464. Minckley, W. L., D. G. Buth, and R. L. Mayden. 1989. Origin of brood stock and allozyme variation in hatchery-reared Bonytail, an endangered North American cyprinid fish. Transactions of the American Fisheries Society 118:131–137. Minckley, W. L., and L. H. Carufel. 1967. The Little Colorado River spinedace, Lepidomeda vittata, in Arizona. The Southwestern Naturalist 12:291–302. Minckley, W. L., and M. E. Douglas. 1991. Discovery and extinction of western fishes: a blink of the eye in geologic time, p. 7–17. In Battle Against Extinction: Native Fish Management in the American West. W. L. Minckley and J. E. Deacon (eds.). The University of Arizona Press, Tucson. Minckley, W. L., and E. S. Gustafson. 1982. Early development of the Razorback Sucker, Xyrauchen texanus (Abbott). Great Basin Naturalist 42:533–561. Minckley, W. L., D. A. Hendrickson, and C. E. Bond. 1986. Geography of western North American freshwater fishes: description and relationships to intracontinental tectonism, p. 519–613. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. Minckley, W. L., J. E. Johnson, J. N. Rinne, and S. E. Willoughby. 1970. Foods of buffalofishes, genus Ictiobus, in central Arizona reservoirs. Transactions of the American Fisheries Society 99:333–342. Minckley, W. L., P. C. Marsh, J. E. Brooks, J. E. Johnson, and B. L. Jensen. 1991. Management toward recovery of the Razorback Sucker, p. 283–357. In Battle Against Extinction: Native Fish Management in the American West. W. L. Minckley and J. E. Deacon (eds.). The University of Arizona Press, Tucson. Minckley, W. L., P. C. Marsh, J. E. Deacon, T. E. Dowling, P. W. Hedrick, W. J. Matthews, and G. Mueller. 2003. A conservation plan for native fishes of the lower Colorado River. BioScience 53:219–234. Minckley, W. L., and G. K. Meffee. 1987. Differential selection by flooding in stream-fish communities of the arid American southwest, p. 93–110. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Minckley, W. L., and S. P. Vives. 1990. Cavity nesting and male nest defense by Ornate Minnow, Codoma ornata (Pisces: Cyprinidae). Copeia 1990:219–221. Minegishi, Y., J. Aoyama, J. G. Inoue, M. Miya, M. Nishida, and K. Tsukamoto. 2005. Molecular phylogeny and evolution of the Freshwater Eels genus Anguilla based on the whole mitochon-

drial genome sequences. Molecular Phylogenetics and Evolution 34:134–146. Minkkinen, S. P., J. L. Devers, W. A. Lellis, and H. S. Galbraith. 2010. Experimental stocking of American Eels in the Susquehanna River watershed. Report to the City of Sunbury. U.S. Fish and Wildlife Ser vice, Maryland Fishery Resources Office (MFRO). Annapolis, Maryland. Mirza, R. S., and D. P. Chivers. 2000. Predator-recognition training enhances survival of Brook Trout: evidence from laboratory and field enclosure studies. Canadian Journal of Zoology 78:2198–2208. Mirza, R. S., and D. P. Chivers. 2001a. Learned recognition of heterospecific alarm signals: the importance of a mixed predator diet. Ethology 107:1007–1018. Mirza, R. S., and D. P. Chivers. 2001b. Are chemical alarm cues conserved within salmonid fishes? Journal of Chemical Ecology 27:1641–1655. Mirza, R. S., and D. P. Chivers. 2002. Brook char (Salvelinus fontinalis) can differentiate chemical cues produced by different age/ size classes of conspecifics. Journal of Chemical Ecology 28:555–564. Mirza, R. S., and D. P. Chivers. 2003a. Predator diet cues and the assessment of predation risk by juvenile brook charr: do diet cues enhance survival? Canadian Journal of Zoology 81:126–132. Mirza, R. S., and D. P. Chivers. 2003b. Response of juvenile Rainbow Trout to varying concentrations of chemical alarm cue: response thresholds and survival during encounters with predators. Canadian Journal of Zoology 81:88–95. Mirza, R. S., and D. P. Chivers. 2003c. Fathead Minnows learn to recognize heterospecific alarm cues they detect in the diet of a known predator. Behaviour 140:1359–1369. Mirza, R. S., and D. P. Chivers. 2003d. Influence of body size on the responses of Fathead Minnows, Pimephales promelas, to damselfly alarm cues. Ethology 109:691–699. Mirza, R. S., S. A. Fisher, and D. P. Chivers. 2003. Assessment of predation risk by juvenile Yellow Perch, Perca flavescens: responses to alarm cues from conspecific and prey guild members. Environmental Biology of Fishes 66:321–327. Mirza, R. S., W. W. Green, S. Connor, A. C. W. Weeks, C. M. Wood, and G. C. Pyle. 2009. Do you smell what I smell? Olfactory impairment in wild Yellow Perch from metal-contaminated waters. Ecotoxicology and Environmental Safety 72:677–683. Mirza, R. S., J. J. Scott, and D. P. Chivers. 2001. Differential responses of male and female Red Swordtails to chemical alarm cues. Journal of Fish Biology 59:716–728. Mittelbach, G. G. 1988. Competition among refuging Sunfishes and effects of fish density on littoral zone invertebrates. Ecology 69:614–623. Mittelbach, G. G., A. M. Turner, D. J. Hall, and J. E. Rettig. 1995. Perturbation and resilience: a long-term, whole-lake study of predator extinction and reintroduction. Ecology 76:2347–2360. Mitzner, L. 1978. Evaluation of biological control of nuisance aquatic vegetation by Grass Carp. Transactions of the American Fisheries Society 107:135–145. Mizelle, J. D., and H. D. McDougal. 1970. Studies on monogenetic trematodes. XLV. The genus Dactylogyrus in North America. Key to species, host-parasite and parasite-host lists, localities, emendations, and description of D. kritskyi sp. n. American Midland Naturalist 84:444–462.

LITERATURE CITED

Mjølnerød, I. B., I. A. Fleming, U. H. Refseth, and K. Hindar. 1998. Mate and sperm competition during multiple-male spawnings of Atlantic Salmon. Canadian Journal of Zoology 76:70–75. Mochioka, N., M. Iwamizu, and T. Kanda. 1993. Leptocephalus Eel larvae will feed in aquaria. Environmental Biology of Fishes 36:381–384. Mock, K. E., R. P. Evans, M. Crawford, B. L. Cardall, S. U. Janecke, and M. P. Miller. 2006. Rangewide molecular structuring in the Utah Sucker (Catostomus ardens). Molecular Ecology 15:2223–2238. Modde, T. 1980. Growth and residency of juvenile fishes within a surf zone habitat in the Gulf of Mexico. Gulf Research Reports 6:377–385. Modde, T., and D. B. Irving. 1998. Use of multiple spawning sites and seasonal movement by Razorback Suckers in the Middle Green River, Utah. North American Journal of Fisheries Management 18:318–326. Modde, T., and J. C. Schmulbach. 1977. Food and feeding behavior of the Shovelnose Sturgeon, Scaphirhynchus platorynchus, in the unchannelized Missouri River, South Dakota. Transactions of the American Fisheries Society 106:602–608. Moen, C. T., D. L. Scarnecchia, and J. S. Ramsey. 1992. Paddlefish movements and habitat use in Pool 13 of the upper Mississippi River during abnormally low river stages and discharges. North American Journal of Fisheries Management 12:744–751. Moerchen, R. 1973. Specific isolating mechanisms of the Orangethroat Darter, Etheostoma spectabile (Agassiz), and the Rainbow Darter, Etheostoma caeruleum Storer, in central Ohio, with considerations of their hybridizations. Unpubl. Ph.D. diss., Ohio State University, Columbus. Monaco, M. E., D. M. Nelson, T. E. Czapla, and M. E. Pattillo. 1989. Distribution and abundance of fishes and invertebrates in Texas estuaries. United States Department of Commerce NOAA Estuarine Living Marine Resources Project. Monette, S. N., and C. B. Renaud. 2005. The gular pouch in northern hemisphere parasitic Lampreys (Petromyzontidae). Canadian Journal of Zoology 83:527–535. Mongeau, J. R., L. Dumont, and L. Cloutier. 1992. La biologie du suceur cuivre (Moxostoma hubbsi) comparee a celle de quatre autres especes de Moxostoma (M. anisurum, M. carinatum, M. macrolepidotum et M. valenciennesi). Canadian Journal of Zoology 70:1354–1363. Monteleone, D. M. 1992. Seasonality and abundance of ichthyoplankton in Great South Bay, New York. Estuaries 15:230–238. Monteleone, D. M., and L. E. Duguay. 1988. Laboratory studies of predation by the ctenophore Mnemiopsis leidyi on the early stages in the life history of the Bay Anchovy, Anchoa mitchilli. Journal of Plankton Research 10:359–372. Moon, D. N., S. J. Fisher, and D. W. Willis. 1998a. Goldeye recruitment and growth in two Missouri River backwaters. Proceedings of the South Dakota Academy of Science 77:139–144. Moon, D. N., S. J. Fisher, and S. C. Krentz. 1998b. Assessment of larval fish consumption by Goldeye (Hiodon alosoides) in two Missouri River backwaters. Journal of Freshwater Ecology 13:317–321. Moore, A. 1994b. An electrophysiological study on the effects of pH on olfaction in mature male Atlantic Salmon (Salmo salar) parr. Journal of Fish Biology 45:493–502.

579

Moore, A., M. J. Ives, and L. T. Kell. 1994. The role of urine in sibling recognition in Atlantic Salmon Salmo salar (L.) parr. Proceedings of the Royal Society of London B 255:173–180. Moore, A., K. H. Olsén, N. Lower, and H. Kindahl. 2002. The role of F-series prostaglandins as reproductive priming pheromones in the Brown Trout. Journal of Fish Biology 60:613– 624. Moore, A., and A. P. Scott. 1992. 17α,20β-dihydroxy-4 pregnen-3one-20-sulphate is a potent odorant in precocious male Atlantic Salmon (Salmo salar L.) parr which have been pre-exposed to the urine of ovulated females. Proceedings of the Royal Society of London B 249:205–209. Moore, A., and C. P. Waring. 1996a. Electrophysiological and endocrinological evidence that F-series prostaglandins function as priming pheromones in mature Atlantic Salmon (Salmo salar) parr. Journal of Experimental Biology 199:2307–2316. Moore, A., and C. P. Waring. 1996b. Sublethal effects of the pesticide Diazinon on olfactory function in mature Atlantic Salmon parr. Journal of Fish Biology 48:758–775. Moore, A., and C. P. Waring. 1999. Reproductive priming in mature male Atlantic Salmon parr exposed to the sound of redd cutting. Journal of Fish Biology 55:884–887. Moore, A., and C. P. Waring. 2001. The effects of synthetic pyrethroid pesticide on some aspects of reproduction in Atlantic Salmon (Salmo salar L.). Aquatic Toxicology 52:1–12. Moore, G. A. 1944. The retinae of two North American teleosts, with special reference to their tapeta lucida. Journal of Comparative Neurology 80:369–379. Moore, G. A. 1950. The cutaneous sense organs of barbeled minnows adapted to live in the muddy waters of the Great Plains region. Transactions of the American Microscopical Society 69:69–95. Moore, G. A., and R. C. McDougal. 1949. Similarity in the retinae of Amphiodon alosoides and Hiodon tergisus. Copeia 1949:298. Moore, G. A., M. B. Trautman, and M. R. Curd. 1973. A description of postlarval Gar (Lepisosteus spatula Lacepede, Lepisosteidae), with a list of associated species from the Red River, Choctaw County, Oklahoma. The Southwestern Naturalist 18:343–344. Moore, J. W., and F. W. H. Beamish. 1973. Food of larval Sea Lamprey (Petromyzon marinus) and American Brook Lamprey (Lampetra lamottei). Journal of the Fisheries Research Board of Canada 30:7–15. Moore, J. W., and J. M. Mallatt. 1980. Feeding of larval Lamprey. Canadian Journal of Fisheries and Aquatic Sciences 37:1658–1664. Moore, P., and J. Atema. 1988. A model of a temporal filter in chemoreception to extract directional information from a turbulent odor plume. Biological Bulletin 174:355–363. Moore P. A. 1994a. A model of the role of adaptation and disadaptation in olfactory receptor neurons: implications for the coding of temporal and intensity patterns in odor signals. Chemical Senses 19:71–86. Moore, W. S., and F. E. McKaye. 1971. Coexistence in unisexualbisexual species complexes of Poeciliopsis (Pisces: Poeciliidae). Ecology 52:791–799. Mora, E. A., S. T. Lindley, D. L. Erickson, and A. P. Klimley. 2009. Do impassable dams and flow regulation constrain the distribution of Green Sturgeon in the Sacramento River, California? Journal of Applied Ichthyology 25 (Supplement 2):39– 47.

580

LITERATURE CITED

Mora Jamett, M., J. Cabrera Peña, and G. Galeano M. 1997. Reproducción y alimentación del gaspar, Atractosteus tropicus (Pisces: Lepisosteidae) en el Refugio Nacional de Vida Silvestre Caño Negro, Costa Rica. Revista de Biología Tropical 45:861–866. Moravec, F., and G. Salgado-Maldonado. 2003 Cystoopsis atractostei sp. n. (Nematoda: Cystoopsidae) from the subcutaneous tissue of the Tropical Gar, Atractosteus tropicus (Pisces) in Mexico. Journal of Parasitology, 89:137–140. Morbey, Y. E. 2003. Pair formation, pre-spawning waiting, and protandry in Kokanee, Oncorhynchus nerka. Behavioral Ecology and Sociobiology 54:127–135. More, W. G. 1942. Field studies on the oxygen requirements of certain fresh-water fishes. Ecology 23:319–329. Moretz, J. A., and W. Rogers. 2004. An ethological analysis of the breeding behavior of the Fantail Darter, Etheostoma flabellare. American Midland Naturalist 152:140–144. Morgan, P., and C. A. Swanberg. 1985. On the Cenozoic uplift and tectonic stability of the Colorado Plateau. Journal of Geodynamics 3:39–63. Morgan, S. G. 1990. Impact of planktivorous fishes on dispersal, hatching, and morphology of estuarine crab larvae. Ecology 71:1639–1652. Morgan II, R. P., B. M. Baker, and J. H. Howard. 1995. Genetic structure of Bay Anchovy (Anchoa mitchilli) populations in Chesapeake Bay. Estuaries 18:482–493. Mori, S. 1995a. Factors associated with and fitness effects of nestraiding in the three-spined Stickleback, Gasterosteus aculeatus, in a natural situation. Behaviour 132:1011–1023. Mori, S. 1995b. Spatial and temporal variations in nest success and the causes of nest losses in the fresh-water 3-spined Stickleback, Gasterosteus aculeatus. Environmental Biology of Fishes 43:323–328. Moriarty, C. 1978. Eels—A Natural and Unnatural History. Universe Books, New York. Morin, P. J. 1987. Predation, breeding asynchrony, and the outcome of competition among treefrog tadpoles. Ecology 68:675–683. Morisawa, M. 1994. Cell signaling mechanism for sperm motility. Zoological Science 11:647–662. Morisawa, M., K. Suzuki, and S. Morisawa. 1983a. Effects of potassium and osmolality on spermatozoan motility of salmonid fishes. Journal of Experimental Biology 107:105–113. Morisawa, M., K. Suzuki, H. Shimizu, S. Morisawa, and K. Yasuda. 1983b. Effects of osmolality and potassium on motility of spermatozoa from freshwater cyprinid fishes. Journal of Experimental Biology 107:95–103. Morita, Y., M. Tabata, K. Uchida, and M. Samejima. 1992. Pinealdependent locomotor activity of Lamprey, Lampetra japonica, measured in relation to LD cycle and circadian rhythmicity. Journal of Comparative Physiology A 171:555–562. Morizot, D. C., J. H. Williamson, and G. J. Carmichael. 2002. Biochemical genetics of Colorado Pikeminnow. North American Journal of Fisheries Management 22:66–76. Morkert, S. B., W. D. Swink, and J. G. Seelye. 1998. Evidence for early metamorphosis of Sea Lampreys in the Chippewa River, Michigan. North American Journal of Fisheries Management 18:966–971. Morman, R. H. 1987. Relationship of density to growth and metamorphosis of caged larval Sea Lampreys, Petromyzon marinus Linnaeus, in Michigan streams. Journal of Fish Biology 30:173–181.

Morris, D. 1955. The reproductive behavior of the River Bullhead (Cottus gobio) with special reference to the fanning activity. Behaviour 7:1–32. Morris, D. 1958. The reproductive behaviour of the ten-spined Stickleback (Pygosteus pungitius L.). Behaviour Supplement 6:1–154. Morris, L. A. 1965. Age and growth of the River Carpsucker, Carpoides carpio, in the Missouri River. American Midland Naturalist 73:423–429. Morris, M. A., and B. M. Burr. 1982. Breeding tubercles in Ictiobus cyprinellus (Pisces: Catostomidae). American Midland Naturalist 107:199–201. Morris, M. R., P. Batra, and M. J. Ryan. 1992. Male-male competition and access to females in Xiphophorus nigrensis. Copeia 1992:980–986. Morris, M. R., and M. J. Ryan. 1990. Age at sexual maturity of Xiphophorus nigrensis in nature. Copeia 1990:747–751. Morris, R. 1972. Osmoregulation, p. 193–239. In The Biology of Lampreys. Vol. 2. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Morrison, W. E., D. H. Secor, and P. M. Piccoli. 2003. Estuarine habitat use by Hudson River American Eels as determined by otolith strontium:calcium ratios. American Fisheries Society Symposium 33:87–99. Morrow, J. E. 1980. The freshwater fishes of Alaska. Alaska Northwest, Anchorage, Alaska. Morrow, J. V., Jr., J. P. Kirk, K. J. Killgore, and S. G. George. 1998a. Age, growth, and mortality of Shovelnose Sturgeon in the lower Mississippi River. North American Journal of Fisheries Management 18:725–730. Morrow, J. V., Jr., J. P. Kirk, K. J. Killgore, H. Rogillio, and C. Knight. 1998b. Status and recovery potential of Gulf Sturgeon in the Pearl River system, Louisiana-Mississippi. North American Journal of Fisheries Management 18:798–808. Morse, R. S., and R. A. Daniels. 2009. A redescription of Catostomus utawana (Cypriniformes: Catostomidae). Copeia 2009:214–220. Morton, T. 1989. Species profiles: life histories and environmental requirements of coastal fishes and invertebrates (MidAtlantic)—Bay Anchovy. Fish and Wildlife Ser vice Biological Report 82(11.97):1–13. Moser, M. L., and S. T. Lindley. 2007. Use of Washington estuaries by subadult and adult Green Sturgeon. Environmental Biology of Fishes 79:243–253. Moser, M. L., and S. W. Ross. 1995. Habitat use and movements of Shortnose and Atlantic Sturgeons in the lower Cape Fear River, North Carolina. Transactions of the American Fisheries Society 124:225–234. Moshenko, R. W., and J. H. Gee. 1973. Diet, time and place of spawning, and environments occupied by Creek Chub (Semotilus atromaculatus) in the Mink River, Manitoba. Journal of the Fisheries Research Board of Canada 30:357–362. Mosher, T. D. 1999. Sturgeon and Paddlefish sportfishing in North America, p. 51–66. Proceedings of the Symposium on the Harvest, Trade and Conservation of North American Paddlefish and Sturgeon, May 7–8, 1998, Chattanooga, Tennessee. D. F. Williamson, G. W. Benz, and C. M. Hoover (eds.). TRAFFIC North America / World Wildlife Fund, Washington, D.C. Moss, R. E., J. W. Scanlon, and C. S. Anderson. 1983. Observations on the natural history of the Blue Sucker Cycleptus elongatus

LITERATURE CITED

(LeSueur) in the Neosho River. American Midland Naturalist 109:15–22. Moyaho, A., C. Macías Garcia, and E. Ávila-Luna. 2004. Mate choice and visibility in the expression of a sexually dimorphic trait in a goodeid fish (Xenotoca variatus). Can. J. Zool. 82: 1917–1922. Moyer, G. R., M. Osborne, and T. F. Turner. 2005. Genetic and ecological dynamics of species replacement in an arid-land river system. Molecular Ecology 14:1263–1273. Moyer, G. R., B. L. Sloss, B. R. Kreiser, and K. A. Feldheim. 2009. Isolation and characterization of microsatellite loci for Alligator Gar (Atractosteus spatula) and their variability in two other species (Lepisosteus oculatus and L. osseus) of Lepisosteidae. Molecular Ecology Resources 9:1–4. Moyle, P. B. 1973. Ecological segregation among three species of minnows (Cyprinidae) in a Minnesota lake. Transactions of the American Fisheries Society 102:794–805. Moyle, P. B. 1976. Inland Fishes of California. University of California Press, Berkeley. Moyle, P. B. 2002. Inland Fishes of California: Revised and Expanded. University of California Press, Berkeley. Moyle, P. B., L. R. Brown, S. D. Chase, and R. M. Quiñones. 2009. Status and conservation of Lampreys in California. American Fisheries Society Symposium 72:279–292. Moyle, P. B., and J. J. Cech, Jr. 2004. Fishes, an Introduction to Ichthyology. 5th edition. Prentice Hall, Upper Saddle River, New Jersey. Moyle, P. B., and B. Herbold. 1987. Life-history patterns and community structure in stream fishes of western North America: comparisons with eastern North America and Europe, p. 25–32. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Moyle, P. B., and R. A. Leidy. 1992. Loss of biodiversity in aquatic ecosystems: evidence from fish faunas, p. 127–169. In Conservation Biology: The Theory and Practice of Nature Conservation, Preservation and Management. P. L. Fiedler and S. K. Jain (eds.). Chapman & Hall, New York. Moyle, P. B., and H. W. Li. 1979. Community ecology and predator-prey relations in warmwater streams, p. 171–180. In Predator-Prey Systems in Fisheries Management. H. Clepper (ed.). Sport Fishing Institute, Washington, D.C. Moyle, P. B., and T. Light. 1996a. Biological invasions of fresh water: empirical rules and assembly theory. Biological Conservation 78:149–161. Moyle, P. B., and T. Light. 1996b. Fish invasions in California: do abiotic factors determine success? Ecology 77:1666–1670. Moyle, P. B., and A. Marciochi. 1975. Biology of the Modoc Sucker, Catostomus microps, in northeastern California. Copeia 1985. Moyle, P. B., and B. Vondracek. 1985. Persistence and structure of the fish assemblage in a small California stream. Ecology 66:1–13. Mueller, G., M. Horn, J. Kahl, T. Burke, and P. C. Marsh. 1993. Use of larval light traps to capture Razorback Sucker (Xyranchen texanus) in Lake Mohave. The Southwestern Naturalist 38:399–402. Mueller, G. A., J. Carpenter, and D. Thornbrugh. 2006. Bullfrog tadpole (Rana catesbeiana) and Red Swamp Crayfish (Procambarus clarkii) predation on early life stages of endangered Razor-

581

back Sucker (Xyrauchen texanus). The Southwestern Naturalist 51:258–261. Mueller, J. F. 1936. Notes on some parasitic copepods and a mite, chiefly from Florida fresh water fishes. American Midland Naturalist 17:807–815. Muir, W. D., G. T. McCabe, Jr., M. J. Parsley, and S. A. Hinton. 2000. Diet of first-feeding larval and young-of-the-year White Sturgeon in the lower Columbia River. Northwest Science 74:25–33. Mulch, A., S. A. Graham, C. P. Chamberlain. 2006. Hydrogen isotopes in Eocene River gravels and paleoelevation of the Sierra Nevada. Science 313:87–89. Mulch, A., A. M. Sarna-Wojcicki, M. E. Perkins, and C. P. Chamberlain. 2008. A Miocene to Pleistocene climate and elevation record of the Sierra Nevada (California). Proceedings of the National Academy of Sciences 105:6819–6824. Muller, E. H. 1965. Quaternary geology of New York, p. 99–112. In The Quaternary of the United States. H. E. Wright, Jr., and D. G. Frey (eds.). Princeton University Press, Princeton, New Jersey. Müller J. 1845. Über den Bau and die Grenzer der Ganoiden and über das natürliche System dei Fische. Archiv für Naturgeschichte 11:91–141. Mulley, J. F., C. Chiu, and P. W. H. Holland. 2006. Breakup of a homeobox cluster after genome duplication in teleosts. Proceedings of the National Academy of Sciences of the United States of America 103:10369–10372. Mundahl, N. D. 1990. Heat death of fish in shrinking stream pools. American Midland Naturalist 123:40–46. Mundahl, N. D., C. Erickson, M. R. Johnston, G. A. Sayeed, and S. Taubel. 2005. Diet, feeding rate, and assimilation efficiency of American Brook Lamprey larvae. Environmental Biology of Fishes 72:67–72. Mundahl, N. D., C. Melnytschuk, D. K. Spielman, J. P. Harkins, K. Funk, and A. M. Bilicki. 1998. Effectiveness of Bowfin as a predator on Bluegill in a vegetated lake. North American Journal of Fisheries Management 18:286–294. Mundahl, N. D., and R. A. Sagan. 2005. Spawning ecology of the American Brook Lamprey, Lampetra appendix. Environmental Biology of Fishes 73:283–292. Munkittrick, K. R., and D. G. Dixon. 1989. Use of White Sucker (Catostomus commersoni) populations to assess the health of aquatic ecosystems exposed to low-level contaminant stress. Canadian Journal of Fisheries Aquatic Science 46:1455–1462. Munro, J., D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). 2007. Anadromus Sturgeons: Habitat, Threats, and Management. American Fisheries Society Symposium 56, Bethesda, Maryland. Munson, P. J., P. W. Parmalee, and R. A. Yarnell. 1971. Subsistence ecology of Scovill, a terminal middle woodland village. American Antiquity 36:410–431. Muntzing, A. 1940. Further studies on apomixes and sexuality in Poa. Hereditas 26:115–188. Murie, D. J., D. C. Parkyn, L. G. Nico, J. J. Herod, and W. F. Loftus. 2009. Age, differential growth and mortality rates in unexploited populations of Florida Gar, an apex predator in the Florida Everglades. Fisheries Management and Ecology 16:315–322. Murphy, C. A., N. E. Stacey, and L. D. Corkum. 2001. Putative steroidal pheromones in the Round Goby, Neogobius melanostomus:

582 LITERATURE CITED

olfactory and behavioral responses. Journal of Chemical Ecology 27:443–470. Murphy, C. E., J. J. Hoover, S. G. George, and K. J. Killgore. 2007a. Morphometric variation among river Sturgeons (Scaphirhynchus spp.) of the middle and lower Mississippi River. Journal of Applied Ichthyology 23:313–323. Murphy, G. 1942. Relationship of the fresh-water mussel to trout in the Truckee River. California Fish and Game 28:89–102. Murphy, C. E., J. J. Hoover, S. G. George, B. R. Lewis, and K. J. Killgore. 2007b. Types and occurrence of morphological anomalies in Scaphirhynchus spp. of the middle and lower Mississippi River. Journal of Applied Ichthyology 23:354–358. Murray A. M. 2001. The fossil record and biogeography of the Cichlidae. Biological Journal of the Linnean Society 74:517–532. Murray, A. M., and M. V. H. Wilson. 1996. A new Palaeocene genus and species of percopsiform (Teleostei: Paracanthopterygii) from the Paskapoo Formation, Smoky Tower, Alberta. Canadian Journal of Earth Science 33:429–438. Murray, S. A., and J. Y. Christmas. 1968. Growth of the uterine young of the Atlantic Stingray, Dasyatis sabina (LeSueur) with notes on its ecology. Journal of the Mississippi Academy of Sciences 14:128. Muth, R. T., and J. C. Schmulbach. 1984. Downstream transport of fish larvae in a shallow prairie river. Transactions of the American Fisheries Society 113:224–230. Muzzall, P. M. 1980a. Population biology and host-parasite relationships of Triganodistomum attenuatum (Trematoda: Lissorchiidae) infecting the White Sucker, Catostomus commersoni (Lacepede). Journal of Parasitology 66:293–298. Muzzall, P. M. 1980b. Seasonal distribution and ecology of three caryophyllaeid cestode species infecting White Suckers in SE New Hampshire. Journal of Parasitology 66:542–550. Myers, G. S. 1938. Fresh-water fishes and West Indian zoogeography. Annual Report of the Smithsonian Institution 1937:339–364. Myers, G. S. 1949. Salt-tolerance of fresh-water fish groups in relation to zoogeographical problems. Bijdragen tot de Dierkunde 28:315–322. Myers, G. S. 1966. Derivation of the freshwater fish fauna of Central America. Copeia 1966:66–773. Myers, R. A., and J. A. Hutchings. 1987. Mating of anadromous Atlantic Salmon, Salmo salar L., with mature male parr. Journal of Fish Biology 31:143–146. Myrberg, A. A., Jr. 1981. Sound communication and interception in fishes, p. 395–426. In Hearing and Sound Communication in Fishes. W. N. Tavolga, A. N. Popper, and R. R. Fay (eds.). Springer, New York. Nachtrieb, H. F. 1910. The primitive pores of Polyodon spathula (Walbaum). The Journal of Experimental Zoology 9:455–468. Nachtrieb, H. F. 1912. The lateral line system of Polyodon. Proceedings of the 7th International Zoölogical Congress, Boston 1907 (1912):215–220. Nagelkerke, L. A. J., F. A. Sibbing, J. G. M. van den Boogaart, E. H. R. R. Lammens, and J. W. M. Osse. 1994. The barbs (Barbus spp.) of Lake Tana: a forgotten species flock? Environmental Biology of Fishes 39:1–22. Nagle, B. C., and A. M. Simons. 2012. Rapid diversification in the North American minnow genus Nocomis. Molecular Phylogenetics and Evolution 63:639–649.

Nakajima, T. 1987. Development of pharyngeal dentition in the cobitid fishes, Misgurnus anguillicaudatus and Cobitis biwae, with a consideration of evolution of cypriniform dentitions. Copeia 1987:208–213. Naseka, A. M., S. B. Tuniyev, and C. B. Renaud. 2009. Lethenteron ninae, a new nonparasitic Lamprey species from the northeastern Black Sea basin (Petromyzontiformes: Petromyzontidae). Zootaxa 2198:16–26. National Center for Biotechnology Information. 2011. GenBank Database. Available from http://www.ncbi.nlm.nih.gov/gen bank; as of October 2011. NatureServe. 2009. NatureServe Explorer: an online encyclopedia of life [web application]. Version 7.1. NatureServe, Arlington, Virginia. Available from http://www.natureserve.org/explorer; as of March 2010. NatureServe. 2010. NatureServer Explorer: an online encyclopedia of life. Version 7.1. NatureServe, Arlington, Virginia. Available from http://www.natureserve.org/explorer/; as of September 2011. Naud, M., and P. Magnan. 1988. Diel onshore offshore migrations in Northern Redbelly Dace, Phoxinus eos (Cope), in relation to prey distribution in a small oligotrophic lake. Canadian Journal of Zoology 66:1249–1253. Naylor, G. J. P., J. A. Ryburn, O Fedrigo, and T. I. Walker. 2005. Phylogenetic relationships among the major lineages of modern elasmobranchs, p. 1–25. In Reproductive Biology and Phylogeny of Chondrichthyes: Sharks, Batoids and Chimaeras. W. C. Hamlett (ed.). Science Publishers, Enfield, New Hampshire. NCWRC (North Carolina Wildlife Resources Commission). 2008. N.C. freshwater fishing records. Available from http://www.nc wildlife.org/fs_index _03_fishing.htm; as of June 2008. NCWRC (North Carolina Wildlife Resources Commission). 2010. Fishing. North Carolina freshwater fishing records. North Carolina Wildlife Resources Commission. Available from http://www.ncwildlife.org/fishing/Fish_NC_Freshwater_Records .htm; as of September 2010. Near, T. J. 2009. Conflict and resolution between phylogenies inferred from molecular and phenotypic data sets for Hagfish, Lampreys, and gnathostomes. Journal of Experimental Zoology (Molecular and Developmental Evolution) 312B:749–761. Near, T. J., D. I. Bolnick, and P. C. Wainwright. 2005. Fossil calibrations and molecular divergence time estimates in centrarchid fishes (Teleostei: Centrarchidae). Evolution 59:1768–1782. Near, T. J., A. Dornburg, R. I. Eytan, B. P. Keck, W. L. Smith, K. L. Kuhn, J. A. Moore, S. A. Price, F. T. Burbrink, M. Friedman, and P. C. Wainwright. 2013. Phylogeny and tempo of diversification in the superradiation of Spiny-rayed Fishes. Proceedings of the National Academy of Sciences Early Edition 1–6. Available from http://www.pnas.org/cgi/doi/10.1073/pnas.1304661110; as of November 2013. Near, T. J., R. I. Etyan, A. Dornburg, K. L. Kuhn, J. A. Moore, M. P. Davis, P. C. Wainwright, M. Friedman, and W. L. Smith. 2012. Resolution of Ray-finned Fishes phylogeny and timing of diversification. Proceedings of the National Academy of Sciences 109:13698–13703. Near, T. J., and B. P. Keck. 2005. Dispersal, vicariance, and timing of diversification in Nothonotus darters. Molecular Ecology 14:3485–3496. Near, T. J., L. M. Page, and R. L. Mayden. 2001. Intraspecific phylogeography of Percina evides (Percidae: Etheostomatinae): an

LITERATURE CITED

additional test of the Central Highlands pre-Pleistocene vicariance hypothesis. Molecular Ecology 10:2235–2240. Needham, R. G. 1965. Spawning of Paddlefish induced by means of pituitary material. The Progressive Fish-Culturist 27:13–19. Neff, B. D. 2001. Genetic paternity analysis and breeding success in Bluegill Sunfish, (Lepomis macrochirus). Journal of Heredity 92:111–119. Neff, B. D. 2003a. Decisions about parental care in response to perceived paternity. Nature 422:716–717. Neff, B. D. 2003b. Paternity and condition affect cannibalistic behavior in nest-tending Bluegill Sunfish. Behavioral Ecology and Sociobiology 54:377–384. Neff, B. D. 2004. Increased performance of offspring sired by parasitic males in Bluegill Sunfish. Behavioral Ecology 15:327–331. Neff, B. D., L. M. Cargnelli, and I. M. Côté. 2004. Solitary nesting as an alternative breeding tactic in colonial nesting Bluegill Sunfish (Lepomis macrochirus). Behavioral Ecology and Sociobiology 56:381–387. Neff, B. D., P. Fu, and M. R. Gross. 2003. Sperm investment and alternative mating tactics in Bluegill Sunfish (Lepomis macrochirus). Behavioral Ecology 14:634–641. Neff, B. D., and M. R. Gross. 2001. Dynamic adjustment of parental care in response to perceived paternity. Proceedings of the Royal Society of London B 268:1559–1565. Neff, B. D., and P. W. Sherman. 2003. Nestling recognition via direct cues by parental male Bluegill Sunfish (Lepomis macrochirus). Animal Cognition 6:87–92. Neff, N. A. 1975. Fishes of the Kannapolis local fauna (Pleistocene) of Ellsworth County, Kansas University of Michigan, Museum of Paleontology, Papers on Paleontology 12:1–43. Neidert, A. H., G. Panopoulou, and J. A. Langeland. 2000. Amphioxus goosecoid and the evolution of the head organizer and prechordal plate. Evolution and Development 2:303–310. Neidert, A. H., V. Virupannavar, G. W. Hooker, and J. A. Langeland. 2001. Lamprey Dlx genes and early vertebrate evolution. Proceedings of the National Academy of Sciences 98:1665–1670. Neil, W. H., and J. J. Magnuson. 1974. Distributional ecology and behavioral thermoregulation of fishes in relation to heated effluent from a power plant at Lake Monona, Wisconsin. Transactions of the American Fisheries Society 103:663–710. Neill, W. T. 1950. An estivating Bowfin. Copeia 1950:240. Neiman, A., and D. F. Russell. 2001. Stochastic biperiodic oscillations in the electroreceptors of Paddlefish. Physical Review Letters 86:3343–3346. Neiman, A. B., D. F. Russell, X. Pei, W. Wojtenek, J. Twitty, E. Simonotto, B. A. Wettring, E. Wagner, L. A. Wilkens, and F. Moss. 2000. Stochastic synchronization of electroreceptors in Paddlefish. International Journal of Bifurcation and Chaos 10:2499–2517. Nellis, P., J. Munro, D. Hatin, G. Desrosiers, R. D. Simons, and F. Guilbard. 2007a. Macrobenthos assemblages in the St. Lawrence estuarine transition zone and their potential as food for Atlantic Sturgeon and Lake Sturgeon, p. 105–128. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Journal of Bifurcation and Chaos 10:2499–2517. Nellis, P., S. Senneville, J. Munro, G. Drapeau, D. Hatin, G. Desrosiers, and F. J. Saucier. 2007b. Tracking the dumping and bed

583

load transportation of dredged sediments in the St. Lawrence estuarine transition zone and assessing their impacts on macrobenthos in Atlantic Sturgeon habitat, p. 215–234. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Nelson, C. M., and K. Planes. 1993. Female choice of nonmelanistic males in laboratory populations of mosquitofish, Gambusia holbrooki. Copeia 1993:1143–1148. Nelson, E. M. 1948. The comparative morphology of the Weberian apparatus of the Catostomidae and its significance in systematics. Journal of Morphology 83:225–251. Nelson, E. M. 1949. The opercular series of the Catostomidae. Journal of Morphology 85:559–567. Nelson, E. M. 1959. The embryology of the swim bladder in the common Sucker, Catostomus commersoni (Lacepéde). American Midland Naturalist 61:245–252. Nelson, E. M. 1961. The comparative morphology of the definitive swim bladder in the Catostomidae. American Midland Naturalist 65:101–110. Nelson, G. J. 1968. Gill arches of teleostean fishes of the division Osteoglossomorpha. Journal of the Linnean Society (Zoology) 47:261–277. Nelson, G. J. 1969a. Gill arches and the phylogeny of fishes, with notes on the classification of vertebrates. Bulletin of the American Museum of Natural History 141:475–552. Nelson, G. J. 1969b. Infraorbital bones and their bearing on the phylogeny and geography of osteoglossomorph fishes. American Museum Novitates 2394:1–37. Nelson, G. J. 1972. Observations on the gut of Osteoglossomorpha. Copeia 1972:325–329. Nelson, J. S. 1968. Hybridization and isolating mechanisms between Catostomus commersonii and C. macrocheilus (Pisces: Catostomidae). Journal of the Fisheries Research Board of Canada 25:101–150. Nelson, J. S. 1973. Occurrence of hybrids between Longnose Sucker (Catostomus catostomus) and White Sucker (C. commersoni) in upper Kananaskis Reservoir, Alberta. Journal of the Fisheries Research Board of Canada 30:557–560. Nelson, J. S. 1974. Hybridization between Catostomus commersoni (White Sucker) and Catostomus macrocheilus (Largescale Sucker) in Williston Reservoir, British Columbia, with notes on other fishes. Syesis 7:187–194. Nelson, J. S. 1994. Fishes of the World. 3rd edition. John Wiley and Sons, New York. Nelson, J. S. 2006. Fishes of the World. 4th edition. John Wiley and Sons, Inc., Hoboken, New Jersey. Nelson, J. S., E. J. Crossman, H. Espinosa-Perez, L. T. Findley, C. R. Gilbert, R. N. Lea, and J. D. Williams. 2004. Common and scientific names of fishes from the United States, Canada, and Mexico. American Fisheries Society, Special Publication 29, Bethesda, Maryland. Nelson, J. S., W. C. Starnes, and M. L. Warren, Jr. 2002. A capital case for common names of species of fishes-a white crappie or a White Crappie. Fisheries 27:31–33. Nelson, P. A. 2005. Ecology of sympatric catostomid fishes in a glaciated riverine system: habitat, food, and biogeography. Unpubl. Ph.D. diss., University of Manitoba, Winnipeg, Canada.

584 LITERATURE CITED

Nessov, L. A., F. P. Vasil’yevich, and N. I. Udovichenko. 1990. Jurassic, Cretaceous and Palaeogene vertebrates from northeastern Fergan and their significance for making more precise the age of deposits and palaeoenvironments. II. Late Cretaceous and Palaeogene vertebrates. Description of new forms of Jurassic vertebrates. Vestnik Leningradskogo Universiteta, Seriia 7, Geologiia, Geografiia 1:1–18 (in Russian). Netsch, N. F. 1964. Food and feeding habits of Longnose Gar in central Missouri. Proceedings of the Southeastern Association of Game and Fish Commissioners 8:506–511. Netsch, N. F., and A. Witt, Jr. 1962. Contributions to the life history of the Longnose Gar (Lepisosteus osseus) in Missouri. Transactions of the American Fisheries Society 9:251–262. New, J. G. 1962. Hybridization between two cyprinids, Chrosomus eos and Chrosomus neogaeus. Copeia 1962:147–152. New, J. G., and D. Bodznick. 1985. Segregation of electroreceptive and mechanoreceptive lateral line afferents in the hindbrain of chondrosteans fishes. Brain Research 336:89–98. New, J. G., and T. C. Tricas. 2001. Electroreceptors and magnetoreceptors, p. 839–856. In Cell Physiology Sourcebook: A Molecular Approach. 3rd edition. N. Sperlakis (ed.). Academic Press, San Diego, California. Newberger, T. A., and E. D. Houde. 1995. Population biology of Bay Anchovy, Anchoa mitchilli in the mid Chesapeake Bay. Marine Ecology Progress Series 116:25–37. Newbrey, M. G., M. V. H. Wilson, and A. C. Ashworth. 2007. Centrum growth patterns provide evidence for two small taxa of Hiodontidae in the Cretaceous Dinosaur Park Formation. Canadian Journal of Earth Sciences 44:721–732. Newcombe, C., and G. Hartman. 1973. Some chemical signals in the spawning behaviour of Rainbow Trout (Salmo gairdneri). Journal of the Fisheries Research Board of Canada 30:995–997. Nguyen, R. M., and C. E. Crocker. 2007. The effects of substrate composition on foraging behavior and growth rate of larval Green Sturgeon, Acipenser medirostris. Environmental Biology of Fishes 79:231–241. Nicholls, H. 2009. Mouth to mouth. Nature 461:164–166. Nichols, J. T. 1916. A large Polyodon from Iowa. Copeia 1916:65. Nichols, O. C., and P. K. Hamilton. 2004. Occurrence of the parasitic Sea Lamprey, Petromyzon marinus, on western North Atlantic Right Whales, Eubalaena glacialis. Environmental Biology of Fishes 71:413–417. Nichols, S. J., G. Kennedy, E. Crawford, J. Allen, J. French III, G. Black, M. Blouin, J. Hickey, S. Chernyák, R. Haas, and M. Thomas. 2003. Assessment of Lake Sturgeon (Acipenser fulvescens) spawning efforts in the lower St. Clair River, Michigan. Journal of Great Lakes Research 29:383–391. Nico, L. G., J. D. Williams, and H. L. Jelks. 2005. Black Carp: biological synopsis and risk assessment of an introduced fish. American Fisheries Society Special Publication 32. Bethesda, Maryland. Nicol, J. A. C., H. J. Arnott, and A. C. G. Best. 1973. Tapeta lucida in bony fishes (Actinopterygii): a survey. Canadian Journal of Zoology 51:69–81. Nicoletto, P. F., and S. H. Linscomb. 2008. Sound production by the Sheepshead Minnow, Cyprinodon variegatus. Environmental Biology of Fishes 81:15–20. Niemeitz, A., R. Kreutzfeldt, M. Schartl, J. Parzefall, and I. Schlupp. 2002. Male mating behaviour of a molly, Poecilia latipunctata: a third host for the sperm-dependent Amazon Molly, Poecilia formosa. Acta Ethologica 5:45–49.

Nikinmaa, M. 2001. Haemoglobin function in vertebrates: evolutionary changes in cellular regulation in hypoxia. Respiration Physiology 128:317–329. Niklitschek, E. J., and D. H. Secor. 2005. Modeling spatial and temporal variation of suitable nursery habitats for Atlantic Sturgeon in the Chesapeake Bay. Estuarine, Coastal and Shelf Science 64:135–148. Nikolskii, G. V. 1956. [Some data on the period of marine life of the Pacific Lamprey Lampetra japonica (Martens)]. Zoologicheskii Zhurnal 35:588–591 (in Russian). Nilsson, C., C. A. Reidy, M. Dynesius, and C. Revenga. 2005. Fragmentation and flow regulation of the world’s large river systems. Science 308:405–408. Nilsson, S. 1981. On the adrenergic system of ganoid fish: the Florida spotted Gar, Lepisosteus platyrhincus (Holostei). Acta Physiologica Scandinavia 111:447–454. Nishidia, K. 1990. Phylogeny of the superorder Myliobatidodei. Memoirs of the Faculty of Fisheries, Hokaido Univ. 37, Hokkaido, Sapporo, Japan. Nizinski, M. S., and T. A. Munroe. 2003 (dated 2002). Engraulidae, p. 764–794. In The Living Marine Resources of the Western Central Atlantic. FAO Species Identification Guide for Fisheries Purposes. Volume 2. K. E. Carpenter (ed.). Rome, Italy. NMFS (National Marine Fisheries Ser vice). 2006. Endangered and threatened wildlife and plants: threatened status for southern distinct population segment of North American Green Sturgeon. Federal Register 71 (67):17757–17766. NMFS (National Marine Fisheries Ser vice). 2010a. Endangered and threatened wildlife and plants; proposed listing determinations for three distinct population segments of Atlantic Sturgeon in the northeast region. Federal Register 75 (193):61872–61904. NMFS (National Marine Fisheries Ser vice). 2010b. Endangered and threatened wildlife and plants; proposed listing for two distinct population segments of Atlantic Sturgeon (Acipenser oxyrinchus oxyrinchus) in the Southeast. Federal Register 75 (193):61904–61929. Noakes, D. L. G., F. W. H. Beamish, and A. Rossiter. 1999. Conservation implications of behaviour and growth of the Lake Sturgeon, Acipenser fulvescens, in northern Ontario. Environmental Biology of Fishes 55:135–144. Nolf, D. 1985. Otolithi piscium. Handbook of Paleoichthyology, volume 10. H.-P. Schultze (ed.). Gustav Fischer Verlag, Stuttgart. Noltie, D. B., and M. H. A. Keenleyside. 1986. Correlates of reproductive success in stream-dwelling male Rock Bass, Ambloplites rupestris (Centrarchidae). Environmental Biology of Fishes 17:61–70. Noltie, D. B., and R. J. F. Smith. 1988. The Redfin Shiner, Notropis umbratilis, in the Middle Thames River, Ontario, and its association with breeding Longear Sunfish, Lepomis megalotis. Canadian Field-Naturalist 102:533–535. Normark, B. B., A. R. McCune and R. G. Harrison. 1991. Phylogenetic relationships of neopterygian fishes, inferred from mitochondrial DNA sequences. Molecular Biology and Evolution 8:819–834. Norris, H. W. 1923. On the function of the paddle of the Paddlefish. Iowa Academy of Science 30:135–137. Norris, H. W. 1924. The lateral line organs of the shovel-nose Sturgeon, distribution and innervation. Proceedings of the Iowa Academy of Science 31:443–444.

LITERATURE CITED

North, J. A., R. A. Farr, and P. Vecsei. 2002. A comparison of meristic and morphometric characters of Green Sturgeon Acipenser medirostris. Journal of Applied Ichthyology 18:234–239. Northcote, T. G. 1995. Confessions from a four decade affair with Dolly Varden: a synthesis and critique of experimental tests for interactive segregation between Dolly Varden char (Salvelinus malma) and Cutthroat Trout (Oncorhynchus clarki) in British Columbia. Nordic Journal of Freshwater Research 71:49– 67. Northcutt, R. G. 1986. Electroreception in nonteleost bony fishes, p. 257–285. In Electroreception. T. H. Bullock and W. Heiligenberg (eds.). Wiley-Interscience, New York. Noss, R. F. 1996. The naturalists are dying off. Conservation Biology 10:1–3. Novales Flamarique, I., and F. I. Hárosi. 2002. Visual pigments and dichroism of Anchovy cones: a model system for polarization detection. Visual Neuroscience 19:467–473. Nuismer, S. L., R. Gomulkiewicz, and M. T. Morgan. 2003. Coevolution in temporally variable environments. The American Naturalist 162:195–204. Nunez, S., and J. M. Trant. 1999. Regulation of interrenal gland steriodogenesis in the Atlantic Stingray (Dasyatis sabina). Journal of Experimental Zoology 284:517–525. Nurnberger, P. K. 1930. The plant and animal food of the fishes of Big Sandy Lake. Transactions of the American Fisheries Society 60:253–259. Oakey, D. D., M. E. Douglas, and M. R. Douglas. 2004. Small fish in a large landscape: diversification of Rhinichthys osculus (Cyprinidae) in western North America. Copeia 2004:207–221. Oakley, N. C., and J. E. Hightower. 2007. Status of Shortnose Sturgeon in the Neuse River, North Carolina, p. 273–284. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. O’Brien, W. J., B. Evans, and C. Luecke. 1985. Apparent size choice of zooplanktivorous Sunfish—exceptions to the rule. Environmental Biology of Fishes 13:225–233. O’Brien, W. J., N. A. Slade, and G. L. Vinyard. 1976. Apparent size as the determinant of prey selection by Bluegill Sunfish (Lepomis macrochirus). Ecology 57:1304–1310. O’Connell, M. T. 2003. Direct exploitation of prey on an inundated floodplain by Cherryfin Shiners (Lythrurus roseipinnis) in a low order, blackwater stream. Copeia 2003:635–645. O’Connell, M. T., T. D. Sheperd, A. M. U. O’Connell, and R. A. Myers. 2007. Long-term declines in two apex predators, Bull Sharks (Carcharhinus leucas) and Alligator Gar (Atractosteus spatula) in Lake Pontchartrain, an oligohaline estuary in southeastern Louisiana. Estuaries and Coasts 30:567–574. Odum, H. T. 1953. Factors controlling marine invasion into Florida fresh waters. Bulletin Marine Science Gulf and Caribbean 3:134–156. Odum, H. T., and D. K. Caldwell. 1955. Fish respiration in the natural oxygen gradient of an anaerobic spring in Florida. Copeia 1955:104–106. Odum, W. E., and E. J. Heald. 1972. Trophic analyses of an estuarine mangrove community. Bulletin of Marine Science 22:671–738. Ogden, J. C. 1970. Relative abundance, food habits, and age of the American Eel, Anguilla rostrata (LeSueur), in certain New Jersey streams. Transactions of the American Fisheries Society 99:54–59.

585

Ogden, J. C., J. A. Kushlan, and J. T. Tilmant. 1976. Prey selectivity by the Wood Stork. The Condor 78:324–330. Ohno, S. 1970. Evolution by gene duplication. Springer-Verlag, New York. Ohno, S. 1999. Gene duplication and the uniqueness of vertebrate genomes circa 1970–1999. Cell and Developmental Biology 10:517–522. Ohno, S., L. Christian, M. Romero, R. Dofuku, and C. Ivey. 1973. On the question of American Eels, Anguilla rostrata, versus European Eels, Anguilla anguilla. Experientia 29:891. Ohno S., J. Muramoto, C. Steius, L. Christian, and W. A. Kittrell. 1969. Microchromosomes in holocephalian, chondrostean and holostean fishes. Chromosoma (Berl.) 26:35–40. Ohno, S., U. Wolf, and N. B. Atkin. 1968. Evolution from fish to mammals by gene duplication. Hereditas 59:169–187. O’Keefe, D. M., and D. C. Jackson. 2009. Population characteristics of Paddlefish in two Tennessee-Tombigbee Waterway habitats, p. 83–101. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. O’Keefe, D. M., J. C. O’Keefe, and D. C. Jackson. 2007. Factors influencing Paddlefish spawning in the Tombigbee watershed. Southeastern Naturalist 6:321–332. O’Kelley, C. T., and S. L. Powers. 2007. Life history aspects of Hypentelium etowanum (Alabama Hog Sucker) (Actinopterygii: Catostomidae) in Northern Georgia. Southeastern Naturalist 6:479– 490. Olden, J. D. 2006. Biotic homogenization: a new research agenda for conservation biogeography. Journal of Biogeography 33:2027–2039. Oliveira, K. 1997. Movements and growth rates of yellow-phase American Eels in the Annaquatucket River, Rhode Island. Transactions of the American Fisheries Society 126:638–646. Oliveira, K. 1999. Life history characteristics and strategies of American Eel, Anguilla rostrata, in four rivers in Maine, U.S.A. Canadian Journal of Fisheries and Aquatic Sciences 56:795–802. Oliveira, K., and W. E. Hable. 2010. Artificial maturation, fertilization, and early development of the American Eel (Anguilla rostrata). Canadian Journal of Zoology 88:1121–1128. Oliveira, K., and J. D. McCleave. 2000. Variation in population and life history traits of the American Eel, Anguilla rostrata, in four rivers in Maine, U.S.A., Environmental Biology of Fishes 59:141–151. Oliveira, K., J. D. McCleave, and G. S. Wippelhauser. 2001. Regional variation and the effect of lake:river area on sex distribution of American Eels. Journal of Fish Biology 58:943–952. Olney, J. E. 1983. Eggs and early larvae of the Bay Anchovy, Anchoa mitchilli, and the Weakfish, Cynoscion regalis, in lower Chesapeake Bay with notes on associated ichthyoplankton. Estuaries 6:20–35. Olsén, K. H. 1989. Sibling recognition in juvenile Arctic Char, Salvelinus alpinus (L.). Journal of Fish Biology 34:571–581. Olsén, K. H. 1992. Kin recognition in fish mediated by chemical cues, p. 229–248. In Fish Chemoreception, Fish and Fisheries Series 6. T. J. Hara (ed.). Chapman and Hall, London. Olsén, K. H., R. Bjerselius, I. Mayer, and H. Kindhal. 2002b. Both ovarian fluid and female urine increase steroid hormone levels in mature Atlantic Salmon (Salmo salar) male parr. Journal of Chemical Ecology 28:29–40.

586 LITERATURE CITED

Olsén, K. H., R. Bjerselius, E. Petersson, T. Järvi, I. Mayer, and M. Hedenskog. 2000. Lack of species-specific primer effects of odours from female Atlantic Salmon, Salmo salar, and Brown Trout, Salmo trutta. Oikos 88:213–220. Olsén, K. H., M. Grahn, and J. Lohm. 2002a. Influence of mhc on sibling discrimination in Arctic Char, Salvelinus alpinus (L.). Journal of Chemical Ecology 28:783–795. Olsén, K. H., M. Grahn, and J. Lohm. 2003. The influence of dominance and diet on individual odours in MHC identical juvenile in Arctic Char siblings. Journal of Fish Biology 63:855–862. Olsén, K. H., M. Grahn, J. Lohm, and A. Langefors. 1998b. MHC and kin discrimination in juvenile Arctic Char, Salvelinus alpinus (L.). Animal Behaviour 56:319–327. Olsén, K. H., T. Järvi, and A-C. Löf. 1996. Aggressiveness and kinship in Brown Trout (Salmo trutta) parr. Behavioral Ecology 7:445–450. Olsén, K. H., T. Jävi, I. Mayer, E. Petersson, and F. Kroon. 1998a. Spawning behaviour and sex hormone levels in adult and precocious Brown Trout (Salmo trutta L.) males and the effect of anosmia. Chemoecology 8:9–17. Olsén, K. H., A. K. Johansson, R. Bjerselius, I. Mayer, and H. Kindhal. 2001. Mature Atlantic Salmon (Salmo salar L.) male parr are attracted to ovulated female urine but not to ovarian fluid. Journal of Chemical Ecology 27:2337–2349. Olsén, K. H., and N. R. Liley. 1993. The significance of olfaction and social cues in milt availability, sexual hormone status and spawning behaviour of male Rainbow Trout (Oncorhynchus mykiss). General and Comparative Endocrinology 89:107–118. Olsén, K. H., and S. Winberg. 1996. Learning and sibling odor preference in juvenile Arctic Char, Salvelinus alpinus (L.). Journal of Chemical Ecology 22:773–786. Olsen, P. E. 1984. The skull and pectoral girdle of the parasemionotid fish Watsonulus eugnathoides from the Early Triassic Sakamena Group of Madagascar, with comments on the relationships of the holostean fishes. J. Vert. Paleontol. 4:481–499. Olsen, P. E., and A. R. McCune. 1991. Morphology of the Semionotus elegans species group from the Early Jurassic part of the Newark supergroup of eastern America, with comments on the family Semiodontidae (Neopterygii). Journal of Vertebrate Paleontology 11:269–292. Olson, D. E., and W. J. Scidmore. 1963. Homing tendency of spawning White Suckers in Many Point Lake, Minnesota. Transactions of the American Fisheries Society 92:13–16. OMNR (Ontario Ministry of Natural Resources). 1986. Baitfish summary 1985. Fisheries Policy Branch, Toronto, Ontario. Onders, R. J., S. D. Mims, C. Wang, and W. D. Pearson. 2001. Reservoir ranching of Paddlefish. North American Journal of Aquaculture 63:179–190. Ong, T. L., J. Stabile, I. Wirgin, and J. R. Waldman. 1996. Genetic divergence between Acipenser oxyrinchus oxyrinchus and A. o. desotoi as assessed by mitochondrial DNA sequencing analysis. Copeia 1996:464–469. Orlando, E. F., G. A. Binczik, N. D. Denslow, and L. J. Guillette, Jr. 2007. Reproductive seasonality of the female Florida Gar, Lepisosteus platyrhincus. General and Comparative Endocrinology 151:318–324. Orlando, E. F., G. A. Binczik, P. Thomas, and L. J. Guillette, Jr. 2003. Reproductive seasonality of the male Florida Gar, Lepisosteus platyrhincus, Orange Lake, Florida. General and Comparative Endocrinology 131:365–371.

Osborne, L. L., and M. J. Wiley. 1992. Influence of tributary spatial position on the structure of warmwater fish communities. Canadian Journal of Fisheries and Aquatic Science 49:671–681. Osmundson, D. B., R. J. Ryel, M. E. Tucker, B. D. Burdick, W. R. Elmblad, and T. E. Chart. 1998. Dispersal patterns of subadult and adult Colorado squawfish in the Upper Colorado River. Transactions of the American Fisheries Society 127:943–966. Ostrand, K. G., B. J. Braeutigam, and D. H. Wahl. 2004. Consequences of vegetation density and prey species on Spotted Gar foraging. Transactions of the American Fisheries Society 133:794–800. Ostrand, K. G., M. L. Thies, D. D. Hall, and M. Carpenter. 1996. Gar ichthyootoxin: its effect on natural predators and the toxin’s evolutionary function. The Southwestern Naturalist 41:375–377. Ostrand, K. G., and G. R. Wilde. 2001. Temperature, dissolved oxygen, and salinity tolerances of five prairie stream fishes and their role in explaining fish assemblage patterns. Transactions of the American Fisheries Society 130:742–749. OSUDM (Ohio State University Division of Molluscs). 2010. Mussel/Host Database. Available from http://128.146.250.63/Musselhost/; as of various dates. Oswood, M. W., J. B. Reynolds, J. G. Irons, III, and A. M. Milner. 2000. Distributions of freshwater fishes in ecoregions and hydroregions of Alaska. Journal of the North American Benthological Society 19:405–418. Ota, K. G., S. Kuraku, and S. Kuratani. 2007. Hagfish embryology with reference to the evolution of the neural crest. Nature 446:672–675. Otake, T. 1991. Serum composition and nephron structure of freshwater elasmobranchs collected from Australia and Papua New Guinea, p. 55–62. In Studies on Elasmobranchs Collected from Seven River Systems in Northern Australia and Papua New Guinea. Vol. 3. M. Shimizu and T. Taniuchi (eds.). University Museum, University of Tokyo, Tokyo, Japan. Oviatt, C. G., D. R. Currey, and D. Sack. 1992. Radiocarbon chronology of Lake Bonneville, Eastern Great Basin, USA. Palaeogeography, Palaeoclimatology, and Palaeoecology 99:225–241. Owen, L. A., R. C. Finkel, R. A. Minnich, and A. E. Perez. 2003. Extreme southwestern margin of late Quaternary glaciation in North America: timing and controls. Geology 31:729–732. Padilla, R. 1972. Reproduction of carp, Smallmouth Buffalo, and River Carpsucker in Elephant Butte Lake. Unpubl. Master’s thesis, New Mexico State University, Los Cruces. Page, L. M. 1974. The life history of the Spottail Darter, Etheostoma squamiceps, in Big Creek, Illinois, and Ferguson Creek, Kentucky. Illinois Natural History Survey Biological Notes 89:1–20. Page, L. M. 1983. Handbook of Darters. TFH Publications, Neptune City, New Jersey. Page, L. M. 1985. Evolution of reproductive behaviors in percid fishes. Illinois Natural History Survey Biological Notes 33:275–295. Page, L. M., and H. L. Bart, Jr. 1989. Egg-mimics in darters (Pisces: Percidae). Copeia 1989:514–517. Page, L. M., and B. M. Burr. 1991. A Field Guide to Freshwater Fishes, North America North of Mexico. Houghton Mifflin Company, Boston, Massachusetts. Page, L. M., and B. M. Burr. 2011. Peterson Field Guide to Freshwater Fishes of North America North of Mexico. 2nd edition. Houghton Mifflin Harcourt Publishing Co., New York. Page, L. M., and P. A. Ceas. 1989. Egg attachment in Pimephales (Pisces: Cyprinidae). Copeia 1989:1074–1077.

LITERATURE CITED

Page, L. M., P. A. Ceas, D. L. Swofford, and D. G. Buth. 1992. Evolutionary relationships within the Etheostoma squamiceps complex (Percidae; subgenus Catonotus) with descriptions of five new species. Copeia 1992:615–646. Page, L. M., H. Espinosa-Pérez, L. T. Findley, C. R. Gilbert, R. N Lea, N. E. Mandrak, R. L. Mayden, and J. L. Nelson. 2013. Common and scientific names of fishes from the United States, Canada, and Mexico. American Fisheries Society Special Publication 34, Bethesda, Maryland. Page, L. M., M. Hardman, and T. J. Near. 2003. Phylogenetic relationships of barcheek darters (Percidea: Etheostoma, Subgenus Catonotus) with descriptions of two new species. Copeia 2003:512–530. Page, L. M., and C. E. Johnston. 1990. Spawning behavior in the Creek Chubsucker, Erimyzon oblongus, with a review of spawning behavior in Suckers (Catostomdae). Environmental Biology of Fishes 27:265–272. Page, L. M., and C. E. Johnston. 1990a. The breeding behavior of Opsopoeodus emiliae (Cyprinidae) and its phylogenetic implications. Copeia 1990:1176–1180. Page, L. M., and J. H. Knouft. 2000. Variation in egg-mimic size in the Guardian Darter, Etheostoma oophylax (Percidae). Copeia 2000:782–785. Page, L. M., and D. W. Schemske. 1978. The effect of interspecific competition on the distribution and size of darters of the subgenus Catonotus (Percidae: Etheostoma). Copeia 1978:406–412. Page, L. M., and T. P. Simon. 1988. Observations on the reproductive behavior and eggs of four species of darters, with comments on Etheostoma tippecanoe and E. camurum. Transactions of the Illinois Academy of Science 81:205–210. Page, L. M., and D. L. Swofford. 1984. Morphological correlates of ecological specialization in darters. Environmental Biology of Fishes 11:139–159. Palkovacs, E. P., M. C. Marshall, B. A. Lamphere, B. R. Lynch, D. J. Weese, D. F. Fraser, D. N. Reznick, C. M. Pringle, and M. T. Kinnison. 2009. Experimental evaluation of evolution and coevolution as agents of ecosystem change in Trinidadian streams. Philosophical Transactions of the Royal Society B 364:1617–1628. Paller, M. H. 2002. Temporal variability in fish assemblages from disturbed and undisturbed streams. Journal of Aquatic Ecosystem Stress and Recovery 9:149–158. Pankhurst, N. W. 1984. Retinal development in larval and juvenile European Eel, Anguilla anguilla (L.). Canadian Journal of Zoology 62:335–343. Pankhurst, N. W. 1985. Final maturation and ovulation of oocytes of the Goldeye, Hiodon alosoides (Rafinesque), in vitro. Canadian Journal of Zoology 63:1003–1009. Pankhurst, N. W., and P. W. Sorensen. 1984. Degeneration of the alimentary canal in maturing European Anguilla anguilla (L.) and American Eels Anguilla rostrata (LeSueur). Canadian Journal of Zoology 62:1143–1149. Pankhurst, N. W., N. E. Stacey, and G. Van der Kraak. 1986a. Reproductive development and plasma levels of reproductive hormones of Goldeye, Hiodon alosoides (Rafinesque), taken from the North Saskatchewan River during the open-water season. Canadian Journal of Zoology 64:2843–2849. Pankhurst, N. W., G. Van Der Kraak, R. E. Peter, and B. Brenton. 1986b. Effects of (D-Ala6, Pro9-N-ethylamide)-LHRH on plasma levels of gonadotropin, 17alpha, 20beta-dihydroxy-4-pregnen-3-

587

one and testosterone in male Goldeye (Hiodon alosoides Rafinesque). Fish Physiology and Biochemistry 1:163–170. Papoulias, D. M., A. J. DeLonay, M. L. Annis, M. L. Wildhaber, and D. E. Tillitt. 2011. Characterization of environmental cues for initiation of reproductive cycling and spawning in Shovelnose Sturgeon Scaphirhynchus platorynchus in the lower Missouri River, USA. Journal of Applied Ichthyology 27:335–342. Papoulias, D., and W. L. Minckley. 1989. Food limited survival of larval Razorback Sucker, Xyrauchen texanus, in the laboratory. Environmental Biology of Fishes 29:73–78. Pappantoniou, A., Schmidt, R. E., and G. Dale. 1983. Aspects of the life history of the Cutlips Minnow, Exoglossum maxillingua (Pisces: Cyprinidae), from the Titicus River, Westchester County, New York. Annals of the New York Academy of Sciences 435:325–357. Paragamian, V. L., R. McDonald, G. J. Nelson, and G. Barton. 2009. Kootenai River velocities, depth, and White Sturgeon spawning site selection—a mystery unraveled? Journal of Applied Ichthyology 25:640–646. Parauka, F. M., M. S. Duncan, and P. A. Lang. 2011. Winter coastal movement of Gulf of Mexico Sturgeon throughout northwest Florida and southeast Alabama. Journal of Applied Ichthyology 27:343–350. Parauka, F. M., W. J. Troxel, F. A. Chapman, and L. G. McBay. 1991. Hormone-induced ovulation and artificial spawning of Gulf of Mexico Sturgeon (Acipenser oxyrhynchus desotoi). The Progressive Fish-Culturalist 53:113–117. Parenti, L. R. 1981. A phylogenetic and biogeographic analysis of cyprinodontiform fishes (Teleostei, Atherinomorpha). Bulletin of the American Museum of Natural History 168:335–557. Park, D., and C. R. Propper. 2002. Pheromones from female mosquitofish at different stages of reproduction differentially affect male sexual activity. Copeia 2002:1113–1117. Parken, C. K., and D. L. Scarnecchia. 2002. Predation on age-0 Paddlefish by Walleye and Sauger in a Great Plains reservoir. North American Journal of Fisheries Management 22:750–759. Parker, B. J. 1988. Status of the Paddlefish, Polyodon spathula, in Canada. The Canadian Field-Naturalist 102:291–295. Parker, B. R., and W. G. Franzin. 1991. Reproductive biology of the Quillback, Carpiodes cyprinus, in a small prairie river. Canadian Journal of Zoology 69:2133–2139. Parker, G. A. 1974. Assessment strategy and evolution of fighting behavior. Journal of Theoretical Biology 47:223–243. Parker, G. A. 1998. Sperm competition and the evolution of ejaculates: towards a theory base, p. 3–54. In Sperm Competition and Sexual Selection. T. R. Birkhead and A. P. Moller (eds.). Academic Press, New York. Parker, G. A., M. A. Ball, P. Stockley, and M. J. G. Gage. 1996. Sperm competition games: individual assessment of sperm competition intensity by group spawners. Proceedings of the Royal Society of London B 263:1291–1297. Parker, G. A, M. A. Ball, P. Stockley, and M. J. G. Gage. 1997. Sperm competition games: a prospective analysis of risk assessment. Proceedings of the Royal Society of London B 264:1793–1802. Parker, P. S., and R. E. Lennon. 1956. Biology of the Sea Lamprey in its pancreatic phase. United States Fish and Wildlife Ser vice Research Report 44:1–32. Parker, R. S., C. T. Hackney, and M. F. Vidrine. 1984. Ecology and reproductive strategy of a south Louisiana freshwater mussel,

588

LITERATURE CITED

Glebula rotundata (Lamarck) (Unionidae: Lampsilini). Freshwater Invertebrate Biology. 3:53–58. Parker, S. J. 1995. Homing ability and home range of yellow-phase American Eels in a tidally dominated estuary. Journal of the Marine Biological Association of the United Kingdom 75:127–140. Parker, S. J., and J. D. McCleave. 1997. Selective tidal stream transport by American Eels during homing movements and estuarine migration. Journal of the Marine Biological Association of the United Kingdom 77:871–889. Parkyn, D. C., D. J. Murie, J. E. Harris, D. E. Colle, and J. D. Holloway. 2007. Seasonal movements of Gulf of Mexico Sturgeon in the Suwannee River and estuary, p. 51–68. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Parmalee, P. W. 1962. The faunal complex of the Fisher site, Illinois. American Midland Naturalist 68:399–408. Parsley, M. J., and L. G. Beckman. 1994. White Sturgeon spawning and rearing habitat in the lower Columbia River. North American Journal of Fisheries Management 14:812–827. Parsley, M. J., P. J. Anders, A. I. Miller, L. G. Beckman, and G. T. McCabe, Jr. 2002. Recovery of White Sturgeon populations through natural production: understanding the influence of abiotic and biotic factors on spawning and subsequent recruitment, p. 55–66. In Biology, Management, and Protection of North American Sturgeon. W. Van Winkle, P. Anders, D. H. Secor, and D. Dixon (eds.). American Fisheries Society Symposium 28, Bethesda, Maryland. Parsley, M. J., L. G. Beckman, and G. T. McCabe, Jr. 1993. Spawning and rearing habitat use by White Sturgeon in the Columbia River downstream from McNary Dam. Transactions of the American Fisheries Society 122:217–227. Parsley, M. J., C. D. Wright, B. K. van der Leeuw, E. E. Kofoot, C. A. Peery, and M. L. Moser. 2007. White Sturgeon (Acipenser transmontanus) passage at the Dalles Dam, Columbia River, USA. Journal of Applied Ichthyology 23:627–635. Parzefall, J. 1969. Zur vergleichenden Ethologie verschiedener Mollienesia-Arten einoschliesslich einer Hohlenform von M. sphenops. Behaviour 33:1–37. Parzefall, J. 1970. Morphologische Untersuchungen an einer Höhlenform von Mollienesia sphenops (Pisces, Poeciliidae). Zeitschrift für Morphologie der Tiere 68:323–342. Parzefall, J. 1973. Attraction and sexual cycle of the Poeciliidae, p. 177–183. In Genetics and Mutagenesis of Fish. J. H. Schröder (ed.). Springer-Verlag, Berlin, Germany. Parzefall, J. 1974. Rückbildung aggressiver Verhaltensweisen bei einer Höhlenform von Mollienesia sphenops (Pisces, Poeciliidae). Zeitschrift für Tierpsychologie 35:66–84. Parzefall, J. 1979. Zur Genetik und biologischen Bedeutung des Aggressionsverhaltens von Poecilia sphenops (Pisces, Poeciliidae). Zeitschrift für Tierpsychologie 50:399–422. Parzefall, J. 1993. Behavioural ecology of cave-dwelling fishes, p. 573–608. In Behaviour of Teleost Fishes. 2nd edition. Pitcher T. J. (ed.). Chapman and Hall, London. Parzefall, J. 2001.A review of morphological and behavioural changes in the cave molly, Poecilia mexicana, from Tabasco, Mexico. Environmental Biology of Fishes 62:263–275. Parzefall, J., U. Gagelmann, and M. Schartl. 1997. Aggressive behaviour and optomotor response in different populations

of Poecilia mexicana (Pisces, Poeciliidae). Mem. Biospeol. 24:63– 69. Pasch, R. W., and C. M. Alexander. 1986. Effects of commercial fishing on Paddlefish populations, p. 46–53. In The Paddlefish: Status, Management and Propagation. J. G. Dillard, L. K. Graham, and T. R. Russell (eds.). American Fisheries Society Special Publication 7. Pasch, R. W., P. A. Hackney, and J. A. Holbrook, II. 1980. Ecology of Paddlefish in Old Hickory Reservoir, Tennessee, with emphasis on first-year life history. Transactions of the American Fisheries Society 109:157–167. Passarelli, N., and A. Piercy. Biological profiles: Atlantic Stingray. Florida Museum of Natural History Ichthyology Department. (July 2009). Available from http://www.flmnh.ufl.edu/fish/Gallery /Descript/AtlanticStingray/AtlanticStingray.html; as of July 2009. Paterson, H. E. H. 1985. The recognition concept of species, p. 21–29. In Species and Speciation. E. S. Vrba (ed.). Transvaal Museum Monograph, No. 4, Pretoria. Paton, K. R., M. H. Cake, and I. C. Potter. 2001. Muscle glycogen, lactate and glycerol-3-phosphate concentrations of larval and young adult Lampreys in response to exercise. Comparative Biochemistry and Physiology 129B:759–766. Patten, B. G., and D. T. Rodman. 1969. Reproductive behavior of northern squawfish, Ptychocheilus oregonensis. Transactions of the American Fisheries Society 98:108–111. Patterson, C. 1973. Interrelationships of holosteans, p. 233–305. In Interrelationships of Fishes. P. H. Greenwood, R. S. Miles, and C. Patterson (eds.). Academic Press, London. Patterson, C. 1981. The development of the North American fish fauna—a problem of historical biogeography, p. 265–281. In The Evolving Biosphere. P. L. Forey (ed.). Cambridge University Press, Cambridge, United Kingdom. Patterson, C. 1982. Morphology and interrelationships of primitive actinopterygian fishes. American Zoologist 22:241–259. Patterson, C., and D. E. Rosen. 1977. Review of the ichthyodectiform and other Mesozoic fishes and the theory and practice of classifying fossils. Bulletin of the American Museum of Natural History 158:81–172. Paukert, C. P., and W. L. Fisher. 2000. Abiotic factors affecting summer distribution and movement of male Paddlefish, Polyodon spathula, in a prairie reservoir. The Southwestern Naturalist 45:133–140. Paukert, C. P., and W. L. Fisher. 2001a. Characteristics of Paddlefish in a southwestern U.S. reservoir, with comparisons of lentic and lotic populations. Transactions of the American Fisheries Society 130:634–643. Paukert, C. P., and W. L. Fisher. 2001b. Spring movements of Paddlefish in a prairie reservoir system. Journal of Freshwater Ecology 16:113–124. Paukert, C. P., and J. M. Long. 1999. New maximum age of Bigmouth Buffalo, Ictiobus cyprinellus. Proceedings of the Oklahoma Academy of Science 79:85–86. Paukert, C. P., and G. D. Scholten (eds.). 2009. Paddlefish Management, Propagation, and Conservation in the 21st century: Building from 20 Years of Research and Management. American Fisheries Society Symposium 66, Bethesda, Maryland. Paukert, C. P., and D. W. Willis. 2002. Seasonal and diel habitat selection by Bluegills in a shallow natural lake. Transactions of the American Fisheries Society 131:1131–1139.

LITERATURE CITED

Paulin, M. G. 1995. Electroreception and the compass sense of Sharks. Journal of Theoretical Biology 174:325–339. Pavao-Zuckerman, B. 2007. Deerskins and domesticates: creek subsistence and economic strategies in the historic period. American Antiquity 72:5–33. Pearson, J. G., and F. P. Ward. 1972. A new record of the Bowfin, Amia calva Linnaeus, in the Upper Chesapeake Bay. Chesapeake Science 13323–324. Pearson, M. P., and M. C. Healey. 2003. Life-history characteristics of the endangered Salish Sucker (Catostomus sp.) and their implications for management. Copeia 2003:759–768. Pearson, W. D., G. A. Thomas, and A. L. Clark. 1979. Early piscivory and timing of the critical period in postlarval Longnose Gar at Mile 571 of the Ohio River. Transactions of the Kentucky Academy of Science 40:122–128. Pearsons, T. N., H. W. Li, and G. A. Lamberti. 1992. Influence of habitat complexity on resistance to flooding and resilience of stream fish assemblages. Transactions of the American Fisheries Society 121:427–436. Peden, A. E. 1972. The function of gonopodial parts and behaviour pattern during copulation by Gambusia (Poeciliidae). Canadian Journal of Zoology 50:955–968. Peebles, E. B. 2002. Temporal resolution of biological and physical influences on Bay Anchovy, Anchoa mitchilli egg abundance near a river-plume frontal zone. Marine Ecology Progress Series 237:257–269. Peebles, E. B., J. R. Hall, and S. G. Tolley. 1996. Egg production by the Bay Anchovy, Anchoa mitchilli in relation to adult and larval prey fields. Marine Ecology Progress Series 131:61–73. Pegg, M. A., J. H. Chick, and B. M. Pracheil. 2009. Potential effects of invasive species on Paddlefish, p. 185–201. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Penczak, T. 2011. Fish assemblages composition in a natural, then regulated, stream: a quantitative long-term study. Ecological Modelling 222:2103–2118. Peng, Z., A. Ludwig, D. Wang, R. Diogo, Q. Wei, and S. He. 2007. Age and biogeography of major clades in Sturgeons and Paddlefishes. Molecular Phylogenetics and Evolution 42:854–862. Pera, T. P., and J. W. Armbruster. 2006. A new species of Notropis (Cypriniformes: Cyprinidae) from the southeastern United States. Copeia 2006:423–430. Percy, Lord R., and I. C. Potter. 1976. Blood cell formation in the River Lamprey, Lampetra fluviatilis. Journal of Zoology, London 178:319–340. Perkins, D. L., J. Kann, and G. G. Scoppettone. 2000. The role of poor water quality and fish kills in the decline of endangered Lost River and Shortnose Suckers in Upper Klamath Lake, Klamath Falls, Oregon. United States Geological Survey Final Report to Bureau of Reclamation Klamath Falls Project Office, Contract 4-AA-29-12160, Klamath Falls, Oregon. Perkins, D. L., and G. G. Scoppettone. 1996. Spawning and migration of Lost River Suckers (Deltistes luxatus) and Shortnose Suckers (Chasmistes brevirostris) in the Clear Lake drainage, Modoc County, California. Final Report to the California Department of Fish and Game, Contract No. FG1494. Perrin, C. J., L. L. Rempel, and M. L. Rosenau. 2003. White Sturgeon spawning habitat in an unregulated river: Fraser River,

589

Canada. Transactions of the American Fisheries Society 132:154–165. Perrin, N. 1995. Signaling, mating success and paternal investment in Sticklebacks (Gasterosteus aculeatus): a theoretical model. Behaviour 132:1037–1053. Perry, S. F. 1997. The chloride cell: structure and function in the gills of freshwater fishes. Annual Review of Physiology 59:325–347. Perry, S. F., R. J. Wilson, C. Straus, M. B. Harris, and J. E. Remmers. 2001. Which came first, the lung or the breath? Comparative Biochemistry and Physiology. Part A. Molecular and Integrated Physiology 129:37–47. Peters, N., W. Schmidt, and D. Fricke. 1990. Fine-structure of the club cells (alarm substance cells) in the epidermis of Astyanax mexicanus (Filippi 1853) (Characinidae, Pisces) and its cave forms Anoptichthys. Internationale Revue der gesamten Hydrobiologie 75:257–267. Petersen, M. E. 1996. Effects of prey growth, physical structure, and piscivore electivity on the relative prey vulnerability of Gizzard shad (Dorosoma cepedianum) and June Sucker (Chasmistes liorus). Unpubl. Master’s thesis, Utah State University, Logan. Peterson, D., P. Vecsei, and D. L. G. Noakes. 2003. Threatened fishes of the world: Acipenser fulvescens Rafinesque, 1817 (Acipenseridae). Environmental Biology of Fishes 68:174. Peterson, M. S. 2003. A conceptual view of environment-habitatproduction linkages in tidal river estuaries. Reviews in Fisheries Science 11:291–313. Peterson, M. S., and M. R. Meador. 1994. Effects of salinity on freshwater fishes in Coastal Plain drainages in the Southeastern U.S. Reviews in Fisheries Science 2:95–121. Peterson, M. S., and L. C. Nicholson. 1997. Relative abundance, life history and habitat characteristics of the Gulf blue sucker, Cycleptus sp. cf. elongatus in the lower Pascagoula and Pearl River systems in Mississippi. Final Report to United States Fish and Wildlife Ser vice, Jackson, Mississippi. Peterson, M. S., L. C. Nicholson, and G. L. Fulling. 2000. Catchper-unit-effort, environmental conditions and spawning migration of Cycleptus meridionalis Burr and Mayden in two coastal rivers of the northern Gulf of Mexico. American Midland Naturalist 143:414–421. Peterson, M. S., L. C. Nicholson, D. J. Snyder, and G. L. Fulling. 1999. Growth, spawning preparedness, and diet of Cycleptus meridionalis (Catostomidae). Transactions of the American Fisheries Society 128:900–908. Peterson, M. S., M. R. Weber, M. L. Partyka, and S. T. Ross. 2007. Integrating in situ quantitative geographic information tools and size-specific laboratory-based growth zones in a dynamic river-mouth estuary. Aquatic Conservation: Marine and Freshwater Ecosystems 17:602–618. Peterson, R. H. (ed.). 1997. The American Eel in eastern Canada: stock status and management strategies. Canadian Technical Reports of Fisheries and Aquatic Science No. 2196. Petrie, M., and B. Kempenaers. 1998. Extra-pair paternity in birds: explaining variation between species and populations. Trends in Ecology and Evolution 13:52–58. Petty, J. T., and G. D. Grossman. 1996. Patch selection by Mottled Sculpin (Pisces: Cottidae) in a southern Appalachian stream. Freshwater Biology 35:261–276. Petty, J. T., and G. D. Grossman. 2010. Giving-up densities and ideal pre-emptive patch use in a predatory benthic stream fish. Freshwater Biology 55:780–793.

590

LITERATURE CITED

Peyer, B. 1968. Comparative Odontology. University of Chicago Press, Chicago, Illinois. Pfeiffer, W. 1960. Uber die schreckreaktion bei fischen und die herkunft des schreckstoffes. Zeitschrift für Vergleichende Physiologie 43:578–614. Pfeiffer, W. 1962. The fright reaction of fish. Biological Reviews 37:495–511. Pfeiffer, W. 1963a. Alarm substances. Experientia 14:113–123. Pfeiffer, W. 1963b. The fright reaction in North American fish. Canadian Journal of Zoology 41:69–77. Pfeiffer, W. 1977. The distribution of fright reaction and alarm substance cells in fishes. Copeia 1977:653–655. Pfeiffer, W., U. Walz, R. Wolf, and U. Mangold-Wernado. 1985. Effects of steroid hormones and other substances on alarm substance cells and mucous cells in the epidermis of the European minnow, Phoxinus phoxinus (L.), and other Ostariophysi (Pisces). Journal of Fish Biology 27:553–570. Pfeiler E. 1986. Towards an explanation of the developmental strategy in leptocephalus larvae of marine fishes. Environmental Biology of Fishes 15:3–13. Pfennig, K. S., and D. W. Pfennig. 2009. Character displacement: ecological and reproductive responses to a common evolutionary problem. The Quarterly Review of Biology 84:253–276. Pflieger, W. L. 1965. Reproductive behavior of the minnows, Notropis spilopterus and Notropis whipplei. Copeia 1965:1–7. Pflieger, W. L. 1966. Young of Orangethroat Darter (Etheostoma spectabile) in nests of Smallmouth Bass (Micropterus dolomieu). Copeia 1966:139. Pflieger, W. L. 1971. A distributional study of Missouri fishes. Museum of Natural History, University of Kansas 20:225–570. Pflieger, W. L. 1975. The Fishes of Missouri. Missouri Department of Conservation, Jefferson City. Pflieger, W. L., and T. B. Grace. 1987. Changes in the fish fauna of the lower Missouri River, 1940–1983, p. 166–177. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Pflieger, W. L. 1997. The Fishes of Missouri. Missouri Department of Conservation, Jefferson City. Pfrender, M. E., J. Hicks, and M. Lynch. 2004. Biogeographic patterns and current distribution of molecular-genetic variation among populations of Speckled Dace, Rhinichthys osculus (Girard). Molecular Phylogenetics and Evolution 30:490–502. Phelps, Q. E., S. J. Tripp, J. E. Garvey, D. P. Herzog, D. E. Ostendorf, J. W. Ridings, J. W. Crites, and R. A. Hrabik. 2009. Ecology and habitat use of age-0 Paddlefish in the unimpounded middle Mississippi River, p. 423–440. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Phelps, Q. E., S. J. Tripp, J. E. Garvey, D. P. Herzog, D. E. Ostendorf, J. W. Ridings, J. W. Crites, and R. A. Hrabik. 2010a. Habitat use during early life history infers recovery needs for Shovelnose Sturgeon and Pallid Sturgeon in the Middle Mississippi River. Transactions of the American Fisheries Society 139:1060–1068. Phelps, Q. E., S. J. Tripp, W. D. Hintz, J. E. Garvey, D. P. Herzog, D. E. Ostendorf, J. W. Ridings, J. W. Crites, and R. A. Hrabik. 2010b. Water temperature and river stage influence mortality and abundance of naturally occurring Mississippi River Scaphi-

rhynchus Sturgeon. North American Journal of Fisheries Management 30:767–775. Philipp, D. P., and M. R. Gross. 1994. Genetic evidence for cuckoldry in Bluegill, Lepomis macrochirus. Molecular Evolution 3:563–569. Phillips, C. T., and C. E. Johnston. 2008a. Sound production and associated behaviors in Cyprinella galactura. Environmental Biology of Fishes 82:265–275. Phillips, C. T., and C. E. Johnston. 2008b. Geographical divergence of acoustic signals in Cyprinella galactura, the Whitetail Shiner (Cyprinidae). Animal Behaviour 75:617–626. Phillips, C. T., and C. E. Johnston. 2009. Evolution of acoustic signals in Cyprinella: degree of similarity in sister species. Journal of Fish Biology 74:120–132. Phillips, C. T., J. R. Gibson, and J. N. Fries. 2009. Agonistic and courtship behaviors in Dionda diaboli, the Devils River Minnow. The Southwestern Naturalist 54:341–368. Phillips, C. T., C. E. Johnston, and A. R. Henderson. 2010. Sound production and spawning behavior in Cyprinella lepida, the Edwards Plateau Shiner. The Southwestern Naturalist 55:129–135. Piavis, G. W. 1971. Embryology, p. 361–400. In The Biology of Lampreys. Vol. 1. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Piché, J, J. A. Hutchings, and W. Blanchard. 2008. Genetic variation and threshold reaction norms for alternative reproductive tactics in male Atlantic Salmon, Salmo salar. Proceedings of the Royal Society of London B 275:1571–1575. Pielou, E. C. 1991. After the Ice Age, the Return of Life to Glaciated North America. University of Chicago Press, Illinois. Piercy, A., J. Gelsleichter, and F. F. Snelson. 2003. Morphological and histological changes in the genital ducts of the male Atlantic Stingray, Dasyatis sabina, associated with the seasonal reproductive cycle. Fish Physiology and Biochemistry 29:23–35. Piercy, A., J. Gelsleichter, and F. F. Snelson. 2006a. Morphological changes in the clasper gland of the Atlantic Stingray, Dasyatis sabina, associated with the seasonal reproductive cycle. Journal of Morphology 267:109–114. Piercy, A., F. F. Snelson, Jr., and R. D. Grubbs. 2006b. Dasyatis sabina. In IUCN 2011. IUCN Red List of Threatened Species. Version 2011.2. Available from http://www.iucnredlist.org; accessed March 2012. Piermarini, P. M., and D. H. Evans. 1998. Osmoregulation of the Atlantic Stingray (Dasyatis sabina) from the freshwater Lake Jesup of the St. Johns River, Florida. Physiological Zoology 71:553–560. Piermarini, P. M., and D. H. Evans. 2000. Effects of environmental salinity on Na+/K+ -ATPase in the gills and rectal gland of a euryhaline elasmobranch (Dasyatis sabina). Journal of Experimental Biology 203:2957–2966. Piermarini, P. M., and D. H. Evans. 2001. Immunochemical analysis of the vacuolar proton-ATPase B-subunit in the gills of a euryhaline Stingray (Dasyatis sabina): effects of salinity and relation to Na(+)/K(+)-ATPase. Journal of Experimental Biology 204:3251–3259. Piermarini P. M., J. W. Verlander, I. E. Royaux, and D. H. Evans. 2002. Pendrin immunoreactivity in the gill epithelium of a euryhaline elasmobranch. American Journal of Physiology: Regulatory, Integrative and Comparative Physiology 283:R983–R992.

LITERATURE CITED

Pigg, J., and R. Gibbs. 1996. Observations on the propagation of two rare fish species in Oklahoma. Proceeding of the Oklahoma Academy of Science 76:89. Pikitch, E. K., P. Doukakis, L. Lauck, P. Chakrabarty, and D. Erickson. 2005. Status, trends and management of Sturgeon and Paddlefish fisheries. Fish and Fisheries 6:233–265. Pilastro, A., S. Benetton, and A. Bisazza. 2003. Female aggregation and male competition reduce costs of sexual harassment in the mosquitofish, Gambusia holbrooki. Animal Behaviour 65:1161–1167. Pilastro, A., E. Giacomello, and A. Bisazza. 1997. Sexual selection for small male size in male mosquitofish (Gambusia holbrooki). Proceedings of the Royal Society of London B 264:1125–1129. Piller, K. R., H. L. Bart, Jr., and J. A. Tipton. 2003. Spawning in the Black Buffalo, Ictiobus niger (Cypriniformes: Catostomidae). Ichthyological Explorations of Freshwaters 14:145–150. Piller, K. R., H. L. Bart, Jr., and C. A. Walser. 2001. Morphological variation of the Redfin Darter, Etheostoma whipplei, with comments on the status of the subspecific populations. Copeia 2001:802–807. Piller, K. R., and B. M. Burr. 1999. Reproductive biology and spawning habitat supplementation of the Relict Darter, Etheostoma chienense, a federally endangered species. Environmental Biology of Fishes 55:145–155. Pine, W. E., III, M. S. Allen, and V. J. Dreitz. 2001. Population viability of the Gulf of Mexico Sturgeon: inferences from capturerecapture and age-structured models. Transactions of the American Fisheries Society 130:1164–1174. Pisarowicz, J. 2005. Return to Tabasco. Association of Mexican Cave Studies Newsletter 28:27–57. Pitcher, T. J. 1986. Functions of shoaling in teleosts, p. 294–337. In The Behavior of Teleost Fishes. J. T. Pitcher (ed.). John Hopkins University Press, Baltimore, Maryland. Pitman, V. M., and J. O. Parks. 1994. Habitat use and movement of young Paddlefish (Polyodon spathula). Journal of Freshwater Ecology 9:181–189. Platania, S. P., and C. S. Altenbach. 1988. Reproductive strategies and egg types of seven Rio Grande Basin cyprinids. Copeia 1998:559–569. Plath, M. 2008. Male mating behavior and costs of sexual harassment for females in cavernicolous and extremophile populations of Atlantic mollies (Poecilia mexicana). Behaviour 145:73–98. Plath, M., A. Brunner, and I. Schlupp. 2004a. Sexual harassment in a livebearing fish. (Poecilia mexicana): influence of population-specific male mating behaviour. Acta Ethologica 7:65–72. Plath, M., K. U. Heubel, F. J. García de León, and I. Schlupp. 2005. Cave molly females (Poecilia mexicana, Poeciliidae, Teleostei) like well-fed males. Behavioral Ecology and Sociobiology 58:144–151. Plath, M., K. E. Körner, J. Parzefall, and I. Schlupp. 2003c. Persistence of a visually mediated mating preference in the cave molly, Poecilia mexicana (Poeciliidae, Teleostei). Subterranean Biology 1:93–97. Plath, M., K. E. Körner, I. Schlupp, and J. Parzefall. 2001. Sex recognition and female preferences of cave mollies, Poecilia mexicana (Poeciliidae, Teleostei) in light and darkness. Memoires de Biospeologie 28:163–167. Plath, M., A. M. Makowicz, I. Schlupp, and M. Tobler. 2007a. Sexual harassment in live-bearing fishes (Poeciliidae): comparing

591

courting and noncourting species. Behavioral Ecology 18:680–688. Plath, M., J. Parzefall, K. E. Körner, and I. Schlupp. 2004b. Sexual selection in darkness? Female mating preferences in surface- and cave-dwelling Atlantic mollies, Poecilia mexicana (Poeciliidae, Teleostei). Behavioral Ecology and Sociobiology 55:596–601. Plath, M., J. Parzefall, and I. Schlupp. 2003a. The role of sexual harassment in cave and surface dwelling populations of the Atlantic molly, Poecilia mexicana (Poeciliidae, Teleostei). Behavioral Ecology and Sociobiology 54:303–309. Plath, M., I. Schlupp, J. Parzefall, and R. Riesch. 2007b. Female preference for large body size in the cave molly, Poecilia mexicana (Poeciliidae, Teleostei): influence of species- and sexspecific cues. Behaviour 144:1147–1160. Plath, M., M. Tobler, R. Riesch, F. J. García de León, O. Giere, and I. Schlupp. 2007c. Survival in an extreme habitat: the roles of behaviour and energy limitation. Naturwissenschaften 94:991–996. Plath, M., K. Wiedemann, I. Schlupp, and J. Parzefall. 2003b. Sex recognition in surface and cave dwelling male Atlantic mollies, Poecilia mexicana (Poeciliidae, Teleostei). Behaviour 140:765–781. Poff, N. L. 1997. Landscape filters and species traits: towards a mechanistic understanding and prediction in stream ecology. Journal of the North American Benthological Society 16:391–409. Poff, N. L., and J. D. Allan. 1995. Functional organization of stream fish assemblages in relation to hydrological variability. Ecology 76:606–627. Poff, N. L., J. D. Allan, M. B. Bain, J. R. Karr, K. L. Prestegaard, B. D. Richter, R. E. Sparks, J. C. Stromberg. 1997. The natural flow regime. BioScience 74:769–784. Polis, G. A., C. A. Myers, and R. D. Holt. 1989. The ecology and evolution of intraguild competitors that eat each other. Annual Review of Ecology and Systematics 20:297–330. Pollock, M. S., and D. P. Chivers. 2003. Does habitat complexity influence the ability of Fathead Minnows to learn heterospecific chemical alarm cues? Canadian Journal of Zoology 81:923–927. Pollock, M. S., and D. P. Chivers. 2004. The effects of density on the learned recognition of heterospecific alarm cues. Ethology 110:341–349. Pollock, M. S., D. P. Chivers, R. S. Mirza, and B. D. Wisenden. 2003. Fathead Minnows, Pimephales promelas, learn to recognize chemical alarm cues of introduced Brook Stickleback, Culaea inconstans. Environmental Biology of Fishes 66:313–319. Pollock, M. S., R. J. Pollock, and D. P. Chivers. 2006. Social context influences the antipredator behaviour of Fathead Minnows to chemical alarm cues. Ethology 112:801–806. Pollock, M. M., M. Heim, and D. Werner. 2003. Hydrologic and geomorphic effects of Beaver dams and their influence on fishes, p. 213–233. In The Ecology and Management of Wood in World Rivers. S. Gregory, K. Boyer, and A. Gurnell (eds.). American Fisheries Society Symposium 37, Bethesda, Maryland. Poly, W. J., and M. H. Sabaj. 1998. Lack of evidence for the validity of Rhinichthys bowersi (Cyprinidae). Copeia 1998:1081–1085. Pope, K., and G. Wilde. 2003. Variation in Spotted Gar (Lepisosteus oculatus) mass-length relationships in Texas reservoirs. The Texas Journal of Science 55:43–49. Popiel, S. A., A. Pérez-Fuentetaja, D. J. McQueen, and N. C. Collins. 1996. Determinants of nesting success in the Pumpkinseed (Lepomis gibbosus): a comparison of two populations under different risks from predation. Copeia 1996:649–656.

592 LITERATURE CITED

Poplin, C., D. Sotty, and P. Janvier. 2001. A Hagfish (Craniata, Hyperotreti) from the Late Carboniferous Konservat-Lagerstätte of Montceau-les-Mines (Allier, France). Comptes Rendus de l’Académie des Sciences, Earth and Planetary Sciences 332:345–350. Poppe, L. J., H. J. Knebel, Z. J. Mlodzinska, M. E. Hastings, and B. A. Seekins. 2000. Distribution of surficial sediment in Long Island Sound and adjacent waters: texture and total organic carbon. Journal of Coastal Research 16:567–574. Popper, A. N., and R. G. Northcutt. 1983. Structure and innervation of the inner ear of the Bowfin, Amia calva. The Journal of Comparative Neurology 213:279–286. Porter, B. A., A. C. Fiumera, and J. C. Avise. 2002. Egg mimicry and allopaternal care: two mate-attracting tactics by which nesting Striped Darter (Etheostoma virgatum) males enhance reproductive success. Behavioral Ecology and Sociobiology 51:350–359. Porter, H. T., and P. J. Motta. 2004. A comparison of strike and prey capture kinematics of the jaws of three species of piscivorous fishes: Florida Gar (Lepisosteus platyrhincus), Redfin Needlefish (Strongylura notate), and Great Barracuda (Sphyraena barracuda). Marine Biology 145:989–1000. Portz, E. E., and H. M. Tyus. 2004. Fish humps in two Colorado River fishes: a morphological response to cyprinid predation. Environmental Biology of Fishes 71:233–245. Post, J. R., and D. J. McQueen. 1988. Ontogenetic changes in the distribution of larval and juvenile Yellow Perch (Perca flavescens): a response to prey or predators? Canadian Journal of Fisheries and Aquatic Sciences 45:1820–1826. Potter, G. E. 1923. Food of the short-nose gar-pike (Lepidosteus platystomus) in Lake Okoboji, Iowa. Proceedings of the Iowa Academy of Science 30:167–170. Potter, G. E. 1927. Respiratory function of the swim bladder in Lepidosteus. Journal of Experimental Zoology 49:45–67. Potter, I. C. 1970. The life cycles and ecology of Australian Lampreys of the genus Mordacia. Journal of Zoology, London 161:487–511. Potter, I. C. 1980a. Ecology of larval and metamorphosing Lampreys. Canadian Journal of Fisheries and Aquatic Sciences 37:1641–1657. Potter, I. C. 1980b. The Petromyzoniformes with particular reference to paired species. Canadian Journal of Fisheries and Aquatic Sciences 37:1595–1615. Potter, I. C., and J. R. Bailey. 1972. The life cycle of the Tennessee brook Lamprey, Ichthyomyzon hubbsi Raney. Copeia 1972:470–476. Potter, I. C., and F. W. H. Beamish. 1975. Lethal temperatures in ammocoetes of four species of Lampreys. Acta Zoologica 56:85–91. Potter, I. C., and F. W. H. Beamish. 1977. The freshwater biology of adult anadromous Sea Lampreys, Petromyzon marinus. Journal of Zoology, London 181:113–130. Potter, I. C., and F. W. H. Beamish. 1978. Changes in haematocrit and haemoglobin concentration during the life cycle of the anadromous Sea Lamprey, Petromyzon marinus L. Comparative Biochemistry and Physiology 60A:431–434. Potter, I. C., and H. S. Gill. 2003. Adaptive radiation of Lampreys. Journal of Great Lakes Research 29(Suppl. 1):95–112. Potter, I. C., B. J. Hill, and S. Gentleman. 1970. Survival and behaviour of ammocoetes at low oxygen tensions. The Journal of Experimental Biology 53:59–73.

Potter, I. C., and R. W. Hilliard. 1986. Growth and the average duration of larval life in the Southern Hemisphere Lamprey, Geotria australis Gray. Experientia 42:1170–1173. Potter, I. C., and R. W. Hilliard. 1987. A proposal for the functional and phylogenetic significance of differences in the dentition of Lampreys (Agnatha: Petromyzontiformes). Journal of Zoology, London 212:713–737. Potter, I. C., R. W. Hilliard, D. J. Bird, and D. J. Macey. 1983. Quantitative data on morphology and organ weights during the protracted spawning-run period of the Southern Hemisphere Lamprey, Geotria australis. Journal of Zoology, London 200:1–20. Potter, I. C., W. J. R. Lanzing, and R. Strahan. 1968a. Morphometric and meristic studies on populations of Australian Lampreys of the genus Mordacia. Journal of the Linnean Society of London, Zoology 47:533–546. Potter, I. C., D. J. Macey, and A. R. Roberts. 1997. Oxygen uptake and carbon dioxide excretion by the branchial and postbranchial regions of adults of the Lamprey, Geotria australis in air. The Journal of Experimental Zoology 278:290–298. Potter, I. C., D. J. Macey, A. R. Roberts, and P. C. Withers. 1996. Oxygen consumption by ammocoetes of the Lamprey, Geotria australis in air. Journal of Comparative Physiology B 166:331–336. Potter, I. C., and P. I. Nicol. 1968. Electrophoretic studies on the haemoglobins of Australian Lampreys. Australian Journal of Experimental Biology and Medical Science 46:639–641. Potter, I. C., P. A. Prince, and J. P. Croxall. 1979. Data on the adult marine and migratory phases in the life cycle of the Southern Hemisphere Lamprey, Geotria australis Gray. Environmental Biology of Fishes 4:65–69. Potter, I. C., and E. S. Robinson. 1981. New developments in vertebrate cytotaxonomy. V. Cytotaxonomy of Lampreys. Genetica 56:149–151. Potter, I. C., E. S. Robinson, and S. M. Walton. 1968b. The mitotic chromosomes of the Lamprey Mordacia mordax (Agnatha: Petromyzonidae). Experientia 24:966–967. Potter, I. C., and M. J. Rogers. 1972. Oxygen consumption in burrowed and unburrowed ammocoetes of Lampetra planeri (Bloch). Comparative Biochemistry and Physiology 41A:427–432. Potter, I. C., G. M. Wright, and J. H. Youson. 1978. Metamorphosis in the anadromous Sea Lamprey, Petromyzon marinus L. Canadian Journal of Zoology 56:561–570. Pottin, K., C. Hyacinthe, and S. Rétaux. 2010. Conservation, development, and function of a cement gland-like structure in the fish Astyanax mexicanus. Proceedings of the National Academy of Science 107:17256–17261. Potts, W. K., C. J. Manning, and E. K. Wakeland. 1991. Mating patterns in seminatural populations of mice influenced by MHC genotype. Nature 352:619–621. Poulin, R., D. J. Marcogliese, and J. D. McLaughlin. 1999. Skinpenetrating parasites and the release of alarm substances in juvenile Rainbow Trout. Journal of Fish Biology 55:47–53. Power, M., and R. S. McKinley. 1997. Latitudinal variation in Lake Sturgeon size as related to the thermal opportunity for growth. Transactions of the American Fisheries Society 126:549–558. Power, M. E. 1987. Predator avoidance by grazing fishes in temperate and tropical streams: importance of stream depth and prey size, p. 333–351. In Predation: Direct and Indirect Impacts on

LITERATURE CITED

Aquatic Communities. W. C. Kerfoot and A. Sih (eds.). University Press of New England, Hanover, New Hampshire. Power, M. E. 1990. Effects of fish in river food webs. Science 250:811–814. Power, M. E., and W. J. Matthews. 1983. Algae-grazing minnows (Campostoma anomalum), piscivorous bass (Micropterus spp.), and the distribution of attached algae in a small prairie-margin stream. Oecologia 60:328–332. Power, M. E., W. J. Matthews, and A. J. Stewart. 1985. Grazing minnows, piscivorous bass, and stream algae: dynamics of a strong interaction. Ecology 66:328–332. Power, M. E., M. S. Parker, and W. E. Dietrich. 2008. Seasonal reassembly of a river food web: floods, droughts, and impacts of fish. Ecological Monographs 78:263–282. Pracheil, B. M., G. E. Mestl, and P. M. Muzzall. 2005. Metazoan parasites of young-of-the-year Paddlefish from Lewis and Clark Lake. Comparative Parasitology 72:227–229. Pramuk, J. B., M. J. Grose, A. L. Clarke, E. Greenbaum, E. Bonaccorso, J. M. Guayasamin, A. H. Smith-Pardo, B. W. Benz, B. R. Harris, E. Siegfreid, Y. R. Reid, N. Holcroft-Benson, and E. O. Wiley. 2007. Phylogeny of finescale shiners of the genus Lythrurus (Cypriniformes: Cyprinidae) inferred from four mitochondrial genes. Molecular Phylogenetics and Evolution 42:287–297. Prejs, A. 1984. Herbivory by temperate freshwater fishes and its consequences. Environmental Biology of Fishes 10:281–296. Pressley, P. H. 1981. Parental effort and the evolution of nestguarding tactics in the Threespine Stickleback, Gasterosteus aculeatus L. Evolution 35:282–295. Price, C. J., W. M. Tonn, and C. A. Paszkowski. 1991. Intraspecific patterns of resource use by Fathead Minnow in a small boreal lake. Canadian Journal of Zoology 69:2109–2115. Price, P. W. 1984. Communities of specialists: vacant niches in ecological and evolutionary time, p. 510–523. In Ecological Communities—Conceptual Issues and the Evidence. D. R. Strong, Jr., D. Simberloff, L. G. Abele, and A. B. Thistle (eds.). Princeton University Press, Princeton, New Jersey. Probst, R. T., and E. L. Cooper. 1955. Age, growth, and production of the Lake Sturgeon (Acipenser fulvescens) in the Lake Winnebago region, Wisconsin. Transactions of the American Fisheries Society 84:207–227. Prowse, G. A. 1961. The use of fertilizers in fish culture. Proceedings of the Indo-Pacific fisheries council 9:73–75. Psenicka, M., V. Kaspar, S. M. H. Alavi, M. Rodina, D. Gela, P. Li, S. Borishpolets, J. Cosson, O. Linhart, and A. Ciereszko. 2011. Potential role of the acrosome of Sturgeon spermatozoa in the fertilization process. Journal of Applied Ichthyology 27:678–682. Ptacek, M. B., M. J. Childress, and M. M. Kittell. 2005. Characterizing the mating behaviors of the Tamesi Molly, Poecilia latipunctata, a sailfin with shortfin morphology. Animal Behaviour 70:1339–1348. Ptacek, M. B., and J. Travis. 1996. Interpopulation variation in male mating behaviour in the Sailfin Molly, Poecilia latipinna. Animal Behaviour 52:59–71. Ptacek, M. B., and J. Travis. 1997. Male choice in the Sailfin Molly, Poecilia latipinna. Evolution 51:1217–1231. Pullen, R. R., W. W. Bouska, S. W. Campbell, and C. P. Paukert. 2009. Bothriocephalus acheilognathi and other intestinal helminths of Cyprinella lutrensis in Deep Creek, Kansas. Journal of Parasitology 95:1224–1226.

593

Purkett, C. A., Jr. 1961. Reproduction and early development of the Paddlefish. Transactions of the American Fisheries Society 90:125–129. Purkett, C. A., Jr. 1963a. The Paddlefish fishery of the Osage River and the Lake of the Ozarks, Missouri. Transactions of the American Fisheries Society 92:239–244. Purkett, C. A., Jr. 1963b. Artificial propagation of the Paddlefish. The Progressive Fish-Culturist 25:31–33. Purvis, H. A. 1970. Growth, age at metamorphosis and sex ratio of Northern Brook Lamprey in a tributary of southern Lake Superior. Copeia 1970:326–332. Putnam, N. H., T. Butts, D. E. K. Ferrier, R. F. Furlong, U. Hellsten, T. Kawashima, M. Robinson-Rechavi, E. Shoguchi, A. Terry, J.-K.Yu, È. Benito-Gutiérrez, I. Dubchak, J. GarciaFernàndez, J. J. Gibson-Brown, I. V. Grigoriev, A. C. Horton, P. J. de Jong, J. Jurka, V. V. Kapitonov, Y. Kohara, Y. Kuroki, E. Lindquist, S. Lucas, K. Osoegawa, L. A. Pennacchio, A. A. Salamov, Y. Satou, T. Sauka-Spengler, J. Schmutz, T. ShinI, A. Toyoda, M. Bronner-Fraser, A. Fujiyama, L. Z. Holland, P. W. H. Holland, N. Satoh, and D. S. Rokhsar. 2008. The amphioxus genome and the evolution of the chordate karyotype. Nature 453:1064–1071. Puzdrowski, R. L., and R. B. Leonard. 1993. The octolaval systems in the Stingray, Dasyatis sabina. I. Primary projections of the octaval and lateral line nerves. Journal of Comparative Neurology 332:21–37. Puzdrowski, R. L., and R. B. Leonard. 1994. Vestibulo-oculomotor connections in an elasmobranch fish, the Atlantic Stingray, Dasyatis sabina. Journal of Comparative Neurology 339:587–597. Pyron, M. 1995. Mating patterns and a test for female mate choice in Etheostoma spectabile (Pisces: Percidae). Behavioral Ecology and Sociobiology 36:407–412. Pyron, M. 1996. Sexual size dimorphism and phylogeny in North American minnows. Biological Journal of the Linnean Society 57:327–341. Pyron, M. 1999. Relationships between geographical range size, body size, local abundance, and habitat breadth in North American Suckers and Sunfishes. Journal of Biogeography 26:549–558. Pyron, M. 2000. Testes mass and reproductive mode of minnows. Behavioral Ecology and Sociobiology 48:132–136. Pyron, M., T. E. Lauer, and J. R. Gammon. 2006. Stability of the Wabash River fish assemblages from 1974 to 1998. Freshwater Biology 51:1789–1797. Quaggio-Grassiotto, I., J. N. C. Negrão, E. D. Carvalho, and E. Foresti. 2001. Ultrastructure of spermatogenic cells and spermatozoa in Hoplias malabaricus (Teleostei, Characiformes, Erythrinidae). Journal of Fish Biology 59:1494–1502. Quartarone, F. 1995. Historical accounts of upper Colorado River Basin endangered fish. Report produced for the Information and Education Committee of the Recovery Program for Endangered Fish of the Upper Colorado River Basin. C. Young (ed.). Upper Colorado River Endangered Fish Recovery Program, United States Fish and Wildlife Ser vice, Denver, Colorado. Quattro, J. M., T. W. Greig, D. K. Coykendall, B. W. Bowen, and J. D. Baldwin. 2002. Genetic issues in aquatic species management: the Shortnose Sturgeon (Acipenser brevirostrum) in the southeastern United States. Conservation Genetics 3:155–166.

594 LITERATURE CITED

Quinn, J. W. 2009. Harvest of Paddlefish in North America, p. 203–221. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Quinn, J. W., W. R. Posey, II, F. J. Leone, and R. L. Limbird. 2009. Management of the Arkansas River commercial Paddlefish fishery with check stations and special seasons, p. 261–275. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Quinn, S. P., and M. R. Ross. 1985. Non-annual spawning in the White Sucker, Catostomus commersoni. Copeia 1985:613–618. Quinn, T. P., M. D. Adkison, and M. B. Ward. 1996. Behavioral tactics of male Sockeye Salmon (Oncorhynchus nerka) under varying operational sex ratios. Ethology 102:304–322. Quinn, T. P., and C. A. Busack. 1985. Chemosensory recognition of siblings in juvenile Coho Salmon (Oncorhynchus kisutch). Animal Behaviour 33:51–56. Quinn, T. P., and C. J. Foote. 1994. The effects of body size and sexual dimorphism on the reproductive behaviour of Sockeye Salmon, Oncorhynchus nerka. Animal Behaviour 48:751–761. Quinn, T. P., and G. M. Tolson. 1986. Evidence of chemically mediated population recognition in Coho Salmon (Oncorhynchus kisutch). Canadian Journal of Zoology 64:84–87. Quintella, B. R., C. S. Mateus, J. L. Costa, I. Domingos, and P. R. Almeida. 2010. Critical swimming speed of yellow- and silverphase European Eel (Anguilla anguilla, L.). Journal of Applied Ichthyology 26:432–435. Quintella, B. R., I. Póvoa, and P. R. Almeida. 2009. Swimming behaviour of upriver migrating Sea Lamprey assessed by electromyogram telemetry. Journal of Applied Ichthyology 25:46–54. Quist, M. C., M. R. Bower, W. A. Hubert, T. L. Parchman, and D. B. McDonald. 2009. Morphometric and meristic differences among Bluehead Suckers, Flannelmouth Suckers, White Suckers, and their hybrids: tools for the management of native species in the Upper Colorado River Basin. North American Journal of Fisheries Management 29:460–467. Quist, M. C., C. S. Guy, M. A. Pegg, P. J. Braaten, C. L. Pierce, and V. H. Travnichek. 2002. Potential influence of harvest on Shovelnose Sturgeon populations in the Missouri River system. North American Journal of Fisheries Management 22:537–549. Quist, M. C., J. S. Tillma, M. N. Burlingame, and C. S. Guy. 1999. Overwintering habitat use of Shovelnose Sturgeon in the Kansas River. Transactions of the American Fisheries Society 128:522–527. Rab, P., M. Rabova, K. M. Reed, and R. B. Phillips. 1999. Chromosomal characteristics of ribosomal DNA in the primitive semionotiform fish, Longnose Gar Lepisosteus osseus. Chromosome Research 7:475–480. Rafferty, N. E., and J. W. Boughman. 2006. Olfactory mate recognition in a sympatric species pair of three-spined Sticklebacks. Behavioural Ecology 17:965–970. Rafinesque, C. S. 1818. Description of three new genera of fluviatile fish, Pomoxis, Sarchirus and Exoglossum. Journal of the Academy of Natural Sciences of Philadelphia 1:417–422. Rafinesque, C. S. 1819. Prodrome de 70 nouveaux genres d’animaux découverts dans l’intérieur des États-Unis

d’Amérique, durant l’année 1818. Journal de Physique, de Chimie et d’Histoire Naturelle 88:417–429. Rafinesque, C. S. 1820a. Ichthyologia Ohiensis, or natural history of the fishes inhabiting the River Ohio and its tributary streams, preceded by a physical description of the Ohio and its branches. Reprinted in 1899 by The Burrows Brothers Co., Cleveland, Ohio. Rafinesque, C. S. 1820b. Ichthyologia Ohiensis [Part 6]. Western Revue and Miscellaneous Magazine 2:299–307. Rahel, F. J. 1984. Factors structuring fish assemblages along a bog lake successional gradient. Ecology 65:1276–1289. Rahel, F. J. 1990. The hierarchical nature of community persistence: a problem of scale. The American Naturalist 136:328–344. Rahel, F. J. 2000. Homogenization of fish faunas across the United States. Science 288:854–856. Rahel, F. J. 2002. Homogenization of freshwater faunas. Annual Review of Ecology and Systematics 33:291–315. Rahel, F. J. 2010. Homogenization, differentiation, and the widespread alteration of fish faunas, p. 311–326. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. Rahel, F. J., J. D. Lyons, and P. A. Cochran. 1984. Stochastic or deterministic regulation of assemblage structure: it may depend on how the assemblage is defined. The American Naturalist 124:583–589. Rahn, H. K., B. Rahn, B. J. Howell, C. Gans, and S. M. Tenney. 1971. Air breathing of the Garfish (Lepisosteus ossseus). Respiration Physiology 11:285–307. Raikova, E. V. 2002. Polypodium hydriforme infection in the eggs of acipenseriform fishes. Journal of Applied Ichthyology 18:405–415. Railsback, S. F., and B. C. Harvey. 2002. Analysis of habitat-selection rules using an individual-based model. Ecology 83:1817–1830. Rajakaruna, R. S., J. A. Brown, K. H. Kaukinen, and K. M. Miller. 2006. Major histocompatibility complex and kin discrimination in Atlantic Salmon and Brook Trout. Molecular Ecology 15:4569–4575. Rakes, P. L., J. R. Shute, and P. W. Shute. 1999. Reproductive behavior, captive breeding, and restoration ecology of endangered fishes. Environmental Biology of Fishes 55:31–42. Raley, M. E., and R. M. Wood. 2001. Molecular systematics of members of the Notropis dorsalis species group (Actinopterygii: Cyprinidae). Copeia 2001:638–645. Ramaswami, L. S. 1957. Skeleton of cyprinoid fishes in relation to phylogenetic studies. 8. The skull and Weberian ossicles of the Catostomidae. Proceedings of the Zoological Society (Calcutta) (Mookerjee Memorial Volume):293–303. Ramos, K. T., L. T. Fries, C. S. Berk house, and J. N. Fries. 1994. Apparent sunburn of juvenile Paddlefish. The Progressive FishCulturist 56:214–216. Randall, D. J. 1972. Respiration, p. 287–306. In The Biology of Lampreys Vol. 2. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Randall, D. J., J. N. Cameron, C. Daxobeck, and N. Smatresk. 1981. Aspects of bimodal gas exchange in the Bowfin, Amia calva L. (Actinopterygii: Amiiformes). Respiration Physiology 43:339–348. Randall, D. J., W. W. Burggren, A. P. Farrell, and M. S. Haswell. 1981. The evolution of air breathing vertebrates. Cambridge University Press, New York.

LITERATURE CITED

Raney, E. C. 1939a. The breeding habits of the silvery minnow, Hybognathus regius Girard. The American Midland Naturalist 21:674–680. Raney, E. C. 1939b. Observations on the nesting habits of Parexoglossum laurae Hubbs and Trautman. Copeia 1939:112–113. Raney, E. C. 1940. Rhinichthys bowersi from West Virginia a hybrid, Rhinichthys cataractae × Nocomis micropogon. Copeia 1940:270–271. Raney, E. C. 1942. Alligator Gar feeds upon birds in Texas. Copeia 1942:50. Raney, E. C. 1943. Unusual spawning habitat for the common White Sucker, Catostomus c. commersonii. Copeia 1943:256. Raney, E. C. 1947. Nocomis nests used by other breeding cyprinid fishes in Virginia. Zoologica 32:125–132. Raney, E. C., and E. A. Lachner. 1946a. Age and growth of the Rustyside Sucker, Thoburnia rhothoeca (Thoburn). American Midland Naturalist 36:675–681. Raney, E. C., and E. A. Lachner. 1946b. Age, growth and habits of the Hog Sucker, Hypentelium nigricans (Le Sueur), in New York. American Midland Naturalist 36:76–86. Raney, E. C., and E. A. Lachner. 1946c. Thoburnia hamiltoni, a new Sucker from the Upper Roanoke river system in Virginia. Copeia 1946:218–226. Raney, E. C., and D. A. Webster. 1942. The spring migration of the common White Sucker, Catostomus c. commersonii (La Cepede), in Skaneateles Lake Inlet, New York. Copeia 1942:139–148. Rapp, T., D. A. Shuman, B. D. S. Graeb, S. R. Chipps, and E. J. Peters. 2011. Diet composition and feeding patterns of adult Shovelnose Sturgeon (Scaphirhynchus platorynchus) in the lower Platte River, Nebraska, USA. Journal of Applied Ichthyology 27:351–355. Rasmussen, A.-S., A. Janke, and U. Arnason. 1998. The mitochondrial DNA molecule of the Hagfish (Myxine glutinosa) and vertebrate phylogeny. Journal of Molecular Evolution 46:382–388. Rasmussen, J. L. 1999. What is MICRA and why does it work?, p. 218–229. In Proceedings of the Symposium on the Harvest and Conservation of North American Paddlefish and Sturgeon, May 7–8, 1998, Chattanooga, Tennessee. D. F. Williamson, G. W. Benz, and C. M. Hoover (eds.). TRAFFIC North America / World Wildlife Fund, Washington, D.C. Rauchenberger, M., K. D. Kallman, and D. C. Morizot. 1990. Monophyly and geography of the Río Pánuco basin swordtails (genus Xiphophorus) with descriptions of four new species. American Museum Novitates 2975:1–41. Rausch, R. R. 1963. Age and growth of the Rio Grande mountainsucker, Pantosteus plebeius (Baird and Girard). Unpubl. Master’s thesis, University of New Mexico, Albuquerque. Ray, J. M., C. B. Dillman, R. M. Wood, B. R. Kuhajda, and R. L. Mayden. 2007. Microsatellite variation among river Sturgeons of the genus Scaphirhynchus (Actinopterygii: Acipenseridae): a preliminary assessment of hybridization*. Journal of Applied Ichthyology 23:304–312. Raymakers, C. 2002. I. Conservation and broodstock management: international trade in Sturgeon and Paddlefish species— the effect of CITES listing. International Review of Hydrobiology 87:525–537. Raymakers, C., and C. Hoover. 2002. Acipenseriformes: CITES implementation from range states to consumer countries. Journal of Applied Ichthyology 18:629–638. Reash, R. J., G. L. Seegert, and W. L. Goodfellow. 2000. Experimentally-derived upper thermal tolerances for redhorse

595

suckers: revised 316(A) variance conditions at two generating facilities in Ohio. Environmental Science & Policy 3:191–196. Redmond, L. C. 1964. Ecology of the Spotted Gar (Lepisosteus oculatus Winchell) in southeastern Missouri. Unpubl. Master’s thesis, University of Missouri, Columbia. Reebs, S. G. 1996. Time-place learning in Golden Shiners (Pisces: Cyprinidae). Behavioural Processes 36:253–262. Reebs, S. G. 2000. Can a minority of informed leaders determine the foraging movements of a fish shoal? Animal Behaviour 59:403–409. Reed, B. C., W. E. Kelso, and D. A. Rutherford. 1992. Growth, fecundity, and mortality of Paddlefish in Louisiana. Transaction of the American Fisheries Society 121:378–384. Reed, J. R. 1969. Alarm substances and fright reaction in some fishes from the southeastern United States. Transactions of the American Fisheries Society 98:664–668. Reed, R. J. 1971. Biology of the Fallfish, Semotilus corporalis (Pisces, Cyprinidae). Transactions of the American Fisheries Society 100:717–725 Reeves, C. D. 1907. The breeding habits of the Rainbow Darter (Etheostoma caeruleum Storer), a study in sexual selection. Biological Bulletin 14:35–59. Regan, C. T. 1923. The skeleton of Lepidosteus, with remarks on the origin and evolution of the lower neopterygian fishes. Proceedings of the Zoological Society London 1923:445–461. Regan, J. D. 1961. Melanism in the poeciliid fish, Gambusia affinis (Baird and Girard). American Midland Naturalist 65:139–143. Regan, M. D., and C. J. Brauner. 2010a. The evolution of Root effect hemoglobins in the absence of intracellular pH protection of the red blood cell: insights from primitive fishes. Journal of Comparative Physiology B 180:695–706. Regan, M. D., and C. J. Brauner. 2010b. The transition in hemoglobin proton-binding characteristics within the basal actinopterygian fishes. Journal of Comparative Physiology B 180:521–530. Reheis, M. 1999. Extent of Pleistocene lakes in the western Great Basin. U.S. Geological Survey Miscellaneous Field Studies Map MF-2323, U. S. Geological Survey, Denver, Colorado. Available from http://geo-nsdi.er.usgs.gov/metadata/map-mf/2323/metadata.faq.html; accessed September 2012. Rehnberg, B. G., R. J. F. Smith, and B. D. Sloley. 1987. The reaction of Pearl Dace (Pisces, Cyprinidae) to alarm substance: time-course of behavior, brain amines, and stress physiology. Canadian Journal of Zoology 65:2916–2921. Rehwinkel, B. J. 1978. The fishery for Paddlefish at Intake, Montana during 1973 and 1974. Transactions of the American Fisheries Society 107:263–268. Reid, S. M., A. L. Edwards, and B. Cudmore. 2007. Recovery strategy for the Paddlefish (Polyodon spathula) in Canada. Species at Risk Act Recovery Strategy Series, Fisheries and Oceans Canada, Ottawa. Reidenauer, J. A., and D. Thistle. 1981. Response of a soft-bottom harpacticoid community to Stingray (Dasyatis sabina) disturbance. Marine Biology 65:261–267. Reighard, J. 1903. The natural history of Amia calva Linnaeus. Mark Anniversary Vol. 4:59–109, Henry Holt and Company, New York, New York. Reighard, J. 1904. Further observations on the breeding habits and on the function of the pearl organs in several species of Eventognathi. Science 19:211–212.

596

LITERATURE CITED

Reighard, J. E. 1910. Methods of studying the habits of fishes with an account of the breeding habits of the horned dace. Bulletin of the United States Bureau of Fisheries 28:1111–1136. Reighard, J. 1920. The breeding behavior of the Suckers and minnows. Biological Bulletin 38:1–32. Reimchen, T. E. 1989. Loss of nuptial colour in Threespine Sticklebacks (Gasterosteus aculeatus). Evolution 43:450–460. Reimers, P. E., and C. E. Bond. 1967. Distribution of fishes in tributaries of the lower Columbia River. Copeia 1967:541–550. Reis, R. R., and J. M. Dean. 1981. Temporal variation in the utilization of an intertidal creek by the Bay Anchovy (Anchoa mitchilli). Estuaries 4:16–23. Reisman, H. M. 1968. Effects of social stimuli on the secondary sex characters of male three-spined Sticklebacks, Gasterosteus aculeatus. Copeia 1968:816–826. Reis-Santos, P., S. D. McCormick, and J. M. Wilson. 2008. Ionoregulatory changes during metamorphosis and salinity exposure of juvenile Sea Lamprey (Petromyzon marinus L.). The Journal of Experimental Biology 211:978–988. Renaud, C. B. 1997. Conservation status of Northern Hemisphere Lampreys (Petromyzontidae). Journal of Applied Ichthyology 13:143–148. Renaud, C. B. 2002. The Muskellunge, Esox masquinongy, as a host for the Silver Lamprey, Ichthyomyzon unicuspis, in the Ottawa River, Ontario/Québec. Canadian Field-Naturalist 116:433–440. Renaud, C. B. 2008. Petromyzontidae, Entosphenus tridentatus: southern distribution record, Isla Clarión, Revillagigedo Archipelago, Mexico. Check List 4:82–85. Renaud, C. B. 2011. Lampreys of the world. An annotated and illustrated cata logue of lamprey species known to date. Food and Agriculture Organization Species Cata logue for Fishery Purposes No. 5, Rome. Renaud, C. B., M. F. Docker, and N. E. Mandrak. 2009b. Taxonomy, distribution, and conservation of Lampreys in Canada. American Fisheries Society Symposium 72:293–309. Renaud, C. B., and P. S. Economidis. 2010. Eudontomyzon graecus, a new nonparasitic Lamprey species from Greece (Petromyzontiformes: Petromyzontidae). Zootaxa 2477:37–48. Renaud, C. B., H. S. Gill, and I. C. Potter. 2009a. Relationships between the diets and characteristics of the dentition, buccal glands and velar tentacles of the adults of the parasitic species of Lamprey. Journal of Zoology 278:231–242. Renfro, J. L., and L. G. Hill. 1970. Factors influencing the aerial breathing and metabolism of Gars (Lepisosteus). The Southwestern Naturalist 15:25–54. Reno, H. W. 1969. Cephalic lateral-line systems of the cyprinid genus Hybopsis. Copeia 1969:736–773. Reno, H. W. 1971. The lateral-line system of the Silverjaw Minnow, Ericymba buccata Cope. The Southwestern Naturalist 15:347–358. Repka, J., and M. R. Gross. 1995. The evolutionary stable strategy under individual condition and tactic frequency. Journal of Theoretical Biology 176:27–31. Reséndez, M. A., and Salvadores, B. M. 1983. Contribución al conocimiento de la biología del pejelagarto Lepisosteus tropicus (Gill) y la tenguayaca Petenia splendida Günther, del Estado de Tabasco. Biotica 8:413–426. Resh, V. H., A. G. Hildrews, B. Statzner, and C. R. Townsend. 1994. Theoretical habitat templets, species traits, and species richness: a synthesis of long-term ecological research on the Up-

per Rhone River in the context of concurrently developed ecological theory. Freshwater Biology 31:539–554. Resink, J. W., T. W. M. van den Berg, R. van den Hurk, E. A. Huisman, and P. G. W. J. van Oordt. 1989a. Induction of gonadotropin release and ovulation by pheromones in the African Catfish, Clarias gariepinus. Aquaculture 83:167–177. Resink, J. W., P. K. Voorthuis, R. van den Hurk, R. C. Peters, and P. G. W. J. Van Oordt. 1989b. Steroid glucuronides of the seminal vesicle as olfactory stimuli in African Catfish, Clarias gariepinus. Aquaculture 83:153–166. Resink, J. W., P. K. Voorthuis, R. van den Hurk, H. G. B. Vullings, and P. G. W. J. van Oordt. 1989c. Pheromone detection and olfactory pathways in the brain of the female African Catfish, Clarias gariepinus. Cell and Tissue Research 256:337–345. Reusch, T. B. H., M. A. Häberli, P. B. Aeschlimann, and M. Milinski. 2001. Female Sticklebacks count alleles in a strategy of sexual selection explaining Mhc polymorphism. Nature 414:300–302. Reutter, J. M., and C. E. Herdendorf. 1976. Thermal discharge from a nuclear power plant: predicted effects on Lake Erie fish. Ohio Journal of Science 76:39–45. Reutter, K., F. Boudriot, and M. Witt. 2000. Heterogeneity of fish taste bud ultrastructure as demonstrated in the holosteans Amia calva and Lepisosteus oculatus. Philosophical Transactions of the Royal Society of London B. 355:1225–1228. Reutter, K., and A. Hansen. 2005. Subtypes of light and dark elongated taste bud cells in fish, p. 211–230. In Fish Chemosenses. K. Reutter and B. G. Kapoor (eds.). Science Publishers, Inc., Enfield, New Hampshire. Reynolds, J. D., and J. C. Jones. 1999. Female preference for preferred males is reversed under low oxygen conditions in the Common Goby (Pomatoschistus microps). Behavioral Ecology 10:149–154. Reynolds, W. W., M. E. Casterlin, and S. T. Millington. 1978. Circadian rhythm of preferred temperature in the Bowfin Amia calva, a primitive holostean fish. Comparative Biochemistry and Physiology 60A:107–109. Reznick, D., and J. A. Endler. 1982. The impact of predation on life history evolution in Trinidadian Guppies (Poecilia reticulata). Evolution 36:160–177. Rhymer, J. M., and D. Simberloff. 1996. Extinction by hybridization and introgression. Annual Review of Ecology and Systematics 27:83–109. Ricciardi, A. 2007. Are modern biological invasions an unprecedented form of global change? Conservation Biology 21:329–336. Rice, A. N., J. P. Ross, A. R. Woodward, D. A. Carbonneau, and H. F. Percival. 2007. Alligator diet in relation to Alligator mortality on Lake Griffin, FL. Southeastern Naturalist 6:97–110. Richards, C. E., and M. Castagna. 1970. Marine fishes of Virginia’s eastern shore (inlet and marsh, seaside waters). Chesapeake Science 11:235–248. Richmond, A. M., and B. Kynard. 1995. Ontogenetic behavior of Shortnose Sturgeon, Acipenser brevirostrum. Copeia 1995:172–182. Richmond, G. M. 1965. Glaciation of the Rocky Mountains, p. 217–230. In The Quaternary of the United States. H. E. Wright, Jr., and D. G. Frey (eds.). Princeton University Press, Princeton, New Jersey. Ricklefs, R. E. 1987. Community diversity: relative roles of local and regional processes. Science 235:167–171.

LITERATURE CITED

Ricklefs, R. E. 1990. Ecology. 3rd edition. W. H. Freeman and Co., New York. Ricker, W. E. 1946. Production and utilization of fish populations. Ecological Monographs 16:373–391. Rico, C., U. Kuhnlein, and G. J. FitzGerald. 1992. Male reproductive tactics in the Threespine Stickleback: an evaluation by DNA fingerprinting. Molecular Ecology 1:79–87. Ridenour, C. J., W. J. Doyle, and T. D. Hill. 2011. Habitats of age-0 Sturgeon in the lower Missouri River. Transactions of the American Fisheries Society 140:1351–1358. Rider, S. J., and P. Hartfield. 2007. Conservation and collection efforts for the endangered Alabama Sturgeon (Scaphirhynchus suttkusi). Journal of Applied Ichthyology 23:489–493. Ridgway, M. S. 1988. Developmental stage of offspring and brood defense in Smallmouth Bass (Micropterus dolomieu). Canadian Journal of Zoology 66:1722–1728. Ridgway, M. S. 1989. The parental response to brood size manipulation in Smallmouth Bass (Micropterus dolomieu). Ethology 80:47–54. Ridgway, M. S., B. J. Shutter, and E. E. Post. 1991. The relative influence of body size and territorial behaviour on nesting asynchrony in male Smallmouth Bass, Micropterus dolomieu (Pisces: Centrarchidae). Journal of Animal Ecology 60:665–681. Ridley, M. 1978. Paternal care. Animal Behaviour 26:904–932. Ridley, M., and C. Rechten. 1981. Female Sticklebacks prefer to spawn with males whose nests contain eggs. Behaviour 76:152–161. Rien, T. A., and R. C. Beamesderfer. 1994. Accuracy and precision of White Sturgeon age estimates from pectoral fin rays. Transactions of the American Fisheries Society 123:255–265. Riggs, C. D., and G. A. Moore. 1960. Growth of young Gar (Lepisosteus) in aquaria. Proceedings of the Oklahoma Academy of Science 40:44–46. Riley, S. C., E. F. Roseman, S. J. Nichols, T. P. O’Brien, C. S. Kiley, and J. S. Schaeffer. 2008. Deepwater demersal fish community collapse in Lake Huron. Transactions of the American Fisheries Society 137:1879–1890. Rilling, G. C., and E. D. Houde. 1999. Regional and temporal variability in distribution and abundance of Bay Anchovy (Anchoa mitchilli) eggs, larvae, and adult biomass in the Chesapeake Bay. Estuaries 22:1096–1109. Rinne, J. N. 1991. Habitat use by Spikedace, Meda fulgida (Pisces: Cyprinidae) in southwestern streams with reference to probable habitat competition by Red Shiner, Notropis lutrensis (Pisces: Cyprinidae). The Southwestern Naturalist 36:7–13. Rinne, J. N., J. E. Johnson, B. L. Jensen, A. W. Ruger, and R. Sorenson. 1986. The role of hatcheries in the management and recover of threatened and endangered fishes, p. 271–286. In Fish Culture in Fisheries Management. R. H. Stroud (ed.). American Fisheries Society, Bethesda, Maryland. Rinne, J. N., and W. L. Minckley. 1991. Native fishes of arid lands: a dwindling resource of the desert southwest, p. 45. In General Technical Report MN-206. U.S. Department of Agriculture, Forest Ser vice, Rocky Mountain Forest and Range Experiment Station, Fort Collins, Colorado. Rios-Cardenas, O., and M. S. Webster. 2005. Paternity and parental effort in the Pumpkinseed Sunfish. Behavioral Ecology 16:914–921. Ritchie, T. C., and R. B. Leonard. 1983. Immunohistochemical studies on the distribution and origin of candidate peptidergic primary afferent neurotransmitters in the spinal cord of an elas-

597

mobranch fish, the Atlantic Stingray (Dasyatis sabina). Journal of Comparative Neurology 213:414–425. Ritchie, T. C., L. J. Roos, B. J. Williams, and R. B. Leonard. 1984. The descending and intrinsic serotoninergic innervations of an elasmobranch spinal cord. Journal of Comparative Neurology 224:395–406. Roach, K. A., and K. O. Winemiller. 2011. Diel turnover of assemblages of fish and shrimp on sandbanks in a temperate floodplain river. Transactions of the American Fisheries Society 140:84–90. Robbins, L. W., G. D. Hartman, and M. H. Smith. 1987. Dispersal, reproductive strategies, and the maintenance of genetic variability in mosquitofish (Gambusia affinis). Copeia 1987:156–164. Roberge, M., J. M. B. Hume, C. K. Minns, and T. Slaney. 2002. Life history characteristics of freshwater fishes occurring in British Columbia and the Yukon, with major emphasis on stream habitat characteristics. Canadian Manuscript Report of Fisheries and Aquatic Sciences 2611:1–248. Roberts, J., A. Chick, L. Oswald, and P. Thoompson. 1995. Effect of carp, Cyprinus carpio L., an exotic benthivorous fish, on aquatic plants and water quality in experimental ponds. Marine and Freshwater Research 46:1171–1180. Roberts, J. H., and N. P. Hitt. 2010. Longitudinal structure in temperate stream fish communities: evaluating conceptual models with temporal data, p. 281–299. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. Roberts, W. 1989. The Mooneye in Alberta. Alberta Naturalist 19:134–140. Robertson, C. R., S. C. Zeug, and K. O. Winemiller. 2008. Associations between hydrological connectivity and resource partitioning among sympatric Gar species (Lepisosteidae) in a Texas river and associated oxbows. Ecology of Freshwater Fish 17:119–129. Robinet, T. T., and E. E. Feunteun. 2002. Sublethal effects of exposure to chemical compounds: a cause for the decline in Atlantic Eels? Ecotoxicology 11:265–277. Robinette, H. R. 1983. Species profiles: life histories and environmental requirements of coastal fishes and invertebrates (Gulf of Mexico)—Bay Anchovy and Striped Anchovy. Fish and Wildlife Ser vice Biological Report 82(11.14):1–15. Robins, C. R., D. M. Cohen, and C. H. Robins. 1979. The Eels, Anguilla and Histiobranchus, photographed on the floor of the deep Atlantic in the Bahamas. Bulletin of Marine Science 29:401–405. Robins, C. R., and E. C. Raney. 1956. Studies of the catostomid fishes of the genus Moxostoma, with descriptions of two new species. Cornell University Agricultural Experiment Station Memoirs 343:1–56. Robins, C. R., and E. C. Raney. 1957. The systematic status of the Suckers of the genus Moxostoma from Texas, New Mexico and Mexico. Tulane Studies in Zoology and Botany 5:291–318. Robins, C. R., and G. C. Ray. 1986. A Field Guide to Atlantic Coast fishes of North America. Houghton Mifflin Company, Boston, Massachusetts. Robinson, B. W., and D. S. Wilson. 1994. Character release and displacement in fishes: a neglected literature. The American Naturalist 144:596–627. Robinson, B. W., D. S. Wilson, and A. S. Margosian. 2000. A pluralistic analysis of character release in Pumpkinseed Sunfish (Lepomis gibbosus). Ecology 81:2799–2812.

598

LITERATURE CITED

Robinson, B. W., D. S. Wilson, A. S. Margosian, and P. T. Lotito. 1993. Ecological and morphological differentiation of Pumpkinseed Sunfish in lakes without Bluegill Sunfish. Evolutionary Ecology 7:451–464. Robinson, B. W., D. S. Wilson, and G. O. Shea. 1996. Trade-offs of ecological specialization: an intraspecific comparison of Pumpkinseed Sunfish phenotypes. Ecology 77:170–178. Robinson, E. S., and I. C. Potter. 1969. Meiotic chromosomes of Mordacia praecox and a discussion of chromosome numbers in Lampreys. Copeia 1969:824–828. Robinson, E. S., and I. C. Potter. 1981. The chromosomes of the Southern Hemispheric Lamprey, Geotria australis Gray. Experientia 37:239–240. Robinson, E. S., I. C. Potter, and N. B. Atkin. 1975. The nuclear DNA content of Lampreys. Experientia 31:912–913. Robinson, E. S., I. C. Potter, and C. J. Webb. 1974. Homogeneity of holarctic Lamprey karyotypes. Caryologia 27:443–454. Robinson, G. L., and L. A. Jahn. 1980. Some observations of fish parasites in pool 20, Mississippi River. Transactions of the American Microscopical Society 99:206–212. Robinson, J. W. 1966. Observations on the life history, movement, and harvest of the Paddlefish, Polyodon spathula, in Montana. Proceedings of the Montana Academy of Sciences 26:33–44. Robinson, L. K., and W. M. Tonn. 1989. Influence of environmental factors and piscivory in structuring fish assemblages of small Alberta lakes. Canadian Journal of Fisheries and Aquatic Sciences 46:81–89. Robinson, M. R., and M. M. Ferguson. 2004. Genetics of North American Acipenseriformes, p. 217–230. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. Robison, H. W., and T. M. Buchanan. 1988. Fishes of Arkansas. University of Arkansas Press, Fayetteville. Robison, R. R., R. D. Fernald, and N. E. Stacey. 1998. The olfactory system of a Cichlid fish responds to steroidal compounds. Journal of Fish Biology 53:226–229. Robles, F., R. de la Herrán, A. Ludwig, C. R. Rejón, M. R. Rejón, and M. A. Garrido-Ramos. 2004. Evolution of ancient satellite DNAs in Sturgeon genomes. Gene 338:133–142. Robles, F., R. de la Herrán, A. Ludwig, C. R. Rejón, M. R. Rejón, M. A. Garrido-Ramos. 2005. Genomic organization and evolution of the 5S ribosomal DNA in the ancient fish Sturgeon. Genome 48:18–28. Rochard, E., G. Castelnaud, and M. Lepage. 1990. Sturgeons (Pisces: Acipenseridae); threats and prospects. Journal of Fish Biology 37 (Supplement A):123–132. Rodriguez, M. A., P. Magnan, and S. Lacassse. 1993. Fish species composition and lake abiotic variables in relation to the abundance and size structure of cladoceran zooplankton. Canadian Journal of Fisheries and Aquatic Sciences 50:638–647. Rodríguez-Muñoz, R., A. G. Nicieza, and F. Braña. 2001. Effects of temperature on developmental performance, survival and growth of Sea Lamprey embryos. Journal of Fish Biology 58:475–486. Rodríguez-Muñoz, R., A. G. Nicieza, and F. Braña. 2003. Densitydependent growth of Sea Lamprey larvae: evidence for chemical interference. Functional Ecology 17:403–408. Rodríguez-Muñoz, R., J. R. Waldman, C. Grunwald, N. K. Roy, and I. Wirgin. 2004. Absence of shared mitochondrial DNA

haplotypes between Sea Lamprey from North American and Spanish rivers. Journal of Fish Biology 64:783–787. Rodzen, J. A., and B. May. 2002. Inheritance of microsatellite loci in the White Sturgeon (Acipenser transmontanus). Genome 45:1064–1076. Rohwer, S. 1978. Parental cannibalism of offspring and egg raiding as a courtship strategy. American Naturalist 112:429–440. Rolff, J. 2007. Why did the acquired immune system of vertebrates evolve? Developmental and Comparative Immunology 31:476–482. Roll, U., T. Dayan, D. Simberloff, and M. Goren. 2007. Characteristics of the introduced fish fauna of Israel. Biological Invasions 9:813–824. Rose, K. A., J. H. Cowan Jr., M. E. Clark, E. D. Houde, and S. B. Wang. 1999. An individual-based model of Bay Anchovy population dynamics in the mesohaline region of Chesapeake Bay. Marine Ecology Progress Series 185:113–132. Rose, M. R., and G. V. Lauder. 1996. Post-spandrel adaptationism, p. 1–8. In Adaptation. M. R. Rose and G. V. Lauder (eds.). Academic Press, San Diego, California. Roseman, E. F., B. Manny, J. Boase, M. Child, G. Kennedy, J. Craig, K. Soper, and R. Drouin. 2011. Lake Sturgeon response to a spawning reef constructed in the Detroit River. Journal of Applied Ichthyology 27 (Supplement 2):66–76. Roseman, E. F., W. W. Taylor, D. B. Hayes, A. L. Jones, and J. T. Francis. 2009. Predation on Walleye eggs by fish on reefs in western Lake Erie. Journal of Great Lakes Research 32:415–423. Rosen, D. E. 1979. Fishes from the uplands and intermontane basins of Guatemala: revisionary studies and comparative geography. Bulletin of the American Museum of Natural History 162:267–376. Rosen, R. A., and D. C. Hales. 1980. Occurrence of scarred Paddlefish in the Missouri River, South Dakota-Nebraska. The Progressive Fish-Culturist 42:82–85. Rosen, R. A., and D. C. Hales. 1981. Feeding of Paddlefish, Polyodon spathula. Copeia 1981:441–455. Rosen, R. A., and D. C. Hales. 1982. Occurrence of a blind Paddlefish, Polyodon spathula. Copeia 1982:212–214. Rosen, R. A., D. C. Hales, and D. G. Unkenholz. 1982. Biology and exploitation of Paddlefish in the Missouri River below Gavins Point Dam. Transactions of the American Fisheries Society 111:216–222. Rosenberger, L. J. 2001a. Pectoral fin locomotion in batoid fishes: undulation versus oscillation. The Journal of Experimental Biology 204:379–394. Rosenberger, L. J. 2001b. Phylogenetic relationships within the Stingray genus Dasyatis (Chondricthyes: Dasyatidae). Copeia 2001:615–627. Rosenfeld, M. J., and J. A. Wilkinson. 1989. Biochemical genetics of the Colorado River Gila complex (Pisces: Cyprinidae). The Southwestern Naturalist 34:232–244. Rosenfield, J. A. 1998. Detection of natural hybridization between Pink Salmon (Oncorhynchus gorbuscha) and Chinook Salmon (Oncorhynchus tshawytscha) in the Laurentian Great lakes using meristic, morphological, and color evidence. Copeia 1998:706–714. Rosenfield, J. A., and A. Kodric-Brown. 2003. Sexual selection promotes hybridization between Pecos Pupfish, Cyrpinodon pecosensis and Sheepshead Minnow, C. variegatus. Journal of Evolutionary Biology 16:595–606.

LITERATURE CITED

Rosenthal, G. G., and C. S. Evans. 1998. Female preference for swords in Xiphophorus helleri reflects a bias for large apparent size. Proceedings of the National Academy of Sciences of the United States of America 95:4431–4436. Rosenthal, G. G., C. S. Evans, and W. L. Miller. 1996. Female preference for dynamic traits in the Green Swordtail, Xiphophorus helleri. Animal Behaviour 51:811–820. Rosenthal, G. G., T. Y. F. Martinez, F. J. García de León, and M. J. Ryan. 2001. Shared preferences by predators and females for male ornaments in swordtails. American Naturalist 158:146–154. Rosenthal, G. G., W. E. Wagner, and M. J. Ryan. 2002a. Secondary reduction of preference for the sword ornament in the Pygmy Swordtail, Xiphophorus nigrensis (Pisces: Poeciliidae). Animal Behaviour 63:37–45. Rosenthal, H. L. 1952. Observations on the reproduction of the poeciliid Lebistes reticulates. Biological Bulletin 102:30–38. Rosenthal, H., P. Bronzi, D. J. McKenzie, G. Arlati, and R. Rossi (eds.). 1999. Third International Symposium on Sturgeon. Special Issue, Journal of Applied Ichthyology 15:1–349. Blackwell Wissenschafts-Verlag, Berlin, Germany. Rosenthal, H., R. M. Bruch, F. P. Binkowski, and S. I. Doroshov (eds.). 2002b. Fourth International Symposium on Sturgeon. Special Issue, Journal of Applied Ichthyology 18:219–698. Blackwell Verlag, Berlin, Germany. Rosiles, J. R., and R. B. Leonard. 1980. The organization of the extraocular motor nuclei in the Atlantic Stingray, Dasyatis sabina. Journal of Comparative Neurology 193:677–687. Ross, M. R. 1976. Nest-entry behavior of female Creek Chubs (Semotilus atromaculatus) in different habitats. Copeia 1976:378–380. Ross, M. R. 1977a. Function of Creek Chub (Semotilus atromaculatus) nest-building. Ohio Journal of Science 77:36–37. Ross, M. R. 1977b. Aggression as a social mechanism in the Creek Chub (Semotilus atromaculatus). Copeia 1977:393–397. Ross, M. R., and T. M. Cavender. 1977. First report of the natural cyprinid hybrid, Notropis cornutus × Rhinichthys cataractae, from Ohio. Copeia 1977:777–780. Ross, M. R., and R. J. Reed. 1978. The reproductive behavior of the Fallfish Semotilus corporalis. Copeia 1978:215–221. Ross, R. M., and R. M. Bennett. 1997. Comparative behaviour and dietary effects in early life phases of American Sturgeons. Fisheries Management and Ecology 4:17–30. Ross, S. T. 1986. Resource partitioning in fish assemblages: a review of field studies. Copeia 1986:352–388. Ross, S. T. 1991. Mechanisms structuring stream fish assemblages—are there lessons from introduced species? Environmental Biology of Fishes 30:359–3368. Ross, S. T. 2001. Inland Fishes of Mississippi. University Press of Mississippi, Jackson. Ross, S. T. 2013. Ecology of North American Freshwater Fishes. University of California Press, Berkeley. Ross, S. T., J. A. Baker, and K. E. Clark. 1987. Microhabitat partitioning of southeastern stream fishes: temporal and spatial predictability, p. 42–51. In W. J. Matthews and D. C. Heins (eds.). Community and Evolutionary Ecology of North American Stream Fishes. University of Oklahoma Press, Norman. Ross, S. T., J. G. Knight, and S. D. Wilkins. 1990. Longitudinal occurrence of the Bayou Darter (Percidae: Etheostoma rubrum) in Bayou Pierre—a response to stream order or habitat availability? Polish Archives of Hydrobiology 37:221–233.

599

Ross, S. T., J. G. Knight, and S. D. Wilkins. 1992. Distribution and microhabitat dynamics of the threatened Bayou Darter, Etheostoma rubrum. Copeia 1992:658–671. Ross, S. T., W. J. Matthews, and A. A. Echelle. 1985. Persistence of stream fish assemblages: effects of environmental change. The American Naturalist 126:24–40. Ross, S. T., M. T. O’Connell, D. M. Patrick, C. A. Latorre, W. T. Slack, J. G. Knight, and S. D. Wilkins. 2001. Stream erosion and densities of Etheostoma rubrum (Percidae) and associated riffle-inhabiting fishes: biotic stability in a variable habitat. Copeia 2001:916–927. Ross, S. T., W. T. Slack, R. J. Heise, M. A. Dugo, H. Rogillio, B. R. Bowen, P. Mickle, and R. W. Heard. 2009. Estuarine and coastal habitat use of Gulf Sturgeon (Acipenser oxyrinchus desotoi) in the north-central Gulf of Mexico. Estuaries and Coasts 32:360–374. Ross, S. W., and G. H. Burgess. 1980. Dasyatis sabina (LeSueur), Atlantic Stingray, p. 37. In Atlas of North American Freshwater Fishes. D. S. Lee, C. R. Gilbert, C. H. Hocutt, R. E. Jenkins, D. E. McAllister, and J. R. Stauffer (eds.). North Carolina State Museum of Natural History, Raleigh. Rossiter, A., D. L. G. Noakes, and F. W. H. Beamish. 1995. Validation of age estimation for the Lake Sturgeon. Transactions of the American Fisheries Society 124:777–781. Roush, K. D., C. P. Paukert, and W. Stancill. 2003. Distribution and movement of juvenile Paddlefish in a mainstem Missouri River reservoir. Journal of Freshwater Ecology 18:79–87. Rovainen, C. M. 1982. Neurophysiology, p. 1–136. In The Biology of Lampreys Vol. 4A. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Rowe, C. 1999. Receiver psychology and the evolution of multicomponent signals. Animal Behaviour 58:921–931. Rowe, C., and T. Guilford. 1999. The evolution of multimodal warning displays. Evolutionary Ecology 13:655–671. Rowland, W. J. 1974. Reproductive behavior of the Fourspine Stickleback, Apeltes quadracus. Copeia 1974:183—194. Rowland, W. J. 1982. Male mate choice by male Sticklebacks Gasterosteus aculeatus. Animal Behaviour 30:1093–1098. Rowland, W. J. 1989a. Mate choice and supernormality effect in female Sticklebacks (Gasterosteus aculeatus). Behavioral Ecology and Sociobiology 24:433–438. Rowland, W. J. 1989b. The ethological basis of mate choice in male Threespine Sticklebacks, Gasterosteus aculeatus. Animal Behaviour 38:112–120. Rowland, W. J., C. L. Baube, and T. T. Horan. 1991. Signalling of sexual receptivity by pigmentation pattern in female Sticklebacks. Animal Behaviour 42:243–249. RRCC (Robust Redhorse Conservation Committee). 2010. Robust Redhorse, Moxostoma robustum. Robust Redhorse Conservation Committee. Available from http://www.robustredhorse.com/; as of September 2010. Ruelle, R., and K. D. Keenlyne. 1993. Contaminants in Missouri River Pallid Sturgeon. Bulletin of Environmental Contamination and Toxicology 50:898–906. Ruelle, R., and P. L. Hudson. 1977. Paddlefish (Polyodon spathula): growth and food of young of the year and a suggested technique for measuring length. Transactions of the American Fisheries Society 106:609–613. Rundle, H. D., L. Nagel, J. W. Boughman, and D. Schluter. 2000. Natural selection and parallel speciation in sympatric Sticklebacks. Science 287:306–308.

600

LITERATURE CITED

Runstrom, A., R. M. Bruch, D. Reiter, and D. Cox. 2002. Lake Sturgeon (Acipenser fulvescens) on the Menominee Indian Reservation: an effort toward co-management and population restoration. Journal of Applied Ichthyology 18:481–485. Runstrom, A. L., B. Vondracek, and C. A. Jennings. 2001. Population statistics for Paddlefish in the Wisconsin River. Transactions of the American Fisheries Society 130:546–556. Ruppert, J. B., P. B. Holden, and P. D. Abate. 1999. Determining age structure of the Razorback Sucker, Xyrauchen texanus, population in Lake Mead. Desert Fishes Council Proceedings 31:31. Rupprecht, R. J., and L. A. Jahn. 1980. Biological notes on Blue Suckers in the Mississippi River. Transactions of the American Fisheries Society 109:323–326. Rusak, J. A., and T. Mosindy. 1997. Seasonal movements of Lake Sturgeon in Lake of the Woods and the Rainy River, Ontario. Canadian Journal of Zoology 75:383–395. Russell, D. F., L. A. Wilkens, and F. Moss. 1999. Use of behavioural stochastic resonance by paddle fish for feeding. Nature 402:291–294. Russell, T. R. 1986. Biology and life history of the Paddlefish—a review, p. 3–20. In The Paddlefish: Status, Management and Propagation. J. G. Dillard, L. K. Graham, and T. R. Russell (eds.). American Fisheries Society Special Publication 7. Russom, C. L., S. P. Bradbury, S. J. Broderius, D. E. Hammermeister, and D. R. Drummond. 1997. Predicting modes of toxic action from chemical structure: acute toxicity in the Fathead Minnow (Pimephales promelas). Environmental Toxicology and Chemistry 16:948–967. Ryan, M. J. 1988. Phenotype, genotype, swimming endurance, and sexual selection in a swordtail (Xiphophorus nigrensis). Copeia 1988:484–487. Ryan, M. J. 1990a. Sexual selection, sensory systems and sensory exploitation, p. 157–195. In Oxford Surveys in Evolutionary Biology. Vol. 7. D. J. Futuyma and J. Antonovics (eds.). Oxford University Press, Oxford, United Kingdom. Ryan, M. J. 1990b. Signals, species, and sexual selection. American Scientist 78:46–52. Ryan, M. J. 1997. Sexual selection and mate choice, p. 179–202. In Behavioural Ecology: An Evolutionary Approach. J. R. Krebs and N. B. Davies (eds.). Blackwell Science, Oxford, United Kingdom. Ryan, M. J. 1998. Sexual selection, receiver biases, and the evolution of sex differences. Science 281:1999–2003. Ryan, M. J., and B. A. Causey. 1989. “Alternative” mating behaviour in the swordtails, Xiphophorus nigrensis and Xiphophorus pygmaeus (Pisces: Poeciliidae). Behavioral Ecology and Sociobiology 24:341–348. Ryan, M. J., L. A. Dries, P. Batra, and D. M. Hillis. 1996. Male mate preferences in a gynogenetic species complex of Amazon Mollies. Animal Behaviour 52:1225–1236. Ryan, M. J., D. K. Hews, and W. E. Wagner. 1990. Sexual selection on alleles that determine body size in the swordtail Xiphophorus nigrensis. Behavioral Ecology and Sociobiology 26:231–237. Ryan, M. J., and Keddy-Hector, A. 1992. Directional patterns of female mate choice and the role of sensory biases. American Naturalist 139:S4–S35. Ryan, M. J., C. M. Pease, and M. R. Morris. 1992. A genetic polymorphism in the swordtail Xiphophorus nigrensis: testing the prediction of equal fitnesses. American Naturalist 139: 21–31.

Ryan, M. J., and A. S. Rand. 1993. Species recognition and sexual selection as a unitary problem in animal communication. Evolution 47:647–657. Ryan, M. J., and W. E. Wagner. 1987. Asymmetries in mating preferences between species: female swordtails prefer heterospecific males. Science 236:595–597. Ryder, R. A., and S. R. Kerr. 1989. Environmental priorities: placing habitat in hierarchic perspective, p. 1–12. In Proceedings of the National Workshop on Effects of Habitat Alteration on Salmonid Stocks. C. D. Levings, L. B. Holtby, and M. A. Hendersons (eds.). Canadian Special Publications of Fisheries and Aquatic Sciences 105. Ryman, N. 1991. Conservation genetics considerations in fishery management. Journal of Fish Biology. 39 (Supplement A):211–224. Sabaj, M. H., E. G. Maurakis, and W. S. Woolcott. 2000. Spawning behaviors of the Bluehead Chub, Nocomis leptocephalus, River Chub, N. micropogon, and Central Stoneroller, Campostoma anomalum. American Midland Naturalist 144:187–201. Sabat, A. M. 1994a. Costs and benefits of parental effort in a broodguarding fish (Ambloplites rupestris, Centrarchidae). Behavioral Ecology 5:195–201. Sabat, A. M. 1994b. Mating success in brood-guarding male Rock Bass, Ambloplites rupestris—the effect of body size. Environmental Biology of Fishes 39:411–415. Sabins, D. S., and F. M. Truesdale. 1974. Diel and seasonal occurrence of immature fishes in a Louisiana tidal pass. Proceedings of the Southeastern Association of Game and Fish Commissioners 28:161–171. Sabo, M. 2000. Threatened fishes of the world: Notropis bifrenatus (Cope, 1867) (Cyprinidae). Environmental Biology of Fishes 59:384. Saff ron, I. 2002. Caviar. Broadway Books, New York. Sage, M, R. G. Jackson, W. L. Klesch and V. L. deVlaming. 1972. Growth and seasonal distribution of the elasmobranch Dasyatis sabina. Contributions in Marine Science 16:71–74. Saglio, P. 1982. Piègeage d’anguilles (A. anguilla, L.) dans le milieu naturel au moyend’extraits biologiques d’origine intraspécifique. Mise en évidence de l’attractivitéphéromonale du mucus épidermique. Acta Ecologica 3:223–231. Sahagian, D., A. Proussevitch, and W. Carlson. 2002. Timing of Colorado Plateau uplift: initial constraints from vesicular basaltderived paleoelevations. Geology 30:807–810. Saint-Jacques, N., H. H. Harvey, and D. A. Jackson. 2000. Selective foraging in the White Sucker (Catostomus commersoni). Canadian Journal of Zoology 78:1320–1331. Saitoh, K., T. Sado, R. L. Mayden, N. Hanzawa, K. Nakamura, M. Nishida, and M. Miya. 2006. Mitogenomic evolution and interrelationships of the Cypriniformes (Actinopterygii: Ostariophysi): the first evidence toward resolution of higher-level relationships of the world’s largest freshwater fish clade based on 59 whole mitogenome sequences. Journal of Molecular Evolution 63:826– 41. Sakaris, P. C., A. M. Ferrara, K. J. Kleiner, and E. R. Irwin. 2003. Movements and home ranges of Alligator Gar in the MobileTensaw Delta, Alabama. Proceedings of the Southeastern Association of Fish and Wildlife Agencies 57:102–111. Saksena, V. P., and E. D. Houde. 1972. Effect of food level on the growth and survival of laboratory-reared larvae of Bay Anchovy (Anchoa mitchilli Valenciennes) and Scaled Sardine (Harengula

LITERATURE CITED

pensacolae Goode & Bean). Journal of Experimental Marine Biology and Ecology 8:249–258. Saksena, V. P. 1975a. Effects of temperature and light on aerial breathing of the Longnose Gar, Lepisosteus osseus. Ohio Journal of Science 72:58–61. Saksena, V. P. 1975b. Effects of temperature and light on aerial breathing of the Shortnose Gar, Lepisosteus platostomus. Ohio Journal of Science 75:178–181. Sale, P. 1974. Overlap in resource use, and interspecific competition. Oecologia 17:245–256. Salewski, V. 2003. Satellite species in Lampreys: a worldwide trend for ecological speciation in sympatry? Journal of Fish Biology 63:267–279. Salgado-Maldonado, G., F. Moravec, G. Cabañas-Carranza, P. Sánchez-Nava, R. Báez-Valé, R. Aguilar-Aguilar, and T. Scholz. 2004. Helminth parasites of the Tropical Gar, Atractosteus tropicus Gill from Tabasco, Mexico. Comparative Parasitology 90:260–265. Salnikov, V. B. 2009. First finding of Gar Atractosteus sp. (Actinopterygii, Lepisosteiformes, Lepisosteidae) in the Caspian Sea near the coast of Turkmenistan. Russian Journal of Biological Invasions 1:17–20. Sampson, S. J., J. H. Chick, and M. A. Pegg. 2009. Diet overlap among two Asian carp and three native fishes in backwater lakes on the Illinois and Mississippi rivers. Biological Invasions 11:483–496. Sanchez, C. 1970. Life history and ecology of carp, Cyprinus carpio (Linneaus), in Elephant Butte Lake, New Mexico. Unpubl. Master’s thesis, New Mexico State University, Las Cruces. Sandercock, F. K. 1991. Life history of Coho Salmon (Oncorhynchus keta), p. 395–446. In Pacific Salmon Life Histories. C. Groot and C. Margolis (eds.). University of British Columbia Press, Vancouver. Sanderson, S. L., J. J. Cech, Jr., and A. Y. Cheer. 1994. Paddlefish buccal flow velocity during ram suspension feeding and ram ventilation. Journal of Experimental Biology 186:145–156. Sandilands, A. P. 1987. Biology of the Lake Sturgeon (Acipenser fulvescens) in the Kenogami River, Ontario, p. 33–46. In Proceedings of a Workshop on the Lake Sturgeon (Acipenser fulvescens). C. H. Olver (ed.). Ontario Fisheries Technical Report Series, No. 23, Ontario Ministry of Natural Resources, Toronto, Canada. Sandland, G. J., and C. P. Goater. 2000. Development and intensity dependence of Ornithodiplostomum ptychocheilus metacercariae in Fathead Minnows (Pimephales promelas). Journal of Parasitology 86:1056–1060. Sandland, G. J., C. P. Goater, and A. J. Danylchuk. 2001. Population dynamics of Ornithodiplostomum ptychocheilus metacercariae in Fathead Minnows (Pimephales promelas) from four Northern-Alberta lakes. Journal of Parasitology 87:744–748. Sargent, R. C. 1988. Paternal care and egg survival both increase with clutch size in the Fathead Minnow, Pimephales promelas. Behavioral Ecology and Sociobiology 23:33–37. Sargent, R. C. 1989. Alloparental care in the Fathead Minnow, Pimephales promelas: stepfathers discriminate against their adopted eggs. Behavioral Ecology and Sociobiology 25:379–385. Sargent, R. C. 1992. Ecology of filial cannibalism in fish: theoretical perspectives, p. 38–62. In Cannibalism: Ecology and Evolution among Diverse Taxa. M. A. Elgar and B. J. Crespi (eds.). Oxford University Press, Oxford, United Kingdom.

601

Sargent, R. C., and M. R. Gross. 1993. Williams’ principle, and explanation of parental care in teleost fishes, p. 333–361. In Behaviour of Teleost Fishes. 2nd edition. T. J. Pitcher (ed.). Chapman and Hall, London, United Kingdom. Sato, K., and N. Suzuki. 2001. Whole-cell response characteristics of ciliated and microvillous olfactory receptor neurons to amino acids, pheromone candidates and urine in Rainbow Trout. Chemical Senses 26:1145–1156. Sato, Y., and M. Nishida. 2010. Teleost fish with specific genome duplication as unique models of vertebrate evolution. Environmental Biology of Fishes 88:169–188. Satou, M., H.-A. Takeuchi, J. Nishii, M. Tanabe, S. Kitamura, N. Okumoto, and M. Iwata 1994b. Behavioral and electrophysiological evidences that the lateral line is involved in the intersexual vibrational communication of the Himé Salmon (landlocked red Salmon, Oncorhynchus nerka). Journal of Comparative Physiology A 174:539–549. Satou, M., H.-A. Takeuchi, K. Takei, T. Hasegawa, T. Matsushima, and N. Okumoto. 1994a. Characterization of vibrational and visual signals which elicit spawning behavior in the male Himé salmon (landlocked red salmon, Oncorhynchus nerka). Journal of Comparative Physiology A 174:527–537. Saucier, D., and L. Astic. 1995. Morpho-functional alterations in the olfactory system of Rainbow Trout (Oncorhynchus mykiss) and possible acclimation in response to long-lasting exposure to low copper levels. Comparative Biochemistry and Physiology 112A:273–284. Sauka-Spengler, T., and M. Bronner-Fraser. 2008. Insights from a Sea Lamprey into the evolution of neural crest gene regulatory network. Biological Bulletin 214:303–314. Savage, T. 1963. Reproductive behavior of the Mottled Sculpin, Cottus bairdi Girard. Copeia 1963:317–325. Savoy, T. 2007. Prey eaten by Atlantic Sturgeon in Connecticut waters, p. 157–165. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Sawada, Y. 1982. Phylogeny and zoogeography of the superfamily Cobitoidea (Cyprinoidei, Cypriniformes). Memoir of the Faculty of Fisheries Hokkaido University 28:65–223. Scalet, C. G. 1973. Stream movements and population density of the Orangebelly Darter, Etheostoma radiosum cyanorum (Osteichthyes: Percidae). Southwestern Naturalist 17:381–387. Scarnecchia, D. L. 1992. A reappraisal of Gars and Bowfins in fishery management. Fisheries 17:6–12. Scarnecchia, D. L., T. W. Gengerke, and C. T. Moen. 1989. Rational for a harvest slot limit for Paddlefish in the upper Mississippi River. North American Journal of Fisheries Management 9:477–487. Scarnecchia, D. L., B. D. Gordon, J. D. Schooley, L. F. Ryckman, B. J. Schmitz, S. E. Miller, and Y. Lim. 2011. Southern and northern Great Plains (United States) Paddlefish stocks within frameworks of acipenseriform life history and the metabolic theory of ecology. Reviews in Fisheries Science 19:279–298. Scarnecchia, D. L., L. F. Ryckman, Y. Lim, G. J. Power, B. J. Schmitz, and J. A. Firehammer. 2007. Life history and the costs of reproduction in northern Great Plains Paddlefish (Polyodon spathula) as a potential framework for other acipenseriform fishes. Reviews in Fisheries Science 15:211–263.

602

LITERATURE CITED

Scarnecchia, D. L., L. F. Ryckman, Y. Lim, S. E. Miller, B. J. Schmitz, G. J. Power, and S. A. Shefstad. 2009. Riverine and reservoir influences on year class strength and growth of upper Great Plains Paddlefish. Reviews in Fisheries Science 17:241–266. Scarnecchia, D. L., L. F. Ryckman, Y. Lim, G. Power, B. Schmitz, and V. Riggs. 2006. A long-term program for validation and verification of dentaries for age estimation in the YellowstoneSakakawea Paddlefish stock. Transactions of the American Fisheries Society 135:1086–1094. Scarnecchia, D. L., and P. A. Stewart. 1997. Implementation and evaluation of a catch-and-release fishery for Paddlefish. North American Journal of Fisheries Management 17:795–799. Scarnecchia, D. L., P. A. Stewart, and Y. Lim. 1996a. Profile of recreational Paddlefish snaggers on the lower Yellowstone River, Montana. North American Journal of Fisheries Management 16:872–879. Scarnecchia, D. L., P. A. Stewart, and G. L. Power. 1996b. Age structure of the Yellowstone-Sakakawea Paddlefish stock, 1963–1993, in relation to reservoir history. Transactions of the American Fisheries Society 125:291–299. Schaefer, J. 2001. Riffles as barriers to interpool movement by three cyprinids (Notropis boops, Campostoma anomalum and Cyprinella venusta). Freshwater Biology 46:379–388. Schaefer, S. A., and T. M. Cavender. 1986. Geographic variation and subspecific status of Notropis spilopterus (Pisces: Cyprinidae). Copeia 1986:122–130. Schafer, J. P., and J. H. Hartshorn. 1965. The Quaternary of New England, p. 113–127. In The Quaternary of the United States. H. E. Wright, Jr., and D. G. Frey (eds.). Princeton University Press, Princeton, New Jersey. Scharf, F. S., J. A. Buckel, and F. Juanes. 2002. Size-dependent vulnerability of juvenile Bay Anchovy, Anchoa mitchilli to Bluefish predation: does large body size always provide a refuge? Marine Ecology Progress Series 233:241–252. Scharpf, C. 2006. Annotated checklist of North American freshwater fishes, including subspecies and undescribed forms. Part II. Catostomidae through Mugilidae. American Currents 32:1–39. Schartl, M., B. Wilde, I. Schlupp, and J. Parzefall. 1995. Evolutionary origin of a parthenoform, the Amazon Molly, Poecilia formosa, on the basis of a molecular genealogy. Evolution 49:827–835. Scheerer, P. D. 2002. Implications of floodplain isolation and connectivity on the conservation of an endangered minnow, Oregon Chub, in the Willamette River, Oregon. Transactions of the American Fisheries Society 131:1070–1080. Scheerer, P. D., and P. J. McDonald. 2003. Age, growth, and timing of spawning of an endangered minnow, the Oregon Chub (Oregonichthys crameri), in the Willamette Basin, Oregon. Northwestern Naturalist 84:68–79. Schemske, D. W. 1974. Age, length and fecundity of the Creek Chub, Semotilus atromaculatus (Mitchell), in central Illinois. American Midland Naturalist 92:505–509. Schenck, R. A., and R. C. Vrijenhoek. 1989. Coexistence among sexual and asexual forms of Poeciliopsis: foraging behavior and microhabitat selection, p. 39–48. In Evolution and Ecology of Unisexual Vertebrates. R. Dawley and J. Bogart (eds.). New York State Museum, Albany, New York. Schermer, E. R., D. G. Howell, and D. Jones. 1984. The origins of allochthonous terranes: perspectives on the growth and shaping

of continents. Annual Review of Earth and Planetary Sciences 12:107–131. Scheurer, J. A., B. A. Berejikian, F. P. Thrower, E. R. Ammann, and T. A. Flagg. 2007. Innate predator recognition and fright response in related populations of Oncorhynchus mykiss under different predation pressure. Journal of Fish Biology 70:1057–1069. Schiavone, A., Jr. 1982. Age and growth of Bowfin in Butterfield Lake, New York. New York Fish and Game Journal 29:107. Schlosser, I. J. 1982. Fish community structure and function along two habitat gradients in a headwater stream. Ecological Monographs 52:395–414. Schlosser, I. J. 1987. A conceptual framework for fish communities in small warmwater streams, p. 17–24. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Schlosser, I. J. 1988a. Predation rates and the behavioral response of adult Brassy Minnows (Hybognathus hankinsoni) to Creek Chub and Smallmouth Bass predators. Copeia 1988:691–697. Schlosser, I. J. 1988b. Predation risk and habitat selection by two size classes of a stream cyprinid: experimental test of a hypothesis. Oikos 52:36–40. Schlosser, I. J. 1991. Stream fish ecology: a landscape perspective. BioScience 41:704–712. Schlosser, I. J., and K. K. Ebel. 1989. Effects of flow regime and cyprinid predation on a headwater stream. Ecological Monographs 59:41–57. Schlupp, I. 2005. The evolutionary ecology of gynogenesis. Annual Review of Ecology, Evolution and Systematics 36:399– 417. Schlupp, I. 2009. Behavior of fishes in the sexual/unisexual mating system of the Amazon Molly (Poecilia formosa). Advances in the Study of Behavior 39:153–183. Schlupp, I., C. Marler, and M. J. Ryan. 1994. Benefit to male Sailfin Mollies of mating with heterospecific females. Science 263:373–374. Schlupp, I., J. Parzefall, and M. Schartl. 1991. Male mate choice in mixed bisexual/unisexual breeding complexes of Poecilia (Teleostei: Poeciliidae). Ethology 88:215–222. Schlupp, I., and M. J. Ryan. 1997. Male Sailfin Mollies (Poecilia latipinna) copy the mate choice of other males. Behavioral Ecology 8:104–107. Schluter, D. 1994. Experimental evidence that competition promotes divergence in adaptive radiation. Science 266:798–801. Schluter, D. 2000. Ecological character displacement in adaptive radiation. The American Naturalist 156:S4–S16. Schluter, D., and J. D. McPhail. 1992. Ecological character displacement and speciation in Sticklebacks. The American Naturalist 140:85–108. Schluter, D., and J. D. McPhail. 1993. Character displacement and replicate adaptive radiation. Trends in Ecology and Evolution 8:197–200. Schmid, T. H. 1988. Age, growth, and movement patterns of the Atlantic Stingray, Dasyatis sabina, in a Florida coastal lagoon system. Unpubl. Master’s thesis. University of Central Florida, Orlando. Schmidt, J. 1912. The reproduction and spawning-places of the fresh-water Eel (Anguilla vulgaris). Nature 2234 (89):633–235. Schmidt, T. R. 1994. Phylogenetic relationships of the genus Hybognathus (Teleostei: Cyprinidae) Copeia 1994:622–630.

LITERATURE CITED

Schmidt, T. R., and J. R. Gold. 1995. Systematic affinities of Notropis topeka (Topeka Shiner) inferred from sequences of the cytochrome b gene. Copeia 1995:199–204. Schmidt, T. R., T. E. Dowling, and J. R. Gold. 1994. Molecular systematics of the genus Pimephales (Teleostei: Cyprinidae). The Southwestern Naturalist 39:241–248. Schmidt, T. R., J. P. Bielawski, and J. R. Gold. 1998. Molecular phylogenetics and evolution of the cytochrome b gene in the cyprinid genus Lythrurus (Actinopterygii: Cypriniformes). Copeia 1998:14–22. Schneberger, E. 1937. The food of small dogfish, Amia calva. Copeia 1937(1):61. Schneider, C. P., R. W. Owens, R. A. Bergstedt, and R. O’Gorman. 1996. Predation by Sea Lamprey (Petromyzon marinus) on Lake Trout (Salvelinus namaycush) in southern Lake Ontario, 1982–1992. Canadian Journal of Fisheries and Aquatic Sciences 53:1921–1932. Schoener, T. W. 1983. Field experiments on interspecific competition. The American Naturalist 122:240–285. Schofield, P. J., J. D. Williams, L. G. Nico, P. Fuller, and M. R. Thoomas. 2005. Foreign nonindigenous Carps and Minnows (Cyprinidae) in the United States—a guide to their identification, distribution, and biology. United States Geological Survey Scientific Investigations Report 2005-5041. Scholik, A. R., and H. Y. Yan. 2002a. Effects of boat engine noise on the auditory sensitivity of the Fathead Minnow, Pimephales promelas. Environmental Biology of Fishes 63:203–209. Scholik, A. R., and H. Y. Yan. 2002b. Effects of noise on the auditory sensitivity of the Bluegill Sunfish, Lepomis macrochirus. Comparative Biochemistry and Physiology A 133:43–52. Scholik, A. R., and H. Y. Yan. 2001. Effects of underwater noise on auditory sensitivity of a cyprinid fish. Hearing Research 152:17–24. Scholten, G. D. 2009. Management of commercial Paddlefish fisheries in the United States, p. 291–306. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Scholten, G. D., and P. W. Bettoli. 2005. Population characteristics and assessment of overfishing for an exploited Paddlefish population in the lower Tennessee River. Transactions of the American Fisheries Society 134:1285–1298. Scholten, G. D., and P. W. Bettoli. 2007. Lack of size selectivity for Paddlefish captured in hobbled gillnets. Fisheries Research 83:355–359. Schönhuth, S., I. Doadrio, O. Dominguez-Dominguez, D. M. Hillis, and R. L. Mayden. 2008. Molecular evolution of southern North American Cyprinidae (Actinopterygii), with the description of the new genus Tampichthys from Central Mexico. Molecular Phylogenetics and Evolution 47:729–756. Schönhuth, S., and R. L. Mayden. 2010. Phylogenetic relationships in the genus Cyprinella (Actinopterygii: Cyprinidae) based on mitochondrial and nuclear gene sequences. Molecular Phylogenetics and Evolution 55:77–98. Schrank, S. J., C. S. Guy, and J. F. Fairchild. 2003. Competitive interactions between age-0 Bighead Carp and Paddlefish. Transactions of the American Fisheries Society 132:1222–1228. Schreiber, D. C., and W. L. Minckley. 1982. Feeding interrelations of native fishes in a Sonoran Desert stream. Great Basin Naturalist 41:409–426.

603

Schrey, A. W., R. Boley, and E. J. Heist. 2011. Hydridization between Pallid Sturgeon Scaphirhynchus albus and Shovelnose Sturgeon Scaphirhynchus platorynchus. Journal of Fish Biology 79:1828–1850. Schrey, A. W., and E. J. Heist. 2007. Stock structure of Pallid Sturgeon analyzed with microsatellite loci. Journal of Applied Ichthyology 23:298–303. Schrey, A. W., B. L. Sloss, R. J. Sheehan, R. C. Heidinger, and E. J. Heist. 2007. Genetic discrimination of middle Mississippi River Scaphirhynchus Sturgeon into Pallid, Shovelnose, and putative hybrids with microsatellite loci. Conservation Genetics 8:683–693. Schroder, S. L. 1981. The role of sexual selection in determining overall mating patterns and mate choice in Chum Salmon. Unpubl. Ph.D. thesis, University of Washington, Seattle. Schroder, S. L. 1982. The influence of intrasexual competition on the distribution of Chum Salmon in an experimental stream, p. 275–285. In Salmon and Trout Migratory Behavior Symposium. E. L. Brannon and E. O. Salo (eds.). University of Washington, Seattle. Schultz, E. T., R. K. Cowen, K. M. M. Lwiza, and A. M. Gospodarek. 2000. Explaining advection: do larval Bay Anchovy (Anchoa mitchilli) show selective tidal-stream transport? ICES Journal of Marine Science 57:360–371. Schultz, E. T., K. M. M. Lwiza, M. C. Fencil, and J. M. Martin. 2003. Mechanisms promoting upriver transport of larvae of two fish species in the Hudson River Estuary. Marine Ecology Progress Series 251:263–277. Schultz, R. J. 1966. Hybridization experiments with an allfemale fish of the genus Poeciliopsis. Biological Bulletin 130:415– 429. Schultz, R. J. 1967. Gynogenesis and triploidy in the viviparous fish Poeciliopsis. Science 157:1564–1567. Schultz, R. J. 1969. Hybridization, unisexuality and polyploidy in the teleost Poeciliopsis (Poeciliidae) and other vertebrates. American Naturalist 103:605–619. Schultz, R., and R. Hermanutz. 1990. Transfer of toxic concentrations of selenium from parent to progeny in the Fathead Minnow (Pimephales promelas). Bulletin of Environmental Contamination and Toxicology 45:568–438. Schultze, H.-P., and G. Arratia. 1988. Reevaluation of the caudal skeleton of some actinopterygian fishes: II. Hiodon, Elops, and Albula. Journal of Morphology 195:257–303. Schulz, P. D., and D. D. Simons 1973. Fish species diversity in a prehistoric central California Indian midden. California Fish and Game 59:107–113. Schutz, F. 1956. Vergleichende Untersuchungen über die Schreckreaktion bei Fischen und deren Verbreitung. Zeitschrift für vergleichende Physiologie 38:84–135. Schwartz, F. J. 1963. The fresh-water fishes of Maryland. Maryland Conservation 40:19–29. Schwartz, F. J. 1981. Effects of freshwater runoff on fishes occupying the freshwater and estuarine coastal watersheds of North Carolina. United States Fish and Wildlife Ser vice Biological Ser vices Program FWS-OBS-81–04:282–294. Schwartz, F. J. 1990. Mass migratory congregations and movements of several species of cownose rays, genus Rhinoptera: a worldwide review. The Journal of the Elisha Mitchell Scientific Society 106:10–13.

604 LITERATURE CITED

Schwartz, F. J. 1995. Elasmobranchs frequenting fresh and low saline waters of North Carolina during 1971–1991. Journal Aquariculture and Aquatic Science 7:45–51. Schwartz, F. J. 2000. Elasmobranchs of the Cape Fear River, North Carolina. Journal of the Elisha Mitchell Scientific Society 116:206–224. Schwartz, F. J., and M. D. Dahlberg. 1978. Biology and ecology of the Atlantic Stingray, Dasyatis sabina (Pices: Dasyatidae) in North Carolina and Georgia. Northeast Gulf Science 2:1–23. Schwartz, F. J. and M. B. Maddock. 2002. Cytogenetics of the elasmobranch: genome evolution and phylogenetic implications. Marine and Freshwater Research 53:491–502. Schwemm, M. R., and P. J. Unmack. 2002. Native catostomids as prey of introduced Largemouth Bass in a central Arizona stream. In Proceedings of the Desert Fishes Council, 2001. Vol. XXXIII. D. A. Hendrickson and L. T. Findley (eds.). Sul Ross State University, Alpine, Texas. Scoppettone, G. G. 1988. Growth and longevity of the Cui-ui and longevity of other catostomids and cyprinids in western North America. Transactions of the American Fisheries Society 117:301–307. Scoppettone, G. G. 1993. Interactions between native and nonnative fishes of the Upper Muddy River, Nevada. Transactions of the American Fisheries Society 122:599–608. Scoppettone, G. G., M. E. Buettner, and P. H. Rissler. 1993. Effect of four fluctuating temperature regimes on Cui-ui, Chasmistes cujus, survival from egg fertilization to swim-up, and size of larvae produced. Environmental Biology of Fishes 38:373–378. Scoppettone, G. G., M. Coleman, and G. A. Wedemeyer. 1986. Life History and status of the endangered Cui-ui of Pyramid Lake, Nevada. United States Fish and Wildlife Ser vice, Fish and Wildlife Research 1:1–23. Scoppettone, G. G., P. H. Rissler, and M. E. Buettner. 2000. Reproductive longevity and fecundity associated with nonannual spawning in Cui-ui. Transactions of the American Fisheries Society 129:658–669. Scoppettone, G. G., P. H. Rissler, B. M. Nielsen, and J. E. Harvey. 1998. The status of Moapa coriacea and Gila seminuda and status information on other fishes of the Muddy River, Clark County, Nevada. The Southwestern Naturalist 43:115–122. Scoppettone, G. G., and G. L. Vinyard. 1991. Life history and management of four endangered lacustrine Suckers, p. 359–377. In Battle Against Extinction. W. L. Minckley and J. E. Deacon (eds.). The University of Arizona Press, Tuscon. Scoppettone, G. G., G. A. Wedemeyer, M. Coleman, and H. Burge. 1983. Reproduction by the endangered Cui-ui in the lower Truckee River. Transactions of the American Fisheries Society 112:788–793. Scott, A. P., N. R. Liley, and E. L. M. Vermeirssen. 1994. Urine of reproductively mature female Rainbow Trout, Oncorhynchus mykiss (Walbaum), contains a priming pheromone which enhances plasma levels of sex steroids and gonadotrophin II in males. Journal of Fish Biology 44:131–147. Scott, G. R., K. A. Sloman, C. Rouleau, and C. M. Wood. 2003. Cadmium disrupts behavioural and physiological responses to alarm substance in juvenile Rainbow Trout (Oncorhynchus mykiss). Journal of Experimental Biology 206:1779–1790. Scott, N. L. 1987. Seasonal variation of critical thermal maximum in the Redbelly Dace, Phoxinus erythrogaster (Cyprinidae). The Southwestern Naturalist 32:435–438.

Scott, W. B., and E. J. Crossman. 1973. Freshwater Fishes of Canada. Bulletin of the Fisheries Research Board of Canada 184:1–966. Scott, W. B., and M. G. Scott. 1988. Atlantic Fishes of Canada. Canadian Bulletin of Fisheries and Aquatic Sciences 219, University of Toronto Press, Toronto, Canada. Scribner, K. T., K. S. Page, and M. L. Barton. 2000. Hybridization in freshwater fishes: a review of case studies and cytonuclear methods of biological inference. Reviews in Fish Biology and Fisheries 10:293–323. Seale, A. 1896. Notes on Deltistes, a new genus of catostomoid fishes. Proceedings of the California Academy of Sciences (Series 2) 6:269. Seamons, T. R., P. Bentzen, and T. P. Quinn. 2004a. The mating system of steelhead, Oncorhynchus mykiss, inferred by molecular analysis of parents and progeny. Environmental Biology of Fishes 69:333–344. Seamons, T. R., P. Bentzen, and T. P. Quinn. 2004b. The effects of adult length and arrival date on individual reproductive success in wild steelhead trout (Oncorhynchus mykiss). Canadian Journal of Fisheries and Aquatic Sciences 61:193–204. Secor, D. H. 1999. Specifying divergent migrations in the concept of stock: the contingent Hypothesis. Fisheries Research 43:13–34. Secor, D. H., and T. E. Gunderson. 1998. Effects of hypoxia and temperature on survival, growth, and respiration of juvenile Atlantic Sturgeon, Acipenser oxyrinchus. Fishery Bulletin 96:603–613. Secor, D. H., P. J. Anders, W. Van Winkle, and D. A. Dixon. 2002. Can we study Sturgeons to extinction? What we do and don’t know about the conservation of North American Sturgeons. P. 3–10. In Biology, Management, and Protection of North American Sturgeon. W. Van Winkle, P. Anders, D. H. Secor, and D. Dixon (eds.). American Fisheries Society Symposium 28, Bethesda, Maryland. Secor, D. H., E. J. Niklitschek, J. T. Stevenson, T. E. Gunderson, S. P. Minkkinen, B. Richardson, B. Florence, M. Mangold, J. Skjeveland, and A. Henderson-Arzapalo. 2000. Dispersal and growth of yearling Atlantic Sturgeon, Acipenser oxyrinchus, released into Chesapeake Bay. Fishery Bulletin 98:800–810. SEDESOL. 1994. NORMA Oficial Mexicana NOM-059ECOL-1994, que determina las especies y subespecies de flora y fauna silvestres terrestres y aquáticas en peligro de extinción, amenazadas, raras y las sujetas a protección especial, y que establece especificaciones paras su protección. Diario Oficial, 16 May 1994, 488 (10):2–60. México. Segaar, J., and J. P. C. de Bruin. 1985. Behavioural consequences of chemoreception during reproductive cycles of the three-spined Stickleback (Gasterosteus aculeatus L.). Behaviour 93:139–149. Segaar, J., J. P. C. de Bruin, A. P. van der Meché, and M. E. van der Meché-Jacobi. 1983. Influence of chemical receptivity on reproductive behaviour of the male three-spined Stickleback (Gasterosteus aculeatus L.): an ethological analysis of cranial nerve functions regarding nest fanning activity and zigzag dance. Behaviour 86:100–166. Segal, E. 1987. Behavior of juvenile Nerocila acuminata (Isopoda, Cymothoidae) during attack, attachment and feeding on fish prey. Bulletin of Marine Science. 41:351–360. Seibert, J. R., Q. E. Phelps, S. J. Tripp, and J. E. Garvey. 2011. Seasonal diet composition of adult Shovelnose Sturgeon in

LITERATURE CITED

the middle Mississippi River. American Midland Naturalist 165:355–363. Seidensticker, E. P. 1987. Food selection of Alligator Gar and Longnose Gar in a Texas reservoir. Proceedings of the Annual Conference Southeastern Association of Fish and Wildlife Agencies 41:100–104. Seidensticker, E. P., and R. A. Ott, Jr. 1989. Comparison of gill nets and jug lines for selectively harvesting large Gar. Proceedings of the Annual Conference Southeastern Association of Fish and Wildlife Agencies 42:229–233. Self, J. T. 1954. Parasites of the Goldeye, Hiodon alosoides (Raf.), in Lake Texoma. The Journal of Parasitology 40:1–4. Selset, R., and K. B. Døving. 1980. Behavior of mature anadromous char (Salmo alpinus L.) towards odorants produced by smolts of their own population. Acta Physiologica Scandinavica 108:113–122. Semmens, K. J. 1985. Induced spawning of the Blue Sucker (Cycleptus elongatus). Progressive Fish Culturist 47:119–120. Semmens, K. J., and W. L. Shelton. 1986. Opportunities in Paddlefish aquaculture, p. 106–113. In The Paddlefish: Status, Management and Propagation. J. G. Dillard, L. K. Graham, and T. R. Russell (eds.). American Fisheries Society Special Publication 7. Sevenster-Bol, A. C. A. 1962. On the causation of drive reduction after a consummatory act (in Gasterosteus aculeatus L.). Archives Neerlandaises de Zoologie 15:175–236. Sevon, M. 1988. Nevada Department of Wildlife, Federal aid job completion report. Lahontan Reservoir: Job 102-F-20-20. Shaklee, J. B., M. J. Champion, and G. S. Whitt. 1974. Developmental genetics of teleosts: a biochemical analysis of Lake Chubsucker ontogeny. Developmental Biology 38:356–382. Shannon, C. E., and W. Weaver. 1949. The Mathematical Theory of Communication. University of Illinois Press, Champaign. Shao, B. B. 1997a. Nest association of Pumpkinseed, Lepomis gibbosus, and Golden Shiner, Notemigonus crysoleucas. Environmental Biology of Fishes 50:41–48. Shao, B. B. 1997b. Effects of Golden Shiner (Notemigonus crysoleucas) nest association on host Pumpkinseeds (Lepomis gibbosus): evidence for a non-parasitic relationship. Behavioral Ecology and Sociobiology 41:399–406. Sharman, A. C., and P. W. H. Holland. 1998. Estimation of Hox gene cluster number in Lampreys. International Journal of Developmental Biology 42:617–620. Shaw, K. A., E. O. Wiley, and T. A. Titus. 1995. Phylogenetic relationships among members of the Hybopsis amblops species group (Teleostei: Cyprinidae). Occasional Papers of the Museum of Natural History, University of Kansas 172:1–28. Sheldon, A. L. 1968. Species diversity and longitudinal succession in stream fishes. Ecology 49:193–198. Sheldon, A. L. 1987. Rarity: patterns and consequences for stream fishes, p. 203–209. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins (eds.). University of Oklahoma Press, Norman. Sheldon, A. L., and G. K. Meffe. 1993. Multivariate analysis of feeding relationships of fishes in blackwater streams. Environmental Biology of Fishes 37:161–171. Sheldon, W. W. 1974. Elvers in Maine: techniques of location, catching, and holding. Maine Department of Marine Resources, Augusta, Maine. Sheldon, W. W., and J. D. McCleave. 1985. Abundance of glass Eels of the American Eel, Anguilla rostrata, in mid-channel and

605

near shore during estuarine migration. Naturaliste Canadien 112:425–430. Shelford, V. C. 1911. Ecological succession. I. Stream fishes and the method of physiographic analysis. Biological Bulletin 21:9–35. Shiklomanov, I. A. 1993. World fresh water resources, p. 13–24. In Water in Crisis, a Guide to the World’s Fresh Water Resources. P. H. Gleick (ed.). Oxford University Press, New York. Shimeld, S. M., and P. W. H. Holland. 2000. Vertebrate innovations. Proceedings of the National Academy of Sciences of the United States of America 97:4449–4452. Shirai, S. 1996. Phylogenetic interrelationships of neoselachians (Chondrichthyes: Euselachii), p. 9–34. In Interrelationships of Fishes. M. L. Stiassny, L. R. Parenti, and G. D. Johnson (eds.). Academic Press, San Diego, California. Shirakashi, S., and C. P. Goater. 2002. Intensity-dependent alteration of minnow (Pimephales promelas) behavior by a brainencysting trematode. Journal of Parasitology 88:1071–1074. Shirakashi, S., and C. P. Goater. 2005. Chronology of parasiteinduced alteration of fish behaviour: effects of parasite maturation and host experience. Parasitology 130:177–183. Shireman, J. V., D. E. Colle, and D. F. DuRant. 1981. Efficiency of rotenone sampling with large and small block nets in vegetated and open-water habitats. Transactions of the American Fisheries Society 110:77–80. Shirley, D. S. 1983. Spawning ecology and larval development of the June Sucker. Proceedings of the Bonneville Chapter of the American Fisheries Society 1983:18–36. Shoup, D. E., R. E. Carlson, and R. T. Heath. 2004. Diel activity levels of centrarchid fishes in a small Ohio lake. Transactions of the American Fisheries Society 133:1264–1269. Shu, D.-G., S. Conway Morris, J. Han, Z.-F. Zhang, K. Yasui, P. Janvier, L. Chen, X.-L. Zhang, J.-N. Liu, Y. Li, and H.-Q. Liu. 2003. Head and backbone of the Early Cambrian vertebrate Haikouichthys. Nature 421:526–529. Shu, D.-G., H.-L. Luo, S. Conway Morris, X.-L. Zhang, S.-X. Hu, L. Chen, J. Han, M. Zhu, Y. Li, and L.-Z.Chen. 1999. Lower Cambrian vertebrates from south China. Nature 402:42–46. Shuman, D. A., R. A. Klumb, R. H. Wilson, M. E. Jaeger, T. Haddix, W. M. Gardner, W. J. Doyle, P. T. Horner, M. Ruggles, K. D. Steffensen, S. Stukel, and G. A. Wanner. 2011. Pallid Sturgeon size structure, condition, and growth in the Missouri River basin. Journal of Applied Ichthyology 27:269–281. Sibbing, F. A. 1982. Pharyngeal mastication and food transport in the carp (Cyprinus carpio L.): a cineradiographic and electromyographic study. Journal of Morphology 172:223–258. Sibbing, F. A. 1988. Specializations and limitations in the utilization of food resources by the carp, Cyprinus carpio: a study of oral food processing. Environmental Biology of Fishes 22:161–178. Sibbing, F. A. 1991a. Food capture and oral processing, p. 377–412. In Cyprinid Fishes: Systematics, Biology and Exploitation. I. J. Winfield and J. S. Nelson (eds.). Chapman & Hall, London. Sibbing, F. A. 1991b. Food Processing by mastication in cyprinid fish, p. 57–92. In Feeding and the Texture of Food. Vol. Society for Experimental Biology, Seminar series 44. J. F. V. Vincent and P. J. Lillford (eds.). Cambridge University Press, Cambridge, United Kingdom. Sibbing, F. A., and R. Uribe. 1985. Regional specializations in the oro-phayrngeal wall and food processing in the carp (Cyprinus carpio). Netherlands Journal of Zoology 35:377–422.

606

LITERATURE CITED

Sibbing, F. A., J. W. M. Osse, and A. Terlouw. 1986. Food handling in the carp (Cyprinus carpio): its movement patterns, mechanisms and limitations. Journal of Zoology, London A. 210:161–203. Siebert, D. J. 1987. Interrelationships among families of the order Cypriniformes (Teleostei). Unpubl. Ph.D. diss., The City University of New York, New York. Siefert, R. E. 1972. The first food of larval Yellow Perch, White Sucker, Bluegill, Emerald Shiner, and Rainbow Smelt. Transactions of the American Fisheries Society 101:219–225. Siefkes, M. J., R. A. Bergstedt, M. B. Twohey, and W. Li. 2003a. Chemosterilization of male Sea Lampreys (Petromyzon marinus) does not affect sex pheromone release. Canadian Journal of Fisheries and Aquatic Sciences 60:23–31. Siefkes, M. J., and W. Li. 2004. Electrophysiological evidence for detection and discrimination of pheromonal bile acids by the olfactory epithelium of female Sea Lampreys (Petromyzon marinus). Journal of Comparative Physiology A 190:193–199. Siefkes, M. J., S. R. Winterstein, and W. Li. 2005. Evidence that 3-keto petromyzonol sulphate specifically attracts ovulating female Sea Lamprey, Petromyzon marinus. Animal Behaviour 70:1037–1045. Siefkes, M. J., B. Zielinski, A. P. Scott, S.-S. Yun, and W. Li. 2003b. A novel release mechanism for a male Sea Lamprey sex pheromone. Biology of Reproduction 69:125–132. Sigler, W. F., and J. W. Sigler. 1987. Fishes of the Great Basin: A Natural History. University of Nevada Press, Reno. Sigler, W. F., and J. W. Sigler. 1996. Fishes of Utah: A Natural History. University of Utah Press, Salt Lake City. Sigler, W. F., S. Vigg, and M. Bres. 1985. Life history of the Cui-ui, Chasmistes cujus Cope, in Pyramid Lake, Nevada: a review. Western North American Naturalist 45:571–603. Sillman, A. J., and D. A. Dahlin. 2004a. The photoreceptors and visual pigments of Sharks and Sturgeons, p. 31–54. In The Senses of Fish, Adaptations for the Reception of Natural Stimuli. G. von der Emde, J. Mogdans, and B. G. Kapoor (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands, and Narosa Publishing House, New Delhi, India. Sillman, A. J., and D. A. Dahlin. 2004b. Photoreceptor topography in the duplex retina of Paddlefish (Polyodon spathula). Journal of Experimental Zoology 301A:674–681. Sillman, A. J., A. K. Beach, D. A. Dahlin, and E. R. Loew. 2005. Photoreceptors and visual pigments in the retina of the fully anadromous Green Sturgeon (Acipenser medirostris) and the potamodromous Pallid Sturgeon (Scaphirhynchus albus). Journal of Comparative Physiology A 191:799–811. Sillman, A. J., C. J. O’Leary, C. D. Tarantino, and E. R. Loew. 1999. The photoreceptors and visual pigments of two species of Acipenseriformes, the Shovelnose Sturgeon (Scaphirhynchus platorynchus) and the Paddlefish (Polyodon spathula). Journal of Comparative Physiology A 184:37–47. Sillman, E. I. 1962. The life history of Azygia longa (Leidy 1851) (Trematoda: Digenea), and notes on A. acuminata Goldberger 1911. Transactions of the American Microscopical Society 81:43–65. Silver, W. L. 1979. Olfactory responses from a marine elasmobranch, the Atlantic Stingray, Dasyatis sabina. Marine and Freshwater Behaviour and Physiology 6:297–305. Simanek, D. E. 1978. Genetic variability and population structure of Poecilia latipinna. Nature 276:612–614.

Simmons, E. G. 1957. An ecological survey of the upper Laguna Madre of Texas. Publications of the Institute of Marine Science 4:156–200. Simon, D. C., and D. F. Markle. 1997. Interannual abundance of nonnative Fathead Minnows (Pimephales promelas) in Upper Klamath Lake, Oregon. Great Basin Naturalist 57:142–148. Simon, D. C., and D. F. Markle. 1999. Evidence of a relationship between Smallmouth Bass (Micropterus dolomeiu) and decline of Umpqua Chub (Oregonichthys kalawatseti) in the Umpqua Basin, Oregon. Northwestern Naturalist 80:110–113. Simon, D. C., M. R. Terwilliger, P. Murtaugh, and D. F. Markle. 2000. Larval and juvenile ecology of Upper Klamath Lake Suckers: 1995–1998. Report to United States Bureau of Reclamation, Klamath Falls Office, Klamath Falls, Oregon. Simon, T. P. 1990. Family Amiidae, p. 89–97. In Reproductive Biology and Early Life History of Fishes in the Ohio River Drainage. Volume 1: Acipenseridae through Esocidae. R. Wallus and T. P. Simon (eds.). Tennessee Valley Authority, Chattanooga. Simon, T. P. 1997. Ontogeny of the darter subgenus Doration with comments on intrasubgeneric relationships. Copeia 1997: 60–69. Simon, T. P., and R. Wallus. 1989. Contributions to the early life histories of Gar (Actinopterygii: Lepisosteidae) in the Ohio and Tennessee River basins with emphasis on larval development. Transactions of the Kentucky Academy of Science 50:59–74. Simonović, P., S. Marić, and V. Nikolić. 2006. Occurrence of Paddlefish Polyodon spathula (Walbaum, 1792) in the Serbian part of the lower River Danube. Aquatic Invasions 1:183–185. Simons, A. M. 2004. Phylogenetic relationships in the genus Erimystax (Actinopterygii: Cyprinidae) based on the cytochrome b gene. Copeia 2004:351–356. Simons, A. M., P. B. Berendzen, and R. L. Mayden. 2003. Molecular systematics of North American phoxinin genera (Actinopterygii: Cyprinidae) inferred from mitochondrial 12S and 16S ribosomal RNA sequences. Zoological Journal of the Linnean Society 139:63–80. Simons, A. M., K. E. Knott, and R. L. Mayden. 2000. Assessment of the minnow genus Pteronotropis (Teleostei: Cyprinidae). Copeia 2000:1068–1075. Simons, A. M., and R. L. Mayden. 1997. Phylogenetic relationships of the creek chubs and spine-fins: an enigmatic group of North American cyprinid fishes (Actinopterygii: Cyprinidae). Cladistics 13:187–205. Simons, A. M., and R. L. Mayden. 1998. Phylogenetic relationships of the western North American phoxinins (Actinopterygii: Cyprinidae) as inferred from mitochondrial 12S and 16S ribosomal RNA sequences. Molecular Phylogenetics and Evolution 9:308–329. Simons, A. M., and R. L. Mayden. 1999. Phylogenetic relationships of North American cyprinids and assessment of homology of the open posterior myodome. Copeia 1999:13–21. Simons, A. M., R. M. Wood, L. S. Heath, B. R. Kuhajda, and R. L. Mayden. 2001. Phylogenetics of Scaphirhynchus based on mitochondrial DNA sequences. Transactions of the American Fisheries Society 130:359–366. Singer, T. D., and J. S. Ballantyne. 2004. Sturgeon and Paddlefish metabolism, p. 167–194. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands.

LITERATURE CITED

Singer, T. D., V. G. Mahadevappa, and J. S. Ballantyne. 1990. Aspects of the energy metabolism of Lake Sturgeon, Acipenser fulvescens, with special emphasis on lipid and ketone body metabolism. Canadian Journal of Fisheries and Aquatic Sciences 47:873–881. Sisneros, J. S., and T. C. Tricas. 2000. Androgen-induced changes in the response dynamics of ampullary electrosensory primary afferent neurons. Journal of Neuroscience 20:8586–8595. Sisneros, J. S., and T. C. Tricas. 2002. Ontogenetic changes in the response properties of the peripheral electrosensory system in the Atlantic Stingray (Dasyatis sabina). Brain, Behavior and Evolution 59:130–140. Sivak, J. G. 1975. The accommodative significance of the “ramp” retina of the eye of the stingray. Vision Research 16:945–950. Skalski, G. T., J. B. Landis, M. J. Grose, and S. P. Hudman. 2008. Genetic structure of Creek Chub, a headwater minnow, in an impounded river system. Transactions of the American Fisheries Society 137:962–975. Skelton, C. E. 2001. New dace of the genus Phoxinus (Cyprinidae: Cypriniformes) from the Tennessee River drainage, Tennessee. Copeia 2001:118–128. Slack, W. T., S. T. Ross, and J. A. Ewing, III. 2004. Ecology and population structure of the Bayou Darter, Etheostoma rubrum: disjunct riffle habitats and downstream transport of larvae. Environmental Biology of Fishes 71:151–164. Slatkin, M. 1987. Gene flow and the geographic structure of natural populations. Science 236:787–792. Slobodkin, L. B., F. E. Smith, and N. G. Hairston. 1967. Regulation in terrestrial ecosystems, and the implied balance of nature. The American Naturalist 101:109–124. Sloman, K. A., G. R. Scott, Z. Y. Diao, C. Rouleau, C. M. Wood, and D. G. McDonald. 2003. Cadmium affects the social behaviour of Rainbow Trout, Oncorhynchus mykiss. Aquatic Toxicology 65:171–185. Sloss, B. L., R. A. Klumb, and E. J. Heist. 2009. Genetic conservation and Paddlefish propagation, p. 307–327. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Smale, M. A., and C. F. Rabeni. 1995. Hypoxia and hyperthermia tolerances of headwater stream fishes. Transactions of the American Fisheries Society 124:698–710. Smallwood, W. M. 1916. Twenty months of starvation in Amia calva. Biological Bulletin 31:453–464. Smatresk, N. 1994. Respiratory control in the transition from water to air breathing in vertebrates. American Zoologist 34:264–279. Smatresk, N. J., and S. Q. Azizi. 1987. Characteristics of lung mechanoreceptors in Spotted Gar, Lepisosteus oculatus. American Journal of Physiology—Regulatory, Integrative and Comparative Physiology 252:R1066–R1072. Smatresk, N. J., M. L. Burleson, and S. Q. Azizi. 1986. Chemoreflexive responses to hypoxia and NaCN in Longnose Gar: evidence for two chemoreceptor loci. American Journal of Physiology 251:116–125. Smatresk, N. J., and J. N. Cameron. 1982a. Respiration and acidbase physiology of the Spotted Gar, a bimodal breather. I. Normal values and the response to severe hypoxia. Journal of Experimental Biology 96:263–280. Smatresk, N. J., and J. N. Cameron. 1982b. Respiration and acidbase physiology of the Spotted Gar, a bimodal breather. III. Re-

607

sponse to a transfer from freshwater to 50% sea water and control of ventilation. Journal of Experimental Biology 96:295–306. Smatresk, N. J., and J. N. Cameron. 1982c. Respiration and acidbase physiology of the Spotted Gar, a bimodal breather. II. Responses to temperature change and hypercapnia. Journal of Experimental Biology 96:281–293. Smit, H. 1965. Some experiments on the oxygen consumption of Goldfish (Carassius auratus L.) in relation to swimming speed. Canadian Journal of Zoology 43:623–633. Smith, B. G. 1923. Notes on the nesting habits of Cottus. Papers of the Michigan Academy of Sciences, Arts, and Letters 2:221–224. Smith, B. R. 1971. Sea Lampreys in the Great Lakes of North America, p. 207–247. In The Biology of Lampreys. Vol. 1. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Smith, B. R., and J. J. Tibbles. 1980. Sea Lamprey (Petromyzon marinus) in Lakes Huron, Michigan, and Superior: history of invasion and control, 1936–78. Canadian Journal of Fisheries and Aquatic Sciences 37:1780–1801. Smith, C. C., and R. C. Sargent. 2006. Female fitness declines with increasing female density but not male harassment in the Western Mosquitofish, Gambusia affinis. Animal Behaviour 71:401–407. Smith, C. L. 1962. Some Pliocene fishes from Kansas, Oklahoma, and Nebraska. Copeia 1962:505–520. Smith, C. L., and C. R. Powell. 1971. The summer fish communities of Brier Creek, Marshall County, Oklahoma. American Museum Novitates, No. 2458:1–30. Smith, C. T., R. J. Nelson, C. C. Wood, and B. F. Koop. 2001. Glacial biogeography of North American Coho Salmon (Oncorhynchus kisutch) Molecular Ecology 10:2775–2785. Smith, C. T., R. J. Nelson, S. Pollard, E. Rubidge, S. J. McKay, J. Rodzen, B. May, and B. Koop. 2002b. Population genetic analysis of White Sturgeon (Acipenser transmontanus) in the Fraser River. Journal of Applied Ichthyology 18:307–312. Smith, G. R. 1966. Distribution and evolution of the North American catostomid fishes of the subgenus Pantosteus, genus Catostomus. Miscellaneous Publications of the Museum of Zoology University of Michigan 129:1–132. Smith, G. R. 1975. Fishes of the Pliocene Glenns Ferry formation, southwest Idaho. University of Michigan Museum of Paleontology Paper on Paleontology 14:1–68. Smith, G. R. 1981. Late Cenozoic freshwater fishes of North America. Annual Review of Ecology and Systematics 12:163–193. Smith, G. R. 1992. Phylogeny and biogeography of the Catostomidae, freshwater fishes of North America and Asia, p. 778–813. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Smith, G. R., C. Badgley, T. P. Eiting, and P. S. Larson. 2010. Species diversity gradients in relation to geological history in North American freshwater fishes. Evolutionary Ecology Research 12:693–726. Smith, G. R., and T. E. Dowling. 2008. Correlating hydrographic events and divergence times of Speckled Dace (Rhinichthys: Teleostei: Cyprinidae) in the Colorado River drainage, p. 301–317. In Late Cenozoic Drainage History of the Southwestern Great Basin and Lower Colorado River Region: Geologic and Biotic Perspec-

608

LITERATURE CITED

tives. M. C. Reheis, R. Hershler, and D. M. Miller (eds.). Geological Society of America, Special Paper 439, Boulder, Colorado. Smith, G. R., T. E. Dowling, K. W. Gobalet, T. Lugaski, D. K. Shiozawa, and R. P. Evans. 2002. Biogeography and timing of evolutionary events among Great Basin fishes, p. 175–234. In Great Basin Aquatic Systems History. R. Hershler, D. B. Madsen, and D. R. Currey (eds.). Smithsonian Contributions to the Earth Sciences, No. 33. Smith, G. R., J. G. Hall, R. K. Koehn, and D. J. Innes. 1983. Taxonomic relationships of the Zuni Mountain Sucker, Catostomus discobolus yarrowi. Copeia 1983:37–48. Smith, G. R., and J. G. Lundberg. 1972. The Sand Draw fish fauna. Bulletin of the American Museum of Natural History 148:40–54. Smith, G. R., R. R. Miller, and W. D. Sable. 1979. Species relationships among fishes of the genus Gila in the upper Colorado River drainage. Proceedings of the First Conference on Scientific Research in the National Parks 1:613–623. Smith, G. R., and R. F. Stearley 1989. The classification and scientific names of Rainbow and cutthroat trouts. Fisheries 14:4–10. Smith, G. R., and T. N. Todd. 1984. Evolution of species flocks in northern temperate lakes, p. 45–68. In Evolution of Species Flocks. A. A. Echelle and I. Kornfield (eds.). University of Maine at Orono Press. Smith, H. M. 1898. Statistics of the fisheries of the interior waters of the United States. Appendix 11, p. 489–574. In U.S. Commission of Fish and Fisheries. Part XXII. Report of the Commissioner for the Year Ending June 30, 1896. A Report to the 55th Congress, 1st Session, House of Representatives Document 32. Government Printing Office, Washington, D.C. Smith, J. J., F. Antonacci, E E. Eichler, and C. T. Amemiya. 2009a. Programmed loss of millions of base pairs from a vertebrate genome. Proceedings of the National Academy of Sciences, 106: 11212–11217. Smith, J. J., A. O. H. C. Leduc, and G. E. Brown. 2008. Chemically mediated learning in juvenile Rainbow Trout. Does predator odour pH influence intensity and retention of acquired predator recognition? Journal of Fish Biology 72:1750–1760. Smith, J. P. 1997. Nesting season food habits of 4 species of herons and egrets at Lake Okeechobee, Florida. Colonial Waterbirds 20:198–220. Smith, K. M., and D. K. King. 2005a. Dynamics and extent of larval Lake Sturgeon Acipenser fulvescens drift in the upper Black River, Michigan. Journal of Applied Ichthyology 21:161–168. Smith, K. M., and D. K. King. 2005b. Movement and habitat use of yearling and juvenile Lake Sturgeon in Black River, Michigan. Transactions of the American Fisheries Society 134:1159–1172. Smith, L. L., D. A. Steen, J. M. Stober, M. C. Freeman, S. W. Golladay, L. M. Conner, and J. Cochrane. 2006. The vertebrate fauna of Ichauway, Baker County, GA. Southeastern Naturalist 5:599–620. Smith, M. L., T. M. Cavender, and R. R. Miller. 1975. Climatic and biogeographic significance of a fish fauna from the late Pliocene-early Pleistocene of the Lake Chapala Basin (Jalisco, Mexico). University of Michigan Museum of Paleontology Paper on Paleontology 12:29–38. Smith, M. W., and J. W. Saunders. 1955. The American Eel in certain freshwaters of the maritime provinces of Canada. Journal of the Fisheries Research Board of Canada 12:238–269.

Smith, N. A., R. E. Condrey, and B. C. Reed. 2009b. The feeding ecology of Paddlefish in the Mermentau River, Louisiana, p. 51– 62. In Paddlefish Management, Propagation, and Conservation in the 21st Century: Building from 20 Years of Research and Management. C. P. Paukert and G. D. Scholten (eds.). American Fisheries Society Symposium 66, Bethesda, Maryland. Smith, O. A. 2008. Reproductive potential and life history of Spotted Gar (Lepisosteus oculatus) in the Upper Barataria Estuary. Unpubl. Master’s thesis, Nicholl’s State University, Thibodaux, Louisiana. Smith, P. W. 1979. The Fishes of Illinois. University of Illinois Press, Urbana. Smith, R. J. F. 1973. Testosterone eliminates alarm substance in male Fathead Minnows. Canadian Journal of Zoology 51:875–876. Smith, R. J. F. 1974. Effects of 17α-methyltestosterone on the dorsal pad and tubercles of Fathead Minnows (Pimephales promelas). Canadian Journal of Zoology 52:1031–1038. Smith, R. J. F. 1976a. Seasonal loss of alarm substance cells in North American cyprinoid fishes and its relation to abrasive spawning behaviour. Canadian Journal of Zoology 54:1172–1182. Smith, R. J. F. 1976b. Male Fathead Minnows (Pimephales promelas Rafinesque) retain their fright reaction to alarm substance during the breeding season. Canadian Journal of Zoology 54:2230–2231. Smith, R. J. F. 1978. Seasonal changes in histology of gonads and dorsal skin of Fathead Minnow, Pimephales promelas. Canadian Journal of Zoology 56:2103–2109. Smith, R. J. F. 1979. Alarm reaction of Iowa and Johnny Darters (Etheostoma, Percidae, Pisces) to chemicals from injured conspecifics. Canadian Journal of Zoology 57:1278–1282. Smith, R. J. F. 1981. Effect of food deprivation on the reaction of Iowa Darters (Etheostoma exile) to skin extract. Canadian Journal of Zoology 59:558–560. Smith, R. J. F. 1982. Reaction of Percina nigrofasciata, Ammocrypta beanii, and Etheostoma swaini (Percidae, Pisces) to conspecific and intergeneric skin extracts. Canadian Journal of Zoology 60:1067–1072. Smith, R. J. F. 1992. Alarm signals in fishes. Reviews in Fish Biology and Fisheries 2:33–63. Smith, R. J. F., and B. D. Murphy. 1974. Functional morphology of dorsal pad in Fathead Minnows (Pimephales promelas Rafinesque). Transactions of the American Fisheries Society 103:65–72. Smith, R. J. F., and J. D. Smith. 1983. Seasonal loss of alarm substance cells in Chrosomus neogaeus, Notropis venustus and N. whipplei. Copeia 1983:822–826. Smith, R. K. 1999. Differential stability of spawning microhabitats of warmwater stream fishes. Unpubl. Master’s thesis, Virginia Tech, Blacksburg. Smith, R. S., and D. L. Kramer. 1986. The effect of apparent predation risk on the respiratory behavior of the Florida Gar (Lepisosteus platyrhincus). Canadian Journal of Zoology 64:2133–2136. Smith, T. I. J. 1985. The fishery, biology, and management of Atlantic Sturgeon, Acipenser oxyrhynchus, in North America, p. 61–72. In North American Sturgeons: Biology and Aquaculture Potential. F. P. Binkowski and S. I. Doroshov (eds.). Dr. W. Junk Publishers, Dordrecht, The Netherlands.

LITERATURE CITED

Smith, T. I. J., E. K. Dingley, and D. E. Marchette. 1980. Induced spawning and culture of Atlantic Sturgeon. The Progressive Fish-Culturalist 42:147–151. Smith, T. I. J., D. E. Marchette, and G. F. Ulrich. 1984. The Atlantic Sturgeon fishery in South Carolina. North American Journal of Fisheries Management 4:164–176. Smith, T. I. J., J. W. McCord, M. R. Collins, and W. C. Post. 2002a. Occurrence of stocked Shortnose Sturgeon Acipenser brevirostrum in non-target rivers. Journal of Applied Ichthyology 18:470–474. Smogor, R. A., P. L. Angermeier, and C. K. Gaylord. 1995. Distribution and abundance of American Eels in Virginia streams: tests of null models across spatial scales. Transactions of the American Fisheries Society 124:789–803. Smolka, A. J., E. R. Lacy, L. Luciano, and E. Reale. 1994. Identification of gastric H,K-ATPase in an early vertebrate, the Atlantic Stingray Dasyatis sabina. The Journal of Histochemistry and Cytochemistry 42:1323–1332. Snedden, G. A., W. E. Kelso, and D. A. Rutherford. 1999. Diel and seasonal patterns of Spotted Gar movement and habitat use in the lower Atchafalaya River Basin, Louisiana. Transactions of the American Fisheries Society 128:144–154. Snelson, F. F. 1976. A study of a diverse coastal ecosystem on the Atlantic coast of Florida. NASA report NGR 10-019-004. Snelson, F. F., Jr., L. E. L. Rasmussen, M. R. Johnson, and D. L. Hess. 1997. Serum concentrations of steroid hormones during reproduction in the Atlantic Stingray, Dasyatis sabina. General and Comparative Endocrinology 108:67–79. Snelson, F. F., Jr., and S. E. Williams. 1981. Notes on the occurrence, distribution, and biology of elasmobranch fishes in the Indian River Lagoon system, Florida. Estuaries 4:110–120. Snelson, F. F. Jr., S. E. Williams-Hooper, and T. H. Schmid. 1988. Reproduction and ecology of the Atlantic Stingray, Dasyatis sabina, in Florida coastal lagoons. Copeia 1988:729–739. Snyder, D. E. 2002. Pallid and Shovelnose Sturgeon larvae— morphological description and identification. Journal of Applied Ichthyology 18:240–265. Snyder, D. E., and S. C. Douglas. 1978. Description and identification of Mooneye, Hiodon tergisus, protolarvae. Transactions of the American Fisheries Society 107:590–594. Snyder, D. E., and S. M. Meismer. 1997. Effectiveness of light traps for capture and retention of larval and early juvenile Xyrauchen texanus and larval Ptychocheilus lucius and Gila elegans. United States National Park Ser vice, Final report to Cooperative Parks Study Unit, Fort Collins, Colorado. Snyder, D. E., and R. T. Muth. 2004. Catostomid fish larvae and early juveniles of the Upper Colorado River basin—morphological descriptions, comparisons, and computer-interactive key. Colorado Division of Wildlife, Technical Publication 42, Fort Collins. Snyder, J. O. 1917. The fishes of the Lahontan system of Nevada and northeastern California. Bulletin of the United States Bureau of Fisheries 35:33–86. Sola, C. 1995. Chemoattraction of upstream migrating glass Eels Anguilla anguilla to earthy and green odorants. Environmental Biology of Fishes 43:179–185. Sola, C., and L. Tosi. 1993. Bile salts and taurine as chemical stimuli for glass Eels, Anguilla anguilla: a behavioural study. Environmental Biology of Fishes 37:197–204.

609

Sola, C., and P. Tongiorgi. 1998. Behavioural responses of glass Eels of Anguilla anguilla to non-protein amino acids. Journal of Fish Biology 53:1253–1262. Sola, L., S. Cataudella and E. Capanna. 1981. New developments in vertebrate cytotaxonomy III. Karyology of bony fishes: a review. Genetica 54:285–328. Sola, L., G. Gentili, and S. Cataudella. 1980. Eel chromosomes: cytotaxonomical interrelationships and sex chromosomes. Copeia 1980:911–913. Solomon, D. J., and M. H. Beach. 2004. Fish pass design for Eel and elver. Environment Agency (UK) R&D Technical Report W2-070/TR1. Soltis, D. E., A. B. Morris, J. S. McLachlan, P. S. Manos, and P. S. Soltis. 2006. Comparative phylogeography of unglaciated eastern North America. Molecular Ecology 15:4261–4293. Sommer, T., R. Baxter, and B. Herbold. 1997. Resilience of Splittail in the Sacramento-San Joaquin Estuary. Transactions of the American Fisheries Society 126:961–976. Sommer, T., W. C. Harrell, M. Zoltan, and F. Frederick. 2008. Habitat associations and behavior of adult and juvenile Splittail (Cyprinidae: Pogonichthys macrolepidotus) in a managed seasonal floodplain wetland. San Francisco Estuary and Watershed Science 6:1–16. Sommerfeld, R. D., T. Boehm, and M. Milinski. 2008. Desynchronizing male and female seasonality: dynamics of male MHCindependent olfactory attractiveness in Sticklebacks. Ethology, Ecology and Evolution 20:325–336. Song, J. K., and R. G. Northcutt. 1991a. The primary projections of the lateral-line nerves of the Florida Gar, Lepisosteus platyrhincus. Brain, Behavior and Evolution 37:38–63. Song, J. K., and R. G. Northcutt. 1991b. Morphology, distribution and innervation of the lateral-line receptors of the Florida Gar, Lepisosteus platyrhincus. Brain, Behavior and Evolution 37:10–37. Sorensen, P. W. 1986. Origins of the freshwater attractant(s) of migrating elvers of the American Eel, Anguilla rostrata. Environmental Biology of Fishes 17:185–200. Sorensen, P. W. 2009. Stream water creates a discernible odor gradient that migratory juvenile American Eels may follow inshore. American Fisheries Society Symposium 69:841–844. Sorensen, P. W., and M. L. Bianchini. 1986. Environmental correlates of the freshwater migration of elvers of the American Eel in a Rhode Island brook. Transactions of the American Fisheries Society 115:258–268. Sorensen, P. W., J. R. Cardwell, T. E. Essington, and D. E. Weigel. 1995b. Reproductive interactions between sympatric Brook and Brown Trout in a small Minnesota stream. Canadian Journal of Fisheries and Aquatic Sciences 52:1958–1965. Sorensen, P. W., J. M. Fine, V. Dvornikovs, C. S. Jeff rey, F. Shao, J. Wang, L. A. Vrieze, K. R. Anderson, and T. R. Hoye. 2005. Mixture of new sulfated steroids functions as a migratory pheromone in the Sea Lamprey. Nature Chemical Biology 1:324–328. Sorensen, P. W., and T. R. Hoye. 2007. A critical review of the discovery and application of a migratory pheromone in an invasive fish, the Sea Lamprey Petromyzon marinus L. Journal of Fish Biology 71(Suppl. D):100–114. Sorensen, P. W., C. A. Murphy, K. Loomis, P. Maniak, and P. Thomas. 2004. Evidence that 4-pregnen-17,20α,21-triol-3-one functions as a maturation-inducing hormone and pheromonal

610

LITERATURE CITED

precursor in the percid fish, Gymnocephalus cernuus. General and Comparative Endocrinology 139:1–11. Sorensen, P. W., and J. Caprio. 1998. Chemoreception. In The Physiology of Fishes. D. H. Evans (ed.). CRC Press, Boca Raton, Florida. Sorenson, P. W., and A. P. Scott. 1994. The evolution of hormonal sex pheromones in teleost fish: poor correlation between the pattern of steroid release by Goldfish and olfactory sensitivity suggests that these cues evolved as a result of chemical spying rather than signal specialization. Acta Physiologica Scandinavica 152:191–205. Sorensen, P. W., A. P. Scott, N. E. Stacey, and L. Bowdin. 1995a. Sulfated 17,20(-Dihydroxy-4-pregnen-3-one functions as a potent and specific olfactory stimulant with pheromonal actions in the Goldfish. General and Comparative Endocrinology 100:128–142. Sorensen, P. W., and L. A. Vrieze. 2003. The chemical ecology and potential application of the Sea Lamprey migratory pheromone. Journal of Great Lakes Research 29(Suppl. 1):66–84. Sorenson, P. W., and H. E. Winn. 1984. The induction of maturation and ovulation in American Eels, Anguilla rostrata (LeSueur), and the relevance of chemical and visual cues to male spawning behaviour. Journal of Fish Biology 25:261–268. Southall, P. D., and W. A. Hubert. 1984. Habitat use by adult Paddlefish in the upper Mississippi River. Transactions of the American Fisheries Society 113:125–131. Southwood, T. R. E., 1977. Habitat, the templet for ecological strategies. Journal of Animal Ecology 46:337–365. Southwood, T. R. E. 1988. Tactics, strategies and templets. Oikos 52:3–18. Sower, S. A. 2003. The endocrinology of reproduction in Lampreys and applications for male Lamprey sterilization. Journal of Great Lakes Research 29(Suppl. 1):50–65. Sower, S. A., M. Freamat, and S. I. Kavanaugh. 2009. Insight from lamprey genomics: brain and pituitary reproductive hormones of lampreys, p. 57–70. In Biology, Management, and Conservation of Lampreys in North America. L. R. Brown, S. D. Chase, M. G. Mesa, R. J. Beamish, and P. B. Moyle (eds.). American Fisheries Society, Symposium 72, Bethesda, Maryland. Sparrowe, R. D. 1986. Threats to Paddlefish habitat, p. 36–45. In The Paddlefish: Status, Management and Propagation. J. G. Dillard, L. K. Graham, and T. R. Russell (eds.). American Fisheries Society Special Publication 7. Spear, B. J. 2007. U.S. management of Atlantic Sturgeon, p. 339– 346. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Speers-Roesch, B., Y. K. Ip and J. S. Ballantyne. 2006. Metabolic organization of freshwater, euryhaline, and marine elasmobranchs: implications for the evolution of energy metabolism in Sharks and Rays. The Journal of Experimental Biology 209:2495–2508. Spitzer, M. 2010. Season of the Gar: Adventures in Pursuit of America’s Most Misunderstood Fish. University of Arkansas Press, Fayetteville. Spotila, J. R., K. M. Terpin, R. R. Koons, and R. L. Benati. 1979. Temperature requirements of fishes from eastern Lake Erie and

the Upper Niagara River: a review of the literature. Environmental Biology of Fishes 4:281–307. Springer, S. 1960. Natural history of the Sandbar Shark, Eulamia milberti. United States Fish and Wildlife Ser vice, Fisheries Bulletin 61:1–38. Sprules, W. M. 1947. A management program for Goldeye (Amphiodon alosoides) in Manitoba’s marsh regions. Canadian FishCult. 2:9–11. Stabell, O. B. 1982. Detection of natural odorants by Atlantic Salmon parr using positive rheotaxis olfactometry, p. 71–78. In Proceedings of the Salmon and Trout Migratory Behaviour Symposium 1981. E. L. Brannon and E. O. Salo (eds.). School of Fisheries, University of Washington, Seattle. Stabell, O. B. 1987. Intraspecific pheromone discrimination and substrate marking by Atlantic Salmon parr. Journal of Chemical Ecology 13:1625–1643. Stacy, G., R. O. Smitherman, and J. W. Avault, Jr. 1970. Food habits of the Bowfin in Lacassine National Wildlife Refuge and other locations in southern Louisiana. The Progressive Fish Culturist 32:153–157. Stacey, N. E., J. R. Cardwell, N. R. Liley, A. P. Scott, and P. W. Sorensen. 1994. Hormones as sex pheromones in fish, p. 438–448. In Perspectives in Comparative Endocrinology. K. G. Davey, R. E. Peter, and S. S. Tobe (eds.). National Research Council, Ottawa. Stacey, N. E., D. S. MacKenzie, T. Marchant, A. L. Kyle, and R. E. Peter. 1984. Endocrine changes during natural spawning in the White Sucker, Catostomus commersoni. I. Gonadotropin, growth hormone, and thyroid hormones. General and Comparative Endocrinology 56:333–348. Stacey, N. E., and P. W. Sorensen. 1991. Function and evolution of fish hormonal pheromones, p. 109–135. In Biochemistry and Molecular Biology of Fishes. Vol. 1. P. L. Hochachka and T. P. Mommsen (eds.). Elsevier Science Publishers, Amsterdam, The Netherlands. Stachowicz, J. J. 2001. Mutualism, facilitation, and the structure of ecological communities. BioScience 51:235–246. Staddon, J. E. R. 1975. Note on evolutionary significance of supernormal stimuli. American Naturalist 109:541–545. Stafford, C. R., R. L. Richards, and C. M. Anslinger. 2000. The bluegrass fauna and changes in middle Holocene huntergatherer foraging in the southern Midwest. American Antiquity 65:317–336. Stahl, M. T., G. W. Whitledge, and A. M. Kelly. 2009. Reproductive biology of middle Mississippi River Shovelnose Sturgeon: insights from seasonal and age variation in plasma sex steroid and calcium concentrations. Journal of Applied Ichthyology 25(Supplement 2):75–82. Stahlberg, S., and P. Peckmann. 1978. The critical swimming speed of small teleost fish species in a flume. Archiv für Hydrobiologie 110:179–193. Stamford, M. D., and E. B. Taylor. 2004. Phylogeographical lineages of Arctic Grayling (Thymallus arcticus) in North America: divergence, origins and affinities with Eurasian Thymallus. Molecular Ecology 13:1533–1549. Stancill, W., G. R. Jordan, and C. P. Paukert. 2002. Seasonal migration patterns and site fidelity of adult Paddlefish in Lake Francis Case, Missouri River. North American Journal of Fisheries Management 22:815–824.

LITERATURE CITED

Starnes, L. B., and W. C. Starnes. 1981. Biology of the Blackside Dace Phoxinus cumberlandensis. American Midland Naturalist 106:360–371. Starnes, W. C., and R. E. Jenkins. 1988. A new cyprinid fish of the genus Phoxinus (Pisces: Cyprinodontiformes) from the Tennessee River drainage with comments on relationships and biogeography. Proceedings of the Biological Society of Washington 101:517–529. Starostka, V. J., and R. L. Applegate. 1970. Food selectivity of Bigmouth Buffalo, Ictiobus cyprinellus, in Lake Poinsett, South Dakota. Transactions of the American Fisheries Society 99:571–576. Stasiak, R. H. 1978. Reproduction, age and growth of the Finescale Dace, Chrosomus neogaeus, in Minnesota. Transactions of the American Fisheries Society 107:720–723. Stauffer, J. R., Jr., C. H. Hocutt, and R. F. Denoncourt. 1979. Status and distribution of the hybrid Nocomis micropogon × Rhinichthys cataractae, with a discussion of hybridization as a viable mode of vertebrate speciation. American Midland Naturalist 101:355–365. Stauffer, J. R., Jr., C. H. Hocutt, and R. L. Mayden. 1997. Pararhinichthys, a new monotypic genus of minnows (Teleostei: Cyprinidae) of hybrid origin from eastern North America. Ichthyological Exploration of Freshwaters 7:327–336. Stearns, S. C. 1983. The evolution of life-history traits in mosquitofish since their introduction to Hawaii in 1905: rates of evolution, heritabilities, and developmental plasticity. American Zoologist 23:65–75. Steele, R. G., and M. H. A. Keenleyside. 1971. Mate selection in two species of Sunfish (Lepomis gibbosus and L. megalotis peltastes). Canadian Journal of Zoology 49:1541–1548. Steffensen, K. D., L. A. Powell, and J. D. Koch. 2010. Assessment of hatchery-reared Pallid Sturgeon survival in the lower Missouri River. North American Journal of Fisheries Management 30:671–678. Stefferud, J. A., K. B. Gido, and D. L. Propst. 2011. Spatially variable response of native fish assemblages to discharge, predators and habitat characteristics in an arid-land river. Freshwater Biology 56:1403–1416. Stein, A. B., K. D. Friedland, and M. Sutherland. 2004. Atlantic Sturgeon marine bycatch and mortality on the Continental Shelf of the northeast United States. North American Journal of Fisheries Management 24:171–183. Stephens, C. M., and R. L. Mayden. 1998. Description of antagonistic and courtship behaviors of the Tricolor Shiner, Cyprinella trichroistia, and the Tallapoosa Shiner, Cyprinella gibbsi, with recommendations for the conservation of the Blue Shiner, Cyprinella caerulea. The Journal of the Elisha Mitchell Scientific Society 114:209–214. Stephens, C. M., and R. L. Mayden. 1999. Threatened fishes of the world: Cyprinella caerulea Jordan, 1877 (Cyprinidae). Environmental Biology of Fishes 55:264. Sterner, R., and N. B. George. 2000. Carbon, nitrogen, and phosphorus stoichiometry of cyprinid fishes. Ecology 81:127–140. Stevenson, M. M., and T. M. Buchanan. 1973. An analysis of hybridization between the cyprinodont fishes Cyprinodon variegatus and C. elegans. Copeia 1973:682–692. Stevenson, R. A., Jr. 1958. A biology of the Anchovies, Anchoa mitchilli mitchilli Cuvier and Valenciennes 1848 and Anchoa hep-

611

setus hepsetus (Linnaeus 1758). Unpubl. Master’s thesis, University of Delaware, Newark. not seen Stevenson, J. T., and D. H. Secor. 1999. Age determination and growth of Hudson River Atlantic Sturgeon, Acipenser oxyrinchus. Fishery Bulletin 97:153–166. Stewart, A. R., S. N. Luoma, C. E. Schlekat, M. A. Doblin, K. A. Hieb. 2004. Food web pathway determines how selenium affects aquatic ecosystems: a San Francisco Bay case study. Environmental Science and Technology 38:4519–4526. Stewart, D. J., J. F. Kitchell, and L. B. Crowder. 1981. Forage fishes and their salmonid predators in Lake Michigan. Transactions of the American Fisheries Society 110:751–763. Stewart, J. G., C. S. Schieble, R. C. Cashner, and V. A. Barko. 2005. Long-term trends in the Bogue Chitto River fish assemblage: a 27-year perspective. Southeastern Naturalist 4: 261–272. Stewart, N. H. 1926. Development, growth and food habits of the White Sucker (Catostomus commersonii). United States Fisheries Bulletin 42:147–184. Stock, D. W., and G. S. Whitt. 1992. Evidence from 18S ribosomal RNA sequences that Lampreys and Hagfishes form a natural group. Science 257:787–789. Stockard, C. R. 1907. Observations on the natural history of Polyodon spathula. The American Naturalist 41:753–766. Stockley, P., J. B. Searle, D. W. MacDonald, and C. S. Jones. 1993. Female multiple mating behaviour in the Common Shrew as a strategy to reduce inbreeding. Proceedings of the Royal Society of London B 254:173–179. Stone, A. R., and D. L. Hawksworth (eds.). 1986. Coevolution and systematics. The Systematics Association Special Volume 32. Clarendon Press-Oxford University Press, New York. Stone, L., T. Dayan, and D. Simberloff. 2000. On desert rodents, favored states, and unresolved issues: scaling up and down regional assemblages and local communities. The American Naturalist 156:322–328. Stone, N., L. Dorman, and H. Thomforde. 2009. Baitfish industry. The Encyclopedia of Arkansas History & Culture. Central Arkansas Library System. Available from http://www.encyclopediaofarkansas.net/encyclopedia; as of June 2010. Stout, J. F. 1959. The reproductive behavior and sound production of the Satinfin Shiner. Anatomical Record 134:643–64. Stout, J. F. 1960. The significance of sound production during the reproductive behavior of Notropis analostanus. Anatomical Record 138:384–385. Stout, J. F. 1963. The significance of sound production during the reproductive behaviour of Notropis analostanus (Family Cyprinidae). Animal Behaviour 11:83–92. Stout, J. F. 1975. Sound communication during the reproductive behavior of Notropis analostanus (Pisces: Cyprinidae). American Midland Naturalist 94:296–325. Stout, J., and H. E. Winn. 1958. The reproductive behavior and sound production of the Satinfin Shiner. Anatomical Record 132:511. Strahler, A. N. 1957. Quantitative analysis of watershed geomorphology. Transactions of the American Geophysical Union 38:913–920. Strange, E. M., and T. C. Foin. 2001. Interaction of physical and biological processes in the assembly of stream fish communities, p. 311–337. In Ecological Assembly Rules. E. Weiher and

612

LITERATURE CITED

P. Keddy (eds.). Cambridge University Press, Cambridge, United Kingdom. Strange, E. M., P. B. Moyle, and T. C. Foin. 1992. Interactions between stochastic and deterministic processes in stream fish community assembly. Environmental Biology of Fishes 36:1–15. Strange, R. M. 2001. Female preference and the maintenance of male fin ornamentation in three egg-mimic darters (Pisces: Percidae). Journal of Freshwater Ecology 16:267–271. Strange, R. M., and B. M. Burr. 1997. Intraspecific phylogeography of North American Highland fishes: a test of the Pleistocene vicariance hypothesis. Evolution 51:885–897. Strange, R. M., and R. L. Mayden. 2009. Phylogenetic relationships and a revised taxonomy for North American cyprinids currently assigned to Phoxinus (Actinopterygii: Cyprinidae). Copeia 2009:494–501. Strauss, R. E., and F. L. Bookstein. 1982. The truss: body form reconstructions in morphometrics. Systematic Zoology 31:113–135. Strecker, U., and A. Kodric-Brown. 1999. Mate recognition systems in a species flock of Mexican pupfish. Journal of Evolutionary Biology 12:927–935. Stringer, G. L. 1992. Late Pleistocene-early Holocene teleostean otoliths from a Mississippi River mudlump. Journal of Vertebrate Paleontology 12:33–41. Strubberg, A. 1913. The metamorphosis of elvers as influenced by outward conditions. Meddelelser fra Kommissionen for Danmarks Fiskeri-og Havundersøgelser 4:1–11. Sturani, C. 1973. A fossil Eel (Anguilla sp.) from the Messinian of Alba (Tertiary Piedmontese Basin); palaeoenvironmental and palaeogeographic implications, p. 243–255. In Messinian Events in the Mediterranean. C. W. Drooger (ed.). North-Holland Publishing Co., Amsterdam, The Netherlands. Stunkard, H. W. 1956. The morphology and life-history of the digenetic trematodes, Azygia sebago Ward, 1910. Biological Bulletin 111:248–268. Sublette, J. E., M. D. Hatch, and M. Sublette. 1990. The Fishes of New Mexico. University of New Mexico Press, Albuquerque. Suchy, M. D. 2009. Effects of salinity on growth and survival of larval and juvenile Alligator Gar Atractosteus spatula, and on plasma osmolality of non-teleost actinopterygiian fishes. Unpubl. Master’s thesis. Nicholls State University, Thibodaux, Louisiana. Suk, H. Y., and J. C. Choe. 2002. The presence of eggs in the nest and female choice in common freshwater Gobies (Rhinogobius brunneus). Behavioral Ecology and Sociobiology 52:211–215. Sulak, K. J., R. A. Brooks, and M. T. Randall. 2007. Seasonal refugia and trophic dormancy in Gulf Sturgeon: test and refutation of the thermal barrier hypothesis, p. 19–49. In Anadromus Sturgeons: Habitat, Threats, and Management. J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron (eds.). American Fisheries Society Symposium 56, Bethesda, Maryland. Sulak, K. J., and J. P. Clugston. 1998. Early life history stages of Gulf Sturgeon in the Suwannee River, Florida. Transactions of the American Fisheries Society 127:758–771. Sulak, K. J., and J. P. Clugston. 1999. Recent advances in life history of Gulf of Mexico Sturgeon, Acipenser oxyrinchus desotoi, in the Suwannee River, Florida, USA: a synopsis. Journal of Applied Ichthyology 15:116–128.

Sulak, K. J., and M. Randall. 2002. Understanding Sturgeon life history: enigmas, myths, and insights from scientific studies. Journal of Applied Ichthyology 18:519–528. Sulak, K. J., R. E. Edwards, G. W. Hill, and M. T. Randall. 2002. Why do Sturgeons jump? Insights from acoustic investigations of the Gulf Sturgeon in the Suwannee River, Florida, USA. Journal of Applied Ichthyology 18:617–620. Sulak, K. J., M. T. Randall, R. E. Edwards, T. M. Summers, K. E. Luke, W. T. Smith, A. D. Norem, W. M. Harden, R. H. Lukens, F. Parauka, S. Bolden, and R. Lehnert. 2009. Defining winter trophic habitat of juvenile Gulf Sturgeon in the Suwanne and Apalachicola rivemouth estuaries, acoustic telemetry investigations. Journal of Applied Ichthyology 25:505–515. Sullivan, A. B., H. I. Jager, and R. Myers. 2003. Modeling White Sturgeon movement in a reservoir: the effect of water quality and Sturgeon density. Ecological Modelling 167:97–114. Sumner, T., J. Travis, and C. D. Johnson. 1994. Methods of fertility advertisement and variation among males in responsiveness in the Sailfin Molly (Poecilia latipinna). Copeia 1994:27–34. Sun, Y., L. Siyang, G. Zhao, S. He, Q. Wu, N. Taniguchi, and Q. Yu. 2004. Genetic structure of Chinese Sucker population Myxocyprinus asiaticus in the Yangtze River based on mitochondrial DNA marker. Fisheries Science 70:412–420. Sun, Y. H., C. X. Xie, W. M. Wang, S. Y. Liu, T. Treer, and M. M. Chang. 2007. The genetic variation and biogeography of catostomid fishes based on mitochondrial and nucleic DNA sequences. Journal of Fish Biology 70:291–309. Surber, T. 1913. Notes on the natural hosts of fresh-water mussels. Bulletin of the United States Bureau of Fisheries (Document 778) 32:101–115. Surber, T. 1915. Identification of the glochidia of fresh-water mussels. Report of the U.S. Commissioner of Fisheries for 1914. Appendix 5:1–9. Sures, B., and K. Knopf. 2004. Parasites as a threat to Freshwater Eels? Science 304:209–211. Sutherland, A. B. 2007. Effects of increased suspended sediment on the reproductive success of an upland crevice-spawning minnow. Transactions of the American Fisheries Society 136:416–422. Sutherland, A. B., J. Maki, and V. Vaughan. 2008. Effects of suspended sediment on whole-body cortisol stress response of two southern Appalachian minnows, Erimonax monachus and Cyprinella galactura. Copeia 2008:234–244. Sutherland, A. B., and J. L. Meyer. 2007. Effects of increased suspended sediment on growth rate and gill condition of two southern Appalachian minnows. Environmental Biology of Fishes 80:389–403. Suttkus, R. D. 1963. Order Lepisostei. Fishes of the western North Atlantic, part 3. Memoirs of the Sears Foundation for Marine Research 1:61–88. Suttkus, R. D., and S. Mettee. 2001. Analysis of four species of Notropis included in the subgenus Pteronotropis Fowler, with comments on relations, origins, and dispersion. Geological Survey of Alabama Bulletin 170:1–50. Sutton, T. M., and S. H. Bowen. 1994. Significance of organic detritus in the diet of larval Lampreys in the Great Lakes basin. Canadian Journal of Fisheries and Aquatic Sciences 51:2380–2387. Sutton, T. M., A. C. Grier, and L. D. Frankland. 2009. Stock structure and dynamics of Longnose Gar and Shortnose Gar in the

LITERATURE CITED

Wabash River, Indiana-Illinois. Journal of Freshwater Ecology 24:657–666. Suzuki, M., K. Kubokawa, H. Nagasawa, and A. Urano. 1995. Sequence analysis of vasotocin cDNAs of the Lamprey, Lampetra japonica, and the Hagfish, Eptatretus burgeri: evolution of cyclostome vasotocin precursors. Journal of Molecular Endocrinology 14:67–77. Suzuki, N., and T. Hibiya. 1985. Pharyngeal teeth and masticatory process of the basioccipital bone in Japanese bitterlings. Japanese Journal of Ichthyology 32:180–188. Sveinsson, T., and T. J. Hara. 1990. Olfactory receptors in Arctic Charr (Salvelinus alpinus) with high sensitivity and specificity for prostaglandin F-2 α. Chemical Senses 15:645–646. Sveinsson T., and T. J. Hara. 2000. Olfactory sensitivity and specificity of Arctic charr, Salvelinus alpinus, to a putative male pheromone, prostaglandin F-2α. Physiology and Behavior 69:301–307. Sveinsson, T., and T. J. Hara. 1995a. Mature males of Arctic charr, Salvelinus alpinus, release F-type prostaglandins to attract conspecific mature females and stimulate their spawning behaviour. Environmental Biology of Fishes 42:253–266. Sveinsson, T., and T. J. Hara. 1995b. Olfactory sensitivity and specificity of Arctic charr, Salvelinus alpinus, to a putative male pheromone, prostaglandin F-2 α. Physiology and Behavior 69:301–307. Swift, C. C. 1970. A review of the eastern North American cyprinid fishes of the Notropis texanus species group (subgenus Alburnops), with a definition of the subgenus Hydrophlox, and material for a revision of the subgenus Alburnops. Unpubl. Ph.D. diss., Florida State University, Tallahassee. Swift, C. C., C. R. Gilbert, S. A. Bortone, G. H. Burgess, and R. W. Yerger. 1986. Zoogeography of the freshwater fishes of the southeastern United States: Savannah River to Lake Pontchartrain, p. 213–265. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. Swift, C. C., R. W. Yerger, and P. R. Parrish. 1977. Distribution and natural history of the fresh and brackish-water fishes of the Ochlockonee River, Florida and Georgia. Bulletin Tall Timbers Research Station 20:1–111. Sytchevskaya, E. K. 1986. Paleogene freshwater fish fauna of the USSR and Mongolia. Joint Soviet-Mongolian Paleontology Expedition Transactions 29:157. Tabit, C. R., and G. M. Johnson. 2002. Influence of urbanization on the distribution of fishes in a southeastern Upper Piedmont drainage. Southeastern Naturalist 1:253–268. Taborsky, M. 1994. Sneakers, satellites, and helpers: parasitic and cooperative behavior in fish reproduction. Advances in the Study of Behavior 23:1–100. Tafanelli, R., P. E. Mauck, and G. Mensinger. 1970. Food habits of the Bigmouth and Smallmouth Buffalo from four Oklahoma reservoirs. Proceedings of the Southeastern Association of Game and Fish Commissioners 1970:649–658. Tagatz, M. E. 1968. Fishes of the St. Johns River, Florida. Journal of the Florida Academy of Sciences 30:25–50. Takai, H., and M. Morisawa. 1995. Change in intracellular K+ concentration caused by external osmolality change regulates sperm motility of marine and fresh water teleosts. Journal of Cell Science 108:1175–1181. Takezaki, N., F. Figueroa, Z. Zaleska-Rutczynska, and J. Klein. 2003. Molecular phylogeny of early vertebrates: monophyly of

613

the agnathans as revealed by sequences of 35 genes. Molecular Biology and Evolution 20:287–292. Tanaka, H., H. Kagawa, and H. Ohta. 2001. Production of leptocephali of Japanese Eel (Anguilla japonica) in captivity. Aquaculture 201:51–60. Tang, Q., H. Liu, R. Mayden, and B. Xiong. 2005. Comparison of evolutionary rates in the mitochondrial DNA cytochrome b gene and control region and their implications for phylogeny of the Cobitoidea (Teleostei: Cypriniformes). Molecular Phylogenetics and Evolution 39:347–57. Taniuchi, T. 1979. Freshwater elasmobranchs from Lake Naujan, Perak River, and Indragiri River, Southeast Asia. Japanese Journal of Ichthyology 25:273–277. Taverne, L. 1977. Ostéologie, phylogénèse et systématique des Téléosteens fossiles et actuels du super-ordre des Ostéoglossomorphes, premiere partie. Ostéologie des genres Hiodon, Eohiodon, Lycoptera, Osteoglossum, Scleropages, Heterotis et Arapaima. Mémoires de la Classe des Sciences, Académie Royale de Belgique 42:1–235. Taverne, L. 1979. Ostéologie, phylogénèse et systématique des Téléosteens fossiles et actuels du super-ordre des Ostéoglossomorphes, troisieme partie. Évolution des structures ostéologiques et conclusions générales relatives à la phylogénèse et à la systématique du super-order. Addendum. Mémoires de la Classe des Sciences, Académie Royale de Belgique 43:1–168. Tavolga, W. N. 1971. Sound production and detection, p. 135–205. In Fish Physiology. Vol. 5. W. S. Hoar and D. S. Randall (eds.). Academic Press, New York. Tavolga, W. N. 1976. Chemical stimuli in reproductive behavior in fish: communication. Experientia 32:1093–1095. Taylor, C. M. 1996. Abundance and distribution within a guild of benthic stream fishes: local processes and regional patterns. Freshwater Biology 36:385–396. Taylor, C. M. 1997. Fish species richness and incidence patterns in isolated and connected stream pools: effects of pool volume and spatial position. Oecologia 110:560–566. Taylor, C. M. 2000. A large-scale comparative analysis of riffle and pool fish communities in an upland stream system. Environmental Biology of Fishes 58:89–95. Taylor, C. M. 2010. Covariation among plains stream fish assemblages, flow regimes, and patterns of water use, p. 447–459. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. Taylor, C. M., and N. J. Gotelli. 1994. The macroecology of Cyprinella: correlates of phylogeny, body size, and geographic range. The American Naturalist 144:549–569. Taylor, C. M., and P. W. Lienesch. 1996. Regional parapatry of the congeneric cyprinids Lythrurus snelsoni and L. umbratilis: species replacement along a complex environmental gradient. Copeia 1996:493–497. Taylor, C. M., and M. L. Warren, Jr. 2001. Dynamics in species composition of stream fish assemblages: environmental variability and nested subsets. Ecology 82:2320–2330. Taylor, C. M., M. R. Winston, and W. J. Matthews. 1993. Fish species-environment and abundance relationships in a Great Plains river system. Ecography 16:16–23. Taylor, C. M., M. R. Winston, and W. J. Matthews. 1996a. Temporal variation in tributary and mainstream fish assemblages in a Great Plains stream system. Copeia 1996:280–289.

614

LITERATURE CITED

Taylor, E. B., and J. D. McPhail. 1999. Evolutionary history of an adaptive radiation in species pairs of Threespine Sticklebacks (Gasterosteus): insights from mitochondrial DNA. Biological Journal of the Linnean Society 66:271–291. Taylor, J. C., and P. S. Rand. 2003. Spatial overlap and distribution of Anchovies (Anchoa spp.) and copepods in a shallow stratified estuary. Aquatic Living Resources 16:191–196. Taylor, J. C., P. S. Rand, and J. Jenkins. 2007. Swimming behavior of juvenile Anchovies (Anchoa spp.) in an episodically hypoxic estuary: implications for individual energetics and trophic dynamics. Marine Biology 152:939–957. Taylor, P. W. 1978. Macroderoides trilobatus sp. n. (Digenea: Macroderoididae) from the Bowfin, Amia calva, and emendation of the genus. The Journal of Parasitology 64:393–394. Taylor, S. A., E. Burt, G. Hammond, and K. Releya. 1996b. Female mosquitofish (Gambusia affinis holbrooki) prefer normally pigmented males to melanistic males. Journal of Comparative Psychology 110:260–266. Teaf, C. M. 1980. A study of tidally-oriented movements of the Atlantic Stingray, Dasyatis sabina (LeSueur) in Apalachee Bay, Florida. Unpubl. Master’s thesis. Florida State University, Tallahassee. Teaf, C. M., and T. C. Lewis. 1987. Seasonal occurrence of multiple caudal spines in the Atlantic Stingray, Dasyatis sabina (Pisces: Dasyatidae). Copeia 1987:224–227. Teeter, J. 1980. Pheromone communication in Sea Lampreys (Petromyzon marinus) implications for population management. Canadian Journal of Fisheries and Aquatic Sciences 37:2123–2132. Teeter, J. H., R. B. Szamier, and M. V. L. Bennett. 1980. Ampullary electroreceptors in the Sturgeon Scaphirhynchus platorynchus (Rafinesque). Journal of Comparative Physiology 138:213–223. Teh, S. J., D. Xin, D. Dong-Fang, T. Foo-Ching, S. S. O. Hung, F. T. W.-M., J. Liu, and R. M. Higashi. 2004. Chronic effects of dietary selenium on juvenile Sacramento Splittail (Pogonichthys macrolepidotus). Environmental Science & Technology 38: 6085–6093. Teichmann, H. 1959. Vergleichende Untersuchungen an der Nase der Fische. Zeitschrift für Morphologie der Tiere 43:171–212. Terwilliger, M. R., D. F. Markle, and J. Kann. 2003. Associations between water quality and daily growth of juvenile Shortnose and Lost River Suckers in Upper Klamath Lake, Oregon. Transactions of the American Fisheries Society 132:691–708. Tesch, F.-W. 2003. The Eel. 3rd edition. Blackwell Publishing, Oxford, England. Tesch, F.-W., and N. Rohlf. 2003. Migration from continental waters to the spawning grounds, p. 223–234. In Eel Biology. K. Aida, K. Tsukamoto, and K. Yamauchi (eds.). Springer, Tokyo. Thibault, I., J. J. Dodson, and F. Caron. 2007. Yellow-stage American Eel movements determined by microtagging and acoustic telemetry in the St. Jean River watershed, Gaspé, Quebec, Canada. Journal of Fish Biology 71:1095–1112. Thibault, R. E. 1974. Genetics of cannibalism in a viviparous fish and its relationship to population density. Nature 251:138–140. Thinès, G., and J.-M. LeGrain. 1973. Effects de la substance d’alarme sur le comportement des poisons cavernicoles Anoptichthys jordani (Characidae) et Caecobarbus geertsi (Cyprinidae). Annales de Speleologie 28:291–297. Thiyagarajah, A, M. B. Anderson, and W. R. Hartley. 2000. Gonadal cysts in Spotted Gar (Lepisosteus oculatus) from Bayou

Trepagnier, Louisiana, USA. Marine Environmental Research 50:279–282. Thomas, D. A., and D. B. Hayes. 2006. A comparison of fish community composition of headwater and adventitious streams in a coldwater river system. Journal of Freshwater Ecology 21:265–275. Thomas, J. 2006. American Eel behavioral patterns in Silver Lake, Dover, Delaware. Unpubl. Master’s thesis. Delaware State University, Dover. Thomas, L. J. 1930. Notes on the life history of Haplobothrium globuliforme Cooper, a tapeworm of Amia calva L. The Journal of Parasitology 16:140–145. Thomas, L. P., VI. 1983. Fine structure of the tentacles and associated microanatomy of Haplobothrium globuliforme (Cestoda: Pseudophyllidea). The Journal of Parasitology 69:719–730. Thomas, M. V., and R. C. Haas. 2002. Abundance, age structure, and spatial distribution of Lake Sturgeon, Acipenser fulvescens, in the St. Clair system. Journal of Applied Ichthyology 18:495–501. Thomaz, D., E. Beall, and T. Burke. 1997. Alternative reproductive tactics in Atlantic Salmon—factors affecting mature parr success. Proceedings of the Royal Society of London B 264: 219–226. Thompson, C. E., W. R. Poole, M. A. Matthews, and A. Ferguson. 1998. Comparison, using minisatellite DNA profiling, of secondary male contribution in the fertilisation of wild and ranched Atlantic Salmon (Salmo salar) ova. Canadian Journal of Fisheries and Aquatic Sciences 55:2011–2018. Thompson, D. B. A., and M. L. P. Thompson. 1985. Early warning and mixed species association: the Plover’s-Page revisited. Ibis 127:559–562. Thompson, D. H. 1933. The finding of very young Polyodon. Copeia 1933:31–33. Thompson, D. H. 1934. Relative growth in Polyodon. Natural History Survey of Illinois Biological Notes No. 2:1–8. Thompson, J. N. 1982. Interaction and coevolution. John Wiley and Sons, New York. Thompson, J. N. 1988. Variations in interspecific interactions. Annual Review of Ecology and Systematics 19:65–87. Thompson, J. N. 1994. The Coevolutionary Process. University of Chicago Press, Chicago, Illinois. Thompson, J. N. 1998. Rapid evolution as an ecological process. Trends in Ecology and Evolution 13:329–332. Thompson, J. N. 1999a. The evolution of species interactions. Science 284:2116–2118. Thompson, J. N. 1999b. Specific hypotheses on the geographic mosaic of coevolution. The American Naturalist 153 (Supplement): S1–S14. Thompson, J. N. 2005. The Geographic Mosaic of Coevolution. The University of Chicago Press, Chicago, Illinois. Thompson, J. N., and B. M. Cunningham. 2002. Geographic structure and dynamics of coevolutionary selection. Nature 417:735–738. Thorson, T. B., and D. R. Watson. 1975. Reassignment of the African freshwater Stingray, Potamotrygon garouaensis to the genus Dasyatis, on physiologic and morphologic grounds. Copeia 1975:701–712. Threader, R. W., and C. S. Brousseau. 1986. Biology and management of the Lake Sturgeon in the Moose River, Ontario. North American Journal of Fisheries Management 6:383–390.

LITERATURE CITED

Thuemler, T. F. 1988. Movements of young Lake Sturgeon stocked in the Menominee River, Wisconsin, p. 104–109. In 11th Annual Larval Fish Conference. R. D. Hoyt (ed.). American Fisheries Society Symposium 5, Bethesda, Maryland. Tibbets, C. A., and T. E. Dowling. 1996. Effects of intrinsic and extrinsic factors on population fragmentation in three species of North American minnows (Teleostei: Cyprinidae). Evolution 50:1280–1292. Tibbets, C. A., A. C. Weibel, and T. E. Dowling. 2001. Population genetics of Lepidomeda vittata, the Little Colorado River Spinedace. Copeia 2001:813–819. Tidwell, J. H., and S. D. Mims. 1990. Survival of Paddlefish fingerlings stocked with large Channel Catfish. The Progressive FishCulturist 52:273–274. Timmons, T. J., and T. A. Hughbanks. 2000. Exploitation and mortality of Paddlefish in the lower Tennessee and Cumberland rivers. Transactions of the American Fisheries Society 129:1171–1180. Timmons, T. J., and J. S. Ramsey. 1983. Life history and habitat of the Blackfin Sucker, Moxostoma atripinne (Osteichthyes: Catostomidae). Copeia 1983:538–541. Timms, A. M., and H. Kleerekoper. 1972. The locomotor responses of male Ictalurus punctatus, the Channel Catfish, to the pheromone released by the ripe female of the species. Transactions of the American Fisheries Society 1972:302–310. Tinbergen, N. 1948. Social releasers and the experimental method for their study. Wilson Bulletin 60:6–51. Tinbergen, N. 1951. The Study of Instinct. Oxford University Press, Oxford, England. Tinbergen, N. 1952. ‘Derived’ activities; their causation, biological significance, origin, and emancipation during evolution. Quarterly Review of Biology 27:1–32. Tipton, M. L., S. Gignoux-Wolfsohn, P. Stonebraker, and B. Chernoff. 2011. Postglacial recolonization of eastern Blacknose Dace, Rhinichthys atratulus (Teleostei: Cyprinidae), through the gateway of New England. Ecology and Evolution 1:343–358. Titus, J. E., D. Grise, G. Sullivan, and M. D. Stephens. 2004. Monitoring submersed vegetation in a mesotrophic lake: correlation of two spatio-temporal scales of change. Aquatic Botany 79:33–50. Tkach, V. V., E. E. Pulis, and R. M. Overstreet. 2010. A new Paramacroderoides species (Digenea: Macroderoididae) from two species of Gar in the southeastern United States. Journal of Parasitology 96:1002–1006. Tkach, V. V. E. J. Strand, and L. Froese. 2008. Macroderoides texanus n. sp. (Digenea: Macroderoididae) from Alligator Gar, Atractosteus spatula in Texas. Parasitological Research 104:27–33. Tobler, M., R. Riesch, F. J. García de León, I. Schlupp, and M. Plath. 2007. A new and morphologically distinct population of cavernicolous Poecilia mexicana (Poeciliidae: Teleostei). Environmental Biology of Fishes 82:101–108. Tobler, M. and I. Schlupp. 2009. Threatened fishes of the world: Poecilia latipunctata Meek, 1904 (Poeciliidae). Environmental Biology of Fishes 85:31–32. Tobler, M. and I. Schlupp. 2010. Differential susceptibility to food stress in neonates of sexual and asexual mollies (Poecilia, Poeciliidae). Evolutionary Ecology 24:39–47. Tobler, M., I. Schlupp, K. U. Heubel, R. Riesch, F. J. García de León, O. Giere, and M. Plath. 2006. Life on the edge: hydrogen sulfide and the fish communities of a Mexican cave and surrounding waters. Extremophiles 10:577–585.

615

Todd, J. H., J. Atema, and J. E. Bardach. 1967. Chemical communication in social behavior of a fish, the Yellow Bullhead (Ictalurus natalis). Science 158:672–673. Todd, R. M. 1999. Sturgeon and Paddlefish commercial fishery in North America, p. 42–50. In Proceedings of the Symposium on the Harvest, Trade and Conservation of North American Paddlefish and Sturgeon, May 7–8, 1998, Chattanooga, Tennessee. D. F. Williamson, G. W. Benz, and C. M. Hoover (eds.). TRAFFIC North America / World Wildlife Fund, Washington, D.C. Toepfer, C., and M. Barton. 1992. Influence of salinity on the rates of oxygen consumption in two species of freshwater fishes, Phoxinus erythrogaster (family Cyprinidae), and Fundulus catenatus (family Fundulidae). Hydrobiologia 242:149–154. Tomelleri, J. R., and M. E. Eberle. 1990. Fishes of the Central United States. University Press of Kansas, Lawrence, Kansas. Tonn, W. M., and J. J. Magnuson. 1982. Patterns in the species composition and richness of fish assemblages in northern Wisconsin lakes. Ecology 63:1149–1166. Tonn, W. M., J. J. Magnuson, M. Rask, and J. Toivonen. 1990. Intercontinental comparison of small-lake fish assemblages: the balance between local and regional processes. The American Naturalist 136:345–375. Torgersen, C. E., C. V. Baxter, H. W. Li, and B. A. McIntosh. 2006. Landscape influences on longitudinal patters of river fishes: spatially continuous analysis of fish-habitat relationships. American Fisheries Society Symposium 48:473–492. Tosi, L. and C. Sola. 1993. Role of geosmin, a typical inland water odour, in guiding glass Eel Anguilla anguilla (L.) migration. Ethology 95:177–185. Tosi, L., A. Spampanato, C. Sola, and P. Tongiorgi. 1990. Relation of water odour, salinity and temperature to ascent of glass-eels, Anguilla anguilla (L.): a laboratory study. Journal of Fish Biology 36:327–340. Townsend, C. H. 1902. Statistics of the fisheries of the Mississippi River and tributaries. Report of the United States Commission of Fish and Fisheries, Appendix 15, 27(1901):659–740. Townsend, C. R., and A. G. Hildrew. 1994. Species traits in relation to a habitat templet for river systems. Freshwater Biology 31:265–275. Trainor, B. C., and A. L. Basolo. 2000. An evaluation of video playback using Xiphophorus helleri. Animal Behaviour 59:83–89. Tranah, G. J., J. J. Agresti, and B. May. 2001b. New microsatellite loci for Suckers (Catostomidae): primer homology in Catostomus, Chasmistes and Deltistes. Molecular Ecology Notes 1:55–60. Tranah, G. J., M. Bagley, J. J. Agresti, and B. May. 2003. Development of codominant markers for identifying species hybrids. Conservation Genetics 4:537–541. Tranah, G., D. E. Campton, and B. May. 2004. Genetic evidence for hybridization of Pallid and Shovelnose Sturgeon. Journal of Heredity 95:474–480. Tranah, G. J., H. L. Kincaid, C. C. Krueger, D. E. Campton, and B. May. 2001a. Reproductive isolation in sympatric populations of Pallid and Shovelnose Sturgeon. North American Journal of Fisheries Management 21:367–373. Tranah, G. J., and B. May. 2006. Patterns of intra- and interspecies genetic diversity in Klamath River Basin Suckers. Transactions of the American Fisheries Society 135:306–316. Traquair, R. H. 1877. The ganoid fishes of the British Carboniferous formations, part 1, Paleoniscidae. Monographs Paleontographical Society 31:1–60.

616

LITERATURE CITED

Trautman, M. B. 1956. Carpoides cyprinus hinei, a new subspecies of carpsucker from the Ohio and Upper Mississippi River systems. Ohio Journal of Science 56:33–40. Trautman, M. B. 1957. The Fishes of Ohio. Ohio State University Press, Baltimore, Maryland. Trautman, M. B. 1981. The Fishes of Ohio. The Ohio State University Press, Columbus, Ohio. Travis, J. 1994. Size-dependent behavioral variation and its genetic control within and among populations, p. 165–187. In Quantitative Genetic Studies of Behavioral Evolution. C. M. Boake (ed.). University of Chicago Press, Chicago, Illinois. Travis, J., J. A. Farr, M. McManus, and J. C. Trexler. 1989. Environmental effects on adult growth patterns in the male Sailfin Molly (Poecilia latipinna, Poeciliidae). Environmental Biology of Fishes 26:119–127. Travis, J., and J. C. Trexler. 1987. Regional variation in habitat requirements of the Sailfin Molly, with special reference to the Florida Keys. Nongame Wildlife Technical Report Number 3, Florida Game and Freshwater Fish Commission, Tallahassee. Travis, J., J. C. Trexler, and M. Mulvey. 1990. Multiple paternity and its correlates in female Poecilia latipinna (Poeciliidae). Copeia 1990:722–729. Travis, J., and B. D. Woodward. 1989. Social context and courtship flexibility in male Sailfin Mollies, Poecilia latipinna (Pisces: Poeciliidae). Animal Behaviour 38:1001–1011. Treberg, J. R., B. Speers-Roesch, P. M. Piermarini, Y. K. Ip, J. S. Ballantyne, and W. R. Driedzic. 2006. The accumulation of methylamine counteracting solutes in elasmobranchs with differing levels of urea: a comparison of marine and freshwater species. Journal of Experimental Biology 209:860–870. Trebitz, A. S., J. C. Brazner, V. J. Brady, R. Axler, and D. K. Tanner. 2007. Turbidity tolerances of Great Lakes coastal wetlands fishes. North American Journal of Fisheries Management 27:619–633. Tregenza, T., and N. Wedell. 1998. Benefits of multiple mates in the cricket Gryllus bimaculatus. Evolution 52:1726–1730. Tremblay, S., and P. Magnan. 1991. Interactions between two distantly related species, Brook Trout (Salvelinus fontinalis) and White Sucker (Catostomus commersoni). Canadian Journal of Fisheries and Aquatic Sciences 48:857–867. Tremblay, V. 2009. Reproductive strategy of female American Eels among five subpopulations in the St. Lawrence River watershed, p. 85–201. In Eels at the Edge: Science, Status, and Conservation Concerns. J. M. Casselman and D. K. Cairns (eds.). American Fisheries Society Symposium 58. Bethesda, Maryland. Trested, D. G., K. Ware, R. Bakal, and J. J. Isely. 2011. Microhabitat use and seasonal movements of hatchery-reared and wild Shortnose Sturgeon in the Savannah River, South Carolina— Georgia. Journal of Applied Ichthyology 27:454–461. Trexler, J. C. 1995. Restoration of the Kissimmee River: a conceptual model of past and present fish communities and its consequences for evaluating restoration success. Restoration Ecology 3:195–210. Trexler, J. C., and J. Travis. 1990. Phenotypic plasticity in the Sailfin Molly, Poecilia latipinna (Pisces: Poeciliidae). I. Field experiments. Evolution 44:143–156. Trexler, J. C., J. Travis, and A. Dinep. 1997. Variation among populations of the Sailfin Molly in the rate of concurrent multiple paternity and its implications for mating-system evolution. Behavioral Ecology and Sociobiology 40:297–305.

Trexler, J. C., J. Travis, and M. Trexler. 1990. Phenotypic plasticity in the Sailfin Molly, Poecilia latipinna (Pisces: Poeciliidae). II. Laboratory experiment. Evolution 44:157–167. Tricas, T. C. 1980. Courtship and mating-related behaviors in myliobatid rays. Copeia 1980:553–556. Tricas, T. C., K. P Maruska, and E. Rasmussen. 2000. Annual cycles of steroid hormone production, gonad development, and reproductive behavior in the Atlantic Stingray. General and Comparative Endocrinology 118:209–225. Trilles, J. P. 2007. Olencira praegustator (Crustacea, Isopoda, Cymothoidae) parasitic on Brevoortia species (Pisces, Clupeidae) from the southeastern coasts of North America: review and redescription. Marine Biology Research. 3:296–311. Tripp, S. J., R. E. Colombo, and J. E. Garvey. 2009a. Declining recruitment and growth of Shovelnose Sturgeon in the middle Mississippi River: implications for conservation. Transactions of the American Fisheries Society 138:416–422. Tripp, S. J., Q. E. Phelps, R. E. Colombo, J. E. Garvey, B. M. Burr, D. P. Herzog, and R. A. Hrabik. 2009b. Maturation and reproduction of Shovelnose Sturgeon in the middle Mississippi River. North American Journal of Fisheries Management 29:730–738. Trippel, E. A., and H. H. Harvey. 1987a. Abundance, growth, and food supply of White Suckers (Catostomus commersoni) in relation to lake morphometry and pH. Canadian Journal of Zoology 65:558–564. Trippel, E. A., and H. H. Harvey. 1987b. Reproductive responses of five White Sucker (Catostomus commersoni) populations in relation to lake acidity. Canadian Journal of Fisheries and Aquatic Sciences 44:1018–1023. Trippel, E. A., and H. H. Harvey. 1989. Missing opportunities to reproduce: an energy dependant or fecundity gaining strategy in White Sucker (Catostomus commersoni)? Canadian Journal of Zoology 67:2180–2188. Trautman, M. B. 1981. The Fishes of Ohio. Ohio State University Press, Columbus. Tsoi, S. C. M., S. C. Lee, and W. C. Chao. 1989. Duplicate gene expression and diploidization in an Asian tetraploid catostomid, Myxocyprinus asiaticus (Cypriniformes, Catostomidae). Comparative Biochemistry and Physiology 93:27–32. Tsukamoto, K., J. Aoyama, and M. Miller. 2002. Migration, speciation and the evolution of diadromy in anguillid Eels. Canadian Journal of Fisheries and Aquatic Science 59:1989–1998. Tsukamoto, K., I. Nakai, and F-W. Tesch. 1998. Do all Freshwater Eels migrate? Nature 396:635–636. Tucker, J. W., Jr. 1989. Energy utilization in Bay Anchovy, Anchoa mitchilli, and Black Seabass, Centropristis striata striata, eggs and larvae. Fishery Bulletin 87:279–293. Turner, B. J. 1982. The evolutionary genetics of a unisexual fish, Poecilia Formosa, p. 265–305. In Mechanisms of Speciation. C. Barigozzi (ed.). Alan R. Liss, New York. Turner, B. J., B. H. Brett, and R. R. Miller. 1980. Interspecific hybridization and the evolutionary origin of a gynogenetic fish, Poecilia formosa. Evolution 34:917–922. Turner, E. and R. Montgomerie. 2002. Ovarian fluid enhances sperm movement in Arctic Charr. Journal of Fish Biology 60:1570–1579. Turner, T. F., T. J. Krabbenhoft, and A. S. Burdett. 2010. Reproductive phenology and fish community structure in an arid-land river system, p. 427–446. In Community Ecology of Stream Fishes: Concepts, Approaches, and Techniques. K. B. Gido and

LITERATURE CITED

D. A. Jackson (eds.). American Fisheries Society Symposium 73. Bethesda, Maryland. Twohey, M. B., J. W. Heinrich, J. G. Seelye, K. T. Fredricks, R. A. Bergstedt, C. A. Kaye, R. J. Scholefield, R. B. McDonald, and G. C. Christie. 2003a. The sterile-male-release technique in Great Lakes Sea Lamprey management. Journal of Great Lakes Research 29(Suppl. 1):410–423. Twohey, M. B., P. W. Sorensen, and W. Li. 2003b. Possible applications of pheromones in an integrated Sea Lamprey management program. Journal of Great Lakes Research 29(Suppl. 1):794–800. Tyler, J. D. 1994. Albinistic Spotted Gar, Lepisosteus oculatus, in Oklahoma. Proceedings of the Oklahoma Academy of Science 74:39. Tyler, J. D., and M. N. Granger. 1984. Notes on food habits, size, and spawning behavior of Spotted Gar in Lake Lawton, Oklahoma. Proceedings of the Oklahoma Academy of Science 64:8–10. Tyler, J. D., J. R. Webb, T. R. Wright, J. D. Hargett, K. J. Mask, and D. R. Schucker. 1994. Food habits, sex ratios, and size of Longnose Gar in southwestern Oklahoma. Proceedings of the Oklahoma Academy of Science 74:41–42. Tyus, H. M. 1986. Life strategies in the evolution of the Colorado Squawfish (Ptychocheilus lucius). Great Basin Naturalist 46:656–661. Tyus, H. M. 1987. Distribution, reproduction, and habitat use of Razorback Sucker in the Green River, Utah, 1979–1986. Transactions of the American Fisheries Society 116:111–116. Tyus, H. M., and C. A. Karp. 1990. Spawning and movements of Razorback Sucker, Xyrauchen texanus, in the Green River basin of Colorado and Utah. Southwestern Naturalist 35:427–433. Ueno, T. 1985. Peculiarity of karyotype variations in taxa of fishes. (In Japanese) Marine Science Monthly 176:125–130. Uglem, G. L., and S. M. Beck. 1972. Habitat specificity and correlated aminopeptidase activity in the acanthocephalans, Neoechinorhynchus cristatus and N. crassus. Journal of Parasitology 58:911–920. Underhill, J. C. 1986. The fish fauna of the Laurentian Great Lakes, the St. Lawrence lowlands, Newfoundland and Labrador, p. 105–159. In The Zoogeography of North American Freshwater Fishes. C. H. Hocutt and E. O. Wiley (eds.). John Wiley and Sons, New York. Unger, I. M., and R. C. Sargent. 1988. Alloparental care in the Fathead Minnow, Pimephales promelas: females prefer males with eggs. Behavioral Ecology and Sociobiology 23:27–32. Unkenholz, D. G. 1986. Effects of dams and other habitat alterations on Paddlefish sport fisheries, p. 54–61. In The Paddlefish: Status, Management and Propagation. J. G. Dillard, L. K. Graham, and T. R. Russell (eds.). American Fisheries Society Special Publication 7. USFWS (United States Fish and Wildlife Ser vice). 1990. Endangered and threatened wildlife and plants: finding on petition to list the Paddlefish. Federal Register 55(80):17473–17475. USFWS (United States Fish and Wildlife Ser vice). 1992. Endangered and threatened wildlife and plants: finding on petition to list the Paddlefish. Federal Register 57(184):43676–43682. USFWS (United States Fish and Wildlife Ser vice). 1993. Pallid Sturgeon recovery plan. U.S. Fish and Wildlife Ser vice, Bismarck, North Dakota. USFWS (United States Fish and Wildlife Ser vice). 1996. Endangered and threatened wildlife and plants: notice of final decision on identification of candidates for listing as endangered or threatened. Federal Register 61(235):64481–64485.

617

USFWS (United States Fish and Wildlife Ser vice). 2000. Endangered and threatened wildlife and plants: final rule to list the Alabama Sturgeon as endangered. Federal Register 65 (88): 26438–26461. USFWS (United States Fish and Wildlife Ser vice). 2009. Endangered and threatened wildlife and plants; designation of critical habitat for Alabama Sturgeon (Scaphirhynchus suttkusi). Federal Register 74 (104):26488–2651073. USFWS (United States Fish and Wildlife Ser vice). 2010a. Endangered and threatened wildlife and plants; threatened status for Shovelnose Sturgeon under the similarity of appearance provision of the endangered species act. Federal Register 75 (169): 53598–53606. USFWS (United States Fish and Wildlife Ser vice). 2010b. Endangered species program. United States Fish and Wildlife Ser vice. Available from http://www.fws.gov/endangered/species/us-species.html; as of February 2010. USFWS (United States Fish and Wildlife Ser vice). 2011. Endangered species ad hoc search. Available from http://ecos.fws.gov /tess_public/SpeciesReport.do?groups=E&listingType=L&map status=1; as of December 2010. USGS (United States Geological Survey). 2010. American Fisheries Society imperiled freshwater and diadromous fishes of North America. Southeast Ecological Science Center, USGS, Gainesville, Florida. Available from http://fl.biology.usgs.gov/afs _fish/index.html. Upchurch, S. B., and A. F. Randazzo. 1997. Environmental Geology of Florida, p. 217–249. In The Geology of Florida. A. F. Randazzo and D. S. Jones (eds.). University Press of Florida, Gainesville. Urbach, D., I. Folstad, and G. Rudolfsen. 2005. Effects of ovarian fluid on sperm velocity in Arctic Charr (Salvelinus alpinus). Behavioral Ecology and Sociobiology 57:438–444. Uyeno, T. 1961a. Late Cenozoic cyprinid fishes from Idaho with notes on other fossil minnows in North America. Paper of the Michigan Academy of Science, Arts, and Letters 46:329–344. Uyeno, T. 1961b. Osteology and phylogeny of the American cyprinid fishes allied to the genus Gila. Unpubl. Ph.D. diss., The University of Michigan, Ann Arbor. Uyeno, T. 1973. A comparative study of chromosomes in the teleostean fish order Osteoglossiformes. Japanese Journal of Ichthyology 20:211–217. Uyeno, T., and R. R. Miller. 1962. Late Pleistocene fishes from a Trinity River Terrace, Texas. Copeia 1962:338–345. Uyeno, T., and R. R. Miller. 1963. Summary of late Cenozoic freshwater fish records for North America. Occasional Papers of the Museum of Zoology, University of Michigan, Ann Arbor, Michigan. Uyeno, T., and G. R. Smith. 1972. Tetraploid origin of the karyotype of catostomid fishes. Science 175:644–646. Vainio, P. W. 1973. Feeding and food habits of the Smallmouth Buffalo, Ictiobus bubalus (Rafinesque), in Elephant Butte Lake, New Mexico. Unpubl. Master’s thesis, New Mexico State University, Las Cruces. Valdez, R. A., T. A. Hoff nagle, C. C. McIvor, T. McKinney, and W. C. Leibfreid. 2001. Effects of a test flood on fishes of the Colorado River in Grand Canyon, Arizona. Ecological Applications 11:686–700. Valenciennes, A. 1848. Histoire naturelle des poissons, volume 21. P. Bertrand, Paris, France. not seen

618

LITERATURE CITED

Valentincic, T., S. Wegert, and J. Caprio. 1994. Learned olfactory discrimination versus innate taste responses to amino acids in Channel Catfish (Ictalurus punctatus). Physiology and Behavior 55:865–873. Valentine, J. M., Jr., J. R. Walther, K. M. McCartney, and L. M. Ivy. 1972. Alligator diets on the Sabine National Wildlife Refuge, Louisiana. Journal of Wildlife Management 36:809–815. Vallowe, H. H. 1953. Some physiological aspects of reproduction in Xiphophorus maculates. Biological Bulletin 104:240–249. VanBlaricom, G. R. 1982. Experimental analyses of structural regulation in a marine sand community exposed to oceanic swell. Ecological Monographs 52:283–305. van den Berghe, E. P., and M. R. Gross. 1986. Length of breeding life of Coho Salmon (Oncorhynchus kisutch). Canadian Journal of Zoology 64:1482–1486. van den Hurk, R., and J. G. D. Lambert. 1983. Ovarian steroid glucuronides function as sex pheromones for male Zebrafish, Brachydanio rerio. Canadian Journal of Zoology 61:2381–2387. van den Hurk, R., W. G. E. J. Schoonen, G. A. van Zoelen, and J. G. D. Lambert. 1987. The biosynthesis of steroid glucuronides in the testis of the Zebrafish, Brachydanio rerio, and their pheromonal function as ovulation inducers. General and Comparative Endocrinology 68:179–188. van den Thillart, G., V. van Ginneken, F. Körner, R. Heijmans, R. van der Linden, and A. Gluvers. 2004. Endurance swimming of European Eel. Journal of Fish Biology. 65:312–318. van Duzer, E. M. 1939. Observations on the breeding habits of the cut-lip minnow, Exoglossus maxillingua. Copeia 1939:65–75. Van Eenennaam, J. P., F. A. Chapman, and P. L. Jarvis. 2004. Aquaculture, p. 277–311. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. H. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. Van Eenennaam, J. P., and S. I. Doroshov. 1998. Effects of age and body size on gonadal development of Atlantic Sturgeon. Journal of Fish Biology 53:624–637. Van Eenennaam, J. P., S. I. Doroshov, G. P. Moberg, J. G. Watson, D. S. Moore, and J. Linares. 1996. Reproductive conditions of the Atlantic Sturgeon (Acipenser oxyrinchus) in the Hudson River. Estuaries 19:769–777. Van Eenennaam, J. P., J. Linares-Casenave, X. Deng, and S. I. Doroshov. 2005. Effects of incubation temperature on Green Sturgeon embryos, Acipenser medirostris. Environmental Biology of Fishes 72:145–154. Van Eenennaam, J. P., J. Linares, S. I. Doroshov, D. C. Hillemeier, T. E. Wilson, and A. A. Nova. 2006. Reproductive conditions of the Klamath River Green Sturgeon. Transactions of the American Fisheries Society 135:151–163. Van Eenennaam, J. P., M. A. H. Webb, X. Deng, S. I. Doroshov, R. B. Mayfield, J. J. Cech, Jr., D. C. Hillemeier, and T. E. Wilson. 2001. Artificial spawning and larval rearing of Klamath River Green Sturgeon. Transactions of the American Fisheries Society 130:159–165. van Ginneken, V. J. T. and Maes, G. E. 2005 The European Eel (Anguilla anguilla, Linnaeus), its lifecycle, evolution and reproduction: a literature review. Reviews in Fish Biology and Fisheries 15:367–398. van Iersel, J. J. A. 1953. An analysis of the parental behaviour of the male three-spined Stickleback (Gasterosteus aculeatus L.). Behaviour Supplement 3:1–159.

van Oordt, G. J. 1928. The duration of life of the spermatozoa in the fertilized female of Xiphophorus helleri Regan. Nederl. Dierkundige Vereenig. 1:77–80. Van Oosten, J. 1961. Records, ages, and growth of the Mooneye, Hiodon tergisus, of the Great Lakes. Transactions of the American Fisheries Society 90:170–174. Van Oosten, J., and H. J. Deason. 1957. History of Red Lake fisheries, 1917–1938, with observations on population status. U.S. Fish and Wildlife Ser vice, Special Scientific Report 229:1– 63. Van Snik Gray, E., W. A. Lellis, J. C. Cole, and C. S. Johnson. 2002. Host identification for Strophitus undulatus (Bivalvia: Unionidae), the Creeper, in upper Susquehanna River Basin, Pennsylvania. The American Midland Naturalist 147:153–161. van Weerd, J. H., M. Sukkel, and C. J. J. Richter. 1988. An analysis of sex-stimuli enhancing ovarian growth in puberal African Catfish, Clarias gariepinus. Aquaculture 75:181–191. Van Winkle, W., P. Anders, D. H. Secor, and D. Dixon (eds.). 2002. Biology, Management, and Protection of North American Sturgeon, American Fisheries Society Symposium 28, Bethesda, Maryland. Vanicek, D. 1961. Life history of the Quillback and Highfin Carpsuckers in the Des Moines River. Proceedings of the Iowa Academy of Sciences 68:238–246. Vannote, R. L., G. W. Minshall, K. W. Cummins, J. R. Sedell, and C. E. Cushing. 1980. The river continuum concept. Canadian Journal of Fisheries and Aquatic Sciences 37:130–137. Vasil’ev, V. P. 1980. Chromosome numbers in fish-like vertebrates and fish. Journal of Ichthyology 20(3):1–38. Vasil’ev, V. P. 1999. Polyploidization by reticular speciation in Acipenseriformes evolution: a working hypothesis. Journal of Applied Ichthyology 15:29–31. Vasil’ev, V. P. 2009. Mechanisms of polyploidy evolution: polyploidy in Sturgeons, p. 97–117. In Biology, Conservation and Sustainable Development of Sturgeons. R. Carmona, A. Domezain, M. García-Gallego, J. A. Hernando, F. Rodríguez, and M. Ruiz-Rejón (eds.). Springer Science + Business Media B. V. Fish & Fisheries Series 29. Vasil’eva, E. D. 1999. Some morphological characteristics of acipenserids fishes: considerations of their variability and utility in taxonomy. Journal of Applied Ichthyology 15:32–34. Vaughn, C. C., F. P. Gelwick, and W. J. Matthews. 1993. Effects of algivorous minnows on production of grazing stream invertebrates. Oikos. 66:1191–28. Vavrek, M. A., and G. E. Brown. 2009. Threat-sensitive responses to disturbance cues in juvenile Convict Cichlids and Rainbow Trout. Annales Zoologici Fennici 46:171–180. Vecsei, P. 1999. Illustrating Sturgeon species: the art of science in highlighting morphological features. Journal of Applied Ichthyology 15:7–8. Vecsei, P., and D. Peterson. 2004. Sturgeon ecomorphology: a descriptive approach, p. 103–133. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. H. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. Vedrasco, A., V. Lobchenko, and R. Billard. 2001. Introductions et élevage du poisson-spatule Polyodon spathula en Europe. Aquatic Living Resources 14:383–390. Venkatesh, B., M. V. Erdman, and S. Brenner. 2001. Molecular synapomorphies resolve evolutionary relationships of extant

LITERATURE CITED

jawed vertebrates. Proceedings of the National Academy of Sciences USA 98:11382–11387. Verheijen, F. J. 1963. Alarm substance in some North American cyprinid fishes. Copeia 1963:174–176. Vermeirssen, E. L. M., and A. P. Scott. 2001. Male priming pheromone is present in bile, as well as urine, of female Rainbow Trout. Journal of Fish Biology 58:1039–1045. Vermeirssen, E. L. M., A. P. Scott, and N. R. Liley. 1997. Female Rainbow Trout urine contains a pheromone which causes a rapid rise in plasma 17,20βα-dihydroxy-4-pregnen-3-one levels and milt amounts in males. Journal of Fish Biology 50:107–119. Verspoor, E., and J. Hammar. 1991. Introgressive hybridization in fishes: the biochemical evidence. Journal of Fish Biology 39 (Supplement A):309–334. Vessel, M. F., and S. Eddy. 1941. A preliminary study of the egg production of certain Minnesota fishes. Fisheries Research Report 26, Minnesota Department of Conservation, Division of Game and Fish. Vinikour, W. S. 1977. Incidence of Neascus rhinichthysi (Trematoda: Diplostomatidae) on longnose dace, Rhinichthys cataractae (Pisces: Cyprinidae), related to fish size and capture location. Transactions of the American Fisheries Society 106:83–88. Vinyard, G. L. 1996. Distribution of a thermal endemic minnow, the Desert Dace (Eremichthys acros), and observations of impacts of water diversion on its population. Great Basin Naturalist 56:360–368. Vinyard, G. L. 1997. Threatened fishes of the world: Eremichthys acros Hubbs and Miller, 1948 (Cyprinidae). Environmental Biology of Fishes 5:370. Vinyard, G. L., and A. C. Yuan. 1996. Effects of turbidity on feeding rates of Lahontan Cutthroat Trout (Onchorhynchus clarki henshawi) and Lahontan Redside Shiner (Richardsonius egrerius). Great Basin Naturalist 56:157–191. Vives, S. P. 1990. Nesting ecology and behavior of Hornyhead Chub, Nocomis biguttatus, a keystone species in Allequash Creek, Wisconsin. American Midland Naturalist 124:46–56. Vives, S. P. 1993. Choice of spawning substrate in Red Shiner with comments on crevice spawning in Cyprinella. Copeia 1993:229–232. Vives, S. P., and W. L. Minckley. 1990. Autumn spawning and other reproductive notes on Loach Minnow, a threatened cyprinid fish of the American Southwest. The Southwestern Naturalist 35:451–454. Vladic, T. V., and T. Järvi. 2001. Sperm quality in the alternative reproductive tactics of Atlantic Salmon: the importance of the loaded raffle mechanism. Proceedings of the Royal Society of London B 268:2375–2381. Vladykov, V. D. 1955. Eel fisheries of Quebec. Quebec Department of Fisheries Album No. 6:1–12. Vladykov, V. D. 1966. Remarks on the American Eel Anguilla rostrata (LeSueur): size of elvers entering streams; the relative abundance of adult males and females; and present economic importance of Eels in North America. Verhandlungen der Internationalen Vereinigung für Theoretische und Angewandte Limnologie 16:1007–1017. Vladykov, V. D. 1971. Homing of the American Eel, Anguilla rostrata, as evidenced by returns of transplanted tagged Eels in New Brunswick. Canadian Field Naturalist 85:241–248. Vladykov, V. D., and J. R. Greeley. 1963. Order Acipenseroidei, p. 24–60. In Fishes of the Western North Atlantic, No. 1, Pt. 3.

619

Y. H. Olsen (ed.). Sears Foundation for Marine Research, Yale University, New Haven, Connecticut. Vladykov, V. D., and E. Kott. 1976. Is Okkelbergia Creaser and Hubbs, 1922 (Petromyzonidae) a distinct taxon? Canadian Journal of Zoology 54:421–425. Vladykov, V. D., and E. Kott. 1978. A new nonparasitic species of the holarctic Lamprey genus Lethenteron Creaser and Hubbs, 1922, (Petromyzonidae) from northwestern North America with notes on other species of the same genus. Biological Papers of the University of Alaska No. 19:1–74. Vladykov, V. D., and E. Kott. 1979a. List of Northern Hemisphere Lampreys (Petromyzonidae) and their distribution. Fisheries and Oceans Miscellaneous Special Publication 42:1–30. Vladykov, V. D., and E. Kott. 1979b. Satellite species among the holarctic Lampreys (Petromyzonidae). Canadian Journal of Zoology 57:860–867. Vokoun, J. C. 2000. Shortnose Gar (Lepisosteus platostomus) foraging on periodical cicadas (Magicicada spp.): territorial defense of profitable pool positions. American Midland Naturalist 143:261–265. Vokoun, J. C., and D. B. Noltie. 2002. Evidence for the inheritance of alarm substance recognition in Johnny Darter (Etheostoma nigrum). American Midland Naturalist 147:400–403. Volkoff, H., J. P. Wourms, E. Amesbury, and F. F. Snelson. 1999. Structure of the thyroid gland, serum thyroid hormones, and the reproductive cycle of the Atlantic Stingray, Dasyatis sabina. Journal of Experimental Zoology 284:505–516. Vøllestad, L. A., B. Jonsson, N. A. Hvidsten, T. Næsje, Ø. Haraldstad, and J. Ruud-Hansen. 1986. Environmental factors regulating the seaward migration of European silver Eels (Anguilla anguilla). Canadian Journal of Fisheries and Aquatic Sciences 43:1909–1916. Vondracek, B., J. J. Cech, and D. Longanecker. 1982. Effect of cycling and constant temperatures on the respiratory metabolism of the Tahoe Sucker, Catostomus tahoensis (Pisces: Catostomidae). Comparative Biochemistry and Physiology 73A:11–14. von Frisch, K. 1938. Zur psychologie des Fisch-Schwarmes. Naturwissenschaften 26:601–606. von Frisch, K. 1941. Unber einen Schreckstoff der Fischhaut und seine biologische Bedeutung. Zeitschrift für vergleichende Physiologie 29:46–146. Vouglitois, J. J., K. W. Able, R. J. Kurtz, and K. A. Tighe. 1987. Life history and population dynamics of the Bay Anchovy in New Jersey. Transactions of the American Fisheries Society 116:141–153. Vrijenhoek, R. C. 1972. Genetic relationships of unisexual-hybrid fishes to their progenitors using lactate dehydrogenase isozymes as gene markers (Poeciliopsis, Poeciliidae). American Naturalist 106:754–766. Vrijenhoek, R. C. 1979. Factors affecting clonal diversity and coexistence. American Zoologist 19:787–797. Vrijenhoek, R. C. 1984. The evolution of clonal diversity in Poeciliopsis, p. 399–429. In Evolutionary Genetics of Fishes. B. J. Turner (ed.). Plenum Press, New York. Vrijenhoek, R. C. 1989. Genetic and ecological constraints on the origins and establishment of unisexual vertebrates, p. 19–24. In Evolution and Ecology of Unisexual Vertebrates. R. M. Dawley and J. P. Bogart (eds.). New York State Museum, Albany. Vrijenhoek, R. C., R. A. Angus, and R. J. Schultz. 1978. Variation and clonal structure in a unisexual fish. American Naturalist 112:41–55.

620

LITERATURE CITED

Vrijenhoek, R. C., E. Pfeiler, and J. D. Wetherington. 1992. Balancing selection in a desert stream-dwelling fish, Poeciliopsis monacha. Evolution 46:1642–1657. Waas, J. R., and P. W. Colgan. 1992. Chemical cues associated with visually elaborate aggressive displays of three-spine Sticklebacks. Journal of Chemical Ecology 18:2277–2284. Wagner, C. C., and E. L. Cooper. 1963. Population density, growth and fecundity of the Creek Chubsucker, Erimyzon oblongus. Copeia 1963:350–357. Wagner, C. M., M. L. Jones, M. B. Twohey, and P. W. Sorensen. 2006. A field test verifies that pheromones can be useful for Sea Lamprey (Petromyzon marinus) control in the Great Lakes. Canadian Journal of Fisheries and Aquatic Science 63:475–479. Wagner, C. M., M. B. Twohey, and J. M. Fine. 2009. Conspecific cueing in the Sea Lamprey: do reproductive migrations consistently follow the most intense larval odour? Animal Behaviour 78:593–599. Wagner, G. 1908. Note on the fish fauna of Lake Pepin. Transactions of the Wisconsin Academy of Science, Arts and Letters 16:23–37. Walbaum, J. J. 1792. Petri Artedi Sueci Generaa piscium. In quibus systema totum ichthyologiae proponitur cum classibus, ordinibus, generum, characteribus, specierum differentiis, observationibus plurimis. Redactis speciebus 242 ad genetra 52. Ichthylologiae. Pars iii, pt. 3:1–723. Walburg, C. H., and W. R. Nelson. 1966. Carp, River Carpsucker, Smallmouth Buffalo, and Bigmouth Buffalo in Lewis and Clark Lake, Missouri River. United States Bureau of Sport Fisheries and Wildlife Research Report 69, Washington, D.C. Waldman, B. 1987. Mechanisms of kin recognition. Journal of Theoretical Biology 128:159–185. Waldman, J., R. Daniels, M. Hickerson, and I. Wirgin. 2009. Mitochondrial DNA analysis indicates Sea Lampreys are indigenous to Lake Ontario: response to comment. Transactions of the American Fisheries Society 138:1190–1197. Waldman, J. R., P. Doukakis, and I. Wirgin. 2008b. Molecular analysis as a conservation tool for monitoring the trade of North American Sturgeons and Paddlefish. Journal of Applied Ichthyology 24 (Supplement 1):20–28. Waldman, J. R., C. Grunwald, N. K. Roy, and I. I. Wirgin. 2004. Mitochondrial DNA analysis indicates Sea Lampreys are indigenous to Lake Ontario. Transactions of the American Fisheries Society 133:950–960. Waldman, J. R., C. Grunwald, J. Stabile, and I. Wirgin. 2002. Impacts of life history and biogeography on the genetic stock structure of Atlantic Sturgeon Acipenser oxyrinchus oxyrinchus, Gulf Sturgeon A. oxyrinchus desotoi, and Shortnose Sturgeon A. brevirostrum. Journal of Applied Ichthyology 18:509–518. Waldman, J., C. Grunwald, and I. Wirgin. 2008a. Sea Lamprey, Petromyzon marinus: an exception to the rule of homing in anadromous fishes. Biological Letters 4:659–662. Walker, J. D., and J. W. Geissman, compilers. 2009. Geologic Time Scale: Geological Society of America. Available from http:// www.geosociety.org/science/timescale/timescl.pdf; accessed February 2010. Walker, P. 1993. A list of the endemic and introduced fishes of Colorado—March, 1993. Colorado Division of Wildlife, Aquatic Resources Unit, Unpubl. Manuscript, Fort Collins, Colorado. Walker, R. L., P. R. H. Wilkes, and C. M. Wood. 1989. The effects of hypersaline exposure on oxygen-affinity of the blood of the

freshwater teleost Catostomus commersoni. Journal of Experimental Biology 142:125–142. Wallace, C. R. 1970. The use of pheromones and sound during reproduction by the Black Bullhead, Ictalurus melas. State of Nebraska Game and Parks Commission Job Progress Reports, March 1, 1969–February 28, 1970:33–38. Wallin, J. E. 1989. Bluehead Chub (Nocomis leptocephalus) nest used by Yellowfin Shiners (Notropis lutipinnis). Copeia 1989:1077–1080. Wallin, J. E. 1992. The symbiotic nest association of Yellowfin Shiners, Notropis lutipinnis, and Bluehead Chubs, Nocomis leptocephalus. Environmental Biology of Fishes 33:287–292. Wallman, H. L., and W. A. Bennett. 2006. Effects of parturition and feeding on the thermal preference of Atlantic Stingray, Dasyatis sabina (LeSueur). Environmental Biology of Fishes 75:259–257. Wallus, R. 1986a. Paddlefish reproduction in the Cumberland and Tennessee River systems. Transactions of the American Fisheries Society 115:424–428. Wallus, R. 1986b. Larval development of Hiodon tergisus (LeSueur) with comparisons to Hiodon alosoides (Rafinesque). Journal of the Tennessee Academy of Sciences 61:77–80. Wallus, R. 1990a. Family Acipenseridae, p. 30–46. In Reproductive Biology and Early Life History of Fishes in the Ohio River Drainage. Volume 1: Acipenseridae through Esocidae. R. Wallus, T. P. Simon, and B. L. Yeager (eds.). Tennessee Valley Authority, Chattanooga, Tennessee. Wallus, R. 1990b. Family Hiodontidae, p. 153–166. In Reproductive Biology and Early Life History of Fishes in the Ohio River Drainage. Volume 1: Acipenseridae through Esocidae. Wallus, R., T. P. Simon, and B. L. Yeager (eds.). Tennessee Valley Authority, Chattanooga. Wallus, R., and J. P. Buchanan. 1989. Contribution to the reproductive biology and early life ecology of Mooneye in the Tennessee and Cumberland Rivers. American Midland Naturalist 122:204–207. Walser, C. A., B. Falterman, and H. L. Bart, Jr. 2000. Impact of introduced Rough Shiner (Notropis baileyi) on the native fish community in the Chattahoochee River system. American Midland Naturalist 144:393–405. Walsh, M. G., M. B. Bain, T. Squiers, Jr., J. R. Waldman, and I. Wirgin. 2001. Morphological and genetic variation among Shortnose Sturgeon Acipenser brevirostrum from adjacent and distant rivers. Estuaries 24:41–48. Walsh, P. J., G. D. Foster, and T. W. Moon. 1983. The effects of temperature on the metabolism of the American Eel Anguilla rostrata (LeSueur): compensation in the summer and torpor in the winter. Physiological Zoology 56:532–540. Walsh, S. J., D. C. Haney, C. M. Timmerman, and R. M. Dorazio. 1998. Physiological tolerances of juvenile Robust Redhorse, Moxoxtoma robustum: conservation implications for an imperiled species. Environmental Biology of Fishes 51:429–444. Walton, A. G., and P. Moller. 2010. Maze learning and recall in a weakly electric fish, Mormyrus rume proboscirostris Boulenger (Momyridae, Telostei). Ethology 116:909–919. Waltz, E. C. 1982. Alternative mating tactics and the law of diminishing returns: the satellite threshold model. Behavioral Ecology and Sociobiology 10:75–83. Wang, C., S. Zhang, and Y. Zhang. 2003. The karyotype of amphioxus Branchiostoma belcheri tsingtauense (Cephalochordata).

LITERATURE CITED

Journal of the Marine Biological Association of the United Kingdom 83:189–191. Wang, S. B., J. H. Cowan, Jr., K. A. Rose, and E. D. Houde. 1997. Individual-based modelling of recruitment variability and biomass production of Bay Anchovy in mid-Chesapeake Bay. Journal of Fish Biology 51:101–120. Wang, S. B., and E. D. Houde. 1994. Energy storage and dynamics in Bay Anchovy, Anchoa mitchilli. Marine Biology 121:219–227. Wang, S. B., and E. D. Houde. 1995. Distribution, relative abundance, biomass and production of Bay Anchovy, Anchoa mitchilli in the Chesapeake Bay. Marine Ecology Progress Series 121:27–38. Wang, Y., and J. M. Conlon. 1994. Purification and characterization of galanin from the phylogenetically ancient fish, the Bowfin (Amia calva) and dogfish (Scyliorhinus canicula). Peptides 15:981–986. Wang, Y., and J. M. Conlon. 1995. Purification and structural characterization of vasoactive intestinal polypeptide from the trout and Bowfin. General and Comparative Endocrinology 98:94–101. Wang, Y., J. H. Youson, and J. M. Conlon. 1993. Prosomatostatin-I is processed to somatostatin-26 and somatostatin-14 in the pancreas of the Bowfin, Amia calva. Regulatory Peptides 47:33–39. Wang, Y. L., F. P. Binkowski, and S. I. Doroshov. 1985. Effect of temperature on early development of White and Lake Sturgeon, Acipenser transmontanus and A. fulvescens, p. 43–50. In North American Sturgeons: Biology and Aquaculture Potential. F. P. Binkowski and S. I. Doroshov (eds.). Dr. W. Junk Publishers, Dordrecht, The Netherlands. Wantanabe, S. 2003. Taxonomy of the Freshwater Eels, genus Anguilla Shrank, 1798, p. 3–18. In Eel Biology. K. Aida, K. Tsukamoto, and K. Yamauchi (eds.). Springer, Tokyo. Ward, D. J., and Wiest, R. L., 1990. A checklist of Paleocene and Eocene sharks and rays (Chondrichthyes) from the Pamunkey Group, Maryland and Virginia, USA. Tertiary Research 12:81–88. Ward, H. B., and T. B. Magath. 1916. Notes on some nematodes from fresh-water fishes. The Journal of Parasitology 3:57–64. Ward, J. L., and D. A. McLennan. 2006. The relative influences of sexual and natural selection upon the evolution of male nuptial colouration in the Brook Stickleback, Culaea inconstans. Behaviour 143:483–510. Waring, C. P., and A. Moore. 1997. Sublethal effects of a carbamate pesticide on pheromonal mediated endocrine function in mature male Atlantic Salmon (Salmo salar L.) parr. Fish Physiology and Biochemistry 17:203–211. Waring, C. P., A. Moore, and A. P. Scott. 1996. Milt and endocrine responses of mature male Atlantic Salmon (Salmo salar L.) parr to water-borne testosterone, 17,20 ß dihydroxy-4-pregnen-one 20-sulfate and the urines from adult female and male salmon. General and Comparative Endocrinology 103:142–149. Warren, M. L., Jr. 2009. Centrarchid identification and natural history, p. 375–533. In Centrarchid Fishes: Diversity, Biology, and Conservation. S. J. Cooke and D. P. Philipp (eds.). WileyBlackwell, Chichester, West Sussex, United Kingdom. Warren, M. L., Jr., and B. M. Burr. 1994. Status of freshwater fishes of the United States: overview of an imperiled fauna. Fisheries 19:6–18. Warren, M. L., Jr., B. M. Burr, and J. M. Grady. 1994. Notropis albizonatus, a new cyprinid fish endemic to the Tennessee and

621

Cumberland river drainages, with a phylogeny of the Notropis procne species group. Copeia 1994:868–886. Warren, M. L., Jr., and M. G. Pardew. 1998. Road crossings as barriers to small-stream fish movement. Transactions of the American Fisheries Society 127:637–644. Warren, M. L., Jr., B. M. Burr, S. J. Walsh, H. L. Bart, Jr., R. C. Cashner, D. A. Etnier, B. J. Freeman, B. R. Kuhajda, R. L. Mayden, H. W. Robison, S. T. Ross, and W. C. Starnes. 2000. Diversity, distribution, and conservation status of the native freshwater fishes of the southern United States. Fisheries 25:7–31. Warren, M. L., A. L. Sheldon, and W. R. Haag. 2009. Constructed microhabitat bundles for sampling fishes and crayfishes in coastal plain streams. North American Journal of Fisheries Management 29:330–342. Watling, L., J. Lindsay, R. Smith, and D. Maurer 1974. The distribution of isopoda in the Delaware Bay region. Internationale Revue der gesamten Hydrobiologie und Hydrographie. 59: 343–351. Wayne, L. M. 1979. Ecology of the Roundnose Minnow, Dionda episcopa (Osteichthys: Cyprinidae) from three central Texas springs. Unpubl. Master’s thesis, Southwest Texas State University, San Marcos. Wayne, L. M., and B. G. Whiteside. 1985. Reproduction data on Dionda episcopa from Fessenden Spring, Texas. Texas Journal of Science 37:321–328. Wayne, W. J., and J. H. Zumberge. 1965. Pleistocene geology of Indiana and Michigan, p. 63–84. In The Quaternary of the United States. H. E. Wright, Jr., and D. G. Frey (eds.). Princeton University Press, Princeton, New Jersey. Webb, M. A. H., and D. L. Erickson. 2007. Reproductive structure of the adult Green Sturgeon, Acipenser medirostris, population in the Rogue River, Oregon. Environmental Biology of Fishes 79:305–314. Webb, M. A. H., J. P. Van Eenennaam, S. I. Doroshov, and G. P. Moberg. 1999. Preliminary observations on the effects of holding temperature on reproductive performance of female White Sturgeon, Acipenser transmontanus Richardson. Aquaculture 176:315–329. Webb, M. A. H., J. E. Williams, and L. R. Hildebrand. 2005. Recovery program review for endangered Pallid Sturgeon in the upper Missouri River basin. Reviews in Fisheries Science 13:165–176. Webb, P. W. 2008. The impacts of changes in water level and human development on forage fish assemblages in Great Lakes coastal marshes. Journal of Great Lakes Research 34:615–630. Webb, P. W., D. H. Hardy, and V. L. Mehl. 1992. The effect of armored skin on the swimming of Longnose Gar, Lepisosteus osseus. Canadian Journal of Zoology 70:1173–1179. Webb, S. A., J. A. Graves, C. Macias-Garcia, A. E. Magurran, D. O. Foighil, and M. G. Ritchie. 2004. Molecular phylogeny of the livebearing Goodeidae (Cyprinodontiformes). Molecular Phylogenetics and Evolution 30:527–544. Webber, H. M., and T. A. Haines. 2003. Mercury effects on predator avoidance of a forage fish, Golden Shiner (Notemigonus crysoleucas). Environmental Toxicology and Chemistry 22:1556–1561. Weber, R. E., B. Sullivan, J. Bonaventura, and C. Bonaventura. 1976. The hemoglobin system of the primitive fish, Amia calva: isolation and functional characterization of the individual

622 LITERATURE CITED

hemoglobin components. Biochemica et Biophysica Acta 434:18–31. Weech, S. A., A. M. Scheuhammer, J. E. Elliott, and K. M. Cheng. 2004. Mercury in fish from the Pinchi Lake region, British Columbia, Canada. Environmental Pollution. 131:275–286. Weed, A. C. 1925. Feeding the Paddlefish. Copeia 1925:67–68. Weeks, S. C., and O. E. Gaggiotti. 1993. Patterns of offspring size at birth in clonal and sexual strains of Poeciliopsis (Poeciliidae). Behavioral Ecology and Sociobiology 30:1–6. Weeks, S. C., O. E. Gaggiotti, R. A. Schenck, K. P. Spindler, and R. C. Vrijenhoek. 1992. Feeding-behavior in sexual and clonal strains of Poeciliopsis. Behavioral Ecology and Sociobiology 30:1–6. Wegner, K. M., M. Kalbe, M. Milinski, and T. B. H. Reusch. 2008. Mortality selection during the 2003 European heat wave in three-spined Sticklebacks: effects of parasites and MHC genotype. BMC Evolutionary Biology 8:124. Wegner, K. M., M. Kalbe, J. Kurtz, T. B. H. Reusch, and M. Milinski. 2003b. Parasite selection for immunogenetic optimality. Science 301:1343. Wegner, K. M., M. Kalbe, G. Rauch, J. Kurtz, H. Schaschl, and T. B. H. Reusch. 2006. Genetic variation in MHC class II expression and interactions with MHC sequence polymorphism in three-spined Sticklebacks. Molecular Ecology 15:1153–1164. Wegner, K. M., T. B. A. Reusch, and M. Kalbe. 2003a. Multiple parasites are driving major histocompatibility complex polymorphism in the wild. Journal of Evolutionary Biology 16:233–241. Weinstein, M. P. 1979. Shallow marsh habitats as primary nurseries for fishes and shellfish, Cape Fear River, North Carolina. Fishery Bulletin 77:339–357. Weisel, G. F. 1960. The osteocranium of the catostomid fish, Catostomus macrocheilus. A study in adaptation and natural relationship. Journal of Morphology 106:109–130. Weisel, G. F. 1973. Anatomy and histology of the digestive system of the Paddlefish (Polyodon spathula). Journal of Morphology 140:243–256. Weisel, G. F. 1975. The integument of the Paddlefish, Polyodon spathula. Journal of Morphology 145:143–150. Weisel, G. F. 1978. The integument and caudal filament of the Shovelnose Sturgeon, Scaphirhynchus platorynchus. The American Midland Naturalist 100:179–189. Weiss, J. L., and J. B. Layzer. 1995. Infestations of glochidia on fishes in the Barren River, Kentucky. American Malacological Bulletin 11:153–159. Welch, D. W., S. Turo, and S. D. Batten. 2006. Large-scale marine and freshwater movement of White Sturgeon. Transactions of the American Fisheries Society 135:386–389. Welcomme, R. L. 1988. International introductions of Inland aquatic species. FAO Fisheries Technical Paper 294:1–318. Welker, T. L., and D. L. Scarnecchia. 2003. Differences in species composition and feeding ecology of catostomid fishes in two distinct segments of the Missouri River, North Dakota, U.S.A. Environmental Biology of Fishes 68:129–141. Wenner, C. A. 1973. Occurrence of American Eels, Anguilla rostrata, in water overlying the eastern North America Continental Shelf. Journal of the Fisheries Research Board of Canada 30:1752–1755. Wenner, C. A., and J. A. Musick. 1974. Fecundity and gonad observations of the American Eel, Anguilla rostrata, migrating from

Chesapeake Bay, Virginia. Journal of the Fisheries Research Board of Canada 31:1387–1391. Werner, E. E., and J. F. Gilliam. 1984. The ontogenetic niche and species interactions in size-structured populations. Annual Review of Ecology and Systematics 15:393–425. Werner, E. E., J. F. Gilliam, D. J. Hall, and G. G. Mittelbach. 1983b. An experimental test of the effects of predation risk on habitat use in fish. Ecology 64:1540–1548. Werner, E. E., and D. J. Hall. 1976. Niche shifts in Sunfishes: experimental evidence and significance. Science 191:404–406. Werner, E. E., and D. J. Hall. 1988. Ontogenetic habitat shifts in Bluegill: the foraging rate—predation risk trade-off. Ecology 69:1352–1366. Werner, E. E., D. J. Hall, D. R. Laughlin, D. J. Wagner, L. A. Wilsmann, and F. C. Funk. 1977. Habitat partitioning in a freshwater fish community. Journal of the Fisheries Research Board of Canada 34:360–370. Werner, E. E., D. J. Hall, and M. D. Werner. 1978. Littoral zone fish communities of two Florida lakes and a comparison with Michigan lakes. Environmental Biology of Fishes 3:163–172. Werner, E. E., G. G. Mittelbach, D. J. Hall, and J. F. Gilliam. 1983a. Experimental tests of optimal habitat use in fish: the role of relative habitat profitability. Ecology 64:1525–1539. Werner, R. 1979. Homing mechanism of spawning White Suckers in Wolf Lake, New York. New York Game and Fish Journal 26:48–58. Westers, H., and R. R. Stickney. 1993. Northern Pike and Muskellunge, p. 199–213. In Culture of Nonsalmonid Freshwater Fishes. R. R. Stickney (ed.). CRC Press, Boca Raton, Florida. Wetherington, J. D., S. C. Weeks, K. E. Kotora, and R. C. Vrijenhoek. 1989. Genotypic and environmental components of variation in growth and reproduction of fish hemiclones (Poeciliopsis: Poeciliidae). Evolution 43:635–645. Whitaker, J. O., Jr. 1976. Fish community changes at one Vigo Country, Indiana locality over a twelve year period. Proceedings of the Indiana Academy of Science 85:191–207. Whitaker, J. O., Jr., and R. A. Schlueter. 1975. Occurence of the crustacean parasite, Lernaea cyprinacea, on fishes from the White River at Petersburg, Indiana. American Midland Naturalist 93:446–450. White, D. S. 1974. The biology of Minytrema melanops (Rafinesque), the Spotted Sucker. Unpubl. Ph.D. diss., University of Louisville, Louisville, Kentucky. White, D. S., and K. H. Haag. 1977. Foods and feeding habits of the Spotted Sucker, Minytrema melanops (Rafinesque). American Midland Naturalist 98:137–146. White, E. M., and B. Knights. 1997. Environmental factors affecting migration of the European Eel in the Rivers Severn and Avon, England. Journal of Fish Biology 50:1104–1116. White, J. L., and B. C. Harvey. 2001. Effects of an introduced piscivorous fish on native benthic fishes in a coastal river. Freshwater Biology 46:987–995. White, J. L., and B. C. Harvey. 2003. Basin-scale patterns in the drift of embryonic and larval fishes and lamprey ammocoetes in two coastal rivers. Environmental Biology of Fishes 67:369–378. White, M. M. 1987. Genetic variation and population structuring in the Rosyside Dace, Clinostomus funduloides, in Ohio. Ohio Journal of Science 88:114–116.

LITERATURE CITED

White, M. W., and N. Aspinwall. 1984. Habitat partitioning among five species of darters (Percidae: Etheostoma), p. 55–60. In Environmental Biology of Darters. D. G. Lindquist and L. M. Page (eds.). Dr. W. Junk Publishers, The Hague. Whitehead, D. L. 2002. Ampullary organs and electroreceptors in freshwater Carcharhinus leucas. Journal of Physiology 96:391–395. Whitehead, D. R. 1973. Late-Wisconsin vegetational changes in unglaciated eastern North America. Quaternary Research 3:621–631. Whitehead, P. J. P., G. J. Nelson, and T. Wongratana. 1988. FAO species cata logue: clupeoid fishes of the world (Suborder Clupeoidei)—Part 2 Engraulidae. FAO Fisheries Synopses 125(7.2):305–579. Whitehurst, D. K. 1981. Seasonal movements of fishes in an eastern North Carolina swamp stream, p. 182–190. In The Warmwater Streams Symposium. L. A. Krumholz (ed.). Southern Division, American Fisheries Society, Allen Press, Lawrence, Kansas. Whiteman, K. W., V. H. Travnichek, M. L. Wildhaber, A. DeLonay, D. Papoulias, and D. Tillett. 2004. Age estimation for Shovelnose Sturgeon: a cautionary note based on annulus formation in pectoral fin rays. North American Journal of Fisheries Management 24:731–734. Whitman, C. O., and A. C. Eycleshymer. 1897. The egg of Amia and its cleavage. Journal of Morphology 12:309–355. Whitmore, D. H. 1983. Introgressive hybridization of Smallmouth Bass (Micropterus dolomieu) and Guadalupe Bass (M. treculii). Copeia 1983:672–679. Whitney, M., and M. C. Belk. 2000. Threatened fishes of the world: Chasmistes liorus Jordan, 1878 (Catostomidae). Environmental Biology of Fishes 57:362. Wiegmann, D. D., and J. R. Baylis. 1995. Male body size and paternal behaviour in Smallmouth Bass, Micropterus dolomieu (Pisces: Centrarchidae). Animal Behaviour 50:1543–1555. Wiegmann, D. D., J. R. Baylis, and M. H. Hoff. 1992. Sexual selection and fitness variation in a population of Smallmouth Bass, Micropterus dolomieu (Pisces: Centrarchidae). Evolution 46:1740–1753. Wikramanayake, E. D., and P. B. Moyle. 1989. Ecological structure of tropical fish assemblages in wet-zone streams of Sri Lanka. Journal of Zoology, London 218:503–526. Wilde, G. R., and B. W. Durham. 2008. Daily survival rates for juveniles of six Great Plains cyprinid species. Transactions of the American Fisheries Society 137:830–833. Wilder, B. G. 1877. Garpikes, old and young. Popular Science Monthly 11:1–12 Wildhaber, M. L., and L. B. Crowder. 1990. Testing a bioenergeticsbased habitat choice model: Bluegill (Lepomis macrochirus) responses to food availability and temperature. Canadian Journal of Fisheries and Aquatic Sciences 47:1664–1671. Wildhaber, M. L., D. M. Papoulias, A. J. DeLonay, D. E. Tillitt, J. L. Bryan, and M. L. Annis. 2007. Physical and hormonal examination of Missouri River Shovelnose Sturgeon reproductive stage: a reference guide. Journal of Applied Ichthyology 23:382–401. Wiley, E. O. 1976. The phylogeny and biogeography of fossil and recent gars (Actinopterygii: Lepisosteidae). Miscellaneous Publication—Museum of Natural History, University of Kansas 64:1–111.

623

Wiley, E. O., and R. L. Mayden. 1985. Species and speciation in phylogenetic systematics, with examples from the North American fish fauna. Annals of the Missouri Botanical Garden 72: 596–635. Wiley, E. O., and H.-P. Schultze. 1984. Family Lepisosteidae (Gars) as living fossils, p. 160–165. In Living Fossils. N. Eldredge and S. M. Stanley (eds.). Springer-Verlag, New York. Wiley, M. L., and B. B. Collette. 1970. Breeding tubercles and contact organs in fishes: their occurence, structure, and significance. Bulletin of the American Museum of Natural History 143:143–216. Wilga, C. D., and G. V. Lauder. 1999. Locomotion in Sturgeon: function of the pectoral fins. The Journal of Experimental Biology 202:2413–2432. Wilimovsky, N. J. 1956. Protoscaphirhynchus squamosus, a new Sturgeon from the Upper Cretaceous of Montana. Journal of Paleontology 30:1205–1209. Wilkens, H. 1971. Genetic interpretation of regressive evolutionary processes: studies on hybrid eyes of two Astyanax cave populations (Characidae, Pisces). Evolution 25:530–544. Wilkens, L. A., and M. H. Hofmann. 2007. The Paddlefish rostrum as an electrosensory organ: a novel adaptation for plankton feeding. Bioscience 57:399–407. Wilkens, L. A., M. H. Hofmann, and W. Wojtenek. 2002. The electric sense of the Paddlefish: a passive system for the detection and capture of zooplankton prey. Journal of Physiology—Paris 96:363–377. Wilkens, L. A., D. F. Russell, X. Pei, and C. Gurgens. 1997. The Paddlefish rostrum functions as an electrosensory antenna in plankton feeding. Proceedings of the Royal Society of London Series B 264:1723–1729. Wilkens, L. A., B. Wettring, E. Wagner, W. Wojtenek, and D. Russell. 2001. Prey detection in selective plankton feeding by the Paddlefish: is the electric sense sufficient? The Journal of Experimental Biology 204:1381–1389. Wilkes, P. R. H., and B. R. McMahon. 1986a. Responses of a stenohaline freshwater teleost (Catostomus commersoni) to hypersaline exposure I. The dependence of plasma pH and bicarbonate concentration on electrolyte regulation. Journal of Experimental Biology 121:77–94. Wilkes, P. R. H., and B. R. McMahon. 1986b. Responses of a stenohaline freshwater teleost (Catostomus commersoni) to hypersaline exposure II. Transepithelial flux of sodium, chloride and “acidic equilavents”. Journal of Experimental Biology 121:95–113. Wilkie, M. P., P. G. Bradshaw, V. Ioanis, J. E. Claude, and S. L. Swindell. 2001. Rapid metabolic recovery following vigorous exercise in burrow-dwelling larval Sea Lampreys (Petromyzon marinus). Physiological and Biochemical Zoology 74:261–272. Wilkie, M. P., J. Couturier, and B. L. Tufts. 1998. Mechanisms of acid-base regulation in migrant Sea Lampreys (Petromyzon marinus) following exhaustive exercise. The Journal of Experimental Biology 201:1473–1482. Wilkie, M. P., S. Turnbull, J. Bird, Y. S. Wang, J. F. Claude, and J. H. Youson. 2004. Lamprey parasitism of sharks and teleosts: high capacity urea excretion in an extant vertebrate relic. Comparative Biochemistry and Physiology 138A:485– 492.

624

LITERATURE CITED

Wilkins, D. S. 1992. Life history of the Least Madtom (Ictaluridae: Noturus hildebrandi) from Bayou Pierre, Mississippi. Unpubl. Master’s thesis, University of Southern Mississippi, Hattiesburg. Williams, E. H., Jr., and L. B. Williams. 1978. Cymothoid isopods of some marine fishes from the northern Gulf of Mexico. Northeast Gulf Science 2:122–124. Williams, J. D. 1975. Systematics of the percid fishes of the subgenus Ammocrypta, genus Ammocrypta, with descriptions of two new species. Unpubl. Ph.D. diss., The University of Alabama, Tuscaloosa. Williams, J. D., A. E. Bogan, and J. T. Garner. 2008. Freshwater mussels of Alabama and the Mobile Basin in Georgia, Mississippi and Tennessee. University of Alabama Press, Tuscaloosa. Williams, J. D., and G. H. Clemmer. 1991. Scaphirhynchus suttkusi, a new Sturgeon (Pisces: Acipenseridae) from the Mobile Basin of Alabama and Mississippi. Bulletin Alabama Museum of Natural History 10:17–31. Williams, J. E. 1995. Threatened fishes of the world: Catostomus warnerensis Snyder, 1908 (Catostomidae). Environmental Biology of Fishes 44:346. Williams, J. E., J. E. Johnson, D. A. Hendrickson, S. ContrerasBalderas, J. D. Williams, M. Navarro-Mendoza, D. E. McAllister, and J. E. Deacon. 1989. Fishes of North America endangered, threatened, or of special concern: 1989. Fisheries 14:2–20. Williams, J. E., and C. E. Bond. 1983. Status and life history notes on the native fishes of the Alvord Basin, Oregon and Nevada. Great Basin Naturalist 43:409–420. Williams, J. E., and C. D. Williams. 1980. Feeding ecology of Gila boraxobius (Osteichthyes: Cyprinidae) endemic to a thermal lake in southeastern Oregon. The Great Basin Naturalist 40:101–114. Williamson, D. F. 2003. Caviar and Conservation: Status, Management, and Trade of North American Sturgeon and Paddlefish. TRAFFIC North America / World Wildlife Fund, Washington, D.C. Williamson, D. F., G. W. Benz, and C. M. Hoover (eds.). 1999. Proceedings of the Symposium on the Harvest and Conservation of North American Paddlefish and Sturgeon, May 7–8, 1998, Chattanooga, TN. TRAFFIC North America / World Wildlife Fund, Washington, D.C. Williot, P. (ed.). 1991. Acipenser: Actes du Premier Colloque International sur l’Esturgeon, Bordeaux, 3–6 octobre 1989. CEMAGREF-DICOVA, Bordeaux, France. Willis, M. S. 2001. Population biology of Allocreadium lobatum Wallin, 1909 (Digenea: Allocreadiidae) in the Creek Chub, Semotilus atromaculatus, Mitchill (Osteichthyes: Cyprinidae), in a Nebraska Creek, USA. Memórias do Instituto Oswaldo Cruz 96:331–338. Wilson, C. B. 1916. Copepod parasites of fresh-water fishes and their economic relations to mussel glochidia. Bulletin of the United States Bureau of Fisheries (Document 824) 34:333–374. Wilson, J. A., and R. S. McKinley. 2004. Distribution, habitat, and movements, p. 40–72. In Sturgeons and Paddlefish of North America. G. T. O. LeBreton, F. W. Beamish, and R. S. McKinley (eds.). Kluwer Academic Publishers, Dordrecht, The Netherlands. Wilson, J. B. 1999. Guilds, functional types and ecological groups. Oikos 86:507–522. Wilson, K. A., and K. Ronald. 1967. Parasite fauna of the Sea Lamprey (Petromyzon marinus von Linné) in the Great Lakes region. Canadian Journal of Zoology 45:1083–1092.

Wilson, M. V. H. 1977. Middle Eocene freshwater fishes from British Columbia. Life Science Contributions of the Royal Ontario Museum 113:1–61. Wilson, M. V. H. 1978. Eohiodon woodruffi n. sp. (Teleostei, Hiodontidae) from the middle Eocene Klondike Mountain Formation near Republic, Washington. Canadian Journal of Earth Sciences 15:679–686. Wilson, M. V. H. 1980. Oldest known Esox (Pisces: Esocidae), part of a new Paleocene teleosts fauna from western Canada. Canadian Journal of Earth Science 17:307–312. Wilson, M. V. H. 1984. Year classes and sexual dimorphism in the Eocene catostomid fish Amyzon aggregatum. Journal of Vertebrate Paleontology 3:137–142. Wilson, M. V. H., and A. M. Murray. 2008. Osteoglossomorpha: phylogeny, biogeography, and fossil record and the significance of key African and Chinese fossil taxa, p. 185–219. In Fishes and the Break-up of Pangaea. L. Cavin, A. Longbottom, and M. Richter (eds.). Geological Society, London. Wilson, M. V. H., and R. R. G. Williams. 1992. Phylogenetic, biogeographic, and ecological significance of early fossil records of North American freshwater teleostean fishes, p. 224–244. In Systematics, Historical Ecology, and North American Freshwater Fishes. R. L. Mayden (ed.). Stanford University Press, Stanford, California. Wilson, R. J. A., M. B. Harris, J. E. Remmers, and S. F. Perry. 2000. Evolution of air-breathing and central CO2/H+ respiratory chemosensitivity: new insights from an old fish. Journal of Experimental Biology 203:3505–3512. Wilson, R. J. A., B. E. Taylor, and M. B. Harris. 2009. Evolution of vertebrate respiratory neural control, p. 67–75. In Encyclopedia of Neuroscience. G. Adelman and B. H. Smith (eds.). Elsevier Ltd., London. Winberg, S., and K. H. Olsén. 1992. The influence of rearing conditions on the sibling odour preference of juvenile Arctic Char, Salvelinus alpinus L. Animal Behaviour 44:157–164. Winfield, G. F. 1929. Plesiocreadium typicum, a new trematodes from Amia calva. The Journal of Parasitology 16:81–87. Winfield, I. J. and J. S. Nelson. 1991. Cyprinid Fishes: Systematics, Biology and Exploitation. Chapman and Hall, London. Winge, Ö. 1937. Succession of broods in Lebistes. Nature 140:467. Winkelman, D. L. 1996. Reproduction under predatory threat: trade-offs between nest guarding and predator avoidance in male Dollar Sunfish (Lepomis marginatus). Copeia 1996:845–851. Winn, H. E. 1958. Comparative reproductive behavior and ecology of fourteen species of darters (Pisces: Percidae). Ecological Monographs 28:155–191. Winn, H. E., W. A. Richkus, and L. K. Winn. 1975. Sexual dimorphism and natural movements of the American Eel (Anguilla rostrata) in Rhode Island streams and estuaries. Helgoländer Meeresuntersuchungen 27:156–166. Winn, H. E., and J. F. Stout. 1960. Sound production by the Satinfin Shiner, Notropis analostanus, and related fishes. Science 132:222–223. Winston, M. R., C. M. Taylor, and J. Pigg. 1991. Upstream extirpation of four minnow species due to damming of a prairie stream. Transactions of the American Fisheries Society 120:98–105. Winston, M. R. 1995. Co-occurrence of morphologically similar species of stream fishes. The American Naturalist 145:527–545.

LITERATURE CITED

Winston, M. R. 1995. Co-occurrence of morphologically similar species of stream fishes. The American Naturalist 145:527–545. Winston, W. D., Jr. 1967. Effects of temperature and light on the rate of aerial breathing of the Alligator Gar, Lepisosteus spatula. Unpubl. Master’s thesis, University of Oklahoma, Norman. Winterbottom, R. 1974. A descriptive synonymy of the striated muscles of the Teleostei. Proceedings of the Academy of Natural Sciences of Philadelphia 125:225–317. Wippelhauser, G. S., and J. D. McCleave. 1987. Precision of behavior of migrating juvenile American Eels (Anguilla rostrata) utilizing selective tidal stream transport. ICES Journal of Marine Science 44:80–89. Wirgin, I., D. D. Currie, J. Stabile, and C. A. Jennings. 2004. Development and use of a simple DNA test to distinguish larval Redhorse species in the Oconee River, Georgia. North American Journal of Fisheries Management 24:293–298. Wirgin, I., C. Grunwald, E. Carlson, J. Stabile, D. L. Peterson, and J. Waldman. 2005. Range-wide population structure of Shortnose Sturgeon Acipenser brevirostrum based on sequence analysis of the mitochondrial DNA control region. Estuaries 28:406–421. Wirgin, I., T. Oppermann, and J. Stabile. 2001. Genetic divergence of Robust Redhorse Moxostoma robustum (Cypriniformes Catostomidae) from the Oconee River and the Savannah River based of mitochondrial DNA control region sequences. Copeia 2001:526–530. Wirgin, I. I., J. E. Stabile, and J. R. Waldman. 1997. Molecular analysis in the conservation of Sturgeons and Paddlefish. Environmental Biology of Fishes 48:385–398. Wirgin, I., J. R. Waldman, J. Rosko, R. Gross, M. R. Collins, S. G. Rogers, and J. E. Stabile. 2000. Genetic structure of Atlantic Sturgeon populations based on mitochondrial DNA control region sequences. Transactions of the American Fisheries Society 129:476–486. Wirgin, I., J. Waldman, J. Stabile, B. Lubinski, and T. King. 2002. Comparison of mitochondrial DNA control region sequence and microsatellite DNA analyses in estimating population structure and gene flow rates in Atlantic Sturgeon Acipenser oxyrinchus. Journal of Applied Ichthyology 18:313–319. Wisenden, B. D. 2008. Active space of chemical alarm cue in natural fish populations. Behaviour 145:391–407. Wisenden, B. D., and K. Barbour. 2005. Antipredator responses to skin extract of Redbelly Dace, Phoxinus eos, by free-ranging populations of Redbelly Dace and Fathead Minnows, Pimephales promelas. Environmental Biology of Fishes 72:227–233. Wisenden, B. D., D. P. Chivers, G. E. Brown, and R. J. F. Smith. 1995b. The role of experience in risk assessment: avoidance of areas chemically labelled with Fathead Minnow alarm pheromone by conspecifics and heterospecifics. Ecoscience 2:116–122. Wisenden, B. D., D. P. Chivers, and R. J. F. Smith. 1994. Risksensitive habitat use by Brook Stickleback (Culaea inconstans) in areas associated with minnow alarm pheromone. Journal of Chemical Ecology 20:2975–2982. Wisenden, B. D., D. P. Chivers, and R. J. F. Smith. 1995a. Early warning of risk in the predation sequence: a disturbance pheromone in Iowa Darters (Etheostoma exile). Journal of Chemical Ecology 21:1469–1480. Wisenden, B. D., J. Karst, J. Miller, S. Miller, and L. Fuselier. 2008a. Anti-predator behaviour in response to conspecific

625

chemical alarm cues in an esociform fish, Umbra limi (Kirtland 1840). Environmental Biology of Fishes 82:85–92. Wisenden, B. D., J. Pogatshnik, D. Gibson, L. Bonacci, A. Schumacher, and A. Willett. 2008b. Sound the alarm: learned association of predation risk with novel auditory stimuli by Fathead Minnows (Pimephales promelas) and Glowlight Tetras (Hemigrammus erythrozonus) after single simultaneous pairings with conspecific chemical alarm cues. Environmental Biology of Fishes 81:141–147. Wisenden, B. D., M. S. Pollock, R. J. Tremaine, J. M. Webb, M. E. Wismer, and D. P. Chivers. 2003. Synergistic interactions between alarm cues and the presence of conspecific and heterospecific fish shoals. Behavioral Ecology and Sociobiology 54:485–490. Wisenden, B. D., M. L. Rugg, N. L. Korpi, and L. C. Fuselier. 2009a. Lab and field estimates of active time of chemical alarm cues of a cyprinid fish and an amphipod crustacean. Behaviour 146:1423–1442. Wisenden, B. D., and R. J. F. Smith. 1997. The effect of physical condition and shoalmate familiarity on the proliferation of alarm substance cells in the epidermis of Fathead Minnows. Journal of Fish Biology 50:799–808. Wisenden, B. D., and T. A. Thiel. 2002. Field verification of predator attraction to minnow alarm substance. Journal of Chemical Ecology 28:433–438. Wisenden, B. D., K. A. Vollbrecht, and J. L. Brown. 2004. Is there a fish alarm cue? Affirming evidence from a wild study. Animal Behaviour 67:59–67. Wisenden, B. D., A. Unruh, A. Morantes, S. Bury, B. Curry, R. Driscoll, M. Hussein, and S. Markegard. 2009b. Functional constraints on nest characteristics of pebble mounds of breeding male Hornyhead Chub, Nocomis biguttatus. Journal of Fish Biology 75:1577–1585. Witte, K., and R. Massmann. 2003. Female Sailfin Mollies, Poecilia latipinna, remember males and copy the choice of others after 1 day. Animal Behaviour 65:1151–1159. Witte, K., and B. Noltemeier. 2002. The role of information in mate-choice copying in female Sailfin Mollies (Poecilia latipinna). Behavioral Ecology and Sociobiology 52:194–202. Witte, K., and M. J. Ryan. 1998. Male body length influences mate-choice copying in the Sailfin Molly, Poecilia latipinna. Behavioral Ecology 9:534–539. Witte, K., and M. J. Ryan. 2002. Mate-choice copying in the Sailfin Molly (Poecilia latipinna) in the wild. Animal Behaviour 63:943–949. Witte, K., and K. Ueding. 2003. Sailfin Molly females (Poecilia latipinna) copy the rejection of a male. Behavioral Ecology and Sociobiology 14:389–395. Witte, C. C., M. L. Wildhaber, A. Arab, and D. B. Noltie. 2009. Substrate choice of territorial male Topeka Shiners (Notropis topeka) in the absence of Sunfish (Lepomis sp.). Ecology of Freshwater Fish 18:350–359. Wittenberg, J. B., and B. A. Wittenberg. 1974. Choroid rete mirabile of fish eye. 1. Oxygen secretion and structure—comparison with swimbladder rete mirabile. Biological Bulletin 146:116–136. Witting, D. A., K. W. Able, and M. P. Fahay. 1999. Larval fishes of a middle Atlantic Bight estuary: assemblage structure and temporal stability. Canadian Journal of Fisheries and Aquatic Sciences 56:222–230.

626

LITERATURE CITED

Wojtenek, W., X. Pei, and L. A. Wilkens. 2001. Paddlefish strike at artificial dipoles simulating the weak electric fields of planktonic prey. The Journal of Experimental Biology 204:1391–1399. Wolf, C., P. Hübner, and J. Lüthy. 1999. Differentiation of Sturgeon species by PCR-RFLP. Food Research International 32:699–705. Wolter, C., and F. Röhr. 2010. Distribution history of non-native freshwater fish species in Germany: how invasive are they? Journal of Applied Ichthyology 26 (Supplement 2):19–27. Wong, B. B. M., U. Candolin, and K. Lindström. 2007. Environmental deterioration compromises socially-enforced signals of male quality in three-spined Sticklebacks. American Naturalist 170:184–189. Wong, B. B. M., H. S. Fisher, and G. G. Rosenthal. 2005. Species recognition by male swordtails via chemical cues. Behavioral Ecology 16:818–822. Wood, C. M., J. D. Turner, and M. S. Graham. 1983. Why do fish die after severe exercise? Journal of Fish Biology 22:189–201. Wood, R. M., R. L. Mayden, R. H. Matson, B. R. Kuhajda, and S. R. Layman. 2002. Systematics and biogeography of the Notropis rubellus species group (Teleostei: Cyprinidae). Bulletin of the Alabama Museum of Natural History 22:37–80. Woodling, J. 1985. Colorado’s Little Fish: A Guide to the Minnows and Other Lesser Known Fishes in the State of Colorado. Colorado Division of Wildlife, Denver. Woodward, R. L., and T. E. Wissing. 1976. Age, growth, and fecundity of the Quillback (Carpiodes cyprinus) and Highfin (C. velifer) Carpsuckers in an Ohio stream. Transactions of the American Fisheries Society 105:411–415. Woolcott, W. S., and W. L. Kirk. 1976. Melanism in Lepisosteus osseus from the James River, Virginia. Copeia 1976:815–817. Wooley, C. M. 1985. Evaluation of morphometric characters used in taxonomic separation of Gulf of Mexico Sturgeon, Acipenser oxyrhynchus desotoi, p. 97–103. In North American Sturgeons: Biology and Aquaculture Potential. F. P. Binkowski and S. I. Doroshov (eds.). Dr. W. Junk Publishers, Dordrecht, The Netherlands. Wootton, R. J. 1971. A note on the nest-raiding behavior of male Sticklebacks. Canadian Journal of Zoology 49:960–962. Wrege, B. M., M. S. Duncan, and J. J. Isely. 2011. Diel activity of Gulf of Mexico Sturgeon in a northwest Florida bay. Journal of Applied Ichthyology 27:322–326. Wrenn, W. B. 1968. Life history aspects of Smallmouth Buffalo and Freshwater Drum in Wheeler Reservoir, Alabama. Proceedings of the Southeastern Association of Game and Fish Commisioners 22:479–495. Wright, J. J., S. R. David, and T. J. Near. 2012. Gene trees, species trees, and morphology converge on a similar phylogeny of living gars (Actinopterygii: Holostei: Lepisosteidae), an ancient clade of Ray-Finned Fishes. Molecular Phylogenetics and Evolution 63:848–856. Wright, P. A. 2007. Ionic, osmotic, and nitrogenous waste regulation. Fish Physiology 26:283–318. Wu, H.-W., Y. Chen, X. Chen, and J. Chen. 1981. A taxonomical system and phylogenetic relationship of the families of the suborder Cyprinoidei. Scientia Sinica 24:563–572. Wund, M. A., J. A. Baker, B. Clancy, J. L. Golub, and S. A. Foster. 2008. A test of the “flexible stem” model of evolution: ancestral plasticity, genetic accommodation, and morphological divergence in the Threespine Stickleback radiation. The American Naturalist 172:449–462.

Wydoski, R. S., and R. R. Whitney. 2003. Inland Fishes of Washington. 2nd edition. University of Washington Press, Seattle. Wydoski, R. G., and R. S. Wydoski. 2002. Age, growth and reproduction of Mountain Suckers in Lost Creek Reservoir, Utah. Transactions of the American Fisheries Society 131:320–328. Wysocki, L. E., and F. Ladich. 2002. Can fishes resolve temporal characteristics of sounds? New insights using auditory brainstem responses. Hearing Research 169:36–46. Yakovlev, V. N. 1977. Phylogenesis of Acipenseriformes, p. 116– 143. In Essays on Phylogenies and Systematics of Fossil Fishes and Agnathans. V. V. Menner (ed.). USSR Academy of Sciences, Moscow, USSR (in Russian). Yamazaki, Y., N. Fukutomi, K. Takeda, and A. Iwata. 2003. Embryonic development of the Pacific Lamprey, Entosphenus tridentatus. Zoological Science 20:1095–1098. Yamazaki, Y., and A. Goto. 1998. Genetic structure and differentiation of four Lethenteron taxa from the Far East, deduced from allozyme analysis. Environmental Biology of Fishes 52:149–161. Yambe, H. 2001. A releaser pheromone that attracts methyltestosterone treated immature fish in the urine of ovulated female Rainbow Trout. Fisheries Science 67:214–220. Yambe, H., and F. Yamazaki. 2001a. A releaser pheromone in the urine of mature female Rainbow Trout demonstrated by methyltestosterone treated immature fish. Fisheries Science 67:214–220. Yambe, H., and F. Yamazaki. 2001b. Species-specific releaser effect of urine from ovulated female Masu Salmon and Rainbow Trout. Journal of Fish Biology 59:1455–1464. Yan, H. Y., M. L. Fine, N. S. Horn, and W. E. Colón. 2000. Variability in the role of the gasbladder in fish audition. Journal of Comparative Physiology A 186:435–445. Yant, P. R., J. R. Karr, and P. L. Angermeier. 1984. Stochasticity in stream fish communities: an alternative interpretation. The American Naturalist 124:573–582. Yap, M. R., and S. H. Bowen. 2003. Feeding by Northern Brook Lamprey (Ichthyomyzon fossor) on sestonic biofilm fragments: habitat selection results in ingestion of a higher quality diet. Journal of Great Lakes Research 29(Suppl. 1):15–25. Yavno, S. and L. D. Corkum. 2010. Reproductive female Round Gobies (Neogobius melanostomus) are attracted to visual male models at a nest rather than to olfactory stimuli in urine of reproductive males. Behaviour 147:121–132. Yeager, B. L. 1980. Early development of the genus Carpiodes (Osteichthyes: Catostomidae). Unpubl. Master’s thesis, University of Tennessee, Knoxville. Yeager, B. L., and J. M. Baker. 1982. Early development of the genus Ictiobus (Catostomidae), p. 63–69. In Proceedings of the Fifth Annual Larval Fish Conference, 2–3 March 1981. C. F. Bryan, J. V. Conner, and F. M. Truesdale (eds.). Louisiana State University Press, Baton Rouge. Yeager, B. L., and K. J. Semmens. 1987. Early development of the Blue Sucker Cycleptus elongatus. Copeia 1987:312–316. Yeager, B. L., and R. Wallus. 1982. Development of larval Polyodon spathula (Walbaum) from the Cumberland River in Tennessee, p. 73–77. In Proceedings of the Fifth Annual Larval Fish Conference, 2–3 March 1981. C. F. Bryan, J. V. Conner, and F. M. Truesdale (eds.). Louisiana State University Press, Baton Rouge. Yeager, B. L., and R. Wallus. 1990. Family Polyodontidae, p. 49–55. In Reproductive Biology and Early Life History of Fishes in the Ohio River Drainage. Volume 1: Acipenseridae through Esoci-

LITERATURE CITED

dae. R. Wallus, T. P. Simon, and B. L. Yeager (eds.). Tennessee Valley Authority, Chattanooga. Yeager, L. E. 1936. An observation of spawning Buffalofish in Mississippi. Copeia 1936:238–239. Yoakim, E. G., and J. M. Grizzle. 1982. Ultrastructure of alarm substance cells in the epidermis of the Channel Catfish, Ictalurus punctatus (Rafinesque). Journal of Fish Biology 20: 213–221. Young, B., J. S. Lowe, A. Stevens, and J. W. Heath. 2006. Wheater’s Functional Histology: A Text and Colour Atlas. 5th edition. Elsevier Limited, Philadelphia, Pennsylvania. Young, J. Z. 1935. The photoreceptors of Lampreys. I. Light sensitive fibres in the lateral line nerves. The Journal of Experimental Biology 12:229–238. Young, P. S., and J. C. C. Cech. 1996. Environmental tolerances and requirements of Splittail. Transactions of the American Fisheries Society 125:664–678. Young, P. S., C. Swanson, and J. C. C. Cech. 2004. Photophase and illumination effects on the swimming performance and behavior of five California estuarine fishes. Copeia 2004:479–487. Youson, J. H. 1980. Morphology and physiology of Lamprey metamorphosis. Canadian Journal of Fisheries and Aquatic Sciences 37:1687–1710. Youson, J. H. 1981. The alimentary canal, p. 95–189. In The Biology of Lampreys. Vol. 3. M. W. Hardisty and I. C. Potter (eds.). Academic Press, London. Youson, J. H. 1988. First metamorphosis, p. 135–196. In Fish Physiology. Vol. 11B. W. S. Hoar and D. J. Randall (eds.). Academic Press, New York. Youson, J. H. 2003. The biology of metamorphosis in Sea Lampreys: endocrine, environmental, and physiological cues and events, and their potential application to Lamprey control. Journal of Great Lakes Research 29(Suppl. 1):26–49. Youson, J. H. 2004. The impact of environmental and hormonal cues on the evolution of fish metamorphosis, p. 239–278. In Environment, Development, and Evolution. Toward a Synthesis. B. K. Hall, R. D. Pearson, and G. B. Müller (eds.). The Massachusetts Institute of Technology Press, Cambridge. Youson, J. H., A. A. Al-Mahrouki, D. Naumovski, and J. M. Conlon. 2001. The endocrine cells in the gastroenteropancreatic system of the Bowfin, Amia calva L.: an immunohistochemical, ultrastructural, and immunocytochemical analysis. Journal of Morphology 250:208–224 Youson, J. H., and R. J. Beamish. 1991. Comparison of the internal morphology of adults of a population of Lampreys that contains a nonparasitic life-history type, Lampetra richardsoni, and a potentially parasitic form, L. richardsoni var. marifuga. Canadian Journal of Zoology 69:628–637. Youson, J. H., and K. L. Connelly. 1978. Development of longitudinal mucosal folds in the intestine of the anadromous Sea Lamprey, Petromyzon marinus L., during metamorphosis. Canadian Journal of Zoology 56:2364–2371. Youson, J. H., J. A. Holmes, J. A. Guchardi, J. G. Seelye, R. E. Beaver, J. E. Gersmehl, S. A. Sower, and F. W. H. Beamish. 1993. Importance of condition factor and the influence of water temperature and photoperiod on metamorphosis of Sea Lamprey, Petromyzon marinus. Canadian Journal of Fisheries and Aquatic Sciences 50:2448–2456. Youson, J. H., J. Lee, and I. C. Potter. 1979. The distribution of fat in larval, metamorphosing, and young adult anadromous Sea

627

Lampreys, Petromyzon marinus L. Canadian Journal of Zoology 57:237–246. Youson, J. H., and I. C. Potter. 1979. A description of the stages in the metamorphosis of the anadromous Sea Lamprey, Petromyzon marinus L. Canadian Journal of Zoology 57:1808–1817. Youson, J. H., G. M. Wright, and E. C. Ooi. 1977. The timing of changes in several internal organs during metamorphosis of anadromous larval Lamprey, Petromyzon marinus L. Canadian Journal of Zoology 55:469–473. Yu, S-L, and E. J. Peters. 2002. Diel and seasonal habitat use by Red Shiner (Cyprinella lutrensis). Zoological Studies 41:229–235. Yun, S-S, A. P. Scott, and W. Li. 2003. Pheromones of the male Sea Lamprey, Petromyzon marinus L: structural studies on a new compound, 3-keto allocholic acid, and 3-keto petromyzonol sulfate. Steroids 68:297–304. Yun, S-S, A. P. Scott, M. J. Siefkes, and W. M. Li. 2002. Development and application of an ELISA for a sex pheromone released by the male Sea Lamprey (Petromyzon marinus L.). General and Comparative Endocrinology 129:163–170. Yunker, W. K., D. E. Wein, and B. D. Wisenden. 1999. Conditioned alarm behavior in Fathead Minnows (Pimephales promelas) resulting from association of chemical alarm pheromone with a nonbiological visual stimulus. Journal of Chemical Ecology 25:2677–2686. Zahl, P. A., and D. D. Davis. 1932. Effects of gonadectomy on the secondary sexual characteristics in the ganoid fish, Amia calva Linnaeus. Journal of Experimental Zoology 63:291–307. Zanatta, D. T., and D. A. Woolnough. 2011. Confirmation of Obovaria olivaria, Hickorynut Mussel (Bivalvia: Unionidae), in the Mississagi River, Ontario, Canada. Northeastern Naturalist 18:1–6. Zane, L., W. S. Nelson, A. G. Jones, and J. C. Avise. 1999. Microsatellite assessment of multiple paternity in natural populations of a live bearing fish, Gambusia holbrooki. Journal of Evolutionary Biology 12:61–69. Zarnescu, O. 2005. Ultrastructure study of spermatozoa of the Paddlefish, Polyodon spathula. Zygote 13:241–247. Zastrow, C. E., E. D. Houde, and L. G. Morin. 1991. Spawning, fecundity, hatch-date frequency and young-of-the-year growth of Bay Anchovy, Anchoa mitchilli in mid-Chesapeake Bay. Marine Ecology Progress Series 73:161–171. Zawodny, J. F. 1975. Osmoregulation in the Florida Spotted Gar, Lepisosteus platyrhincus. Unpubl. Master’s thesis, University of Miami, Coral Gables, Florida. Zbinden, M., C. R. Largiadèr, and T. C. M. Bakker. 2001. Sperm allocation in the three-spined Stickleback. Journal of Fish Biology 59:1287–1297. Zbinden, M., C. R. Largiadèr, and T. C. M. Bakker. 2004. Body size of virtual rivals affects ejaculate size in Sticklebacks. Behavioral Ecology 15:137–140. Zbinden, M., D. Mazzi, R. Künzler, C. R. Largiadèr, and T. C. M. Bakker. 2003. Courting virtual rivals increase ejaculate size in Sticklebacks (Gasterosteus aculeatus). Behavioral Ecology and Sociobiology 54:205–209. Zeiske E. 1968. Prädispositionen bei Mollienesia sphenops (Pisces, Poeciliidae) für einen Übergang zum Leben in subterranean Gewässern. Zeitschrift für vergleichende Physiologie 58:190–222. Zelditch, M. L., D. L. Swiderski, and W. L. Fink. 2000. Discovery of phylogenetic characters in morphometric data, p. 37–83. In Phylogenetic Analysis of Morphological Data. J. J. Wiens (ed.). Smithsonian Institution Press, Washington.

628 LITERATURE CITED

Zhang, J.-Y. 2006. Phylogeny of the Osteoglossomorpha. Vertebrata PalAsiatica 44:43–59. Zhang, Y., S. Doroshov, T. Famula, F. Conte, D. Kueltz, J. LinaresCasenave, J. Van Eenennaam, P. Struffenegger, K. Beer, and K. Murata. 2011. Egg quality and plasma testosterone (T) and estradiol-17β (E2) in White Sturgeon (Acipenser transmontanus) farmed for caviar. Journal of Applied Ichthyology 27:558–564. Zhuang, P., B. Kynard, L. Zhang, T. Zhang, and W. Cao. 2002. Ontogenetic behavior and migration of Chinese Sturgeon, Acipenser sinensis. Environmental Biology of Fishes 65:83–97. Zhuang, P., B. Kynard, L. Zhang, T. Zhang, and W. Cao. 2003. Comparative ontogenetic behavior and migration of Kaluga, Huso dauricus, and Amur Sturgeon, Acipenser schrenckii, from the Amur River. Environmental Biology of Fishes 66:37–48. Zielinski, B., W. Arbuckle, A. Belanger, L. D. Corkum, W. Li, and A. P. Scott. 2003. Evidence for the release of sex pheromones by male Round Gobies (Neogobius melanstomus). Fish Physiology and Biochemistry 28:237–239. Zigler, S. J., M. R. Dewey, and B. C. Knights. 1999. Diel movement and habitat use by Paddlefish in navigation Pool 8 of the upper Mississippi River. North American Journal of Fisheries Management 19:180–187. Zigler, S. J., M. R. Dewey, B. C. Knights, A. L. Runstrom, and M. T. Steingraeber. 2003. Movement and habitat use by radiotagged Paddlefish in the upper Mississippi River and tributaries. North American Journal of Fisheries Management 23:189–205. Zigler, S. J., M. R. Dewey, B. C. Knights, A. L. Runstrom, and M. T. Steingraeber. 2004. Hydrologic and hydraulic factors affect-

ing passage of Paddlefish through dams in the upper Mississippi River. Transactions of the American Fisheries Society 133:160–172. Zimmer, K. D., M. A. Hanson, M. G. Butler, and W. G. Duff y. 2001a. Influences of Fathead Minnows and aquatic macrophytes on nutrient partitioning and ecosystem structure in two prairie wetlands. Archiv für Hydrobiologie 150:411–433. Zimmer, K. D., M. A. Hanson, M. G. Butler, and W. G. Duff y. 2001b. Size distribution of aquatic invertebrates in two prairie wetlands, with and without fish, with implications for community production. Freshwater Biology 46:1373–1386. Zimmerer, E. J., and K. D. Kallman. 1989. Genetic basis for alternative reproductive tactics in the Pygmy Swordtail, Xiphophorus nigrensis. Evolution 43:1298–1307. Zimmerman, E. G. 1987. Relationships between genetic parameters and life-history characteristics of stream fish, p. 239–244. In Community and Evolutionary Ecology of North American Stream Fishes. W. J. Matthews and D. C. Heins, (eds.). University of Oklahoma Press. Zimmerman, E. G., and M. C. Richmond. 1981. Increased heterozygosity at the MDH-B locus in fish inhabiting a rapidly fluctuating thermal environment. Transactions of the American Fisheries Society 110:410–416. Zu, J. B. 1939. Haff krankheit [Haff disease]. Ergebnisse in der inneren Medizin 57:138–182. Zyznar, E. S., F. B. Cross, and J. A. C. Nicol. 1978. Uric acid in the tapetum lucidum of Mooneyes Hiodon (Hiodontidae Teleostei). Proceedings of the Royal Society of London. Series B, Biological Sciences 201:1–6.

Index of Scientific Names

Page numbers followed by f indicate figures and those followed by p indicate plates. †Acipenser albertensis, 166 Acipenser baerii (Siberian Sturgeon), 163, 185 Acipenser brevirostrum (Shortnose Sturgeon), 161, 162, 163, 167f, 168f, 170, 173, 174, 175, 176, 177, 179, 181, 182, 184, 185, 187, 188, 189, 191, 192, 193, 194, 195, 196, 197, 198, 200, 201, 202, 206, 228 †Acipenser eruciferus, 166 Acipenser fulvescens (Lake Sturgeon), 135, 161, 162, 173, 175, 176, 177, 179, 180, 181, 182, 183, 184, 185, 186, 186f, 187, 188, 189, 191, 192, 194, 195, 195–196, 196, 197, 198, 199, 200, 201, 203, 204, 204–205, 205, 205f, 206, 228, 240, 254, 259 Acipenser medirostris (Green Sturgeon), 160–161, 162, 173, 175, 176, 177, 178, 179, 180, 181, 182, 183, 185, 188, 189, 191, 192, 194, 197, 199, 200, 203, 205 Acipenser mikadoi (Sakhalin Sturgeon), 161 †Acipenser ornatus, 166 Acipenser oxyrinchus desotoi (Gulf Sturgeon), 161, 162, 174, 176, 179, 180, 181, 183, 184, 185, 187, 189, 191–192, 192, 194, 194–195, 195, 196, 197, 200 Acipenser oxyrinchus oxyrinchus (Atlantic Sturgeon), 161, 173, 174, 174–175, 176, 177, 179, 179–180, 180, 181, 182, 183, 184, 185, 187–188, 188, 189, 191, 192, 192f, 193, 194, 196, 197, 199, 200, 204, 204–205, 205, 206, 240 Acipenser ruthenus (Sterlet), 163 Acipenser stellatus (Stellate Sturgeon), 165, 178 Acipenser sturio (European Sturgeon), 163

Acipenser transmontanus (White Sturgeon), 160, 160p, 162, 163, 170, 173, 175, 176, 177, 178, 179, 180, 181, 184, 185, 188, 191, 192, 193, 194, 195, 197, 198, 199, 199f, 200, 200f, 201, 203, 204, 205, 205f, 206 Acrocheilus alutaceus (Chiselmouth), 387p, 443 †Acrocheilus latus, 404 Agosia chrysogaster (Longfin Dace), 396p, 410f, 423, 432, 433–434 Alburnops baileyi (Rough Shiner), 436 Alburnops bairdi (Red River Shiner), 445 Alburnops blennius (River Shiner), 379, 431 Alburnops buccula (Smalleye Shiner), 426 Alburnops chalybaeus (Ironcolor Shiner), 379 Alburnops chlorocephalus (Greenhead Shiner), 415f Alburnops chrosomus (Rainbow Shiner), 379, 436 Alburnops lutipinnis (Yellowfin Shiner), 415f, 427, 436 Alburnops potteri (Chub Shiner), 445 Alburnops texanus (Weed Shiner), 378p, 379, 430, 432, 437 Algansea monticola (Mountain Chub), 409 Algansea popoche (Popoche Chub), 406 Alosa kessleri (Caspian Anadromous Shad), 340 Alosa pseudoharengus (Alewife), 34, 134, 239, 340 Alosa sapidissima (American Shad), 134, 200, 342 Amazonsprattus scintilla (Rio Negro Pygmy Anchovy), 337 Ambloplites rupestris (Rock Bass), 19, 20, 82, 51, 494

Ameiurus melas (Black Bullhead), 428 Ameiurus nebulosus (Brown Bullhead), 19, 20, 82 Amia calva (Bowfin), 87, 246, 279–297 passim, 280f, 281f, 281p, 285f, 286, 290, 290f, 291f, 292f, 293f, 294f, 297f, 436 †“Amia” hesperia, 283 †Amia scutata, 283f †Amyzon aggregatum, 463 †Amyzon brevipinne, 463 †Amyzon commune, 463 †Amyzon fusiforme, 463 †Amyzon gosiutensis, 463 †Amyzon huanensis, 463 †Amyzon mentale, 463 †Amyzon pendatum, 463 Anchoa analis (Longfin Pacific Anchovy), 333 Anchoa belizensis (Belize Anchovy), 333 Anchoa delicatissima (Slough Anchovy), 336–337 Anchoa hepsetus (Striped Anchovy), 333 Anchoa lamprotaenia (Bigeye Anchovy), 344 Anchoa lyolepis (Dusky Anchovy), 344 Anchoa mitchilli (Bay Anchovy), 158, 332–352 passim, 333p, 347f Anchoa mundeoloides (Northern Gulf Anchovy), 335 †Anchoa nitida, 337 Anchoa parva (Little Anchovy), 333, 336 Anguilla anguilla (European Eel), 313, 314–315, 317, 319, 320, 324 †Anguilla annosa, 318 Anguilla japonica (Japanese Eel), 313, 326 †Anguilla rectangularis, 318 Anguilla rostrata (American Eel), 313–330 passim, 313p, 316f, 318f, 319f, 320f, 325f, 326f, 327f, 494

630

INDEX OF SCIENTIFIC NAMES

†Anguilla rouxi, 318 Apeltes quadracus (Fourspine Stickleback), 56 Aphredoderus sayanus (Pirate Perch), 102 Aplodinotus grunniens (Freshwater Drum), 20, 272, 273, 412 Arapaima gigas (Pirarucú), 299 Ariopsis felis (Hardhead Catfish), 272 †Asiacipenser kotelnikovi, 167 Astyanax fasciatus (Banded Astyanax), 273 Astyanax mexicanus (Mexican Tetra), 52, 58 †Atractosteus atrox, 248 †Atractosteus falipoui, 248 †Atractosteus messelensis, 248 †Atractosteus simplex, 248, 249f Atractosteus spatula (Alligator Gar), 243, 245, 245f, 246, 246p, 247, 250, 250f, 252, 253, 254, 255, 257, 258, 259, 260, 261, 262, 263f, 264, 264–265, 266, 267, 268, 270, 271, 272, 272–273, 273, 273–274, 274, 275, 277, 277–278, 278 Atractosteus tristoechus (Cuban Gar), 245, 245f, 247, 258, 270 Atractosteus tropicus (Tropical Gar), 245, 245f, 247, 250, 254, 258, 264, 266, 267, 268, 270, 272, 273, 273–274, 275, 277, 277–278, 278 Aztecula calientis (Yellow Shiner), 401 Brevoortia patronus (Menhaden), 272 Campostoma anomalum (Central Stoneroller), 19, 21, 32, 37, 41–42, 42, 46, 47, 380f, 406f, 407f, 408f, 418, 430, 435, 438, 439, 440, 487 Campostoma oligolepis (Largescale Stoneroller), 37, 87f, 414f, 438, 442 Campostoma ornatum (Mexican Stoneroller), 393p, 409 Carassius auratus (Goldfish), 64, 65f, 288, 354, 383, 432, 448, 457, 493 Carpiodes carpio (River Carpsucker), 20, 260, 485, 487, 488, 489, 492, 493 Carpiodes cyprinus (Quillback), 453p, 459, 485, 487, 488, 492 Carpiodes velifer (Highfin Carpsucker), 468, 484–485, 485, 487, 488, 492 Caspiomyzon wagneri (Caspian Lamprey), 114, 115 Catostomus ardens (Utah Sucker), 470–471 Catostomus catostomus (Longnose Sucker), 59, 452p, 454, 457f, 471, 472, 478, 479, 493, 496, 497 Catostomus clarkii (Desert Sucker), 471, 490, 493 Catostomus columbianus (Bridgelip Sucker), 32, 98, 487, 489, 496 Catostomus commersonii (White Sucker), 45, 121f, 134, 434, 451, 454, 457, 459,

468, 471, 472, 473, 476, 476–477, 478, 482, 484, 485, 486f, 494, 486, 488, 489, 492, 495, 496, 500 Catostomus discobolus (Bluehead Sucker), 8, 460, 471, 485 Catostomus discobolus yarrowi (Zuni Bluehead Sucker), 471 Catostomus insignis (Sonora Sucker), 471, 490, 493, 497 Catostomus latipinnis (Flannelmouth Sucker), 8, 485, 488, 494 Catostomus macrocheilus (Largescale Sucker), 98, 451, 468, 471, 472, 474–475, 476, 478, 488, 489, 496 Catostomus microps (Moduc Sucker), 493, 497 Catostomus occidentalis (Sacramento Sucker), 43, 47, 476, 488 Catostomus platyrhyncus (Mountain Sucker), 460, 484, 485, 488, 492–493 Catostomus plebeius (Rio Grande Sucker), 460, 471, 493 Catostomus rimiculus (Klamath Smallscale Sucker), 464, 471, 472 Catostomus santaanae (Santa Ana Sucker), 493 Catostomus snyderi (Klamath Largescale Sucker), 471, 472, 473, 474 Catostomus tahoensis (Tahoe Sucker), 25, 470, 475–476, 488, 496 Catostomus tsiltcoosensis (Tyee Sucker), 451 Catostomus utawana (Summer Sucker), 451 Catostomus warnerensis (Warner Sucker), 480, 495, 497 Catostomus wigginsi (Opata Sucker), 493 Cetengraulis mysticetus (Anchoveta), 335 Cetorhinus maximus (Basking Shark), 135 Chasmistes brevirostris (Shortnose Sucker), 471, 472, 473, 474, 488, 489, 492, 495, 497 Chasmistes cujus (Cui-ui), 470, 475, 483, 485, 487, 489, 492, 497 Chasmistes liorus (June Sucker), 453p, 470–471, 492, 493, 497 Chasmistes muriei (Snake River Sucker), 497 Chrosomus cumberlandensis (Blackside Dace), 98, 380f, 434, 436 Chrosomus eos (Redbelly Dace), 20, 99, 388p, 419–420, 441 Chrosomus erythrogaster (Southern Redbelly Dace), 13, 88f, 417f, 418, 426, 426–427 Chrosomus neogaeus (Finescale Dace), 99, 103, 405, 419–420, 425, 441, 449 Chrosomus tennesseensis (Tennessee Dace), 98 Clinostomus funduloides (Redside Dace), 37, 39–40, 47, 380f, 384p, 427, 439 Clupea pallasii (Pacific Herring), 123f, 135

Codoma ornata (Ornate Shiner), 62, 398, 405 Coilia brachygnathus (Yangtse Grenadier Anchovy), 340 Coilia nasus (Japanese Grenadier Anchovy), 337 †Coilia planate, 337 Coregonus alpenae (Longjaw Cisco), 101 Coregonus artedii (Lake Herring), 34 Coregonus clupeaformis (Lake Whitefish), 13 Coregonus hoyi (Bloater), 34 Coregonus nasus (Broad Whitefish), 59 Cottus asper (Prickly Sculpin), 196 Cottus bairdii (Mottled Sculpin), 14, 19, 40, 52–53 Cottus beldingii (Paiute Sculpin), 25 Cottus carolinae (Banded Sculpin), 9, 13, 18, 24–25, 39 Cottus hangiongensis (Kankyo-kajika), 53 Cottus perplexus (Reticulate Sculpin), 72 Cottus ricei (Spoonhead Sculpin), 59 Couesius plumbeus (Lake Chub), 59, 390p, 479 †Coupatezia woutersi, 143 †Crossopholis magnicaudatus, 211, 212, 212f Ctenopharyngodon idella (Grass Carp), 384, 448, 493 Culaea inconstans (Brook Stickleback), 41, 56, 70, 430 †Cuneatus cuneatus, 249, 249f †Cuneatus wileyi, 249 Cycleptus elongatus (Blue Sucker), 455p, 465, 466, 472, 487, 490f, 492 Cycleptus meridionalis (Southeastern Blue Sucker), 482, 492 Cynoscion nebulosus (Spotted Seatrout), 352 Cynoscion regalis (Weakfish), 351 Cynoscion urenarius (Sand Seatrout), 352 Cyprinella analostanus (Satinfin Shiner), 61, 432 Cyprinella caerulea (Blue Shiner), 439, 447 Cyprinella callisema (Ocmulgee Shiner), 61–62 Cyprinella callistia (Alabama Shiner), 378p, 413f, 442 Cyprinella camura (Bluntface Shiner), 431 Cyprinella galactura (Whitetail Shiner), 61–62, 398, 413f, 427, 432 Cyprinella garmani (Gibbous Shiner), 411f Cyprinella gibbsi (Tallapoosa Shiner), 61–62 Cyprinella labrosa (Thicklip Chub), 398, 400, 401 Cyprinella leedsi (Bannerfin Shiner), 432 Cyprinella lepida (Edwards Plateau Shiner), 61–62 Cyprinella lutrensis (Red Shiner), 23, 25, 38, 40, 41, 48, 62, 259, 272, 379, 383,

INDEX OF SCIENTIFIC NAMES

384, 418, 422, 425, 426, 429, 432, 437, 447, 447–448 Cyprinella proserpina (Proserpine Shiner), 446 Cyprinella spiloptera (Spotfin Shiner), 62, 379, 422–423, 432, 433 Cyprinella trichroistia (Tricolor Shiner), 61–62, 61f, 445–446 Cyprinella venusta (Blacktail Shiner), 20, 46, 62, 78, 379, 418, 431, 437, 442 Cyprinella whipplei (Steelcolor Shiner), 379 Cyprinella zanema (Santee Chub), 398, 400, 401 Cyprinodon bifasciatus (Cuatro Cienegas Pupfish), 60, 60f Cyprinodon diabolis (Devils Hole Pupfish), 14 Cyprinodon nevadensis calidae (Tecopa Pupfish), 101 Cyprinodon pecosensis (Pecos Pupfish), 55–56, 82, 88 Cyprinodon variegatus (Sheepshead Minnow), 60 Cyprinus carpio (Common Carp), 122f, 203, 259, 271, 354, 383, 493 Danio rerio (Zebrafish), 354, 417, 432 Dasyatis centroura (Roughtail Stingray), 140, 145 Dasyatis sabina (Atlantic Stingray), 140–159 passim, 141p, 144f, 147f, 151f, 152f, 155f Dasyatis say (Bluntnose Ray), 148 Deltistes luxatus (Lost River Sucker), 451, 455f, 471, 472, 488, 492, 495, 497, 499–500 Denticeps cupeoides (Denticle Herring), 335 †Dentilepisosteus kemkemensis, 249 †Dentilepisosteus laevis, 249 Dimidiochromis compressiceps (Malawi Eyebiter), 431 Dionda episcopa (Roundnose Minnow), 396p Dorosoma cepedianum (Gizzard Shad), 20, 273, 493 Dorosoma petenense (Threadfin Shad), 34, 36, 239, 273 Elassoma evergladei (Everglades Pygmy Sunfish), 56 Elops saurus (Ladyfish), 150 †Engraulis brevipinnis, 337 †Engraulis evolans, 337 Engraulis japonicus (Japanese Anchovy), 340 †Engraulis longipinnis, 337 †Engraulis macrocephalus, 337 Engraulis mordax (Northern Anchovy), 335, 340, 340f †Engraulis tethensis, 337 †Engraulites remifer, 337

Entosphenus folletti (Northern California Brook Lamprey), 115 Entosphenus hubbsi (Kern Brook Lamprey), 111, 115, 137 Entosphenus lethophagus (Pit-Klamath Brook Lamprey), 111, 115 Entosphenus macrostomus (Lake Lamprey), 111, 135, 137 Entosphenus minimus (Miller Lake Lamprey), 111, 119, 137 Entosphenus similis (Klamath Lamprey), 111, 115 Entosphenus tridentatus (Pacific Lamprey), 111, 112, 113p, 114, 115, 118, 123, 123f, 129f, 130–131, 131, 133, 135, 136, 137, 138 Eptatretus burgeri (Inshore Hagfish), 124 Eremichthys acros (Desert Dace), 388p, 406, 448 Ericymba amplamala (Longjaw Minnow), 37, 412 Ericymba buccata (Silverjaw Minnow), 37, 400, 412, 412f Erimonax monachus (Spotfin Chub), 380f, 398, 399p, 427, 446–447 Erimystax cahni (Slender Chub), 446 Erimystax dissimilis (Streamline Chub), 406, 411f, 421 Erimystax insignis (Blotched Chub), 410f Erimystax x-punctatus (Gravel Chub), 393p, 424 Erimyzon oblongus (Creek Chubsucker), 435, 454p, 457f, 482, 483, 483–484, 487, 494 Erimyzon sucetta (Lake Chubsucker), 87, 435, 470, 487, 497 Esox lucius (Northern Pike), 20, 41, 59, 78, 135, 289, 430, 431, 447, 494, 500 Esox masquinongy (Muskellunge), 135, 136, 494, 500 Etheostoma basilare (Corrugated Darter), 83 Etheostoma blennioides (Greenside Darter), 18, 20, 494 Etheostoma caeruleum (Rainbow Darter), 13, 18, 19, 55, 55f Etheostoma chienense (Relict Darter), 55 Etheostoma crossopterum (Fringed Darter), 60–61, 60f Etheostoma exile (Iowa Darter), 103, 430 Etheostoma flabellare (Fantail Darter), 18, 19, 55, 81, 83f Etheostoma neopterum (Lollipop Darter), 83 Etheostoma nigripinne (Blackfin Darter), 60–61 Etheostoma nigrum (Johnny Darter), 83f, 98 Etheostoma olmstedi (Tessellated Darter), 81

631

Etheostoma oophylax (Guardian Darter), 83, 83f, 84 Etheostoma podostemone (Riverweed Darter), 19 Etheostoma pseudovulatum (Egg-mimic Darter), 83 Etheostoma punctulatum (Stippled Darter), 98 Etheostoma radiosum (Orangebelly Darter), 14 Etheostoma rufilineatum (Redline Darter), 20, 494 Etheostoma simoterum (Snubnose Darter), 20 Etheostoma spectabile (Orangethroat Darter), 18, 18–19, 24–25, 32, 39, 55, 97–98 Etheostoma squamiceps (Spottail Darter), 55, 82, 83 Etheostoma tetrazonum (Saddled Darter), 18, 19 Etheostoma virgatum (Striped Darter), 55, 82, 83, 83f Etheostoma vulneratum (Wounded Darter), 20 Etheostoma zonale (Banded Darter), 18 Eucinostomus harengulus (Tidewater Mojarra), 150 Eudontomyzon danfordi (Carpathian Lamprey), 115 Eudontomyzon graecus (Epirus Brook Lamprey), 105, 115 Eudontomyzon hellenicus (Macedonia Brook Lamprey), 115 Eudontomyzon mariae (Ukrainian Brook Lamprey), 115 Eudontomyzon stankokaramani (Drin Brook Lamprey), 105, 115 Exoglossum laurae (Tonguetied Minnow), 406 Exoglossum maxillinguae (Cutlips Minnow), 406, 406f, 431 Fundulus catenatus (Northern Studfish), 9, 426–427 Fundulus chrysotus (Golden Topminnow), 60 Fundulus diaphanus (Banded Killifish), 20 Fundulus grandis (Gulf Killifish), 60 Fundulus kansae (Northern Plains Killifish), 60 Fundulus notatus (Blackstripe Topminnow), 32 Fundulus olivaceus (Blackspotted Topminnow), 78 Fundulus pulvereus (Bayou Killifish), 60 Fundulus zebrinus (Plains Killifish), 60 Gambusia affinis (Western Mosquitofish), 41, 48, 69, 78, 94, 96, 271, 272

632

INDEX OF SCIENTIFIC NAMES

Gambusia amistadensis (Amistad Gambusia), 101 Gambusia holbrooki (Eastern Mosquitofish), 94, 94–96 Gambusia hurtadoi (Crescent Gambusia), 56 Gasterosteus aculeatus (Threespine Stickleback), 38–39, 43, 51, 56, 58, 70, 71, 81, 82, 95, 96, 103 Gasterosteus wheatlandi (Blackspotted Stickleback), 56 Geotria australis (Pouched Lamprey), 107, 112, 114, 121, 123–124, 127, 128, 131 Giardichthys multiradiatus (Darkedged Splitfin), 57–58, 57f Gila atraria (Utah Chub), 44, 380, 414, 423 Gila conspersa (Nazas Chub), 380 Gila coriacea (Moapa Dace), 387, 448 Gila crassicauda (Thicktail Chub), 445 Gila cypha (Humpback Chub), 8, 378p, 380, 415, 416, 421, 444 Gila elegans (Bonytail Chub), 8, 380, 380f, 405, 416, 420, 421, 444, 494 Gila orcutti (Arroyo Chub), 418 Gila purpurea (Yaqui Chub), 410f Gila robusta (Roundtail Chub), 8, 415, 420, 447 Gila seminuda (Virgin River Chub), 420, 421 Gobio gobio (Gudgeon), 46 Gobiosoma bosc (Naked Goby), 150 Graodus boucardi (Balsas Shiner), 401 Gymnocephalus cernuus (Ruffe), 412 Gymnotus cylindricus, 273 †Haikouichythys ercaicunensis, 118 †Hardistiella montanensis, 118 †Heliobatis radians, 143 Hemitremia flammea (Flame Chub), 391p Herichthys cyanoguttatum (Rio Grande Cichlid), 97 Hesperoleucas symmetricus (California Roach), 388p, 421 Hiodon alosoides (Goldeye), 299–311 passim, 301p, 303f, 307f †Hiodon consteniorum, 302 †Hiodon falcatus, 302, 302f, 307f †Hiodon lirellus, 302 †Hiodon rosei, 302 Hiodon tergisus (Mooneye), 299–311 passim, 301p, 302f, 307f †Hiodon woodruffi, 302 Himantura signifer (White-edged Freshwater Whip Ray), 149 Hudsonius altipinnis (Highfin Shiner), 401 Hudsonius cummingsae (Dusky Shiner), 401, 435, 436 Hudsonius hudsonius (Spottail Shiner), 399p, 401, 442, 443, 449 Hybognathus amarus (Rio Grande Silvery Minnow), 40, 399p, 433, 443

Hybognathus argyritis (Western Silvery Minnow), 445 Hybognathus hankinsoni (Brassy Minnow), 439 Hybognathus nuchalis (Mississippi Silvery Minnow), 432 Hybognathus placitus (Plains Minnow), 40, 425, 426, 427, 432, 433, 443, 445 Hybognathus regius (Eastern Silvery Minnow), 437, 449 Hybopsis amblops (Bigeye Chub), 398, 399p, 401, 423–424 Hybopsis hypsinotus (Highback Chub), 401 Hybopsis lineapunctatus (Lined Chub), 401 Hybopsis rubrifrons (Rosyface Chub), 398, 401 Hybopsis winchelli (Clear Chub), 401 Hypentelium etowanum (Alabama Hog Sucker), 488 Hypentelium nigricans (Northern Hog Sucker), 45, 46, 454p, 465f, 472, 482, 483, 490, 491, 491f, 492, 496 Hypentelium roanokense (Roanoke Hog Sucker), 456, 464 Hypopthalmichthys molitrix (Silver Carp), 239, 383, 493 Hypopthalmichthys nobilis (Bighead Carp), 239, 383, 493 Ichthyomyzon bdellium (Ohio Lamprey), 106f, 110, 135, 496 Ichthyomyzon castaneus (Chestnut Lamprey), 110, 113p, 116, 122f, 129f, 135, 136, 296, 496 Ichthyomyzon fossor (Northern Brook Lamprey), 111, 132, 133 Ichthyomyzon gagei (Southern Brook Lamprey), 111, 113f, 116, 131, 133 Ichthyomyzon greeleyi (Mountain Brook Lamprey), 111, 125 Ichthyomyzon unicuspis (Silver Lamprey), 110, 120f, 131, 132, 135, 136, 197, 235, 296 Ictalurus furcatus (Blue Catfish), 3, 21 Ictalurus punctatus (Channel Catfish), 3, 21, 196, 234, 260, 273, 500 †Ictiobus aguilerai, 463 Ictiobus bubalus (Smallmouth Buffalo), 453p, 454, 459, 463, 464, 471, 483, 487, 488, 489, 493, 496 Ictiobus cyprinellus (Bigmouth Buffalo), 459, 463, 471, 483, 487, 488, 489, 492, 495, 500 Ictiobus niger (Black Buffalo), 459, 463, 466f, 471, 482, 483, 488, 489, 494 Iotichthys phlegethontis (Least Chub), 414 Labidesthes sicculus (Brook Silverside), 271 Lagodon rhomboides (Pinfish), 352

Lampetra aepyptera (Least Brook Lamprey), 112, 114f, 115, 116, 132, 133 Lampetra ayresii (American River Lamprey), 111, 112, 114, 115, 118–119, 120f, 121, 130, 135, 136, 138 Lampetra fluviatilis (River Lamprey), 114, 115, 116, 116f, 124, 125, 132, 136 Lampetra lanceolata (Turkish Brook Lamprey), 115 Lampetra pacifica (Pacific Brook Lamprey), 112, 115 Lampetra planeri (European Brook Lamprey), 107, 114, 115, 116, 116f, 125, 128 Lampetra richardsoni (Western Brook Lamprey), 113p, 115, 117, 131, 136, 137 Lampetra zanandreai (Po Brook Lamprey), 115 Lavinia exilicauda (Hitch), 421 Lepidomeda albivallis (White River Spinedace), 390, 390p Lepidomeda aliciae (Southern Leatherside Chub) 390 Lepidomeda copei (Leatherside Chub), 387, 390 Lepidomeda mollispinis (Virgin Spinedace), 390, 430 Lepidomeda vittata (Little Colorado Spinedace), 390, 423, 447 †Lepisosteus bemisi, 248, 249f †Lepisosteus indicus, 248 Lepisosteus oculatus (Spotted Gar), 243, 245f, 246, 246p, 247, 249, 250, 251, 253, 254, 255, 256, 257, 258, 258–259, 259, 260, 262, 263, 264, 265, 266, 266–267, 267, 268, 270, 270f, 271, 272, 273, 274, 275, 277, 278 Lepisosteus osseus (Longnose Gar), 84, 243, 245f, 247, 249, 250, 251, 252, 253, 253–254, 254, 255, 256, 257, 258, 259, 261, 262, 263f, 264, 265, 266, 267, 267f, 268, 270, 270f, 271, 272, 273, 274, 275, 277, 278, 287, 289 Lepisosteus platostomus (Shortnose Gar), 243, 245, 245f, 247, 250, 253, 254, 255, 257, 262, 263–264, 264, 270, 271, 272, 274, 277 Lepisosteus platyrhincus (Florida Gar), 243, 245f, 246, 247, 249, 250, 251, 253, 254, 255, 258, 259, 260, 263, 264, 267, 270, 271, 272, 273, 274, 275, 277, 278 Lepomis auritus (Redbreast Sunfish), 62, 84, 89, 436 Lepomis cyanellus (Green Sunfish), 39, 47, 71–72, 260, 295, 435, 436 Lepomis gibbosus (Pumpkinseed), 19, 21, 39, 62, 82, 89, 90, 103, 288, 295, 428 Lepomis humilis (Orangespotted Sunfish), 62, 63

INDEX OF SCIENTIFIC NAMES

Lepomis macrochirus (Bluegill), 19, 20, 21, 21–22, 23, 36, 39, 42, 43, 45, 62, 78, 89, 89–90, 92, 260, 272, 289, 295 Lepomis marginatus (Dollar Sunfish), 82 Lepomis megalotis (Longear Sunfish), 32, 47, 63–64, 90–91, 98 Lepomis microlophus (Redear Sunfish), 98 Lepomis peltastes (Northern Longear Sunfish), 90 Lepomis punctatus (Spotted Sunfish), 90 Lethenteron alaskense (Alaskan Brook Lamprey), 111, 115 Lethenteron appendix (American Brook Lamprey), 107, 113p, 114, 115, 117, 133, 136 Lethenteron camtschaticum (Arctic Lamprey), 107, 111, 114, 115, 117, 120f, 121, 132, 136 Lethenteron kessleri (Siberian Brook Lamprey), 115, 117 Lethenteron ninae (Western Transcaucasian Brook Lamprey), 105, 115 Lethenteron reissneri (Far Eastern Brook Lamprey), 115, 117 Leuciscus leuciscus (Dace), 46 Limia perugiae (Perugia’s Limia), 53, 93 Lophius americanus (Goosefish), 200 Lota lota (Burbot), 59, 494, 496 Lucania goodei (Bluefin Killifish), 54–55, 54f, 60 Lutjanus griseus (Gray Snapper), 150 Luxilus albeolus (White Shiner), 434, 435–436 Luxilus cardinalis (Cardinal Shiner), 413f Luxilus cerasinus (Crescent Shiner), 401, 434, 436 Luxilus chrysocephalus (Striped Shiner), 42, 122f, 382, 382p, 400, 413f, 418, 418–419, 419, 434, 439 Luxilus coccogenis (Warpaint Shiner), 19, 37, 87f, 401, 434 Luxilus cornutus (Common Shiner), 45, 87, 382, 406, 411, 418, 418–419, 419, 427, 434 Luxilus pilsbryi (Duskystripe Shiner), 20, 434 Luxilus zonatus (Bleeding Shiner), 380f, 434, 438 Luxilus zonistius (Bandfin Shiner), 434 Lythrurus ardens (Rosefin Shiner), 430 Lythrurus fasciolaris (Scarlet Shiner), 381f, 416f Lythrurus fumeus (Ribbon Shiner), 410f Lythrurus roseipinnis (Cherryfin Shiner), 440 Lythrurus snelsoni (Ouachita Mountain Shiner), 381 Lythrurus umbratilis (Redfin Shiner), 47, 381, 435, 436 Macrhybopsis aestivalis (Speckled Chub), 383, 392, 410f, 412f, 433 Macrhybopsis australis (Prairie Chub), 445

Macrhybopsis gelida (Sturgeon Chub), 392, 410, 445 Macrhybopsis hyostoma (Shoal Chub), 411 Macrhybopsis meeki (Sicklefin Chub), 392, 410, 410f Macrhybopsis storeriana (Silver Chub), 381, 410f, 431, 432 Macrhybopsis tetranema (Peppered Chub), 392, 410f, 411, 425, 443 Margariscus margarita (Pearl Dace), 62, 391p, 441, 449 †Masillosteus janeae, 249, 249f †Mayomyzon pieckoensis, 117–118, 118f Meda fulgida (Spikedace), 38, 391p, 423, 433, 440, 447 Melanogrammus aeglefinus (Haddock), 327 Membras martinica (Rough Silverside), 150 Menidia audens (Mississippi Silverside), 20, 271 Menidia clarkhubbsi (Texas Silverside), 101 Merluccius bilinearis (Silver Hake), 134 †Mesomyzon mengae, 118, 119f Microgobius gulosus (Clown Goby), 150 Micropogonias undulatus (Atlantic Croaker), 150, 158 Micropterus dolomieu (Smallmouth Bass), 20, 41, 43, 51, 82, 84, 87, 89, 196, 265, 439, 447 Micropterus floridanus (Florida Bass), 87, 487 Micropterus punctulatus (Spotted Bass), 42, 78 Micropterus salmoides (Largemouth Bass), 19, 34, 36, 42, 43–44, 45, 87, 89, 96–97, 122f, 234, 289, 295, 435, 440, 447, 448, 487, 493 Miniellus heterodon (Blackchin Shiner), 402 Miniellus procne (Swallowtail Shiner), 402 Miniellus stramineus (Sand Shiner), 402, 449 Miniellus topeka (Topeka Shiner), 400p, 402, 431 Minytrema melanops (Spotted Sucker), 451, 455p, 463, 465–466, 470, 488, 492 Mordacia mordax (Short-headed Lamprey), 115, 116, 124, 127, 128 Mordacia praecox (Precocious Lamprey), 115, 116, 124 Morone americana (White Perch), 239 Morone chrysops (White Bass), 20 Morone saxatilis (Striped Bass), 20, 34, 327, 351 Moxostoma anisurum (Silver Redhorse), 452p, 463, 465f, 481f, 488–489 Moxostoma austrinum (Mexican Redhorse), 456 Moxostoma carinatum (River Redhorse), 468, 469, 487, 490, 500 Moxostoma cervinum (Blacktip Redhorse), 492

633

Moxostoma congestum (Gray Redhorse), 497 Moxostoma duquesnei (Black Redhorse), 460, 484, 487 Moxostoma erythrurum (Golden Redhorse), 465f, 467, 486f, 487, 488–489 Moxostoma hubbsi (Copper Redhorse), 468, 469, 472, 490, 492, 497 Moxostoma lacerum (Harelip Sucker), 451, 465, 465f, 468, 469, 469f, 490, 497 Moxostoma lachneri (Greater Jumprock), 470 Moxostoma macrolepidotum (Shorthead Redhorse), 465f, 470, 484, 485, 487, 488–489, 491f Moxostoma mascotae (Mascota Jumprock), 456 Moxostoma pisolabrum (Pealip Redhorse), 482 Moxostoma robustum (Robust Redhorse), 466f, 473, 474, 476, 490 Moxostoma valenciennesi (Greater Redhorse), 487, 490, 494 Mugil cephalus (Striped Mullet), 183 Mugil curema (White Mullet), 150 Mylocheilus caurinus (Peamouth Chub), 395p, 408, 409, 426, 430, 431, 449 †Mylocheilus inflexus, 404 †Mylocheilus robustus, 404 Mylopharyngodon conocephalus (Hardhead), 388p Mylopharnyngodon piceus (Black Carp), 383–384, 493 †Myxineidus gonorum, 106 Myxocyprinus asiaticus (Chinese Sucker), 451, 454, 456p, 463, 464f, 470, 497 Neogobius melanostomus (Round Goby), 203, 239 Nocomis asper (Redspot Chub), 394p Nocomis biguttatus (Hornyhead Chub), 37, 43, 46, 88f, 412, 435, 438, 439, 449 Nocomis leptocephalus (Bluehead Chub), 47, 380f, 409f, 410, 415f, 418, 423, 424, 430, 435, 436, 439 Nocomis micropogon (River Chub), 19, 87f, 412, 420, 494 Nocomis raneyi (Bull Chub), 409f Notemigonus crysoleucas (Golden Shiner), 20, 82, 87, 295, 296, 383, 384, 385, 385p, 403, 409, 411, 428, 429, 431, 432, 435, 436, 441, 449 Nothonotus rubrum (Bayou Darter), 18, 21 “Notropis” alborus (Whitemouth Shiner), 401 “Notropis” ammophilus (Orangefin Shiner), 432 Notropis atherinoides (Emerald Shiner), 378p, 379, 406, 418, 427 “Notropis” bifrenatus (Bridle Shiner), 401, 405, 446

634 INDEX OF SCIENTIFIC NAMES

“Notropis” boops (Bigeye Shiner), 32, 37, 46, 438 Notropis buchanani (Ghost Shiner), 401 Notropis chorocephalus (Greenhead Shiner), 47 Notropis cummingsae (Dusky Shiner), 84, 435 “Notropis” dorsalis (Bigmouth Shiner), 397 Notropis girardi (Arkansas River Shiner), 425, 427, 432, 433, 443–444 Notropis harperi (Redeye Chub), 89 Notropis heterodon (Blackchin Shiner), 19 “Notropis” heterolepis (Blacknose Shiner), 19, 427, 430 Notropis jemezanus (Rio Grande Shiner), 433, 443 Notropis leuciodus (Tennessee Shiner), 87f, 416f “Notropis” longirostris (Longnose Shiner), 37, 431 Notropis lutipinnis (Yellowfin Shiner), 47 “Notropis” maculatus (Taillight Shiner), 431, 432 “Notropis” micropteryx (Highland Shiner), 424 “Notropis” nazas (Nazas Shiner), 397 “Notropis” nubilus (Ozark Minnow), 37, 46, 438 “Notropis” oxyrhynchus (Sharpnose Shiner), 426 Notropis percobromus (Carmine Shiner), 424 “Notropis” photogenis (Silver Shiner) 400 “Notropis” rafinesquei (Yazoo Shiner), 437 Notropis rubellus (Rosyface Shiner), 418, 424, 428 Notropis rubricroceus (Saff ron Shiner), 87f “Notropis” scepticus (Sandbar Shiner), 397 “Notropis” simus (Bluntnose Shiner), 433, 443 “Notropis” telescopus (Telescope Shiner), 37, 400 Notropis topeka (Topeka Shiner), 47, 435 “Notropis” tropicus (Pygmy Shiner), 379 Notropis volucellus (Mimic Shiner), 20, 45, 418, 430 Notropis wickliffi (Channel Shiner), 431 Noturus exilis (Slender Madtom), 18 Noturus funebris (Black Madtom), 20 Noturus gyrinus (Tadpole Madtom), 20 Noturus hildebrandi (Least Madtom), 20 Noturus leptacanthus (Speckled Madtom), 20 Noturus miurus (Brindled Madtom), 272 Noturus phaeus (Brown Madtom), 20 †Obaichthys africanus, 249 †Obaichthys decoratus, 249 Oncorhynchus clarkii (Cutthroat Trout), 19, 20, 21, 34, 40, 44 Oncorhynchus gorbuscha (Pink Salmon), 135

Oncorhynchus keta (Chum Salmon), 91, 123 Oncorhynchus kisutch (Coho Salmon), 34, 39, 41, 71, 91, 92, 93, 135, 138, 478 Oncorhynchus masou (Masu Salmon), 64, 92 Oncorhynchus mykiss (Rainbow Trout), 19, 21, 32, 34, 39–40, 42, 43, 45, 47, 64, 71, 91–92, 103, 135, 430, 447 Oncorhynchus nerka (Sockeye Salmon), 21, 91, 92, 135 Oncorhynchus tshawytscha (Chinook Salmon), 34, 41, 45, 138 Opisthonema oglinum (Atlantic Thread Herring), 150 Opsopoeodus emiliae (Pugnose Minnow), 402, 405, 406, 408f, 409, 435 Oregonichthys crameri (Oregon Chub), 394p, 445 Oregonichthys kalawatseti (Umpqua Chub), 447 Oreochromis aureus (Blue Tilapia), 100, 448 †Orthodon hadrognathus, 404 Orthodon microlepidotus (Sacramento Blackfish), 389p, 406–407, 407f, 441, 449 Osmerus mordax (Rainbow Smelt), 34 †Paleoosephurus wilsoni, 211, 212 Pantodon buchholzi (African Butterfly Fish), 303, 307 Parachromis managuense (Jaguar Guapote), 273 Paralichthys dentatus (Summer Flounder), 351 Paralichthys lethostigma (Southern Flounder), 150 Pastinachus sephen (Cowtail Stingray), 140 Perca flavescens (Yellow Perch), 20, 21, 41, 272, 430, 494 Percina aurantiaca (Tangerine Darter), 20 Percina burtoni (Blotchside Logperch), 20, 46 Percina caprodes (Logperch), 20, 45, 46, 97–98, 272 Percina evides (Gilt Darter), 9, 20, 46 Percina roanoka (Roanoke Logperch), 19 Percina williamsi (Sickle Darter), 20 Percopsis omiscomaycus (Trout-perch), 59, 412 Petromyzon marinus (Sea Lamprey), 34, 107, 110, 111p, 112, 114, 116, 118, 120f, 121, 121f, 124, 125, 127, 128–130, 128f, 131, 132, 133, 134, 134–135, 135, 135–136, 136, 138, 254, 203 Phenacobius mirabilis (Suckermouth Minnow), 406f, 407f Phenacobius teretulus (Kanawha Minnow), 405 Phenacobius uranops (Stargazing Minnow), 393p, 406

Phoxinus phoxinus (Eurasian Minnow), 46, 72, 403 Pimephales notatus (Bluntnose Minnow), 45, 61f, 62, 400p, 427, 430, 432, 433, 439 Pimephales promelas (Fathead Minnow), 41, 71, 72, 78, 81, 89, 103, 202, 272, 295, 383, 384, 408, 408f, 424, 425, 427, 428, 429, 430, 430– 432, 433, 440– 441, 442, 443, 449, 497 Pimephales vigilax (Bullhead Minnow), 78 †Pipscius zangerli, 118 Plagopterus argentissimus (Woundfin), 411f, 447, 448 Platygobio gracilis (Flathead Chub), 393p, 411f, 425, 432, 444, 445 †Plesiolycoptera daqingensis, 302 Poecilia formosa (Amazon Molly), 99–100, 99f Poecilia gilli (Costa Rican Molly), 273 Poecilia latipinna (Sailfin Molly), 51–52, 51f, 53, 65, 93, 94, 96, 99–100 Poecilia latipunctata (Broadspotted Molly, Tarnesi Molly), 65, 100 Poecilia mexicana (Shortfin Molly), 52, 64–65, 79–80, 79f, 99–100, 448 Poecilia reticulata (Guppy), 53 Poecilia velifera (Yucatan Molly), 64–65 Poeciliopsis lucida (Clearfin Livebearer), 100–101 Poeciliopsis monacha (Headwater Livebearer), 100–101 Poeciliopsis monacha-lucida, 100–101 Poeciliopsis occidentalis (Gila Topminnow), 41 Pogonichthys macrolepidotus (Splittail), 396p, 426, 429, 430, 431, 444, 448 Polyodon spathula (North American Paddlefish), 135, 136, 207–242 passim, 207p, 208f, 210f, 213f, 214f, 215f, 216f, 217f, 218f, 236f, 259, 383, 493 †Polyodon tuberculata, 211, 211–212, 212f Pomatomus saltatrix (Bluefish), 327, 351 Pomatoschistus microps (Common Goby), 81, 92 Pomoxis annularis (White Crappy), 20 Pomoxis nigromaculatus (Black Crappy), 19 Priapella olmecae (Olmec Priapella), 51 †Priscomyzon riniensis, 118 †Priscosturion longipinnis, 165, 166f, 167 †Protopsephurus liui, 211, 211f, 212f, 212–213 †Protoscaphirhynchus squamosus, 165, 166f Psephurus gladius (Chinese Paddlefish), 207, 207–208, 210, 212, 214 Pteronotropis euryzonus (Broadstripe Shiner), 402 Pteronotropis grandipinnis (Apalachee Shiner), 382p Pteronotropis harperi (Redeye Chub), 382, 402

INDEX OF SCIENTIFIC NAMES

Pteronotropis hubbsi (Bluehead Shiner), 88–89, 380f, 382, 402, 413, 435, 448–449 Pteronotropis hypselopterus (Sailfin Shiner), 402 Pteronotropis merlini (Orangetail Shiner), 402 Pteronotropis signipinnis (Flagfin Shiner), 402 Pteronotropis welaka (Bluenose Shiner), 89, 402, 413, 435 Pteroplatytrygon violacea (Pelagic Stingray), 140, 142 †Ptychocheilus arciferus, 404 Ptychocheilus grandis (Sacramento Pikeminnow), 43, 383, 430, 439, 441 Ptychocheilus lucius (Colorado Pikeminnow), 8, 41, 384, 387, 389p, 405, 416, 421, 439, 440, 445, 446f Ptychocheilus oregonensis (Northern Pikeminnow), 41, 196, 425, 431, 432, 433, 443, 447 Ptychocheilus umpquae (Umpqua Pikeminnow), 447 Pungitius pungitius (Ninespine Stickleback), 56, 59 Puntius titteya (Cherry Barb), 354 Pylodictis olivaris (Flathead Catfish), 21, 493 Rasbora heteromorpha (Harlequin Rasbora), 354 Relictus solitarius (Relict Dace), 389p Rhamdia guatemalensis (South American Catfish), 273 Rhinichthys atratulus (Blacknose Dace), 12–13, 21, 43, 411f, 423, 425, 433, 434, 438 Rhinichthys bowersi (Cheat Minnow), 420–421 Rhinichthys cataractae (Longnose Dace), 19, 44, 45, 394p, 406f, 420, 422, 443 Rhinichthys osculus (Speckled Dace), 8, 32, 383, 414, 423, 430 Rhodeus sericeus (Bitterling), 384 Richardsonius balteatus (Redside Shiner), 395p Richardsonius egregius (Lahontan Redside), 427 Roeboides guatemalensis (Guatemalan Headstander), 273 Salmo salar (Atlantic Salmon), 21, 71, 90, 91, 92, 93, 103, 121f, 134, 288 Salmo trutta (Brown Trout), 20, 25, 32, 34, 39

Salvelinus alpinus (Arctic Charr), 34, 71, 96 Salvelinus confluentus (Bull Trout), 19, 20, 21 Salvelinus fontinalis (Brook Trout), 34, 39, 103, 288, 494 Salvelinus malma (Dolly Varden), 40 Salvelinus malma miyabei (Miyabe Charr), 92 Salvelinus namaycush (Lake Trout), 34, 135, 138 Sander canadensis (Sauger), 234 Sander vitreus (Walleye), 3, 234, 311 Sarda sarda (Atlantic Bonito), 279 Sardinella aurita (Spanish Sardine), 344 Scaphirhynchus albus (Pallid Sturgeon), 63, 161, 170, 174, 175, 177, 178, 179, 180, 181, 182, 183, 185, 186, 188, 189, 191, 192, 195, 196, 197, 198, 200, 201, 202, 203 Scaphirhynchus platorynchus (Shovelnose Sturgeon), 63, 161, 161p, 162, 173, 174, 175, 176, 177, 178, 179, 180, 181, 182, 183, 184, 185, 186, 187, 188, 189, 191, 192, 193, 195, 196, 197, 199, 200, 202, 203, 205–206, 210, 288 Scaphirhynchus suttkusi (Alabama Sturgeon), 161, 162, 170, 175, 193, 196, 197, 198, 201–202, 204 Scardinius erythropthalamus (Rudd), 384 Scartomyzon ariommus (Bigeye Jumprock), 460, 465f, 467 Scartomyzon cervinus (Blacktip Jumprock), 460 Scomber scombrus (Atlantic Mackerel), 134 Semotilus atromaculatus (Creek Chub), 13, 19, 42, 43, 45, 46–47, 98, 391p, 406f, 407f, 408, 408f, 410f, 418, 422, 429, 432, 435, 438, 442, 449, 487 Semotilus corporalis (Fallfish), 433, 434, 437, 494 †Setipinna retusa, 337 Siphateles bicolor (Tui Chub), 406, 427, 473 Siphateles bicolor mohavensis (Mohave Tui Chub), 389p, 418, 426 Siphateles boraxobius (Borax Chub), 448 †Stolephorus furculus, 337 †Stolephorus lemoinei, 337 †Stolephorus productus, 337 Strongylura marina (Atlantic Needlefish), 352 Stypodon signifer (Stumptooth Minnow), 408 Syngnathus scovelli (Gulf Pipfish), 150 Synodus foetens (Inshore Lizardfish), 352

635

Tampichthys rasconis (Blackstripe Minnow), 405 Tetrapleurodon geminis (Mexican Brook Lamprey), 111, 115, 132 Tetrapleurodon spadiceus (Mexican Lamprey), 111, 114, 115, 122, 132, 136, 137 Thoburnia atripinnis (Blackfin Sucker), 462, 466 Thoburnia hamiltoni (Rustyside Sucker), 465f Thoburnia rhothoeca (Torrent Sucker), 454p, 488 Thunnus thynnus (Bluefin Tuna), 135 Tiaroga cobitis (Loach Minnow), 423 Tinca tinca (Tench), 384, 493 Trachemys scripta (Red-eared Slider), 260–261 Trachinotus falcatus (Permit), 352 Umbra limi (Central Mudminnow), 41 Xenotoca variata (Jeweled Splitfin), 58–59 Xiphias gladius (Swordfish), 134–135, 327–328 Xiphophorus birchmanni (Sheepshead Swordtail), 52 Xiphophorus continens (Short-sword Platyfish), 70 Xiphophorus cortezi (Delicate Swordtail), 57, 70 Xiphophorus helleri (Green Swordtail), 51, 51f, 52, 64–65, 430 Xiphophorus maculatus (Southern Platyfish), 51, 96 Xiphophorus malinche (Highland Swordtail), 58 Xiphophorus montezumae (Montezuma Swordtail), 70 Xiphophorus multilineatus (Barred Swordtail), 57, 58f, 93 Xiphophorus nezahualcoytl (Mountain Swordtail), 57, 58f Xiphophorus nigrensis (Panuco Swordtail), 51, 57, 57–58, 70, 93 Xiphophorus pygmaeus (Pygmy Swordtail), 70 Xiphophorus variatus (Variable Platyfish), 51, 93 †Xiphytrygon acutidens, 143 Xyrauchen texanus (Razorback Sucker), 8, 451, 456p, 463, 474, 479, 484, 485–486, 487, 490, 492, 493–494, 494, 495, 497 †Yabbiania wangqingica, 302

General Index

Page numbers followed by f indicate figures and those followed by t indicate tables. acidification, of water bodies, 103, 427 acousticolateralis system. See lateral-line system Acuña, S., 429 adjustment stability, 34; and elasticity, 34 agonistic behavior, 97; aggressive intent expressed in ritualized displays, 97; and resource holding potential, 97 agriculture, effects of: on Carps and Minnows, 445–447f; on Sturgeons, 202 alarm substance (Schreckstoff ) system, 71–72, 78–79, 430–431, 478; and breeding, 72, 78; in Carps and Minnows, 424–425, 430, 431; in nonostariophysan fishes, 72; in North American freshwater fishes, 73–78t; in Ostariophysi, 71–72; reduction of activity in presence of, 72; response of predators in presence of, 78; in Suckers, 478, 480 Allan, J. R., 46 Allen, J. D., 15–16 Alligator Gar Technical Committee of the Southern Division of the American Fisheries Society, 277 alloparental care, 81–82; effect of parental effort on clutch survival, 82; foster fathers, 82; in North American freshwater fishes, 81t. See also alloparental care, and female preference for nests with many eggs alloparental care, and female preference for nests with many eggs: dilution effect hypothesis, 81–82; elevated courtship effect hypothesis, 82; increased parental care effect hypothesis, 82; mate choice hypothesis, 82 alternative mating strategies: female, 96–97; in livebearers, 93–96; male,

87–89, 88f, 484; reproductive success of, 89–90; in salmonids, 91–93; and selection vectors, 94; in Sunfishes, 89–91 Anchovies (Engraulidae), 332–333; Bay Anchovy morphology, 337–340; commercial importance of, 352; conservation of, 352; diel activity in, 343–344; diet of, 349, 349f; dissolved oxygen tolerance of, 341; diversity and distribution of, 332–334, 333t, 334f; ecological limitations on, 352; egg densities of, 346; embryonic development in, 346–347, 347f; energy budget for Bay Anchovy larvae, 342, 342t; fecundity of, 346; feeding behaviors of, 344–345; fossil record of, 337; genetics of, 340; growth and longevity in, 349–350; habitat of, 348–349; as hearing specialists, 342; intraspecific variability, subspecies, and clines in, 334–335; larval and juvenile development in, 347, 347f, 348t; life history traits of Bay Anchovy, 334t; mating system of, 348; meaning of family name Engraulidae, 332; migration of, 342–343, 343f; morphology of, 332, 337, 340f; mortality rates of, 350; as non-natives, 335; as opportunistic life history strategists, 348; parasites of, 352; phylogenetic relationships in, 335–337, 336f, 338–339f; population ecology of, 350–351, 351f; as predators, 351; predators of, 351–352; reproductive allocation in, 346; salinity tolerance of, 341; schooling behavior of, 344; seasonality of reproduction in, 345, 345f; sexual maturity in, 345; spawning in, 346; spawning cues, 345; spawning sites, 345–346; temperature tolerance of,

341; visual structures and photoreception in, 341–342, 341f Anderson, K. A., 59 Angermeier, P. L., 14, 15, 23–24 anguillid Eels. See Freshwater Eels Ankley, G. T., 428 Applegate, V., 138 aquaculture: of Paddlefishes, 240, 241–242; of Sacramento Blackfish, 449; of Sturgeons, 163, 206; of Suckers, 499, 500 Atlantic States Marine Fisheries Commission (ASMFC), 328 Bailey, R. M., 385, 392, 401, 402, 403 Ballard, W. W., 268 Baltz, D. M., 47 Barbour, C. D., 405 Barton, M., 426–427 Bassista, T. P., 346 Becker, G. C., 271 Behnke, R. J., 414 Bemis, W. E., 165, 282, 283, 284–285, 286 benthic cruising, 168, 170 Berendzen, P. B., 402, 472 Bessert, M. L., 472 Bichirs (Polypteriformes), 163–164, 210–211; scale jacket of, 253 Bielawski, J. P., 402 Bluegills: alternative mating strategies in, 89–90; and Largemouth Bass predation, 42; pharyngeal sound production in, 62; and Pumpkinseeds, 39 Bonneville Basin, 11 Boucher, D. H., 44 Bowfins (Amiidae), 279, 280f, 281f; acidity tolerance of, 287–288; age and growth in, 295–296; behavior of, 289;

GENERAL INDEX

commercial importance of, 296–297; conservation of, 296; diet of, 294–295; digestion in, 289; diversity and distribution of, 279, 281f; early radiation of, into fresh water, 4; eggs, embryos, and larval development of, 293, 293f, 294f; fecundity of, 292–293; and fisheries, 296; fossil record of, 283–284, 283f, 284f, 285f; gametes of, 293; genome size and variation in, 286; habitat of, 294; inter- and intraspecific variation in, 279–280; intraspecific genetic variation and phylogeography, 286; karyology of, 286; lack of a dorsal nucleus in, 289; lateral-line system in, 288–289; life history characteristics of, 291t; male parental care in, 290; meaning of name Amia calva, 279; morphology of, 284–286, 285f; mortality rates of, 295; movement and dispersal in, 289–290; as non-natives, 281; olfaction and chemosensation in, 289; opportunistic nature of feeding of, 295; origin of common name of, 279; origin of genus name of, 279; oxygen requirements, respiration, and air breathing in, 286–287; parasites of, 296; parental care in, 279, 290, 292; phylogenetic relationships in, 281–283, 282f, 283f; physiology of, 286; as a pollution-tolerant species, 294; population sizes and densities of, 296; predators of, 296; salinity tolerance of, 287; schooling in, 290, 290f; seasonality of reproduction in, 291; sexual dimorphism in, 290, 291f; sexual maturity in, 290; and sister relationship with Gars, 255; spawning migrations of, 291; spawning mode, behavior, and habitat of, 291–292, 292f; stress responses of, 288; survival of, in extreme environments, 288; thermal capacity and preferred temperature of, 287; use of, in fisheries management, 297, 297f; vernacular names for, 279; vision, photoreceptors, and visual pigments in, 288 Branson, B. A., 466, 467 Breder, C. M., 50 Breitburg, D. L., 344 Brett, J. R., 21 Brier Creek, Oklahoma: effects of flooding on spawning fish in, 32; fish fauna persistence in, 27, 32; large bass and Central Stonerollers in, 42; species-specific responses to drought in, 32; study of fish assemblages in, 35; survival of larval centarchids and cyprinids in, 43–44 Briggs, T., 460 Brooks, D. R., 36 Broughton, R. E., 398, 422 Bruner, J. C., 463

Bufalino, A. P., 386 Burkhead, N. M., 464–465 Burleson, M. L., 256 Burr, B. M., 392, 497 Bussjaeger, C., 460 Buth, D. G., 401, 460 cannibalism, 101; in Alligator Gars, 275; in Anchovies, 344; testing of hypotheses in Poeciliopsis clones, 101 Carps and Minnows (Cyprinidae), 354–355; acclimation in, 428; acidification tolerance of, 427; and agriculture and development on, 445–447; alarm substance system in, 424–425, 430, 431; antipredator behavior of, 430–431; barbels in, 410–411, 410f, 411f; buccal cavity and pharynx morphology of, 406–409, 407f, 408f; clonal lineages and hybrid species in, 419–421; commercial importance of, 449; conservation of, 441; conservation status of, 443; and dams and flow modification, 422, 443–445, 446f; diel activity in, 429–430; digestion in, 425; egg characteristics and clutches of, 437; evolution of spawning modes in, 436–437; eye-picking behavior in, 431; fecundity of, 432–433; fossil record of, 403–405, 404f; generic classification of North American species, 355–356t; genetic variation in, 421–422; genetics of, 417; genome size and base substitution in, 417; growth in, 424; gut morphology of, 409, 409f; hearing in, 411; hybridization in, 417, 417–419; jaw morphology of, 405–406, 406f; karyotype of, 417; larval behavior in, 431; lateral-line system in, 411–412, 412f; learning in, 431; mitochondrial genome of, 422–423; morphology of, 405; morphology in taxonomy and ecomorphology, 414–416; as nest associates, 435–436; and nonnative species, 447–448; as non-natives, 383–384, 418; nuptial structures in, 412–414, 412f, 414f, 415f, 416f, 417f; olfaction and taste in, 410, 410f; oral grasping in, 431; origin of name Cyprinidae, 354; oxygen tolerance of, 427–428; parasites of, 441–443, 442f; phylogenetic relationships in, 384–387, 386f, 390, 390f, 392, 392f, 395–398, 397f, 398f, 400–403; phylogeographic studies of, 423–424; role of, in nutrient cycling, 440–441; salinity tolerance of, 426–427; seasonality of reproduction in, 432; sexual maturity and sexual dimorphism in, 437–438; spatial ecology of, 438–439; spawning cues in, 432; spawning mode and male-male competition in, 435; spawning modes in, 433; substrate

637

preparation for spawning in, 433–435; territorial and courtship sounds in minnows, 61–62, 61f; thermal tolerances of, 426; toxicological tolerances of, 428–429; trophic ecology of, 439–440; turbidity tolerance of, 427; vulnerability of small populations of, 448–449; Weberian apparatus in, 354, 411, 411f, 425 Carps and Minnows (Cyprinidae), diversity and distribution of genera: Alburnops, 357f, 379; Cyprinella, 357f, 379; Gila, 357f, 380, 380f; Luxilus, 357f, 382–383; Lythrurus, 357f, 380–381; Macrhybopsis, 357f, 381–382; Notropis, 355, 357f, 379; Pteronotropis, 357f, 380f, 382 Carps and Minnows (Cyprinidae), geographic range of: Acrocheilus, 358f; Agosia, 358f; Algansea, 358f; Aztecula, 358f; Campostoma, 358f; Chrosomus, 358f; Clinostomus, 358f; Codoma, 358f; Couesius, 359f; Dionda, 359f; Eremichthys, 359f; Ericymba, 359f; Erimonax, 359f; Erimystax, 359f; Exoglossum, 359f; Graodus, 359f; Hemitremia, 360f; Hesperoleucas, 360f; Hudsonius, 360f; Hybognathus, 360f; Hybopsis, 360f; Iotichthys, 360f; Klamathella, 360f; Lavinia, 360f; Lepidomeda, 361f; Margariscus, 361f; Meda, 361f; Miniellus, 361f; Mylocheilus, 361f; Mylopharodon, 361f; Nocomis, 361f; Notemigonus, 361f; Opsopoeodus, 362f; Oregonichthys, 362f; Orthodon, 362f; Phenacobius, 362f; Pimephales, 362f; Plagopterus, 362f; Platygobio, 362f; Pogonichthys, 362f; Ptychocheilus, 363f; Relictus, 359f; Richardsonius, 363f; Rhinichthys, 363f; Semotilus, 363f; Siphatales, 363f; Tampichthys, 359f; Tiaroga, 363f; Yuriria, 363f Carps and Minnows (Cyprinidae), life history data for type species of: Acrocheilus, 376t; Agosia, 370–371t; Alburnops, 370–371t; Aztecula, 370–371t; Campostoma, 366–367t; Chrosomus, 376t; Clinostomus, 366–367t; Codoma, 370–371t; Couesius, 364–365t; Cyprinella, 370–371t; Dionda, 370–371t; Eremichthys, 376t; Ericymba, 370–371t; Erimonax, 372–373t; Erimystax, 366– 367t; Exoglossum, 366–367t; Gila, 376t; Graodus, 372–373t; Hemitremia, 364–365t; Hesperoleucas, 376t; Hudsonius, 372–373t; Hybognathus, 372–373t; Hybopsis, 372–373t; Iotichthys, 366– 367t; Klamathella, 376t; Lavinia, 377t; Lepidomeda, 364–365t; Luxilus, 372–373t; Lythrurus, 372–373t; Macrhybopsis, 366–367t; Margariscus, 364–365t; Meda, 364–365t; Miniellus, 372–373t;

638

GENERAL INDEX

Carps and Minnows (Cyprinidae), life history data for type species of (cont.) Mylocheilus, 366–367t; Mylopharodon, 377t; Nocomis, 366–367t; Notemigonus, 364–365t; Notropis, 374–375t; Opsopoeodus, 374–375t; Oregonichthys, 368–369t; Phenacobius, 368–369t; Pimephales, 374–375t; Plagopterus, 364–365t; Platygobio, 368–369t; Pogonichthys, 368–369t; Pteronotropis, 374–375t; Ptychocheilus, 377t; Orthodon, 377t; Relictus, 377t; Rhinichthys, 368–369t; Richardsonius, 368–369t; Semotilus, 364–365t; Siphatales, 377t; Tampichthys, 374–375t; Tiaroga, 368–369t; Yuriria, 374–375t Cashner, M. F., 400 Cavender, T. M., 386, 400, 401, 401–402, 403, 405 Central Highlands, 8–10, 9f, 10–11; high diversity of, and vicariance hypothesis, 9; ichthyofauna of, 9–10 Chang, M.-M., 463 character displacement, 38 character release, 38 Chen, X.-Y., 403 Clark, H. W., 420 Clements, M. D., 460 Coburn, M. M., 386, 400, 401, 401–402, 403, 405, 411 co-evolution, 47–48; diffuse, 47; geographic mosaic models of, 47; and habitat availability, 48; influence of biology of individual species on, 48; lack of empirical evidence on, 48; tightly coupled, 47 Cohen, D. M., 1 Collette, B. B., 467 Colorado River system, 7–8, 7f; artificial floods in, 444; effects of dams in, on cyprinid fauna, 444; flooding in, 416; origins of fish assemblage in, 7–8; and uplift of Colorado Plateau, 7–8 competition, 27, 36, 36– 40, 38f; between North American Sturgeon species, 194; effects on fish assemblages, 40; effects on specific habitats, 39– 40; loss of species due to, 40; over resources, 38–39 Convention on International Trade of Endangered Species (CITES), 198, 235 Cook, A. G., 471 Cooke, S. J., 497 Cooperman, M. S., 499 Cope, E. D., 403, 404, 461 Cross, F. B., 271 Crossman, E. J., 311 Cueva Luna Azufre, Mexico, 80 Cueva del Azufre, Mexico, troglobitic population of Shortfin Mollies in, 79–80, 79f; female mate choice in, 80; loss of male alternative mating strategies in,

80; simplification of male-male aggressive behavior in, 79–80 Daly, R. J., 333, 344 dams, 422; and on Carps and Minnows, 443– 445, 446f; and cyprinid species, 443– 445; and Freshwater Eels, 329; and genetic variation between populations, 422; and Lampreys, 137; and Paddlefishes, 207, 223, 226, 237–238; and riverine habitat, 443, 445; storage of annual river runoff worldwide, 2; and Sturgeons, 160, 196, 201–202; and turbidity, 444– 445 darters: color in, 55, 55f; egg mimics, 83–84; hybridization in, 55; territorial and courtship sounds in, 60–61, 60f Dawkins, M. S., 79 Dawson, H. A., 133 Dean, B., 292 DeMarais, B. D., 420 de Perera, T. B., 57 Diamond, J. N., 24 Diamond, S. A., 428 Di Dario, F., 335 Dimmick, W. W., 392, 398, 402 Docker, M. F., 117 Doosey, M., 462 Douglas, M. E., 38, 415 Dowling, T. E., 401, 422 Drevnick, P. E., 429 drought, 27, 32 Durham, B. W., 425 Eastman, J. T., 468 Edwards, L. F., 467 egg mimics, 83–84, 83f; benefits of, 83; evolution of egg spot, 83–84 egg stealing and alloparental care, 81–82; and female preference for nests with many eggs, 81–82; in North American freshwater fishes, 81t Ehrlich, P. R., 47 Eisenhour, D. J., 533 Electric Rays (Torpediniformes), 147 Elephantfishes (Mormyridae), 299 environment, 14 Evans-White, M., 440 Evermann, B. W., 499 experiments. See studies/experiments facilitation, 27, 36, 44–45; between fishes and other taxa, 44–45; in fish assemblages, 46–48; mixed-species associations and potential for, 45–46. See also nest associations Fast, A. W., 36 Felley, J. D., 415–416 Ferrara, A. M., 264, 273 Ferris, S. D., 460, 470

Findeis, E. K., 165 Fink, W., 469 fish assemblages, 1, 47; and association of species’ ancestors, 3; colonization potential of fish species, 25; conceptual model of formation of, 3–4, 4f; definition of, 1; and Diamond’s assembly rules, 24; effect of habitat size on, 23–24, 23f; faunal ages of North American fish families, 4–6, 6f; faunal origins of North American fish families, 4, 5f; formation of, 24–25; interactions within (see competition; facilitation; mutualism; predation); local and regional faunal effects on, 16–17, 17f; long-term studies of North American, 28–31, 33t; physico-chemical responses of, 27, 32, 34–35; stability and persistence of, in space and time, 25–27; Tertiary and Quaternary events and, 6–13. See also fish assemblages, local and regional environmental effects on fish assemblages, local and regional environmental effects on, 14–16; habitat template model, 14, 15; landscape filters model, 14, 15–16; river continuum model, 14, 16 fish diversity, 1–2; geography of, in North America, 2f; in southeastern region of United States, 2 fisheries, effects of: on Bowfins, 296; on Paddlefishes, 235–237, 236f, 240–241, 241f; on Sturgeons, 199–201, 199f, 200f fitness: and fertilization success, 90, 92; lifetime fitness, 90 Flecker, A. S., 14–15 floods, 27, 32, 444; artificial, 444 Foin, T. C., 25 Forbes, S. A., 466 foundation species, 44 Fowler, H. W., 466 fresh water, 1, 150; lentic habitats, 2–3; lotic habitats, 2–3; unavailability of, as fish habitat, 1; volume of worldwide in lakes, 2; volume of worldwide in streams, 2 Freshwater Eels (Anguillidae), 313–314; behavior of glass eels and elvers, 322; behavior of leptocephalus larvae, 322; commercial importance of, 330; conservation of, 328–330, 329f; conservation status of, 328–329; and dams, 329; diet of larvae, 326–327; diet of post-larvae, 327; diversity and distribution of, 314–317, 316f; as ecological generalists, 326; eggs, embryos, and larvae of, 325, 326f; elvers, 314, 321, 323, 330; emigrating strategies of, 324; as facultative catadromes, 316; fecundity of, 325; fossil record of, 318; glass Eels, 314, 318–319, 321, 323, 330; habitat of, 325–326;

GENERAL INDEX

hybridization in, 320; as indicators of habitat integrity, 330; intraspecific genetic variation in, 320; karyology of, 320; leptocephalus larvae of, 320–321; life history and migration cycle of, 313– 314, 314f; life history characteristics of, 315t; morphology of, 318–319, 318f, 319f; movement of, homing in, and home range of, 322–323; as non-natives, 317; olfaction in, 321–322; origin of name of, 313; parasites of, 328; phylogenetic relationships in, 317–318, 317f; post-larval respiration in, 321; predators of, 327–328, 327f; rotational feeding in, 327; salinity tolerance of, 321; sexual differentiation in, 323–324; sexual dimorphism in, 319; silver-phase Eels, 314, 319, 323, 324, 330; spawning area of, 324–325; spawning migrations of, 313; spawning migrations to the sea of, 324; spawning mode of, 325, 325f; swimming in, 323; thermal tolerance of, 321; transition to silver and yellow phases in, 319–320, 320f; vision in, 321; yellow-phase Eels, 314, 319, 321, 323, 330 Frimpong, E. A., 14, 15 Frisch, K. von, 478 Fuller, P. L., 281, 459 Futey, L. M., 411 Garant, D., 93 Gardiner, B. G., 165 Gars (Lepisosteidae), 243; adult locomotion, 261; age and growth in, 274; age at maturation, 264; agonistic behavior and feeding territoriality of, 263–264; airbreathing process in, 243, 256–257; alimentary canal and organs of taste and smell in, 251; annual mortality rates of, 275; behavioral regulation of air breathing in, 257–258; behavior of post-larvae, 261; behavior of pre-juveniles, 261; behavior of prolarvae, 261; blood physiology of, in absence of a choroid rete, 254–255; commercial importance of, 278; conservation of, 277–278; conservation status of, 277; demographic variation in, 273; development of, 268–269; diel feeding periodicity in, 273; diet of adults, 272–273; diet of larvae and juveniles, 271–272; digestion in, 258; dissolved oxygen tolerance of, 260; diversity and distribution of, 243, 244f, 245, 245f; early life stages of, 269–270, 269f, 270f; fecundity of, 267; feeding behavior of, 262–263, 263f; fossil record of, 243, 247–249, 248t, 249f; gametes of, 267–268; gar toxin, 260–261; gills and gill surface area in, 252–253; growth rates in larvae and juveniles, 273–274;

habitat of, 270–271; habitat partitioning in, 271; home range and movements of, 262; hybridization in, 254; intraspecific genetic variation in, 254; karyology of, 253–254; lack of dorsal nucleus in, 289; length-weight relationships in, 275; life history characteristics of, 244t; as living fossils with derived characters, 243, 250, 250f; lung in, 252, 255; mechanosensory canals and pit lines in, 251–253, 251f; metabolic organization in, 254; morphological differences between Atractosteus and Lepisosteus, 250–251; morphology of, 249; morphomechanics of jaw in, 252; as non-natives, 245–246; osmoregulation and exchange of carbon dioxide and ammonia in, 258–259; parasites of, 275, 276t, 277; perception of, as undesirable, 243; pH tolerance of, 259; phylogenetic relationships in, 246–247, 247f; pigmentation of, 250–251; pollution tolerance of, 259–260; population sizes and densities of, 275; predators of, 275; prey manipulation of, 263; rate of air breathing in, 257; respiratory control in, 255–256; salinity tolerance of, 259; scale jacket of, 253; seasonality of reproduction in, 264; sexual dimorphism in, 264; significance of bimodal respiration in, 256; sister relationship with Bowfins, 255; size of, 250; spawning behavior of, 266–267, 267f; spawning habitat and cues in, 265–266, 265f; spawning movements of, 264–265; thermoregulation in, 260; urogenital system of, 251; visual system in, 255 Gilbert, C. R., 401 Gill, H. S., 115 glacial refugia, 11–13, 12f, 13f; Atlantic Refugium, 12–13; Beringia Refugium, 13; Cascadia = Pacific Refugium, 13; Mississippi Refugium, 11–12, 13; Missouri Refugium, 12, 13 glaciation, 10–13; limits of glacial advance, 10; number of glacial advances, 10; and reduced species richness, 13; in western states, 10; Wisconsinan, 10. See also glacial refugia Glenn, C. L., 306 global climate change, effects of: on Paddlefishes, 240; on Sturgeons, 203 Goddard, K. A., 99 Gold, J. R., 398, 402 Goldsborough, E. L., 420 Goldstein, R. M., 15 gonad somatic index (GSI), 92, 346 Goodfellow, W. L., Jr., 420 Gorman, O. T., 24, 37, 45 Gould, S. J., 79

639

Grande, L., 248, 282, 283, 284–285, 286, 337 Great Basin, 8 Great Lakes: effects of landlocked Sea Lamprey on fish populations in, 135, 138; fish fauna in, 2–3; use of lampricides in, 137. See also Lake Erie; Lake Huron; Lake Michigan Green River, Kentucky, 424 Green River Fish Control Project, 495 Gregory, W. K., 467–468 Grose, M. J., 401 Gross, M. R., 92–93 Grossman, G. D., 47 Guilford, T., 79 Gunter, G., 150 gynogens, 99–101 Haase, B. L., 266 habitat, 2–3; availability and species co-occurrence, 48; diel shifts in use of, 19–21; degradation and hybridization, 102–103; distinction between environment and, 14; distribution theory, 438; effect of size on fish assemblages, 23–24, 23f; effect of type and quality, 18–23; influence of life history stage on use of, 21; influence of water temperature on selection of, 22–23; influences on selection of, 21; lentic, 2–3 (see also lakes); lotic, 2–3 (see also rivers; streams); meanings of term, 14 habitat template model, 14, 15; flow predictability as a component of, 15; support for, 15; testing predictions of, in Rhône River drainage, 15 Hagfishes (Myxiniformes), 105–107; acrosomal process in sperm of, 189; adult features shared with Lampreys, 106; genome duplications in, 124; hemoglobins of, 126; as iono- and osmoconformers, 105–106, 126; lack of a dorsal nucleus in, 289; marine environment of, 105; rearrangements of genomes of, 124; relationship with Lampreys, 106–107; sodium and chloride concentrations in, 106 Haines, T. A., 429 Hamman, R. L., 421 Harrington, R. W. J., 405 Harris, P. M., 459, 460–461, 461–462, 462 Hartman, K. J., 346 Heins, D. C., 437 heteroplasmy, 173 Hildebrand, S. F., 334–335 Hildrew, A. G., 15 Hilton, E. J., 165, 301 H+ -K+ -ATPases, 148 Hoff man, G. L., 496 Holčík, J., 105

640 GENERAL INDEX

Hollingsworth, P. R., Jr., 392 Houde, E. D., 342 Howes, G., 403 Hox, Sox, Pax, and Dlx families of genes, 124 Hubbs, C. L., 98, 243, 385, 387, 401, 418, 460, 466, 467 Hugueny, B., 14 Hulsey, C. D., 392 Humphries, J., 469 hybridization, 97, 101; and anthropogenic interference, 103; in Carps and Minnows, 417, 417–421; and clonal lineages, 419; in darters, 55; as essential part of species diversification in some lineages, 101; in Freshwater Eels, 320; in Gars, 254; and gynogens, 99–101; and habitat degradation, 102–103; and introgression, 101, 102; natural events of, 97–98; and nest associations, 98, 436; among North American cyprinid species, 417–419; outcomes of, 99; and plesiomorphic behavior, 98; Poecilia latipinna x P. mexicana = P. formosa (Amazon Molly), 99–100, 99f; Poeciliopsis monacha x P. lucida = P. monacha-lucida, 100–101; results of, 99; in Sturgeons, 162, 174; in Suckers, 470–472 hydrologic variability, 32 hypoxia, 126, 427; and increase in Bohr effect, 126; oxygen tolerance of Carps and Minnows, 427–428; and reduction in mean cellular hemoglobin concentration, 126; Sturgeons’ intolerance of, 177, 202; success of Gars in hypoxic conditions, 256 Illick, H. J., 412 Indian River Lagoon, Florida, 151, 153–154, 157, 158 Inebnit, T. E., III, 261 Infante, D. M., 15–16 Interior Highlands region, 17 International Network for Lepisosteid Fish Research, 277 introgression, 101, 102; in Sturgeons, 174; in Suckers, 470–472 invasive species, 27; and colonization potential, 25; invasion sequence, 25; invasion success, 25, 26f; and match between invader and hydrologic regime, 25; and Paddlefishes, 239; and Sturgeons, 203 Jelks, H. L., 136 Jenkins, R. E., 451, 460, 464–465 Jennings, M. J., 90, 98 Johnson, B. L., 266 Johnson, L., 34 Johnson, M. R., 158

Johnston, C. E., 482 Jordan, D. S., 311, 499 Kennedy, W. A., 311 Kettratad, J., 451 killifishes: color in, 54–55, 54f; territorial and courtship sounds in, 60 Kitchell, J. F., 3 Klamath Basin, Oregon and California: columnaris among Suckers in, 495; synergistic effects of threats on Suckers in, 497, 499 Klamath Lake, Oregon, dramatic die-offs of Chasmistes brevirostris and Deltistes luxatus in, 495 Koch, J. D., 295 Kolok, A. S., 428 Kott, E., 114 Kuraku, S., 107 Kuratani, S., 107 Lagler, K. F., 243 Lake Bonneville, 423 Lake Erie: during Pleistocene, 419; entrance of Sea Lamprey into, 138 Lake Huron, introduction of non-native species in, 34 Lake Idaho, fossil fauna of, 404 Lake Michigan: heavy fishing pressure in, 34; introduction of non-native species in, 34; during Pleistocene, 419 Lake Pinchi, British Columbia, 447 lakes: diel shifts in habitat use in, 19–20; influence of local and regional factors on, 16–17; lack of support of large species flocks, 2, 3; non-random habitat use of fishes in, 19; volume of fresh water in worldwide, 2; young age of large North American, 2 Lampreys (Petromyzontidae), 105–107, 106f, 108–110t; acrosomal process in sperm of, 189, 229; adult features shared with Hagfishes, 106; blood vs. flesh feeding in, 135–136; commensalism in, 136; commercial importance of, 137–138; conservation of, 136–137; conservation status of, 137t; and dams, 137; distribution and body size, 112, 114; diversity and distribution of, 107, 110–112, 110f, 111f; ecology of feeding adults, 134–135; embryonic development in, 132–133; family and genetic relationships in, 114–115, 114f; fecundity of, 132; feeding mechanisms of adults of parasitic species, 120–123, 120f, 121f, 122f; feeding mechanisms of larvae, 119, 119f; fossil record of, 117–118, 118f, 119f; gene order of, 124; genetics and craniate evolution, 124; genome duplications in, 124; hemoglobins of, 126;

hypothalamic-pituitary-gonadal axis and reproduction in, 128, 128f; inferior swimming ability of adult, 126, 131; as iono- and osmoregulators, 105, 126–127, 127f, 130; karyology of, 123–124; lack of homing in parasitic species of, 129; larvae of, 105, 107, 118, 133–134; larval phase and metamorphosis in, 133–134; lateral-line system in, 130; lifecycle characteristics of North American, 112t; lifecycle type as a species-specific characteristic, 117; melatonin biosynthesis in, 130; metamorphosis of, 123, 133–134; morphology of, 118; nonparasitic species of, 105, 115–117, 116f; olfaction in, 135; parasites of, 136; parasitic species of, 105, 107; phylogenetic relationships in, 113f; predators of, 136; and relationship with Hagfishes, 106–107; respiration in, 125–126; roots of ordinal name, 106; semelparity of, 105; sequential stages in speciation of nonparasitic, 117; sexual dimorphism in mature adults, 128; size of, 118–119; spawning in, 131–132, 132f; synapomorphies for adult, 114; upstream spawning migration of nonparasitic, 131, 131f; upstream spawning migration of parasitic, 128–131, 129f; vision in, 127–128 landscape fi lters model, 14, 15–16; key elements of, 15; role of landscape in, 15–16 Lang, N. J., 115 lateral-line system, 91, 146–147; in Bowfin, 288–289; in Carps and Minnows, 411– 412; in Lampreys, 130; and mating behavior, 91; in Paddlefish, 217; in Sturgeons, 169; in Whiptail Stingrays, 147, 147f Laurentian Great Lakes. See Great Lakes Lewis, T. C., 157 Li, G.-Q., 301 light: light-sensitive pigments in retinas of freshwater fishes, 54; transmission of, in aquatic ecosystems, 53–54; ultraviolet light, 57; ultraviolet light and courtship in freshwater fishes, 57–58 Light, T., 25 Lindquist, D. G., 84 livebearers, alternative mating strategies in, 93–96 Long, W. L., 268 Lovejoy, N., 142 Macías Garcia, C., 57 Mahy, G., 405 major histocompatibility complex (MHC), 71, 422; and recognition template, 71 Markle, D. F., 451, 499

GENERAL INDEX

mating behavior: diversification of, 102; elevated courtship effect, 82; evolution of, 102; and fertility advertisement, 94, 100; and lateral-line system, 91; stimulus-response chains in, 102. See also alternative mating strategies; olfactory cues in courtship; sound in courtship; vision in courtship Matthews, W. J., 1, 37, 426 Mayden, R. L., 36, 84, 385–386, 387, 390, 392, 396–397, 398, 400, 401, 401–402, 402, 403, 405, 423, 459, 460–461 McEachran, J. D., 142 McElroy, D. M., 415 McIntyre, P. B., 14–15 McKaye, K. R., 436 McLennan, D. A., 36 McPhee, M. V., 472 Meador, M. R., 15 microcystins, 429 Miller, D. L., 414 Miller, R. R., 385, 387, 405, 460 Minckley, W. L., 421 Mississippi Interstate Cooperative Resource Association (MICRA), 198, 235 Missouri River system, 444–445 Mittelbach, G. G., 39 Mooneyes (Hiodontidae), 299; abundance of, 308; age and growth in, 309–310; commercial importance of, 311; conservation of, 311; conservation status of, 311; diet of, 308–309; diversity and distribution of, 299, 300f; egg and larval development in, 307–308, 307f; eye morphology and vision in, 304; fossil record of, 301–302; fragility of, 305; genetics of, 304; habitat of, 308; karyology of, 304; life history traits of, 300t; as living fossils, 299; meaning of name alosoides, 299; meaning of name Hiodon, 299; meaning of name tergisus, 299; mean length of Goldeye, 309t; morphology of, 302–303, 302f; mortality rate of, 310; movement and diel activity in, 305; as non-natives, 299–300; parasites of, 310–311; phylogenetic relationships in, 300–301, 301f; physiology of, 304; predators of, 311; reproductive physiology of, 304–305; separation of species based on meristic data, 303; sexual dimorphism in, 310; sexual maturity in, 305; skeletal features of, 303, 303f; spawning behavior and fecundity of, 306; spawning habitat and timing of, 306; spawning migrations of, 305–306; swim bladder–ear connection in, 303–304; territoriality and dominance in, 305; uptake and secretion of heavy metals in, 305 mosquitofishes, alternative mating strategies in, 94–96

Moyle, P. B., 25, 47 Murie, D. J., 273 mutualism, 36, 44– 45; between fishes and other taxa, 44–45; direct mutualism, 44; in fish assemblages, 46–48; indirect mutualism, 44. See also nest associations Nagle, B. C., 424 Naylor, G. J. P., 401 Near, T. J., 107 Nelson, E. M., 460, 467 Nelson, G., 337 Nelson, J. S., 141–142, 142 nest associations, 46–47, 84, 87, 87f; advantages and disadvantages of, 436; benefits to associate, 84; benefits to nest builder, 84; Carps and Minnows as nest associates, 435–436; costs to nest builder, 84; and egg eating, 84; and hybridization, 98, 436; nest associates and their host species, 85–86t; origin of, as a plesiomorphic side effect, 84; and selfish-herd hypothesis, 436 Netsch, N. F., 265 Nilsson, S., 252 Noltie, D. B., 266 non-native species, introduction of, 3, 24, 27; Anchovies, 335; Bowfin, 281; Carps and Minnows, 384–385, 418; effects of, on Carps and Minnows, 447–448; effects of, on Suckers, 493–494; Freshwater Eels, 317; Gars, 245–246; and loss of species due to competition with, 40; and loss of species due to predation of, 41; Mooneyes, 299–300; Paddlefish, 209–210; Sturgeons, 163; Suckers, 456–457, 459 Northcote, T. G., 40 Ochlockonee River, Florida, 150 Ohno, S. L., 123, 124 olfactory cues in courtship, 64; discrimination of kin from non-kin, 71; and females’ discrimination between conspecific and heterospecific males, 69–71; and major histocompatibility complex genes, 71; males’ use of, to discriminate between receptive and nonreceptive females, 64–65, 69; and medial olfactory tract in teleosts, 70; and olfactory rosette, 64; phenotypic matching, 71; sexual pheromones in North American freshwater fishes, 66–69t; syntactic coding, 70. See also alarm substance (Schreckstoff ) system: and breeding optimal foraging theory, 21–22, 22f Ouachita Highlands, 11 Ouachita River system, 10–11; biogeographic relationships of fishes from, 11t

641

P450arom enzyme, 148 Paddlefishes (Polyodontidae), 208f, 209; acrosomal process in sperm of, 189, 229; age and growth in, 232–234; age at sexual maturity, 225; age of, and interactions among environmental variables, 222; ampullary organs in, 169, 215; ancient body plan of, 168, 213, 213f; annual mortality rates of, 234; aquaculture of, 240, 241–242; artificial propagation and stocking of, 239–240, 241–242; breeding system of, 228; conservation of, 235; conservation status of, 235; daily movement of, 223; and dams, 207, 223, 226, 237–238; derivation of family and genus names of, 207; diel periodicity in, 232; diet of, 231– 232; digestion in, 220–221; diversity and distribution of, 207–208, 208f; early life stages of, 217–218, 218f; egg, sperm, and environment interactions, 229; electrical currents and, 222; electrosensory systems and ram suspension feeding in, 223–224; embryo characteristics and development in, 229, 230t; energetics of, 234; evolutionary considerations regarding, 211; exploratory movement of, 222; fecundity of, 229; filter feeding and electrosensory organs in, 214f, 214–215, 215f, 216f; and fisheries, 235–237, 236f, 240–241, 241f; fossil record of, 211f, 211–213, 212f; global climate change and, 240; habitat of, 231; hearing abilities and inner ear in, 178, 221; higher phylogenetic relationships in, 210–211; industrial use of waterways of, 238–239; inter- and intraspecific variation in, 208–209, 210f; intraspecific genetic variation in, 219; invasive species and, 239; jumping in, 225; karyology of, 219; lack of courtship in, 228; lack of parental care in, 228; lack of resting in, 223; larval development in, 231; larval dispersion in, 222; lateralline system in, 217; life history traits of Polyodon spathula, 209t; low genetic divergence in, 219; metabolism of, 175, 220; navigation in, 222; as non-natives, 209–210; non-territoriality of, 225; original description of, as Sharks, 163, 207; origin of name, 207; other characters, 216–217; oxygen requirements of, 219–220; paedomorphosis in, 218; parasites of, 235; photoreceptors and visual pigments of, 221; phylogenetic relationships in Polyodontidae, 211, 211f; and pollution, 238; population sizes and densities of, 234; predators of, 234; ram ventilation in, 211, 215f, 216, 220, 223, 237; reproductive allocation and spawning

642 GENERAL INDEX

Paddlefishes (Polyodontidae) (cont.) frequency in, 228; salinity tolerance of, 220; schooling behavior of, 225; seasonal feeding periodicity in, 232; seasonality of spawning in, 227; sex ratios in, 228; sexual dimorphism in, 225–226; size of, 214; sound production in, 225; spawning in, 228; spawning cues, 227; spawning migrations of, 226; spawning modes and location, 227–228; spawning site fidelity of, 226; as spoonbill catfishes, 207; swimming of, 216, 217f; swimming performance of, 221; temperature capacity of, 219; tolerance of extreme environments, 222; unique characters of, 213–214, 214f; vision and chemosensory system in, 215–216, 216f paedomorphosis: in Paddlefishes, 218; in Sturgeons, 172–173 Page, L. M., 587 ParaHox genes, 286 Peebles, E. B., 348 peramorphosis, in Sturgeons, 173 Pflieger, W. L., 287 pharyngeal sound production (PSP), 62 phenotypic matching, 71 Philipp, D. P., 90, 98 phylogeography, 286, 423–424 Pigg, J., 25 Piney Creek, Arkansas: effects of flooding on fish fauna in, 32; fish fauna in, 3; fish fauna stability in, 32 Poff, N. L., 15 pollution/pollutants, 103; ammonia, 259–260; Bowfins’ tolerance of, 294; and Carps and Minnows, 428–429; chlordane, 202, 238; copper, 428; DDT, 202; Gars’ tolerance of, 259–260; mercury, 238, 305, 429; and Mooneyes, 305; nitrite, 260; organochlorines, 202, 238; organophosphates, 259; and Paddlefishes, 238; PAHs, 428; PCBs, 202, 328; petroleum, 259; selenium, 429; and Sturgeons, 202–203 Poly, W. J., 420–421 polyploidy, 123, 173 Potter, I. C., 255 predation, 27, 36, 41–44; and alteration of species composition, 41; balancing of risk of, against rewards of foraging, 41–42; costs of predator avoidance, 42; and dilution effect, 95; effects on activity periods, 44; effects on life history attributes, 44; impacts of predation on fish assemblages, 41; predator threat by piscivores, 43; response of predators to presence of alarm substance, 78; risk of, 439; and spatial or temporal shifts in habitat use, 41, 41f; top-down predation, 43

Pumpkinseeds: alternative mating strategies in, 89–90; and Bluegills, 39; pharyngeal sound production in, 62 Pupfishes: alternative male mating strategies of, 88; color in, 55–56; territorial and courtship sounds in, 60, 60f Pyron, M., 435 Rabito, F. G., Jr., 437 Rahel, F. J., 26 Raley, M. E., 400 Ramaswami, L. S., 468 Raney, E. C., 401, 420, 460 Ratajczak, R. E., 47 Raven, P. H., 47 recognition template, 71; and phenotypic matching, 71 Red River system, 10, 10f, 18; testing of efficacy of Diamond’s assembly rules in, 24 Reighard, J., 45, 46, 466, 482 Renaud, C. B., 136 Reno, H. W., 412 resistance, 34 resource partitioning, 36; between Sturgeon species, 194; between sympatric catostomids, 489 Rhymer, J. M., 101 Richardson, R. E., 466 River Analogy, 3 river continuum model, 16 rivers, impoundment of, 2 River Stingrays (Potamotrygonidae), 140; electroreception in, 153 Roberts, W., 311 Robins, C. R., 460 Robinson, B. W., 38 Rolff, J., 124 Rosen, D. E., 50, 97 Rosenberger, L. J., 142 Ryan, M. J., 79 Sabaj, M. H., 420–421 Sage, M., 154 salmonids: alternative mating strategies in, 91–93; male morphs in, 91 salt water, 150 Sandheinrich, M. B., 429 Sawada, Y., 459 Scharpf, C., 451 Schekter, R. C., 342 Schlosser, I. J., 18, 23–24 Schmidt, J., 313 Schmidt, T. R., 401, 402 Schönhuth, S., 396–397, 398, 400, 401 Schultze, H.-P., 243 Scott, W. B., 311 Shaw, K. A., 401 Sibbing, F. A., 407 Siebert, D. J., 459

signals, 79; definition of signal, 79; multiple-signal communication in fishes, 50; signal evolution, 79 Simberloff, D., 101 Simons, A. M., 386, 387, 391, 392, 398, 400, 401, 403, 424 Skalski, G. T., 422 Smith, G. R., 404, 451, 459, 459–460, 460, 461, 468, 471 Smith, R. J. F., 72, 79 Snelson, F. F., Jr., 145, 158 Šorić, V., 105 sound in courtship, 59–60, 63–64; amplification, 50; in darters, 60–61, 60f; frequency, 59; hearing specialists, 59; in minnows, 61–62, 61f; pharyngeal sound production in Pumpkinseeds and Bluegills, 62; production of, 59; pulse parameters of, 62; in pupfishes, 60, 60f; species-specificity of, 59; in Sturgeons, 63; in Sunfishes, 62–63, 63f spawning: broadcast, 227, 228, 433; broadcast, as a precursor to nest association, 436; crevice, 433; crevice, as a precursor to egg clustering, 437; evolution of mode of, 436–437; lithophilic riverine, 185; male simultaneous reproductive parasitism, 483; plasticity in mode of, 436; and water chemistry, 185. See also spawning, and substrate preparation spawning, and substrate preparation: egg clustering, 435; male simultaneous reproductive parasitism pattern, 483; mound building, 434–435; pit building, 434; pit-ridge building, 434; saucer building, 433–434 speciation: alloparapatric speciation, 99; reticulate speciation, 99, 162; speciation by reinforcement, 99 sperm, in freshwater environment, 96 Sprules, W. M., 311 Stauffer, J. R., Jr., 420 Sticklebacks: nuptial color in, 56–57; use of olfactory cues in courtship, 70, 73 St. Johns River, Florida, 150; Atlantic Stingray population in, 157; complex water chemistry of, 150; uniqueness of among major North American rivers, 150 stocking: of Paddlefishes, 239–240; of Sturgeons, 198–199 Strange, E. M., 25 Strange, R. M., 403 streams: adventitious streams, 24; diel shifts in habitat use in, 20–21; forested streams and juvenile salmon, 44–45; impacts of bass species on prey fishes in, 42–43; influence of local and regional factors on, 16; long-term faunal

GENERAL INDEX

shifts in, 35; pools, 18; riffles, 18–19; seasonal changes in habitat in, 19, 21; species richness in pools and riffles, 23; stream order, 16; stream reaches, 18, 19; variation in species composition or functional groups along longitudinal gradients in, 16; volume of fresh water in worldwide, 2 studies/experiments: manipulative field or laboratory experiments, 36, 39; natural experiments, 36, 38; observational field studies, 36, 36–38; studies of character displacement and release, 38; use of null models in evaluating nonexperimental evidence, 24, 37–38 Sturgeons (Acipenseridae), 160; acrosomal process in sperm of, 179, 189, 229; age at sexual maturity, 183–184; age of, 194–195; age of, and interactions among environmental variables in, 179; agriculture and, 202; ampullary organs in, 169; anadromy and diversity in, 162; ancient body plan of, 168, 168f, 213; annual mortality rates of, 196; artificial propagation and stocking of, 198–199; behavioral responses to environmental extremes in, 179; as benthic cruisers, 160, 168–170, 169f; breeding systems of, 188; camouflage coloration in, 182–183; chemosensory systems and feeding in, 182; commercial importance of, 204–206, 205f; competition and resource partitioning in, 194; conservation of, 197–198; conservation status of, 197; courtship in, 186; dams and, 160, 196, 201–202; diel periodicity in, 179–180; diet of, 193; digestion in, 175–176; diversity and distribution of, 160–161, 161f; early life stages of, 170, 172–173, 173f; egg, sperm, and environment interactions, 179; embryo characteristics and development in, 189, 190t, 191; evolutionary rate and diversity in, 162; fasting in, 193, 194; fecundity of, 188–189; feeding periodicity in, 193; feedingrelated sensory systems in, 182; fisheries and, 160, 188, 199–201, 199f, 200f; fossil record of, 165–167, 166f; global climate change and, 203–204; growth in, 195; habitat of, 191–192, 192f; and habitat restoration, 199; hearing abilities and inner ear in, 178, 221; higher phylogenetic relationships in, 163–164, 210; homing capabilities of, 163; hybridization in, 162, 174; identification of species, species boundaries, and hybrids, 174; industrial use of waterways and, 203; intraspecific variation in, 163; introgression in, 174; invasive species and, 203; jumping and sound produc-

tion in, 183, 225; karyology of, 173; lack of vouchering protocols associated with harvesting of tissues, 174; lateral-line system in, 169; life history attributes for genus Acipenser in North America, 171t; life history attributes for genus Scaphirhynchus, 172t; low genetic divergence in, 173–174, 219; metabolism of, 175, 220; morphology of, 167–168, 167f, 168f; movement and non-spawning migrations to optimize feeding and reproductive success, 180–182; natal fidelity of, 184; as non-natives, 163; ontogenetic shifts in habitat use, 192–193; original description of, as Sharks, 160, 163, 210; origin of name of, 160; osmoregulation in, 176; oxygen requirements of, 172; paedomorphosis and peramorphosis in, 172–173; parasites of, 196–197; parental care in, 188; photoreceptors and visual pigments of, 178; phylogenetic evolutionary considerations, 168; phylogenetic relationships in Acipenseridae, 164–165, 164f; pollution and, 202–203; polyploidization and diversity in, 161–162; population genetics of, 174–175; population sizes, densities, and productivity of, 195–196; predators of, 196; relationships among species of, 162–163; salinity tolerance of, 176; seasonality of feeding in, 193; seasonality of reproduction in, 181; sex ratios in, 188; size of, 170; sound production during spawning, 186–187; spawning frequency, 185; spawning in, 187–188; spawning migration patterns in, 184–185; spawning modes and location, 185, 186f; spawning territories of, 185; station-holding techniques of, 178, 221; summer resting in, 182, 191–192; susceptibility of, to overfishing, 188; swimming mechanics in, 170, 170f; swimming performance of, 177–178, 221; techniques for determining sex and reproductive readiness of, 170; temperature capacity of, 176; territorial and courtship sounds in, 63 Suckers (Catostomidae), 451; adult feeding ecology of, 489–491, 490f, 491f; age and growth in, 491–493; alarm substance system in, 478, 480; alternative spawning tactics used by males, 484; aquaculture of, 499, 500; as baitfish, 500; character states of lip textures, 465, 465f; classification of extant Catostomidae, 452t; conservation of, 496–497, 499; conservation status of, 498–499t; derivation of family name, 451; die-offs of, 495; diversity and distribution of, 451; dwarfism in, 491; as ecological indicators, 500; effects of competition,

643

predation, and non-indigenous species on, 493–494; egg characteristics of, 488; environmental pH tolerances of, 472–473; fecundity of, 487–488; and fish kills, 495; fossil record of, 462–463; functional morphology of feeding in, 468– 469, 469f; general morphology of, 463– 464, 464f; gene silencing and duplicate gene expression in, 469– 470; genetic variability in, 470; geographic genetic variation in, 472; habitat preference and environmental tolerance of, 488– 489; and Haff disease, 496; hearing in, 479; higher phylogenetic relationships in, 459–460; as human food, 499–500; hybridization and introgression in, 470– 472; intrafamilial phylogenetic relationships in, 460– 462, 461f; intraspecific morphological variation in, 469; karyology of, 469; larval and juvenile feeding ecology of, 491; larval behavior, 480; migration and homing tendency in, 481– 482; molecular markers in, 472; mouth and lip morphology of, 464– 465, 465f; as mullets, 451; native range of, 454, 456, 457f; nest building, pit spawning, and territoriality in, 487; non-annual spawning in, 486– 487; as non-natives, 456– 457, 459; osteological characters of, 467– 468; parasites of, 495– 496; physiological effects of saline exposure on, 476– 477; pigmentation and breeding tubercles in, 465– 467, 466f, 467f; productivity, recruitment, and drift ecology of, 494– 495; rheotaxis and thigmotaxis in, 480– 481, 481f; schooling behavior of, 480; spawning in, 482– 484; spawning season and conditions, 484– 486, 486f; swim bladder anatomy of, 467; thermal biology and metabolism of, 473– 476; as “trash fish,” 489, 495, 499; vision in, 478– 479; Weberian apparatus in, 479 Suckers (Catostomidae), geographic range of: Carpiodes, 458f; Catostomus, 457f; Chasmistes, 457f; Cycleptus, 458f; Deltistes, 458f; of Erimyzon, 458f; Hypentelium, 458f; Ictiobus, 458f; Minytrema, 458f; Moxostoma, 459f; Thoburnia, 459f; Xyrauchen, 458f Suckers (Catostomidae), life history traits of: Carpiodes, 474t; Catostomus, 475t; Chasmistes, 476t; Cycleptus, 477t; Deltistes, 478t; Erimyzon, 479t; Hypentelium, 480t; Ictiobus, 481t; Minytrema, 482t; Moxostoma, 483t; Thoburnia, 484t; Xyrauchen, 488t Sun, Y., 462

644 GENERAL INDEX

Sunfishes: alternative mating strategies in, 89–91; basic breeding system of, 89; hybridization in, 98; territorial and courtship sounds in, 62–63. See also Bluegills; Pumpkinseeds Suttkus, R. D., 251 swordtails: alternative mating strategies in, 91–92; evolution of body size in, 51–52; and role of ultraviolet radiation in courtship, 57–58, 58f; use of olfactory cues in courtship, 70 synapomorphy, 114, 303 syntactic coding, 70 thermal tolerances: critical thermal maximum, 426; intrinsic thermal acclimation zone, 151; thermal critical maximum, 422; thermal tolerance polygon, 151 Thompson, J. N., 36 Thompson, W. F., 311 Tinbergen, N., 102 Tipton, M. L., 423 Toepfer, C., 426–427 Townsend, C. R., 15 turbidity, 427; Carps’ and Minnows’ tolerance of, 427; effect on mating behavior of freshwater fishes, 103; impact of dams on, 444–445 Turner, T. F., 15 Umpqua River system, 447 Uyeno, T., 405, 460 Villeneuve, D. L., 428 vision in courtship, 50–51, 58–59; and body size, 51–52, 51f; color in darters, 55, 55f; color in killifishes, 54–55, 54f; color in pupfishes, 55–56; exceptions to body-size preference, 52–53; and mate-choice copying, 53; nuptial color-

ation and limits of vision, 53–54; nuptial color in Sticklebacks, 56–57; ultraviolet light and courtship, 57–58, 58f, 59f visual field, 151–152; binocular convergence point, 151–152; convergence distance, 152; measurement of, 151; Type III visual field, 151, 152 Vives, S. J., 398 Vladykov, V. D., 114 Vrba, E. S., 79 Wabash River, Indiana, long-term faunal shifts in, 35 Wallus, R., 306 Warren, M. L., 497 waterways, industrial use of: and Paddlefishes, 238–239; and Sturgeons, 203 Webber, H. M., 429 Weisel, G. F., 468 Whiptail Stingrays (Dasyatidae), 140; age and growth in, 158; commercial importance of, 159; and comparative genomics, 148; conservation of, 159; courtship in, 156; diet and feeding of, 158; diversity and distribution of, 140–141, 143f; electroreception in, 152–153, 152f; embryonic development in, 157; fecundity of, 156–157; fossil record of, 143; genetics of, 148; habitat of, 157; higher phylogenetic relationships in, 141–142; hormone shifts in, 155–156; interspecific phylogenetic relationships in, 142–143, 143f; karyology of, 148; lateralline system in, 147, 147f; life history information for genus Dasyatis, 142t; locomotion in, 146; maximum size of, 145; and mechanotactile hypothesis, 147; metabolism of, 149; morphology of, 143–145, 144f; morphology and mechanosensory function of lateral-line sys-

tem in, 146–147, 146t; movement and diel activity patterns of, 153–154; ontogenetic shifts in habitat use of, 158; origin of family name, 140; osmoregulation in, 149–150; ovoviviparity in, 140, 154; parasites of, 158–159; phylogenetic relationships in Dasyatidae, 142; physiology of, 148–149; predators of, 159; and prey, 158; reproduction in, 154; reproductive behavior of, 156, 156f; reproductive morphology of, 143; salinity tolerance of, 149–150; schooling behavior of, 154; seasonality of reproduction in, 155; sensitivity of, to low conductivity, 150–151; sexual dimorphism and ecomorphological shifts, 154–155, 155f; size at sexual maturity, 154; spawning and parental care in, 15; spine morphology and replacement of, 145–146; swimming kinematics for Atlantic Stingray, 146t; temperature tolerance of, 151; vision in, 147–148; visual field of, 151–152, 151f Whitehead, P. J. P., 339 Whitt, G. S., 460, 470 Wilde, G. R., 425 Wiley, E. O., 243, 401, 423 Wiley, M. L., 623 Williams, R. R. G., 306 Wilson, D. S., 38 Wilson, M. V. H., 301 Wilson, R. J. A., 252 Witt, A., Jr., 265 Wood, C. M., 402 Wood, R. M., 400 Wu, H.-W., 459 Zastrow, C. E., 346 zoogeographic realms: Australian, 1; Ethiopian, 1; Nearctic, 1; Neotropical, 1; Oriental, 1; Palearctic, 1