Amphibians Ecology and Conservation OK
Amphibians Ecology and Conservation OK
Amphibians Ecology and Conservation OK
Edited by
C. Kenneth Dodd, Jr
1
3
Great Clarendon Street, Oxford OX2 6DP
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide in
Oxford New York
Auckland Cape Town Dar es Salaam Hong Kong Karachi
Kuala Lumpur Madrid Melbourne Mexico City Nairobi
New Delhi Shanghai Taipei Toronto
With offices in
Argentina Austria Brazil Chile Czech Republic France Greece
Guatemala Hungary Italy Japan Poland Portugal Singapore
South Korea Switzerland Thailand Turkey Ukraine Vietnam
Oxford is a registered trade mark of Oxford University Press
in the UK and in certain other countries
Published in the United States
by Oxford University Press Inc., New York
© Oxford University Press, 2009
The moral rights of the author have been asserted
Database right Oxford University Press (maker)
First published 2009
All rights reserved. No part of this publication may be reproduced,
stored in a retrieval system, or transmitted, in any form or by any means,
without the prior permission in writing of Oxford University Press,
or as expressly permitted by law, or under terms agreed with the appropriate
reprographics rights organization. Enquiries concerning reproduction
outside the scope of the above should be sent to the Rights Department,
Oxford University Press, at the address above
You must not circulate this book in any other binding or cover
and you must impose the same condition on any acquirer
British Library Cataloguing in Publication Data
Data available
Library of Congress Cataloging in Publication Data
Data available
Typeset by Newgen Imaging Systems (P) Ltd., Chennai, India
Printed in Great Britain
on acid-free paper by
CPI Antony Rowe, Chippenham, Wiltshire
10 9 8 7 6 5 4 3 2 1
Preface
As this volume is completed, more than 6400 amphibian species have been
recognized, with new taxa being described nearly every day. The last few dec-
ades have seen an explosion in systematic research, particularly in the tropics.
Long-recognized centers of diversity have been explored using increasingly
sophisticated sampling techniques, yielding many new taxa. At the same time,
new centers of speciation, such as Sri Lanka and the Western Ghats of India,
have been discovered, while molecular techniques have yielded previously
unsuspected diversity within some well-known taxa, such as the plethodontid
salamanders of southeastern North America and the green toad (Bufo viridis)
complex of Eurasia. For amphibian systematists, these are exciting times.
Unfortunately, amphibians are now at greater peril than at any time in recent
geologic history, a situation chronicled in two recent data-rich books (Lannoo
2005; Stuart et al. 2008). Habitats are being lost at alarming rates because of
expanding human populations and generally favorable economic conditions
fostering development; emerging infectious diseases, particularly amphibian
chytrid fungus (Batrachochytrium dendrobatidis), threaten worldwide impacts;
non-indigenous species proliferate, affecting amphibians and their habitats; and
amphibians, with their permeable skins, diverse life histories, and often biphasic
life cycles requiring both terrestrial and aquatic habitats, are being saturated by
a host of lethal and sublethal toxic substances. New threats, such as the effects
of global climate change, further imperil amphibians, especially those with
limited distributions and dispersal capabilities. Fully one-third of all amphib-
ians are now considered threatened (Stuart et al. 2004), and 168 species have
become extinct within the last two decades. Clearly, these are treacherous times
for many frogs, salamanders, and caecilians.
Amphibians are, quite frankly, engaging animals. Despite Linnaeus’ early
characterization of amphibians in the context of “Terrible are thy works, O
God”, biologists have come to appreciate that their diverse life histories and
shear numbers offer a wealth of material for research on basic ecological princi-
ples, such as trophic interactions, phenotypic plasticity, predator–prey interac-
tions, community structure, mate choice and recognition, water balance, and
many others. In response to threats, conservation biologists have probed these
and other questions in hopes of understanding amphibian biology in order to
prevent declines and extinctions. The basic and applied themes of biology merge
vi | Preface
Margaret Gunzburger, Tibor Hartel, Robert Jehle, Steve Johnson, Y.-C. Kam,
Sarah Kupferberg, Frank Lemckert, Harvey Lillywhite, John Maerz, Joseph
Mitchell, Clinton Moore, Erin Muths, James Petranka, Benedikt Schmidt,
Ulrich Sinsch, Kevin Smith, Lora Smith, Joseph Travis, Susan Walls, and
Matthew Whiles. I greatly appreciate the support from Ian Sherman and Helen
Eaton at Oxford University Press and editorial help from freelance copy-editor
Nik Prowse, and thank series editor, Bill Sutherland, for inviting me to edit the
amphibian volume. This volume is dedicated to all the biologists who take up
the challenge of amphibian ecology and conservation.
C. Kenneth Dodd, Jr
References
Frost, D. R., Grant, T., Faivovich, J., Bain, R. H., Haas, A., Haddad, C. F. B., de Sá, R. O.,
Channing, A.,Wilkinson, M., Donnellan, S. C. et al. (2006). The amphibian tree of
life. Bulletin of the American Museum of Natural History, 297, 1–370.
Gascon, C., Collins, J. P., Moore, R. D., Church, D. R., McKay, J. E., and Mendelson,
III, J. R. (eds) (2007). Amphibian Conservation Action Plan. IUCN/SSC Amphibian
Specialist Group, Gland and Cambridge.
Gent, T. and Gibson, S. (eds) (1998). Herpetofauna Worker’s Manual. Joint Nature
Conservation Committee, Peterborough.
Henle, K. and Veith, M. (eds) (1997). Naturschutzrelevante Methoden der Feldherpetologie.
Mertensiella 7.
Heyer, W. R., Donnelly, M. A., McDiarmid, R. W., Hayek, L.-A., and Foster, M. S.
(eds). (1994). Measuring and Monitoring Biological Diversity. Standard Methods for
Amphibians. Smithsonian Institution Press, Washington DC.
Lannoo, M. J. (ed) (2005). Amphibian Decline. The Conservation Status of United States
Species. University of California Press, Berkeley, CA.
Stuart, S., Chanson, J. S., Cox, N. A., Young, B. E., Rodrigues, A. S. L., Fishman, D. L.,
and Waller, R. W. (2004). Status and trends of amphibian declines and extinctions
worldwide. Science, 306, 1783–6.
Stuart, S., Hoffman, M., Chanson, J. S., Cox, N. A., Berridge, R. J., Ramani, P., and
Young, B. E. (eds) (2008). Threatened Amphibians of the World. Lynx Edicions,
Barcelona; IUCN, Gland; Conservation International, Arlington, VA.
This page intentionally left blank
Contents
Part 1 Introduction 1
1 Amphibian diversity and life history 3
Martha L. Crump
1.1 Introduction 3
1.2 Amphibian species richness and distribution 3
1.3 Amphibian lifestyles and life history diversity 5
1.3.1 Caecilians 6
1.3.1.1 Aquatic 7
1.3.1.2 Combination aquatic and terrestrial 7
1.3.1.3 Terrestrial and/or fossorial 7
1.3.2 Salamanders 7
1.3.2.1 Aquatic 8
1.3.2.2 Combination of aquatic and terrestrial 8
1.3.2.3 Terrestrial 9
1.3.2.4 Fossorial 10
1.3.2.5 Arboreal 10
1.3.3 Anurans 10
1.3.3.1 Aquatic/aquatic 10
1.3.3.2 Terrestrial/aquatic 11
1.3.3.3 Arboreal/aquatic 13
1.3.3.4 Fossorial/aquatic 14
1.3.3.5 Terrestrial/non-aquatic 14
1.3.3.6 Arboreal/non-aquatic 14
1.3.3.7 Fossorial/non-aquatic 15
1.4 Amphibian declines and why they matter 15
1.4.1 Economics 16
1.4.2 Ecosystem function 16
1.4.3 Esthetics 17
1.4.4 Ethics 17
1.5 References 17
Part 2 Larvae 37
3 Morphology of amphibian larvae 39
Roy W. McDiarmid and Ronald Altig
3.1 Background 39
3.2 Larval caecilians 40
3.2.1 Morphology and ontogeny 40
3.2.2 Coloration 42
3.2.3 Diversity 42
3.3 Larval and larviform salamanders 42
3.3.1 Morphology and ontogeny 42
3.3.2 Coloration 44
3.3.3 Diversity 44
3.4 Anuran tadpoles 45
3.4.1 Morphology and ontogeny 45
3.4.2 Coloration 49
3.4.3 Diversity 49
3.5 Summary 50
3.6 References 51
4 Larval sampling 55
David K. Skelly and Jonathan L. Richardson
4.1 Introduction 55
4.1.1 Why sample larvae? 55
4.1.2 Target responses 56
4.1.3 Timing 57
4.1.4 Sampling effort 57
4.2 Sampling techniques 58
4.2.1 Box/pipe sampler 58
Contents | xi
4.2.1.1 Description 58
4.2.1.2 Application 60
4.2.1.3 Considerations 60
4.2.2 Dip net 60
4.2.2.1 Description 60
4.2.2.2 Application 61
4.2.2.3 Considerations 61
4.2.3 Seine 61
4.2.3.1 Description 61
4.2.3.2 Application 62
4.2.3.3 Considerations 62
4.2.4 Leaf litterbags 62
4.2.4.1 Description 62
4.2.4.2 Application 63
4.2.4.3 Considerations 63
4.2.5 Trapping 64
4.2.5.1 Description 64
4.2.5.2 Application 64
4.2.5.3 Considerations 64
4.2.6 Mark–recapture 65
4.2.6.1 Description 65
4.2.6.2 Application 66
4.2.6.3 Considerations 66
4.3 Other techniques 66
4.3.1 Bottom net 67
4.3.2 Electroshocking 67
4.3.3 Visual encounter survey 67
4.4 Conclusions 67
4.5 Acknowledgments 68
4.6 References 68
5.6 Summary 82
5.7 References 83
6 Aquatic mesocosms 87
Raymond D. Semlitsch and Michelle D. Boone
6.1 Introduction 87
6.2 Historical background 88
6.3 Why use mesocosms? 90
6.4 Types of mesocosm 93
6.5 Setting up mesocosms 95
6.6 Common experimental designs 96
6.7 Case studies 98
6.7.1 Community ecology 98
6.7.2 Evolutionary ecology 99
6.7.3 Ecotoxicology 100
6.7.4 Land use and management 100
6.8 Conclusion 101
6.9 References 102
Index 529
This page intentionally left blank
Contributors
Ross A. Alford School of Marine and Tropical Biology, James Cook University,
Townsville, Queensland 4811, Australia. E-mail: [email protected]
Ronald Altig Department of Biological Sciences, Mississippi State University,
Mississippi State, Mississippi 39762-5759, USA. E-mail: [email protected]
Kimberly J. Babbitt Department of Natural Resources and the Environment, University
of New Hampshire, Durham, NH 03824, USA. E-mail: [email protected]
Larissa L. Bailey Department of Fish, Wildlife and Conservation Biology, Colorado
State University, Fort Collins, CO 80523, USA. E-mail: [email protected]
William J. Barichivich Florida Integrated Science Center, US Geological Survey, 7920
NW 71st Street, Gainesville, FL 32653, USA. E-mail: [email protected]
Trevor J. C. Beebee School of Life Sciences, University of Sussex, Falmer, Brighton
BN1 9QG, UK. E-mail: [email protected]
Michelle D. Boone Department of Zoology, 212 Pearson Hall, Miami University,
Oxford, Ohio 45056, USA. E-mail: [email protected]
Andrew J. Crawford Smithsonian Tropical Research Institute, Apartado Postal
0843-03092, Balboa, Ancón, Republic of Panama. E-mail: [email protected]
Dan Cogălniceanu University Ovidius Constant‚a, Faculty of Natural Sciences, Bvd.
Mamaia 124, Constant‚a, Romania. E-mail: [email protected]
Paul Stephen Corn U.S. Geological Survey, Aldo Leopold Wilderness Research
Institute, 790 E. Beckwith Ave., Missoula, MT 59801, USA. E-mail: [email protected]
Martha L. Crump Department of Biological Sciences, Northern Arizona University,
Flagstaff, Arizona 86011, USA. E-mail: marty.crump@ nau.edu
C. Kenneth Dodd, Jr Department of Wildlife Ecology and Conservation, University
of Florida, Gainesville, FL 32611, USA. E-mail: [email protected]
Michael E. Dorcas Department of Biology, Davidson College, Davidson, NC 28035,
USA. E-mail: [email protected]
John W. Ferner Department of Biology, Thomas More College, Crestview Hills,
Kentucky 41017, USA. E-mail: [email protected]
J. Whitfield Gibbons Savannah River Ecology Laboratory, P.O. Drawer E, Aiken,
South Carolina 29802, USA. E-mail: [email protected]
James P. Gibbs Department of Environmental and Forest Biology, State University
of New York, Syracuse, NY 13210, USA. E-mail: [email protected]
Matthew J. Gray Center for Wildlife Health, Department of Forestry, Wildlife and
Fisheries, University of Tennessee, 274 Ellington Plant Sciences Building, Knoxville,
TN 37996, USA. E-mail: [email protected]
xxvi | Contributors
D. Earl Green U.S. Geological Survey, National Wildlife Health Center, 6006
Schroeder Drive, Madison, WI 53711, USA. E-mail: [email protected]
Elizabeth B. Harper State University of New York, College of Environmental
Sciences and Forestry, 1 Forestry Drive, Syracuse, New York 13210, USA. E-mail:
[email protected]
Reid N. Harris Department of Biology, James Madison University, Harrisonburg,
Virginia 22807, USA. E-mail: [email protected]
Lillian M. B. Haywood Department of Biology, Washington and Lee University,
Lexington, VA 24450, USA. E-mail: [email protected]
Kim Howell Department of Zoology and Wildlife Conservation, PO Box 35064,
University of Dar es Salaam, Dar es Salaam, Tanzania. E-mail: [email protected]
Victor S. Lamoureux Binghamton University, PO Box 6000, Binghamton, New
York 13902, USA. E-mail: [email protected]
Harvey B. Lillywhite Department of Zoology, University of Florida, Gainesville, FL
32611, USA. E-mail: [email protected]fl.edu
Dale M. Madison Binghamton University, PO Box 6000, Binghamton, New York
13902, USA. E-mail: [email protected]
David M. Marsh Department of Biology, Washington and Lee University, Lexington,
VA 24450, USA. E-mail: [email protected]
Roy W. McDiarmid Patuxent Wildlife Research Center, US Geological Survey, National
Museum of Natural History, Washington DC 20036, USA. E-mail: [email protected]
Claude Miaud Laboratoire d’ Alpine UMR CNRS 5553, Université de Savoie, 73376
Le Bourget-du-Lac, France. E-mail: [email protected]
Debra L. Miller Veterinary Diagnostic and Investigational Laboratory, The University
of Georgia, College of Veterinary Medicine, 43 Brighton Road, Tifton, GA 31793,
USA. E-mail: [email protected]
Joseph C. Mitchell Mitchell Ecological Research Service, LLC, PO Box 5638,
Gainesville, FL 32627–5638, USA. E-mail: [email protected]
James D. Nichols U.S. Geological Survey, Patuxent Wildlife Research Center, 12100
Beech Forest Road, Laurel, MD 20708, USA. E-mail: [email protected]
Peter W. C. Paton Department of Natural Resources, University of Rhode Island,
Kingston, Rhode Island 02881, USA. E-mail: [email protected]
Joseph H. K. Pechmann Department of Biology, 132 Natural Sciences Building,
Western Carolina University, Cullowhee, North Carolina 28723, USA. E-mail:
[email protected]
Jérôme Pellet A. Mailbach Sàrl, CP 99, Ch. de la Poya 10, CH-1610 Oron-la-Ville,
Switzerland. E-mail : [email protected]
James W. Petranka Department of Biology, University of North Carolina at Asheville,
Asheville, North Carolina 28804, USA. E-mail: [email protected]
Contributors | xxvii
When the first crossopterygian crawled out of the rich Devonian waters and cast the
first envious vertebrate gaze at the terrestrial world, a boundless empire awaited colon-
ization. Although the change from an ungainly lobe-finned locomotion to a terrestrial
walking gait . . . was agonizingly slow, generations succeeded generations, archotypes [sic]
gave way to new evolutionary experiments, and the land became the home for the first
quadrupeds—the amphibians.
William E. Duellman (1970)
1.1 Introduction
Over the past 350 million years, amphibian descendents of lobe-finned fishes
have radiated into most habitats on Earth. In doing so, they have acquired spec-
tacular and sometimes bizarre physiological, morphological, behavioral, and
ecological attributes that mold their innovative life histories. Amphibians have
highly permeable skin, which makes them both vulnerable to losing water and
able to absorb water. Their eggs, covered with jelly capsules rather than hard
shells, lose water rapidly. For these reasons, amphibians require relatively moist
environments.
Many sampling techniques have been developed in North America or Europe where
most amphibians exhibit the complex life cycle of aquatic eggs, aquatic larvae, and
metamorphosis into terrestrial adults that return to water to breed. Not all amphib-
ians fit this stereotype. The spectacular diversity of amphibian life histories provides a
focus for studying their natural history, as well as presents a challenge since researchers
must ensure that field methods are appropriate for the target species.
class gets it name from the Greek words amphi meaning “two” and bios mean-
ing “mode of life” because many species have a diphasic life history: they spend
part of their lives in water and part on land. Biologists divide the class into three
orders: Gymnophiona (caecilians), Urodela (salamanders), and Anura (frogs)
(Figure 1.1).
Gymnophiona, from the Greek words gymnos and ophis meaning “naked ser-
pent,” encompasses 174 species. Caecilians, which resemble large earthworms,
are long, skinny animals with no legs and reduced eyes. Annuli (grooves) encir-
cle their bodies. Their tails are either greatly reduced or absent, and a sensory
tentacle sits between each eye and nostril. Some caecilians have small dermal
scales beneath the surface of their mucus-covered skin. These scales, composed
mainly of collagen fibers and minerals, are not found in salamanders or anurans.
Adult caecilians range in length from a little more than 10 cm to about 1.5 m.
Most are highly specialized for burrowing. Others live on the ground but are
cryptic and secretive. Some are aquatic. Caecilians occur in tropical habitats
around the world except for Madagascar and the Papuan–Australian region
(Pough et al. 2004).
(a) (b)
(c) (d)
Fig. 1.1 Representatives of the three orders of amphibians. (a) Anura: Rana palmipes,
from Ecuador, (b) Anura: Bufo arenarum, from Argentina, (c) Urodela: Phaeognathus
hubrichti, from Alabama, USA, (d) Gymnophiona: Hypogeophis rostratus, Seychelles.
Photographs (a) and (b) by Martha L. Crump; photographs (c) and (d) by C. Kenneth
Dodd, Jr.
1 Amphibian diversity and life history | 5
Five hundred and seventy-one species belong to Urodela, from the Greek
words uro and delos meaning “tail evident.” All salamanders have tails, and adults
have elongate bodies. Most have front and back legs of about the same length;
the limbs of a few aquatic species are greatly reduced or absent. Salamanders
are completely aquatic, terrestrial, combined aquatic and terrestrial, fossor-
ial, or arboreal. Adults range in size from about 30 mm to nearly 2 m. Most
salamander species occur in eastern and western North America and temper-
ate Eurasia, although plethodontids have radiated extensively in Central and
South America. There are no salamanders in sub-Saharan Africa, Australasia,
Australia, or much of tropical Asia, and they are missing from most islands
(Pough et al. 2004).
Frogs, which include toads, make up the order Anura from the Greek words
an and oura, meaning “without tail.” Although tadpoles have tails, adults do
not. Anurans live in the water, on the ground, underground, and in the trees.
Most have long, strong back legs well suited for jumping. Males of most species
call to attract females for mating. Adults of the 5602 recognized species range
in size from about 13 mm to 30 cm. Anurans live almost everywhere except
where restricted by cold temperatures or extremely dry conditions, and except
for many oceanic islands (Pough et al. 2004).
Duellman identified 43 areas worldwide with exceptionally high numbers
of amphibian species, endemic species (those found nowhere else), or both
(Duellman 1999). Nineteen of these high diversity areas are in the western hemi-
sphere; the others are in Eurasia, Africa, and the Papuan–Australian region.
The neotropical region houses 54% of the world’s amphibian species.
Amphibians live in nearly every habitat except for open oceans, most oceanic
islands, polar regions, and some extremely dry deserts (Wells 2007). These
restrictions are imposed on them because of their highly permeable skin that
loses water, and because they are ectothermic: the energy needed to raise their
body temperatures comes from the sun. Thus, amphibians become inactive at
low temperatures. These characteristics, however, work to their advantage as
well. In dry areas and those with seasonal rainfall, amphibians absorb water
through their skin by contacting moist soil. Their low metabolic rates translate
into low energy requirements and allow them to estivate, often underground,
during unfavorable conditions.
of amphibian life histories. In fact, amphibian life histories are often complex,
and many are still poorly understood. Not all species have an aquatic larval stage
and a terrestrial adult stage.
Reproductive mode, a central aspect of amphibian life history, refers to the
site of egg deposition, egg and clutch characteristics, type and duration of
embryonic and larval development, and type of parental care if any (Duellman
and Trueb 1986). Many amphibians are not tied to aquatic habitats for repro-
duction. Instead, they reproduce on land, underground, or in trees, even in
the temperate zones. The following brief discussion of selected lifestyles and
life histories reveals that similar behaviors have evolved in diverse taxonomic
groups and geographical areas: amphibian “experiments” toward greater inde-
pendence from standing or flowing water and perhaps from lower predation
pressure as well.
Readers wishing more information concerning amphibian life histories
should consult Duellman (2007) and Wells (2007). For reviews of reproduct-
ive modes, see Salthe (1969), Salthe and Duellman (1973), Wake (1982, 1992),
Haddad and Prado (2005), and Duellman (2007). For reviews of parental care
see Crump (1995, 1996).
1.3.1 Caecilians
Caecilian lifestyles include aquatic, combined aquatic and terrestrial, terrestrial,
and fossorial. Fully aquatic caecilians generally have compressed bodies with
well-developed dorsal fins on the posterior portion. Fossorial species generally
have blunt heads, used for pushing and compacting the soil while burrowing.
Although details of reproductive biology are unknown for many species, cae-
cilians display two basic modes: oviparous (egg-laying) and viviparous (bear-
ing live young) (Wake 1977, 1992). Oviparous caecilians lay eggs on land. In
some species the eggs hatch into larvae that wriggle to water. Caecilian larvae
exhibit less dramatic metamorphosis than do salamanders or frogs. They hatch
almost fully developed, and the larval period is short. The eggs of some ovip-
arous species undergo direct development. That is, development occurs within
the egg capsule; there is no free-living larval stage. Some oviparous females stay
with their eggs, which probably protects them from predators and from drying
out. Viviparity evolved independently several times in caecilians. Females retain
eggs in their oviducts until development is complete (Wake 1993; Wilkinson
and Nussbaum 1998). After the developing young exhaust their yolk reserves,
they scrape lipid-rich secretions from their mothers’ oviductal epithelium with
their fetal teeth. Gestation lasts for many months, and the newborn are large
relative to their mothers.
1 Amphibian diversity and life history | 7
1.3.1.1 Aquatic
All species in the South American family Typhlonectidae are either fully aquatic
or semi-aquatic. Some in the latter group spend the day in burrows they con-
struct next to water, then emerge at night to feed in shallow water. All typhlonec-
tids are assumed to be viviparous. Soon after the young are born, they shed their
gills and quickly acquire adult morphology.
1.3.2 Salamanders
Salamander lifestyles include fully aquatic, combined aquatic and terrestrial,
terrestrial, arboreal, and fossorial. Body shapes of aquatic salamanders range
from the slender and eel-like sirenids and amphiumids to the flattened and
robust cryptobranchids. Many permanently aquatic species retain external gills
as adults. Most arboreal salamanders are small and have extensively webbed feet;
some have prehensile tails. Burrowing salamanders generally have long slender
bodies and tails, reduced limbs and feet, and small body size.
8 | Amphibian ecology and conservation
1.3.2.1 Aquatic
Some aquatic salamanders lay eggs in still or slowly flowing water. Siren and
Pseudobranchus from southeastern USA and northeastern Mexico live in swamps,
lakes, marshes, and sluggish streams, where they attach their eggs to vegetation.
The larvae develop into eel-like paedomorphic adults: they lack eyelids, have
external gills, and their skin resembles larval skin. When their habitat dries out,
sirenids secrete a mucous cocoon and burrow into the mud where they estivate
until conditions improve. The Mexican axolotl (Ambystoma mexicanum) is per-
manently aquatic and paedomorphic. Its larvae fail to metamorphose fully, and
the gonads mature in the larval body form.
Proteus anguinus, from southeastern Europe, lives in subterranean lakes and
streams of limestone caves where it lays aquatic eggs that hatch into aquatic lar-
vae. Females attend their eggs. In North America, Typhlomolge and Haideotriton
also live in cave waters and lay aquatic eggs that hatch into aquatic larvae. All
these species are paedomorphic.
Some aquatic salamanders oviposit in flowing water of cold streams. Male
hellbenders (Cryptobranchus alleganiensis) from eastern North America con-
struct nests under rocks. More than one female might lay her strings of eggs in
the nest. The male guards the eggs through their early stages. Likewise, male
Andrias japonicus from Japan and male Andrias davidianus from central China
guard their nests. In all three cryptobranchid species, aquatic larvae undergo
incomplete metamorphosis. The adults retain certain larval features such as lid-
less eyes and the absence of a tongue pad.
1.3.2.3 Terrestrial
Many species of salamander live under leaf litter or logs, retreating into crevices
or holes during dry conditions. Terrestrial salamanders, including many in the
temperate zone, have various ways of reproducing independent of water bodies.
The seepage salamander (Desmognathus aeneus) oviposits in a nest near water. The
larvae hatch at an advanced stage and metamorphose within a few days in the
nest without feeding. Other plethodontid salamanders (e.g. Desmognathus wrighti,
Bolitoglossa, and Plethodon) lay large, direct-developing eggs in cavities or inside
hollow logs. In many species, the female remains with the eggs; in a few species
the male remains instead. Montane populations of European fire salamanders
(Salamandra salamandra) often retain eggs in their oviducts. The young absorb
nutrients from their yolk and are born live, although lowland populations usually
have aquatic eggs and larvae. Salamandra atra are viviparous; after the developing
young exhaust their yolk reserves, they obtain nutrients from the female.
10 | Amphibian ecology and conservation
1.3.2.4 Fossorial
Most Oedipina, fossorial or semi-fossorial plethodontid salamanders ranging
from Mexico to northern South America, lay direct-developing eggs under-
ground. Some fossorial plethodontids (e.g. Lineatriton lineola) attend their
direct-developing eggs.
1.3.2.5 Arboreal
Neotropical arboreal plethodontids (e.g. Bolitoglossa and Nototriton) lay their
eggs under mats of mosses and liverworts on tree branches and in bromeliads.
The eggs undergo direct development, and some female Bolitoglossa attend their
eggs. Aneides lugubris from western North America lays direct-developing eggs
as high as 10 m above the ground.
1.3.3 Anurans
Anuran lifestyles include purely aquatic, aquatic and terrestrial, terrestrial, arbor-
eal, and fossorial. A frog with relatively short hind legs is most likely a terrestrial
hopper or fossorial species. One with long hind legs is likely to be aquatic, arbor-
eal, or a jumping terrestrial species. Aquatic anurans tend to have their eyes on the
top of their heads rather than at the sides, and they often have fully webbed feet.
Some have flattened bodies. Arboreal frogs generally have expanded pads on the
ends of their toes. Many fossorial species have small heads with pointed snouts
and depressed bodies. Some have spadelike tubercles on their hind feet used for
burrowing.
Anurans have evolved remarkably diverse life histories, from aquatic eggs to
viviparity. In between those extremes, frogs from numerous families on many
continents lay their eggs out of water yet have aquatic larvae. In the section head-
ings below, the first designation refers to post-metamorphic stages, the second
to egg and/or larval stages.
1.3.3.1 Aquatic/aquatic
Many aquatic frogs lay eggs that hatch into tadpoles. African clawed frogs
(Xenopus) attach their eggs to submerged vegetation in standing water. The
South American paradox frog (Pseudis paradoxa) oviposits among vegetation
in shallow water of ponds and lakes. Aquatic larvae grow to 25 cm—the largest
of any frog—then they metamorphose into relatively small juveniles. Tailed
frogs (Ascaphus truei) from northwestern USA and adjacent Canada live in cold,
torrential streams where they lay eggs under rocks. In some areas larvae require
several years to metamorphose.
1 Amphibian diversity and life history | 11
Some fully aquatic anurans brood their young. Female South American Pipa
carry eggs embedded in their backs. In some species of Pipa the eggs hatch into
tadpoles. In others they undergo direct development. Female Australian gastric-
brooding frogs (Rheobatrachus) swallowed their late-stage eggs or early-stage
larvae, and the tadpoles absorbed yolk reserves while developing in their mothers’
stomachs. No Rheobatrachus have been seen since the 1980s; both species are
assumed extinct.
1.3.3.2 Terrestrial/aquatic
Terrestrial anurans from many families lay aquatic eggs, and the aquatic larvae
metamorphose into terrestrial juveniles. Species that oviposit in standing water
produce clutches in compact masses (some ranids), floating rafts on the water
surface (some ranids), strings (most Bufo and some pelobatids), scattered indi-
vidually or in small packets on the bottom substrate (Bombina and Discoglossus),
attached individually or in small groups onto submerged plants (some species
of Pseudacris, Acris, Hyla, and Spea), or as a film on the water surface (some
microhylids).
Some anurans breed in moderately fast streams or mountain torrents. Female
Atelopus from Central and South America lay their eggs in strings attached to
rocks. The larvae have ventral sucker-like discs that allow them to adhere to
rocks while feeding. Tadpoles of bufonid stream-breeding Ansonia from Asia
and Werneria from Africa have similar suckers. Other stream-adapted frogs,
such as many ranids, lay large eggs in compact masses attached to rocks in areas
where the current is slow and the eggs are less likely to be swept away.
Some terrestrial anurans produce foam nests in which their eggs are sus-
pended. Male Leptodactylus, Physalaemus, and Pleurodema from Central and
South America kick their hind legs during amplexus, whipping the females’
eggs and mucus and their sperm and mucus into foamy masses (Figure 1.2a).
The outermost layer of foam dries quickly and provides some protection against
desiccation and predation. Some foam-nesters produce their nests on the water
surface, others in cavities or holes next to ponds. Leptodactylus bufonius con-
structs mud nests at the margin of temporary ponds and deposits its foam nests
inside (Figure 1.2b). The eggs hatch into tadpoles that remain in the nests until
rains dissolve the nests and flood the area. Some Australian myobatrachids also
construct foam nests on the water surface.
Some terrestrial frogs oviposit in very small bodies of water on land.
Brazilian Bufo castaneoticus lay their eggs in water-fi lled fruit capsules of the
Brazil nut tree, and the tadpoles feed on detritus. Eupsophus from Chile and
12 | Amphibian ecology and conservation
(a) (b)
(c) (d)
Fig. 1.2 Representatives of four anuran modes of reproduction. (a) Pleurodema borelli
pair constructing a foam nest on the surface of water, from Argentina, (b) Leptodactylus
bufonius mud nest by the edge of a depression, from Argentina, (c) Hyla bokermanni eggs
on a leaf above water, from Ecuador, (d) male Rhinoderma darwinii brooding tadpoles in
its vocal sac, from Chile. Photographs by Martha L. Crump.
where they hatch, then transports the larvae to water. In Europe, male midwife
toads (Alytes) carry their eggs wrapped around their hind legs. Eventually they
hop to ponds where the eggs hatch into aquatic larvae.
1.3.3.3 Arboreal/aquatic
Taxonomically diverse arboreal frogs lay their eggs on vegetation overhanging
water. After the eggs hatch, the larvae fall into the water below where they continue
to develop. Agalychnis, Phyllomedusa, and some Hyla from the New World trop-
ics lay their eggs over standing water (Figure 1.2c), and neotropical centrolenids
oviposit over flowing water. In some centrolenids, a parent protects the eggs from
predators and keeps them moist by resting on them. Female Afrixalus from sub-
Saharan Africa oviposit on leaves above water, then fold the leaf edges together and
glue them in place with oviductal secretions. Some arboreal Old World rhacoph-
orids and hyperoliids construct foam nests on vegetation overhanging temporary
pools or slow-moving streams.
Water-filled basins offer oviposition sites that presumably lessen the risk of
eggs getting swept away and reduce predation. Males of several neotropical
gladiator frogs (e.g. Hyla boans and Hyla rosenbergi) construct basins beside
streams or rivers. Water seeps in and fills the nests, and the frogs lay eggs as a
surface film. After developing in the basin, the tadpoles metamorphose into
froglets that take to the trees.
Some arboreal anurans oviposit in water-filled tree holes and axils of aer-
ial plants. The eggs of many of these frogs (e.g. Anodonthyla, Platypelis, and
Plethodonthyla from Madagascar) have large amounts of yolk. The tadpoles typ-
ically lack mouthparts, and they are non-feeding. In contrast, female Osteopilus
brunneus from Jamaica lay eggs in water-filled leaf axils of bromeliads and con-
tinue to deposit about 250 more eggs in the bromeliad every few days throughout
the tadpoles’ development. The tadpoles—up to about 170 in a clutch—feed on
the later-arriving eggs until they metamorphose into arboreal froglets.
Some arboreal frogs attach their eggs to the walls of water-filled cavities in
trees. After hatching, the tadpoles drop into the water. In Chirixalus eiffingeri,
an Asian rhacophorid, the female returns periodically and deposits fresh eggs
for her tadpoles to eat. In others of this reproductive mode, the aquatic larvae
feed on algae and debris.
In the New World tropics, female Flectonotus carry their eggs in dorsal
pouches. After the eggs hatch as advanced tadpoles, the females transport
them to water-filled bromeliads or bamboo where they complete development.
Females of some neotropical Gastrotheca also brood their eggs in dorsal pouches
and transport the tadpoles to aquatic sites.
14 | Amphibian ecology and conservation
1.3.3.4 Fossorial/aquatic
Scaphiopus and Spea, North American spadefoot toads, spend much of their
lives underground but emerge following heavy rains and lay their eggs in newly
formed ponds. The tadpoles develop quickly, which increases the probability
of metamorphosing before the ponds dry. Rhinophrynus dorsalis, from south-
ern Texas to Costa Rica, likewise lives underground and emerges after the first
heavy rains to breed in temporary ponds. Female Hemisis marmoratum from
Africa lay their eggs in subterranean chambers and stay with their eggs until
after they hatch. At that point, the female digs a tunnel into adjacent water for
the tadpoles.
1.3.3.5 Terrestrial/non-aquatic
Terrestrial anurans exhibit diverse life histories that free them from aquatic
breeding sites. Adenomera from South America deposits foam nests under logs
or in terrestrial cavities, and non-feeding tadpoles develop in the nests until they
metamorphose. Neotropical Eleutherodactylus, Oreophrynella from Guyana and
southern Venezuela, and New Guinean microhylids lay direct-developing eggs
under logs or leaf litter. Many attend their eggs.
Some completely terrestrial anurans brood their young. Female Assa dar-
lingtoni from Australia attend their terrestrial eggs. When the eggs are about
12 days old, the father climbs into the egg mass, rupturing the capsules. The
newly hatched tadpoles wriggle into brood pouches, one along each side of
the male’s body. He broods his non-feeding larvae until they metamorph-
ose into froglets. Female Darwin’s frogs (Rhinoderma darwinii) from Chile
and Argentina lay their eggs on moist ground. Just before the eggs hatch the
males gobble them into their mouths and into the vocal sacs (Figure 1.2d).
The young ingest secretions from the vocal-sac lining and emerge from their
fathers’ mouths as froglets.
Several Nectophrynoides, African bufonids, retain eggs in their oviducts and
give birth to live young. In Nectophrynoides occidentalis, after depleting their
yolk reserves the developing embryos feed on “uterine milk” secretion produced
by glands in the mother’s oviduct walls. These frogs live at high elevations,
exposed to long periods of cold and drought. The females have a 9-month gesta-
tion period during which they estivate underground.
1.3.3.6 Arboreal/non-aquatic
Some neotropical Eleutherodactylus lay their direct-developing eggs in tree
holes, bromeliads, moss, or on leaves. Some attend their eggs, others do not.
Eleutherodactylus jasperi from Puerto Rico lived in arboreal bromeliads and gave
1 Amphibian diversity and life history | 15
birth to live young. The direct-developing eggs were retained in the oviducts,
and nutrition came entirely from the embryo’s yolk reserves. This species has not
been seen since 1981 and is assumed extinct.
Female Cryptobatrachus, Stefania, and Hemiphractus, neotropical hylids,
carry their direct-developing eggs exposed on their backs, secured by mucous
gland secretions. Females of some Gastrotheca brood direct-developing eggs in
dorsal pouches that protect the developing embryos from predators and des-
iccation and also function in gaseous exchange between the females and their
embryos.
1.3.3.7 Fossorial/non-aquatic
The burrowing microhylid Synapturanus salseri from Colombia lays its eggs
in burrows just below the root mat on the forest floor. Non-feeding tadpoles
hatch at an advanced stage and absorb their yolk reserves. The Brazilian bur-
rowing leptodactylid Cycloramphus stejnegeri likewise oviposits in underground
nests and has non-feeding tadpoles. Other fossorial anurans, such as Geocrinia
and Arenophryne (Australian myobatrachids), Callulops (New Guinean micro-
hylids), and Breviceps (African microhylids), lay direct-developing eggs in
underground burrows. Female Breviceps stay with their eggs and presumably
keep them moist.
1.4.1 Economics
Selfishly, we should care if we lose amphibians because we use them for our own
benefit, including for food and as pets. We use literally tonnes of frogs each year
in medical research and teaching. We have isolated novel chemical compounds
from granular glands of anuran skin and have used these compounds to develop
new drugs.
1.4.3 Esthetics
Imagine the silence of rainy spring evenings without the lively croaking of male
frogs. The monotonous roads without spring migrations of salamanders. People
worldwide consider frogs to be good luck because of their association with rain.
Amphibians provide inspiration for our artistic endeavors, from literature to
music and the visual arts.
1.4.4 Ethics
Every species is a unique product of evolution. In 1982 the United Nations
General Assembly adopted the World Charter for Nature, which states: “Every
form of life is unique, warranting respect regardless of its worth to man, and,
to accord other organisms such recognition, man must be guided by a moral
code of action” (Noss and Cooperrider 1994). More than 100 nations signed
the charter. Like all other living species, amphibians have intrinsic value and a
right to exist.
Amphibians, amazing descendants of terrestrial pioneers, are fighting for
their lives in a world greatly modified by humans.
1.5 References
Blaustein, A. R. and Wake, D. B. (1990). Declining amphibian populations: a global phe-
nomenon? Trends in Ecology and Evolution, 5, 203–4.
Collins, J. P. and Storfer, A. (2003). Global amphibian declines: sorting the hypotheses.
Diversity and Distributions, 9, 89–98.
Crump, M. L. (1995). Parental care. In H. Heatwole and B. K. Sullivan (eds), Amphibian
Biology, vol. 2: Social Behaviour, pp. 518–67. Surrey Beatty and Sons, Chipping
Norton, NSW.
Crump, M. L. (1996). Parental care among the Amphibia. In J. S. Rosenblatt and
C. T. Snowdon (eds), Advances in the Study of Behavior, vol. 25: Parental Care: Evolution,
Mechanisms, and Adaptive Significance, pp. 109–44. Academic Press, New York.
Daszak, P., Berger, L., Cunningham, A. A., Hyatt, A. D., Green, D. E., and Speare, R.
(1999). Emerging infectious diseases and amphibian population declines. Emerging
Infectious Diseases, 5, 735–48.
Duellman, W. E. (1970). The Hylid Frogs of Middle America, vol. 1. Monograph of the
Museum of Natural History, Number 1. University of Kansas, Lawrence, KA.
Duellman, W. E. (1999). Global distribution of amphibians: patterns, conservation, and
future challenges. In W. E. Duellman (ed.), Patterns of Distribution of Amphibians: a
Global Perspective, pp. 1–30. John Hopkins University Press, Baltimore, MD.
18 | Amphibian ecology and conservation
Duellman, W. E. (2007). Amphibian life histories: their utilization in phylogeny and clas-
sification. In H. Heatwole and M. J. Tyler (eds), Amphibian Biology, vol. 7. Systematics,
pp. 2843–92. Surrey Beatty and Sons, Chipping Norton, NSW.
Duellman, W. E. and Trueb, L. (1986). Biology of Amphibians. McGraw-Hill, NewYork.
Ehrlich, P. R. and Ehrlich, A. H. (1981). Extinction: The Causes and Consequences of the
Disappearance of Species. Random House, New York.
Gallant, A. L., Klaver, R. W., Casper, G. S., and Lannoo, M. J. (2007). Global rates of
habitat loss and implications for amphibian conservation. Copeia, 2007, 967–79.
Groom, M., Meffe, G. K., and Carroll, C. R. (2006). Principles of Conservation Biology,
3rd edn. Sinauer Associates, Sunderland, MA.
Haddad, C. F. B. and Prado, C. P. A. (2005). Reproductive modes in frogs and their unex-
pected diversity in the Atlantic Forest of Brazil. BioScience, 55, 207–17.
IUCN (International Union for the Conservation of Nature) (2008). Red List of
Threatened Species. IUCN, Gland. http://www.iucn.org/themes/ssc/redlist.htm.
Myers, N. (1990). Mass extinctions: what can the past tell us about the present and the
future? Global and Planetary Change, 82, 175–85.
Noss, R. F. and Cooperrider, A. Y. (1994). Saving Nature’s Legacy: Protecting and Restoring
Biodiversity. Island Press, Washington DC.
Phillips, K. (1994). Tracking the Vanishing Frogs: an Ecological Mystery. St. Martin’s Press,
New York.
Pough, F. H., Andrews, R. M., Cadle, J. E., Crump, M. L., Savitzky, A. H., and Wells, K. D.
(2004). Herpetology, 3rd edn. Prentice Hall, Upper Saddle River, NJ.
Raven, P. H. (1990). The politics of preserving biodiversity. BioScience, 40, 769–74.
Salthe, S. N. (1969). Reproductive modes and the number and sizes of ova in the urodeles.
American Midland Naturalist, 81. 467–90.
Salthe, S. N. and Duellman, W. E. (1973). Quantitative constraints associated with
reproductive mode in anurans. In J. L. Vial (ed.), Evolutionary Biology of the Anurans,
pp. 229–49. University of Missouri Press, Columbia, MO.
Skerratt, L. F., Berger, L., Speare, R., Cashins, S., McDonald, K. R., Phillott, A. D.,
Hines, H. B., and Kenyon, N. (2007). Spread of chytridiomycosis has caused the rapid
global decline and extinction of frogs. EcoHealth, 4, 125–34.
Smith, K. F., Sax, D. F., and Lafferty, K. D. (2006). Evidence for the role of infectious
disease in species extinction and endangerment. Conservation Biology, 20, 1349–57.
Stuart, S. N., Chanson, J. S., Cox, N. A., Young, B. E., Rodrigues, A. S.L., Fischman, D. L.,
and Waller, W. (2004). Status and trends of amphibian declines and extinctions world-
wide. Science, 302, 1783–6.
Wake, M. H. (1977). The reproductive biology of caecilians: an evolutionary perspec-
tive. In D. H. Taylor and S. I. Guttman (eds), The Reproductive Biology of Amphibians,
pp. 73–101. Plenum, New York.
Wake, M. H. (1982). Diversity within a framework of constraints. Amphibian repro-
ductive modes. In D. Mossakowski and G. Roth (eds), Environmental Adaptation and
Evolution, pp. 87–106. Gustav Fischer, New York.
Wake, M. H. (1992). Reproduction in caecilians. In W. C. Hamlett (ed.), Reproductive
Biology of South American Vertebrates, pp. 112–20. Springer-Verlag, New York.
1 Amphibian diversity and life history | 19
for order among facts. Fieldwork then serves to test one’s hypothesis, thereby
scientifically demonstrating correlation, causation, or the absence thereof. The
hypothesis must of necessity regard some facts as more significant than others,
based on the researcher’s previous experience, familiarity with the literature,
and individual interpretation.
As scientific knowledge is rarely complete, any list of potential alternative
hypotheses is also unlikely to be complete. Therefore, alternative hypotheses for
the particular problem or question being considered must also be formulated.
Thus, it is necessary to conduct research with both the hypothesis and alter-
native hypothesis in mind, as more than one cause may be contributing to any
single effect (Wolff and Krebs 2008). For example, Jaeger (1972) used enclos-
ures to test the mechanism of interspecific competition between two species of
plethodontid salamanders, hypothesizing that it resulted from either differential
exploitation of food resources or through interference. Both primary and alter-
native hypotheses are formulated and tested so that researchers can determine
which explanations best fit the results obtained. Unsupported hypotheses can be
rejected, but even the most parsimonious hypothesis may not be fully accepted
because there still may be underlying explanations as yet untested (Senar 2004).
Repeated experimentation involving alternative hypotheses eventually allows
researchers to gain a measure of confidence in the validity of empirically sup-
ported hypotheses (Jaeger and Halliday 1998). Both hypotheses and data are
essential for credible science since, as Krebs (1999) concluded, hypotheses with-
out data are not very useful, and data without hypotheses are useful only for an
inductive approach.
Ecology is an empirical science that requires data from the “real” world,
and a field study is a tool which can ultimately lead to the acceptance or rejec-
tion of an hypothesis. After repeated experimentation and validation, data
and information from many field studies are synthesized and organized into
concepts of how nature functions. These concepts are the result of the integra-
tion between what scientists think they know (based on previous research and
observation) and newly acquired data (Ford 2000). Asking the right question
is important, because the type of question asked and the particular techniques
and methods used in hypothesis-driven research strongly influence what is
discovered and the direction of future research. During the course of this
feedback process, scientists come to a better understanding of the phenomena
they study (Figure 2.1).
Scientists present what they want to accomplish during a field study in terms
of goals and objectives. A goal is a statement that explains what the study is
designed to accomplish. It is usually a broad and general statement, inclusive of
2 Setting objectives in field studies | 23
Fig. 2.1 The cycle of scientific investigation and the shift towards the spiral of knowledge.
a long-term direction. The goal is then split into specific, measurable objectives
that indicate how the goal can be achieved within a specified time frame, and
the expected results.
Defining a study’s objectives clearly is the first and probably the most
important single step in research planning, and is a key element in a success-
ful project. The goals of research can be non-applied (i.e. aimed at increasing
or changing existing knowledge) or applied (i.e. focused on solving practical
questions), or both. Rather than addressing an effect, a study should focus
on the cause of the phenomenon in question. Separating effects from causes
is a major challenge when setting objectives. For example, acid rain has well-
known causes, but much research today focuses on the effects or seeks solutions
to limit its impact. In another example, many recent studies have reported on
amphibian declines (Stuart et al. 2004), but many fewer have identified spe-
cific causes (e.g. Becker et al. 2007). It is important that amphibian biologists
do not limit their focus to specific taxa; otherwise, they lose sight of the fact
that similar effects may be occurring in other taxa. Amphibian biologists need
to establish closer links with researchers studying similar problems in other
taxa (Halliday 2005).
24 | Amphibian ecology and conservation
Ideally, the cycle of scientific investigation should proceed from data and infor-
mation gathering and analysis, to quantifying knowledge, and to conceptual
understanding. A logical, stepwise approach of scientific inquiry (Lehner 1996)
should be followed: (1) perceive that a problem/question exists; (2) formulate a
possible explanation (i.e. devise an hypothesis); (3) formulate alternative hypoth-
eses; (4) identify the best approach to test the hypothesis (i.e. theoretical models,
experiments, or field observations); (5) collect and analyze data; (6) support or
reject the hypothesis; and (7) understand the meaning and implications of the
results. The original hypothesis can then be modified, experiments repeated and,
with time, conceptual understanding attained.
be based on an understanding of its life history (i.e. feeding, habitat use, disper-
sal abilities, reproduction, behavior, predators, phylogeography, population gen-
etics) and its suitability in resolving the particular question being asked (Wolff
and Krebs 2008). For example, a species with low detectability is not a good
choice for a mark–recapture study since it will require a considerable sampling
effort and entail potential capture bias (Weir et al. 2005). The role of keystone
species (e.g. Eleutherodactylus bransfordii in Costa Rican forests), umbrella spe-
cies (e.g. Rana sylvatica in the Milwaukee river basin), or flagship species (e.g.
Salamandra lanzai in the Western Alps) can also be considered. Rare or threat-
ened species are often selected because of conservation applications, but research
on them could incur potential risks due to ethical considerations. Researchers
should avoid choosing a rare or threatened species if common or less vulnerable
surrogate species can be selected.
words, it might be better to collect more data than appears necessary at first,
than to discover later that some important parameter was omitted. A pilot study
helps to avoid under- or over-collecting data. Although conceptual models help
researchers identify what to measure, the timing of studies is determined by the
natural history of the species of interest.
• S: The specific part of an objective defines what will be done and where it will
occur. When setting objectives, ask simple questions that can be answered,
and avoid ambiguities, the use of buzzwords or jargon (e.g. “cutting-edge”),
and pompous phrasing.
• M: Measurable is an attribute of an activity or its results. The source of
and mechanism for collecting measurable data are identified, and collec-
tion of these data is determined feasible. If the objective of the study is to
document trends (i.e. an increase or decrease of one or more measurable
variables), then a baseline is required to act as a reference point (e.g. habi-
tat availability, characteristics and use; amphibian community structure;
population size). If a baseline is not yet available then it will be useful to
first have it established.
• A: Attainable refers to the probability of conducting the proposed activ-
ities within the established time frame with the available resources and sup-
port. It also includes the external factors critical to success. Doing research
today, for example, can become difficult due, in part, to increasing admin-
istrative restrictions (Prathapan et al. 2008). Other external factors include
unforeseen costs, shifts in exchange rates, obtaining collecting and access
permits, changes in legislation, and political, security, and health (both
human and animal) issues. In coping with external factors, a risk analysis
is useful because it allows for planning alternative strategies.
• R: Some useful measures for the relevance of the study’s objectives are the
utility and value of the results for practical purpose (e.g. management of
protected areas, better conservation measures, and ecological restoration). It
28 | Amphibian ecology and conservation
may be difficult to determine how relevant the objective may be, especially
since the true relevance may not be apparent until the study is completed
(e.g. if a drought were to affect a long-term study of amphibian breeding).
Perhaps an easy way to evaluate the relevance is to answer the “so what”
question, thus avoiding undertaking studies of limited or no interest.
• T: Finally, the time required to complete the project will depend on the
parameters discussed above.
Mathematical
model
Control and precision
Laboratory
Unpredictability
experiment Simple
Realism
Outdoor
Complexity
experiment
Unrestricted field
experiment
(3) prematurely accepting a solution as the only possible answer; and (4) using
data and information that are either incorrect (e.g. inaccurate information in the
scientific literature) or irrelevant. There are several additional pitfalls that may be
avoided by careful planning and a thorough understanding of the questions to be
asked (Tucker et al. 2005). Four of the most frequent are listed below.
1) The statistical framework might be inadequate, since many techniques
developed in the context of controlled experimentation are sometimes
incorrectly applied to field data, resulting in an inappropriate use of the
null hypothesis (Johnson 1999).
2) Researchers and technicians might differ in their skills, use non-comparable
methods, or have different personal goals. Training prior to the start of field
collection of data and a comparison of each person’s abilities helps to min-
imize these problems.
3) Methods may be changed during a study. This could lead to an incompati-
bility of data sets and limit the interpretation of results.
4) The locations of permanent sample sites are not properly recorded so that
different areas are subsequently revisited or sampled.
When designing fieldwork, researchers need to be aware of potential options
and trade-offs, and try to balance them (Hairston 1989). Examples include:
(1) complexity versus simplicity (e.g. choices ranging from theoretical models to
field experiments), (2) confidence in results versus general application (e.g. high
confidence can be achieved at short temporal and small spatial scales and with
relatively simple goals and conceptual models, but the results will be of limited
value), and (3) replication versus sophistication of experimental design, recog-
nizing that it is impossible to simultaneously maximize precision, realism, and
generality (Levins 1968).
measures, such as equipment disinfection (Chapter 26) and routine checks for
unwanted “passengers” (e.g. seeds) are now widely used (ARG-UK 2008).
Controversies continue regarding the negative effects of some common tech-
niques. For example, toe-clipping, historically one of the most widely used mark-
ing techniques, has recently received much criticism (Chapter 8; May 2004;
McCarthy and Parris 2004). Critics, however, have not provided much evalu-
ation of the impacts of alternative procedures. Apart from dorsal or ventral pat-
tern mapping by photography or computer imaging, which are non-invasive, all
other marking techniques have disadvantages (Phillott et al. 2007). The effects
of toe-clipping vary among species, and therefore must be assessed accordingly.
Other marking methods for amphibians are available, although they are not as
economic or as easy to use, but which have fewer risks (Chapter 8; Ferner 2007).
Toe-clipping also may be prohibited by regulatory constraints; information on
regulations is best obtained and evaluated during the planning stage. Thus, there
may be a trade-off between the risks associated with methodology and the know-
ledge to be gained, even when the species may be benefited (Funk et al. 2005).
Field studies often involve years of hard work. In the end, the results may
provide few insights compared to the amount of effort to acquire them, unless
careful planning precedes the initiation of research activities. Careful planning
optimizes researcher effort and helps ensure that the data recorded will be stat-
istically accurate, with beneficial results in advancing knowledge of amphibian
ecology and conservation biology.
2.5 Acknowledgments
Cristina Vâlcu, Tibor Hartel, Marian Griffey, and Ken Dodd provided helpful
comments on previous versions of this chapter.
2.6 References
Alford, R.A. and Richards, S.J. (1999). Global amphibian declines: a problem in applied
ecology. Annual Review of Ecology and Systematics, 30, 133–65.
ARG-UK (Amphibian and Reptile Groups of the UK) (2008). Amphibian Disease
Precautions: a Guide for UK Fieldworkers. ARG-UK Advice Note 4, pp. 1–5. www.
arg-uk.org.uk/Downloads/ARGUKAdviceNote4.pdf.
Arntzen, J. W., Oldham, R. S., and Latham, D. M. (1995). Cost effective drift fences for
toads and newts. Amphibia-Reptilia, 16, 137–45.
Arntzen, J. W., Goudie, I. B., Halley, J.J., and Jehle, R. (2004). Cost comparison of
marking techniques in long-term population studies: PIT-tags versus pattern maps.
Amphibia-Reptilia, 25, 305–15.
2 Setting objectives in field studies | 33
Maher, W. A., Cullen, P. W., and Norris, R. H. (1994). Framework for designing sam-
pling programs. Environmental Monitoring and Assessment, 30, 139–62.
May, R. M. (2004). Ethics and amphibians. Nature, 431, 403.
McCarthy, M. A. and Parris, K. M. (2004). Clarifying the effect of toe clipping on frogs
with Bayesian statistics. Journal of Applied Ecology, 41, 780–6.
Meyer A. H., Schmidt, B.R., and Grossenbacher, K. (1998). Analysis of three amphibian
populations with quarter-century long time-series. Proceedings of the Royal Society of
London Series B Biological Sciences 265, 523–8.
Michener, W. K. (2000). Research design: translating ideas to data. In W. K. Michener
and J. W. Brunt (eds), Ecological Data: Design, Management and Processing, pp. 1–24.
Blackwell Science, Oxford.
Morrison, M. L. (2002). Wildlife Restoration. Techniques for Habitat Analysis and Animal
Monitoring. Society for Ecological Restoration. Island Press, Washington DC.
Pechmann, J.H.K., Scott, D.E., Semlitsch, R.D., Caldwell, J.P., Vitt, L.J., and Gibbons,
J.W. (1991). Declining amphibian populations: the problem of separating human
impacts from natural fluctuations. Science, 253, 892–5.
Pellet J., Schmidt, B.R., Fivaz, F., Perrin, N., and Grossenbacher, K. (2006). Density, cli-
mate and varying return points: an analysis of long-term population fluctuations in the
threatened European tree frog. Oecologia, 149, 65–71.
Phillott, AD, Skerratt, L. F., McDonald, K. R., Lemckert, F. L., Hines, H. B., Clarke, J. M.,
Alford, R. A., and Speare, R. (2007). Toe-clipping as an acceptable method of identifying
individual anurans in mark recapture studies. Herpetological Review, 38, 305–8.
Piotrow, P. T., Kincaid, D. L., Rimon, J. G., and Rinehart W. (1997). Health
Communication: Lessons from Family Planning and Reproductive Health. Johns Hopkins
School of Public Health, Center for Communication Programs. Praeger Publishers,
Westport, CT.
Pollock, K. H., Nichols, J. D., Simons, T. R., Farnsworth, G. L., Bailey, L. L., and Sauer
J. R. (2002). Large scale wildlife monitoring studies: statistical methods for design and
analysis. Environmetrics, 13, 105–19.
Prathapan, K. D., Rajan, P. D., Narendran, T. C., Viraktamath, A. C., Aravind, N. A., and
Poorani, J. (2008). Death sentence on taxonomy in India. Current Science, 94, 170–1.
Rowe, C. L. and Dunson, W. A. (1994). The value of simulated pond communities in
mesocosms for studies of amphibian ecology and ecotoxicology. Journal of Herpetology,
28, 346–56.
Schneider, D. C. (2001). The rise of the concept of scale in ecology. BioScience, 51,
545–53.
Scott, C. T. and Köhl, M. (1993). A method of comparing sampling design alternatives
for extensive inventories. Mitteilungen der Eidgengessischen Anstalt fuer Wald, Schneeund
Landschaft, Birmensdorf, 69, 1–62.
Senar, J. C. (2004). Mucho mas que plumas. Monografies del Museu de Ciències Naturals
2. Museu de Ciències Naturals i l’Institut Botànic de Barcelona, Barcelona.
Stuart, S. N., Chanson, J. S., Cox, N. A., Young, B. E., Rodrigues, A. S. L., Fischman, D. L.,
and Waller, W. (2004). Status and trends of amphibian declines and extinctions world-
wide. Science, 306, 1783–5.
2 Setting objectives in field studies | 35
Tucker, G., Bubb, P., de Heer, M., Miles, L., Lawrence, A., Bajracharya, S.B., Nepal, R.C.,
Sherchan, R., and Chapagain, N. R. (2005). Guidelines for Biodiversity Assessment and
Monitoring for Protected Areas. KMTNC, Kathmandu, Nepal.
Underwood, A. J. (1997). Experiments in Ecology. Their Logical Design and Interpretation
Using Analysis of Variance. Cambridge University Press, Cambridge.
Weir, L. A., Royle, A., Nanjappa, P., and Jung, R. E. (2005). Modeling anuran detection
and site occupancy on North American Amphibian Monitoring Program (NAAMP)
routes in Maryland. Journal of Herpetology, 39, 627–39.
Whinam, J., Chilcott, N., and Bergstrom, D. M. (2005). Subantarctic hitchhikers: expe-
ditioners as vectors for the introduction of alien organisms. Biological Conservation,
121, 207–19.
Wilbur, H. M. (1997). Experimental ecology of food webs: complex systems in temporary
ponds. Ecology, 78, 2279–2302.
Wolff, J. O. and Krebs, C. J. (2008). Hypothesis testing and the scientific method revis-
ited. Acta Zoologica Sinica, 54, 383–6.
Yoccoz, N. G. (1991). Use, overuse and misuse of significance tests in evolutionary biol-
ogy and ecology. Bulletin of the Ecological Society of America, 72, 106–11.
This page intentionally left blank
Part 2
Larvae
This page intentionally left blank
3
Morphology of amphibian larvae
Roy W. McDiarmid and Ronald Altig
3.1 Background
The larvae of amphibians are non-reproductive and usually aquatic. Most
undergo metamorphosis prior to attaining an adult morphology and sexual
maturity. Species within each amphibian order that develop by other modes (e.g.
direct development; Altig and Johnston 1989; Thibaudeau and Altig 1999) have
non-feeding larvae or embryos, and we do not discuss them in this chapter.
Amphibian larvae have some generalized morphological features that are use-
ful for identification. In contrast to fish, they lack bony supports in the tail fins,
and the vent is a longitudinal slit (not obvious in tadpoles). Amphibian larvae
also lack eyelids, and most have external gills that are visible at some stage in
their ontogeny.
Most caecilians (Gymnophiona) whether terrestrial or aquatic as adults,
have aquatic larvae that look grossly like the legless, elongate adults. The larvae
of salamanders (Caudata) look much like the adults. Unlike the condition in
caecilians and frogs, salamanders may occur in larval (i.e. larval morphology,
non-reproductive, and metamorphic) or larviform (i.e. larval morphology,
reproductive, may or may not metamorphose) states. Larviform salamanders
may exist as pedotypes (i.e. larval relative to the normal developmental tra-
jectory of the taxon, reproductive, will metamorphose if the environmental
conditions change to the detriment of being in the larval environment; some
ambystomatids and salamandrids; terminology of Reilly et al. 1997) or pedo-
morphs (i.e. larval relative to the developmental pattern of their ancestors, do
not metamorphose; all amphiumids, cryptobranchids, proteids, sirenids, and
some plethodontids). Larval and larviform salamanders grossly resemble meta-
morphosed individuals in general body form but retain a number of larval fea-
tures. Pond-adapted forms have a more bulky body and larger gills and tail fins
than the more streamlined, stream-adapted forms.
40 | Amphibian ecology and conservation
The larvae of frogs and toads (Order Anura), called tadpoles, are grossly dif-
ferent from adults and have many developmental (Altig and Johnston 1989) and
morphological (e.g. Altig and McDiarmid 1999a; also various morphologies doc-
umented in staging tables, see Duellman and Trueb 1986, pp. 128–9) features not
seen in other amphibian larvae. Tadpoles live in many kinds of habitats; the most
common types of tadpoles are found in lentic or lotic water, spend most of their
time on the bottom, and feed by rasping material from submerged surfaces.
Because amphibians are ectotherms, their inherent developmental rates are
modified by temperature and other environmental variables; size is thus an
inaccurate estimator of chronological age. Consequently, biologists describe
tadpole ontogeny using a staging table that divides their development into rec-
ognizable stages based on the attainment of specific morphological landmarks.
With such a table the degree of development of morphological features of tad-
poles can be compared among populations and across taxa occurring in the
same or different habitats regardless of chronological age or attainable size.
Larval amphibians are exceptionally variable within and among species, although
the degree and patterns of that variation are poorly documented and their sources
rarely investigated. The many papers on induced morphological changes published
in recent years (e.g. Relyea and Auld 2005, among many others) have made it abun-
dantly clear that every tadpole of a given taxon collected at any site is a variant
within the broad phenotypic range of its taxon. Although results from mesocosm
experiments with controlled combinations and densities of species provide some
insight into understanding phenotypic variation, predicting morphological vari-
ation from random samples of ponds is highly unlikely. The presence of different
sets of predators in natural situations complicates the picture even more, and one
has to keep these factors in mind when evaluating the morphology of tadpoles.
We mention in passing that less is known about amphibian eggs (e.g. Altig
and McDiarmid 2007), hatchlings (Gosner stages 21–24; Altig 1972), and
metamorphs than is known about tadpoles (stages 25–41) and other amphibian
larvae. We urge workers to preserve and describe positively identified samples of
these stages. Here we summarize data on the morphology, ontogeny, and diver-
sity of larvae in each amphibian order.
(a) (b)
(c) (d)
3.2.2 Coloration
Most caecilians are somewhat drab shades of gray, brown, black, or blue. A few
species are more brightly colored and slightly banded or striped (terminology of
Altig and Channing 1993).
3.2.3 Diversity
Larvae of species of caecilians in the families Rhinatrematidae, Ichthyophiidae,
and some Caeciilidae that are known are similar to and grossly resemble adults
in general morphology. The paucity of ontogenetic data, however, makes fur-
ther comparisons impossible.
(a) Gm Df Vf
TL
Gf Cf Cg
V
SVL (d)
Gm
(b) (c)
Lb N
Lf
Gf
Gm
B Gr
Ll
B Is
Fig. 3.2 Measurements and body parts of a larval salamander. (a) Lateral (upper)
and ventral (lower) views of a typical Ambystoma salamander larvae, drawing by
D. Karges; (b) dorsal view of the head and anterior body region of a hatchling
Ambystoma maculatum, photograph by A.M. Richmond; (c) ventral view of head
of A. maculatum; and (d) stylized drawing of gill structure of a larval salamander,
modified from drawing in Pfingsten and Downs (1989). B, balancer; Cf, costal fold;
Cg, costal groove; Df, dorsal fin; Gf, gular fold; Gm, gill ramus with fimbriae attached
to posterior surface; Gr, gill rakers; Is, interbranchial septum; Lb, limb bud; Lf, labial
fold; Ll, lateral line organs (neuromasts); N, naris; SVL, snout–vent length; TL, total
length; V, vent; Vf, ventral fin.
dorsal fin usually starts long before other metamorphic changes are noticeable.
Fleshy flaps occur on the trailing edges of the hind legs of Onychodactylus lar-
vae. Keratinized toe tips are found in a number of taxa but are most common in
stream inhabitants. Sirenids have keratinized jaw sheaths (upper and lower) as
do some ambystomatid larvae (lower).
Hatching occurs before the limbs are fully developed, and the front limbs
usually develop faster than the hind ones. Some hatchlings (e.g. in the families
Ambystomatidae and Salamandridae) have a balancer, a fleshy projection on
each side of the lower part of the head (Figures 3.2b and c), that is lost soon after
hatching. Staging tables (compilation in Duellman and Trueb 1986, p. 128)
are available for several species (e.g. Ambystoma maculatum, Harrison 1969;
Ambystoma mexicanum, Cano-Martinez et al. 1994; and Hynobius nigrescens,
Iwasawa and Yamashita 1991).
3.3.2 Coloration
In contrast to Stereochilus marginatus (Plethodontidae), Rhyacotriton spp.
(Rhyacotritonidae), and most pedomorphs, most of which retain something
similar to the larval coloration as adults, larval salamanders often have a color-
ation distinct from that of the metamorph or the adult. Larval sirenids are jet
black with contrasting stripes and bands of red or yellow, while adults have either
a gray or black ground color usually overlain by speckles and small blotches of
gold to greenish iridophores. Color and pattern (i.e. coloration) in larvae of most
species can vary considerably during ontogeny, throughout a day, and among
sites in response to substrate color, temperature, and water clarity. Colors are
typically muted grays, browns, and blacks, and patterns range from none (uni-
colored), blotched, and mottled through striped (longitudinal or diagonal con-
trasting markings) and banded (transverse contrasting markings). The dorsum
of the tail muscle of small Ambystoma is often banded, and the pattern may
be retained throughout ontogeny (e.g. Ambystoma talpoideum) or change to a
totally different pattern sometime after hatching and then again after metamor-
phosis. In some species and populations, larval Desmognathus (Plethodontidae)
have a distinct pattern that is kept throughout life. Although colors are usually
more muted, the adult patterns in other salamanders (e.g. Ambystoma tigrinum
group, many plethodontids) may appear at metamorphosis or a different pattern
may appear (e.g. most Ambystoma) after metamorphosis and slowly develop into
the adult pattern, which is achieved long before sexual maturity.
3.3.3 Diversity
Larval salamanders show much less ecomorphological diversity than tadpoles.
By definition, pedotypes and pedomorphs retain a larval morphology even
3 Morphology of amphibian larvae | 45
though they become reproductive, and species that metamorphose and are
adapted for either pond or flowing water are the most easily recognized groups.
Cannibal morphotypes with enlarged heads and altered dentition occur in some
species (Ambystomatidae, Hynobiidae). Pond inhabitants often do not over-
winter, whereas some stream inhabitants may grow as larvae for several years
before undergoing metamorphosis. In some parts of their range Notophthalmus
viridescens (Salamandridae) larvae metamorphose into a brilliantly colored eft
that lives on the forest floor for several years before returning to the ponds for an
aquatic existence as a reproductive adult.
(a)
IND
TMW
IOD
BL TaL
TMH MTH
TL
OD SP LB VT TMA
(b) (c)
Fig. 3.3 Measurements and body parts of a tadpole. (a) Dorsal (upper) and lateral
(lower) views of a typical tadpole, drawing of Rana sp. by D. Karges; (b) dorsal (left)
and lateral (right) eye positions of tadpoles, stylized drawings by D. Karges; and
(c) examples of medial (left, Bufo boreas) and dextral (right, Rana catesbeiana) vent
tubes; white arrow, plane of ventral fin; black arrow, outflow of vent tube. BL, body
length; IND, internarial distance; IOD, interorbital distance; LB, hind limb bud; MTH,
maximum tail height; OD, oral disc; SP, spiracle; TMA, tail muscle axis; TMH, tail muscle
height; TMW, tail muscle width; TL, total length; TaL, tail length; VT, vent tube.
associated with the margins. A vent tube extends posteriorly from the midventral
abdomen. Two major types are recognized: dextral, where the aperture lies to the
right of the sagittal plane of the tail fin (e.g. hylids and ranids), and medial, where
the aperture lines parallel with the plane of the tail fin (e.g. bufonids and scaphi-
opodids; Figure 3.3c). As with the spiracular tube configurations, there are many
subtle variations in the shape and position of the vent tube.
The lateral line system (i.e. neuromasts; Hall et al. 2002; Lannoo 1985) is
composed of many depressions in the skin with sensory cells in the center that
3 Morphology of amphibian larvae | 47
signal the patterns of water flow over various parts of the body and tail. The
distribution and arrangement of neuromasts may be useful in distinguish-
ing between closely related species. In darkly pigmented tadpoles these sites
are often pale and obvious, but evaluation of stitch patterns in most tadpoles
requires separating the epidermis from the underlying dermis, clearing in gly-
cerin, and viewing the skin with dark-field illumination (see Lannoo 1985).
The oral apparatus, the composite of upper and lower labia and all kerati-
nized mouthparts, is highly variable across taxa and ecological types. The most
common oral apparatus (Figure 3.4a and c) occurs in many taxa in lentic and
lotic sites. An assembly of the two infralabial and two Meckel’s cartilages with
three joints forms the lower jaw that is surmounted by a serrated, keratinized
jaw sheath. The supralabial cartilage of the upper jaw is surmounted by a simi-
lar keratinized jaw sheath, and during a bite, the lower jaw passes totally behind
the upper (Figure 3.4d). The interactions of the serrated margins of the sheaths
serve as cutting/gouging surfaces when a tadpole feeds. The highly variable
shapes of the jaw sheaths suggest different feeding abilities.
The face of the oral disc has fleshy transverse tooth ridges (Figure 3.4a,
c, and d) surmounted by a row(s) of keratinized labial teeth. In most cases,
several replacement teeth are interdigitated below a presently erupted tooth
(Figure 3.4b, lower right), and they successively move into position as the
erupted tooth wears out. The tooth rows are numbered from the anterior edge
of the disc toward the mouth on the upper labium and from the mouth to
the posterior edge of the disc on the lower labium. A fractional designation
indicates the number of tooth rows on each labium; some rows have naturally
occurring medial gaps denoted parenthetically. For example, a Labial Tooth
Row Formula (LTRF) of 2(2)/3(1) indicates two upper rows with a gap in the
second one and three lower rows with a gap in the first one (Figures 3.4a and
c). Some tadpoles lack tooth rows (i.e. 0/0), and the maximum LTRF known
is 17/21 in a tadpole of an undescribed hylid frog from the Guayana Highlands
of southern Venezuela.
The papillate margins of the oral disc may be complete and encircle the disc
(e.g. tadpoles of Scaphiopodidae, Pelobatidae, and many stream-inhabiting tad-
poles of several families), have a medial dorsal gap (most common; Figure 3.4a),
or have both dorsal and ventral gaps (e.g. Bufonidae and those of some Hylidae,
Mantellidae, Ranidae, and Rhacophoridae; Figure 3.4c). Although the number
of rows of papillae on different parts of the disc margin may vary, the lengths of
the papillae are typically somewhat uniform; several species of Phrynobatrachus
(Ranidae) have exceptionally elongate papillae along the posterior margin of the
disc. Submarginal papillae occur on the face of the disc away from the margin
(a) (b)
UJS LJS
A-1 A-2
G
SRC
MZJ
MP IRC
A-1
UJS
LJS P-1
P-2
MZT P-3
TS
C
H
TR S-1
B
P-1 P-2 P-3 TR SP EM S-2
(c) (d)
1
2
Fig. 3.4 Components of the oral disc of a tadpole. (a) Oral disc of a typical tadpole,
schematic drawing by D. Karges; (b) sagittal sections of the oral apparatus of a common
benthic tadpole (upper) and a tooth ridge (lower left), schematic drawings modified
from ones in Heron-Royer and Van Bambeke (1889); two labial teeth of Hyla chrysoscelis
(lower right) removed from a tooth series and in natural position; (c) scanning electron
photomicrograph of the oral apparatus of a tadpole of Bufo fowleri, actual oral disc
width, about 2.3 mm, photograph by M. Penuel-Matthews; and (d) lateral view of
the mouthparts of H. chrysoscelis (1, upper jaw sheath; 2, lower jaw sheath; 3, lower
tooth rows; 4, upper tooth row). A-1, A-2, anterior tooth rows 1 and 2; P-1, P-2,
P-3, posterior tooth rows 1 to 3; S-1, S-2, sheaths of presently erupted (1) and first
replacement (2) teeth; B, body of first replacement tooth; C, cusps on first replacement
tooth; EM, lateral emargination in oral disc; G, dorsal gap in marginal papillae; H, head of
presently erupted tooth; IRC, infrarostral cartilage; LJS, lower jaw sheath; M, mouth; MP,
marginal papillae; MZJ, mitotic zone for production of jaw sheath; MZT, mitotic zone for
production of labial teeth; SP, submarginal papillae; SRC, suprarostral cartilage; TR, tooth
ridge; TS, tooth series; UJS, upper jaw sheath.
3 Morphology of amphibian larvae | 49
and form various patterns (Figure 3.4a). The margin of the disc may be emar-
ginate (i.e. indented; Figure 3.4a) or not.
Hatchlings (Gosner 1960; stages 21–24) usually have external gills but lack
eyes and limb buds. The forelimb buds develop beneath the operculum after it
closes, and the hind-limb buds grow from the posteroventral intersection of the
body and tail muscle (Figure 3.3a, bottom). Staging tables have been made for
a number of taxa (Duellman and Trueb 1986, pp. 128–129), but using a com-
mon table allows for meaningful comparisons among taxa. Gosner (1960; gen-
eral) and Nieuwkoop and Faber (1956; Xenopus) are the two most commonly
cited. Recent summaries of tadpole morphology and terminology are included
in Altig (2007b) and Altig and McDiarmid (1999a, 1999b).
3.4.2 Coloration
Except for notations in descriptions, surprisingly little has been written
about tadpole coloration. As in other larval amphibians, three basic popula-
tions of pigment-containing cells interact to produce both color and pattern.
Melanophores contain melanins that produce browns and blacks, iridophores
contain reflective guanine crystals that produce whites and silvers, and xan-
thophores contain carotenoids that produce yellows and reds. The pigments
are retained inside the cells and can be dispersed in various patterns under the
influence of temperature and light. Altig and Channing (1993) summarized the
diversity of colorations in tadpoles, and Caldwell (1982) tested the functions of
coloration in tadpoles experimentally.
3.4.3 Diversity
Most morphological characters of tadpoles reflect their ecology. Suctorial tad-
poles in a number of families have streamlined bodies and mouthparts modi-
fied to maintain position in fast-flowing water as they feed. A typical increase
in the number of tooth rows, to a maximum of 17/21, is usually accompanied
by a larger oral disc with complete marginal papillae. Other unusual morpho-
logical structures found in stream-inhabiting tadpoles include a belly modified
as a sucker (a few bufonid and ranid species), and lateral sacs (or lymphatic
sacs) on the ventrolateral parts of the body of other stream-dwelling tadpoles
(Arthroleptidae). Tadpoles of Mertensophryne (Bufonidae) have a hollow crown
on the head that encircles the eyes and nares, and tadpoles of Schismaderma
carens (Bufonidae) have a semicircular, transverse flap of skin behind the eyes.
Suspension-feeding tadpoles in the families Microhylidae, Pipidae, and
Rhinophrynidae have reduced, soft mouthparts that lack keratinized struc-
tures. They usually hang in midwater and capture suspended particles as water
is pumped in through the mouth and out the spiracles. Even so, not all tadpoles
50 | Amphibian ecology and conservation
that lack keratinized mouthparts are suspension feeders, and tadpoles with
keratinized mouthparts that are infected with the amphibian chytrid fungus
(Batrachochytrium dendrobatidis) often lose most or all such structures.
Carnivores and other macrophagous feeders have a diversity of mouth-
parts related to how they feed. For example, tadpoles of the leptodactylid frog
Lepidobatrachus have huge mouths but almost no soft or keratinized mouth-
parts; they engulf entire organisms, including other tadpoles. Tadpoles of
other leptodactylid frogs, Ceratophrys spp., have huge jaws and many tooth
rows and efficiently tear their victims to pieces. Carnivorous tadpoles of Spea
(Scaphiopodidae) feed similarly. A number of other tadpoles (e.g. Hylidae: Hyla
leucophyllata group; Ranidae: Occidozyga) lack all or most tooth rows but have
huge jaw sheaths. Tadpoles that occupy tree holes and bromeliad tanks (e.g.
some species of Dendrobatidae, Hylidae, and Rhacophoridae) are of several
morphological types (Lannoo et al. 1987; Lehtinen et al. 2004) and have differ-
ent diets; some are non-feeding, several are detritovores or macrophagous car-
nivores, many eat frog eggs (fertilized or not) of their own (cannibals) or other
species, and some eat only trophic eggs supplied by their mother. Surface-film
feeders have large oral discs, but keratinized structures are reduced or absent;
the disc is turned upward (i.e. umbelliform) and captures material carried in
the surface film. The oral discs of these surface-feeding tadpoles may be formed
primarily from the lower labium (e.g. Microhylidae) or from parts of both labia
(e.g. Megophryidae). This convergent morphology occurs in six families, and
most tadpoles of this sort occur only in the slow reaches of streams.
Attempts have been made to define ecomorphological guilds or groups of taxa
with suites of common morphological characters that are presumed to reflect
a common ecology (e.g. Altig and Johnston 1989). Because of the lack of eco-
logical data for many species and an incomplete understanding of how some of
their morphologies actually function, we advise caution in assigning species to
specific guilds without some knowledge of their natural history. For example,
the morphologies of Mantidactylus lugubris (Altig and McDiarmid 2006) and
of some taxa that occur in phytotelms suggest that one might find them in fast-
flowing water. In fact, tadpoles of M. lugubris live in leaf packs in slow-flowing
water.
3.5 Summary
Amphibian larvae show considerable morphological diversity from the relatively
conserved forms of caecilians and salamanders to the unusual and often novel struc-
tures found in tadpoles of frogs and toads. The extreme variability of tadpoles is
3 Morphology of amphibian larvae | 51
3.6 References
Altig, R. (1972). Notes on the larvae and premetamorphic tadpoles of four Hyla and
three Rana with notes on tadpole color patterns. Journal of the Elisha Mitchell Scientific
Society, 88, 113–19.
Altig, R. (2007a). Comments on the descriptions and evaluations of tadpole mouthpart
anomalies. Herpetological Conservation and Biology, 2, 1–4.
Altig, R. (2007b). A primer for the morphology of anuran tadpoles. Herpetological
Conservation and Biology, 2, 73–6.
Altig, R. and Channing, A. (1993). Hypothesis: functional significance of colour and
pattern of anuran tadpoles. Herpetological Journal, 3, 73–5.
Altig, R. and Johnston, G. F. (1989). Guilds of anuran larvae: relationships among devel-
opmental modes, morphologies, and habitats. Herpetological Monographs, 3, 81–109.
Altig, R. and McDiarmid, R. W. (1999a). Body plan: development and morphology. In
R. W. McDiarmid and R. Altig (eds), Tadpoles, the Biology of Anuran Larvae, pp. 24–51.
University of Chicago Press, Chicago, IL.
Altig, R. and McDiarmid, R. W. (1999b). Diversity: familial and generic characteriza-
tions. In R. W. McDiarmid and R. Altig (eds), Tadpoles, the Biology of Anuran Larvae,
pp. 295–337. University of Chicago Press, Chicago, IL.
Altig, R. and McDiarmid, R. W. (2006). Descriptions and biological notes on three
unusual mantellid tadpoles (Amphibia: Anura: Mantellidae) from southeastern
Madagascar. Proceedings of the Biological Society of Washington, 119, 418–25.
Altig, R. and McDiarmid, R. W. (2007). Morphological diversity and evolution of egg
and clutch structure in amphibians. Herpetological Monographs, 21, 1–33.
Brauer, A. (1897). Beiträge zur Kenntnis der Entwicklungsgeschichte und Anatomie der
Gymnophionen. Zoologische Jahrbucher Anatomie, 10, 277–472.
Brauer, A. (1899). Beiträge zur Kenntnis der Entwicklung und Anatomie Gymnophionen.
Zoologische Jahrbucher Anatomie, 12, 477–508.
Caldwell, J. P. (1982). Disruptive selection: a tail color polymorphism in Acris tadpoles in
response to differential predation. Canadian Journal of Zoology, 60, 2818–27.
Cano-Martínez, A., Vargas-González, A., and Asai, M. (1994). Metamorphic stages in
Ambystoma mexicanum. Axolotl Newsletter, 23, 64–71.
52 | Amphibian ecology and conservation
Drake, D. L., Altig, R., Grace, J. R., and Walls, S. C. (2007). Occurrence of oral defects
in larval anurans from protected sites. Copeia, 2007, 449–58.
Duellman, W. E. and Trueb, L. (1986). Biology of Amphibians. McGraw-Hill, New York.
Dünker, N., Wake, M. H., and Olson, W. M. (2000). Embryonic and larval development
in the caecilian Ichthyophis kohtaoensis (Amphibia, Gymnophiona): a staging table.
Journal of Morphology, 243, 3–34.
Gosner, K. L. (1960). A simplified table for staging anuran embryos and larvae with notes
on identification. Herpetologica, 16, 183–90.
Haas, A., Hertwig, S., and Das, I. (2006). Extreme tadpoles: the morphology of the fos-
sorial megophryid larva, Leptobrachella mjobergi. Zoology, 109, 26–42.
Hall, J. A., Larsen, Jr, J. H., and Fitzner, R. E. (2002). Morphology of the prometamor-
phic larva of the spadefoot toad, Scaphiopus intermontanus (Anura: Pelobatidae), with
an emphasis on the lateral line system and mouthparts. Journal of Morphology, 252,
114–30.
Handrigan, G. R., Haas, A., and Wassersug, R. J. (2007). Bony-tailed tadpoles: the devel-
opment of supernumerary caudal vertebrae in larval megophryids (Anura). Evolution
and Development, 9, 190–202.
Harrison, R. G. (1969). Harrison stages and description of the normal development
of the spotted salamander, Ambystoma punctatum (Linn.). In R. G. Harrison (ed.),
Organization and Development of the Embryo, pp. 44–6. Yale University Press, New
Haven, CT.
Héron-Royer, L. F. and Van Bambeke, C. (1889). Le vestibule de la bouche chez les
têtards des batraciens anoures d’Europe; sa structure, ses caractères chez les diverses
espèces. Archives de Biologie, 9, 185–309.
Iwasawa, H. I. and Yamashita, K. (1991). Normal stages of development of a hynobiid
salamander, Hynobius nigrescens Stejneger. Japanese Journal of Herpetology, 14, 39–62.
Lannoo, M. J. (1985). Neuromast topography in Ambystoma larvae. Copeia, 1985,
535–9.
Lannoo, M. J. (2008). Malformed Frogs. The Collapse of Aquatic Ecosystems. University of
California Press, Berkeley, CA.
Lannoo, M. J., Townsend, D. S., and Wassersug, R. J. (1987). Larval life in the leaves:
arboreal tadpole types, with special attention to the morphology, ecology, and behavior
of the oophagous Osteopilus brunneus (Hylidae) larva. Fieldiana Zoology, 38, 1–31.
Lehtinen, R. M., Lannoo, M. J., and Wassersug, R. J. (2004). Phytotelm-breeding
anurans: past, present and future research. In R. M. Lehtinen (ed.), Ecology and
Evolution of Phytotelm-Breeding Anurans, pp. 1–9. Miscellaneous Publication 193.
Museum of Zoology, University of Michigan, Ann Arbor, MI.
Nieuwkoop, P. D. and Faber, J. (1956). Normal tables of Xenopus laevis (Daudin). North-
Holland Publishing Company, Amsterdam.
Pfingsten, R. A. and Downs, F. L. (1989). Salamanders of Ohio. Bulletin of the Ohio
Biological Survey, 7, 1–350.
Reilly, S. M., Wiley, E. O., and Meinhardt, D. J. (1997). An integrative approach to
heterochrony: the distinction between interspecific and intraspecific phenomena.
Biological Journal of the Linnean Society, 60, 119–43.
Relyea, R. A. and Auld, J. R. (2005). Predator- and competitor-induced plasticity: how
changes in foraging morphology affect phenotypic trade-offs. Ecology, 86, 1723–9.
3 Morphology of amphibian larvae | 53
Ruibal, R. and Thomas, E. (1988). The obligate carnivorous larvae of the frog,
Lepidobatrachus laevis (Leptodactylidae). Copeia, 1988, 591–604.
Sammouri, R., Renous, S., Exbrayat, J. M., and Lescure, J. (1990). Développement
embryonnaire de Typhlonectes compressicaudus (Amphibia, Gymnophiona). Annales de
Sciences Naturelles-Zoologie et Biologie Animale, Paris, 11, 135–63.
Sarasin, P. and Sarasin, S. (1887–1890). Zur Entwicklungsgeschichte und Anatomie der
Ceylonesischen Blindwühle Ichthyophis glutinosus. In Ergebnisse Naturwissenschaftlicher
Forschungen auf Ceylon in den Jahren 1884–1886. Band II. C. W. Kreidel’s Verlag,
Wiesbaden.
Thibaudeau, D. G. and Altig, R. (1999). Endotrophic anurans: development and evolu-
tion. In R. W. McDiarmid and R. Altig (eds), Tadpoles, the Biology of Anuran Larvae,
pp. 170–88. University of Chicago Press, Chicago, IL.
This page intentionally left blank
4
Larval sampling
David K. Skelly and Jonathan L. Richardson
4.1 Introduction
Most amphibian species are metamorphic, and among those, the majority have
larvae that are fully aquatic (Duellman and Trueb 1986). These larvae are found
in an enormous variety of contexts ranging from bromeliads and tree holes, to
brackish pools and the largest rivers and lakes (Duellman and Trueb 1986). Add
to this mix of environments the fact that amphibian larvae differ dramatically in
their microhabitat use, from the water surface to benthic mud, and researchers
sampling larval amphibians are confronted with a non-trivial challenge of match-
ing techniques with study goals and logistic limits. Fortunately, there is a wealth
of experience that can be tapped when making decisions about where, when, and
how to sample (Shaffer et al. 1994; Olson et al. 1997).
Table 4.1 Larval amphibian sampling methods and resulting response variables
Response Metrics Sampling methods Inferences
4.1.3 Timing
Amphibian larvae range from practically immobile, yolk-laden hatchlings, to
30-cm-long salamander larvae capable of moving extremely rapidly. Deciding
when to sample during the larval period is best done in conjunction with deci-
sions about how to sample. Small larvae move slowly and are often found in shal-
low areas. These attributes can make sampling relatively straightforward using
a variety of techniques. This is something to consider if your study aims allow
flexibility in the developmental stage sampled. It also suggests that effective lar-
val sampling depends on a close knowledge of the life histories of the species to
be sampled and their developmental progress within a given year. Late snowpack
melt, droughts, and a particularly cold or warm spring can all shift the timing of
larval development substantially. In tropical climates the time of year can be far
less important than the onset of wet season rains in triggering breeding and the
timing of larval sampling.
visits, timing of visits) to determine the sensitivity of target responses (e.g. Werner
et al. 2007). There is no shortcut here; rules of thumb can be misleading resulting
in wasted effort and more collection than necessary or, more likely, incomplete
and inaccurate information (Skelly et al. 2003).
If you are new to a system, plan to learn during your initial sampling.
Intentionally trying multiple techniques and different degrees of sampling effort
will provide information that will ultimately save you time and produce more
reliable information. Before you step in the water, be prepared to spend more time
in each habitat if it is needed and to consider using multiple techniques to capture
the range of species and larval stages you intend to study.
(b)
(c)
(d)
4.2.1.2 Application
Box samplers are most effective within vernal ponds with an open water column and
simple bottom substrates such as decaying leaves. Pipe samplers were later developed
to capture the advantages of a box sampler while enabling the sampling of a wider
variety of environments including those where water is interspersed with emergent
vegetation (Skelly 1996). We are unaware of the use of area-based larval amphibian
samplers within stream environments, although the use of comparable samplers for
benthic macroinvertebrates suggests the potential for such an application.
4.2.1.3 Considerations
The major advantage of box and pipe samplers is that sampling of a known area
of wetland bottom provides direct estimates of larval density (individuals/m2). If
depth within each pipe is recorded, estimates of density per unit volume can be
determined. In either case, if wetland bottom area (or volume) is known, research-
ers have extrapolated pipe- and box-sample-based density estimates into estimates
of the size of entire larval cohorts (e.g. Werner et al. 2007).
A major disadvantage of these area-based samplers is their time-intensiveness.
Repeatedly placing and clearing a box or pipe is relatively slow and methodical
work even for experienced users. In addition, when sampling takes place in remote
areas that require hiking into and out of, box and pipe samplers can be inconveni-
ent because of their size.
nets approaching the size of some small seines (S. Cortwright, personal communi-
cation). Regardless of net size, it is common for amphibian biologists to construct
their own nets or to customize store-bought nets. Dip nets used for larval amphib-
ian sampling can have much shallower net bags and tend to require finer mesh
than those used for fish and other larger organisms. A shallow net bag facilitates
rapid processing of each sweep and can speed sampling significantly. Researchers
can also construct net dimensions that fit the structure of the environments they
sample (e.g. narrower nets may be appropriate for dipping out of marshes with
emergent vegetation, small stream pools, or tire ruts).
4.2.2.2 Application
Dip nets are used in most of the places where amphibian larvae are found. Nets
can be of lighter construction in vernal ponds (e.g. using mosquito mesh for the
net bag) compared with those used in streams and other places with abrasive
substrates.
4.2.2.3 Considerations
Dip-netting is fast, requires a minimum of equipment, and can be performed in
a wide range of environments. Nets can be constructed inexpensively to perform
well in particular contexts. On the downside, as much as any technique, dip-
netting effectively relies on experience. An experienced user can vastly outper-
form a beginner working side by side. A second disadvantage relates to population
density estimates. While dip-netting can be used by itself to estimate relative
density in terms of catch per unit effort (where effort is often measured as time
spent sampling), estimates of area- or volume-based density must rely on other
techniques or through calibration from samples collected using other methods in
the same or comparable environments (Werner et al. 2007).
4.2.3 Seine
4.2.3.1 Description
Seines have long been used to collect amphibian larvae (Routman 1984). They are
particularly effective in sampling open, deeper areas that cannot easily be sampled
using box and pipe samplers or dip nets. Seining typically requires at least two
people. One person at each end sets the seine in a line and then begins moving in
an agreed direction until a position is reached where they move together and begin
gathering the ends of the seine up and out of the water, forcing the entrapped lar-
vae down into the middle. This is done most easily if the seine is being gathered
onto the shore. Sometimes this is not possible in which case, some sort of floating
platform such as a foam bucket float can be used. When the captured sample is
62 | Amphibian ecology and conservation
concentrated in the center of the seine, it is picked up and moved onto the shore
where the contents are sorted and the larvae are processed as desired. Most seines
used for amphibian larvae are relatively small, on the order of 3–5 m long and 1 m
or so deep (Shaffer et al. 1994). Larger seines often have a bag sewn in the middle.
The bag can greatly increase sampling effectiveness, but may also require a third
person to keep it from getting caught or rolled up in vegetation.
4.2.3.2 Application
Seines are typically used in open-water areas of ponds and lakes, although large
stream pools may also be sampled using seines in conditions where flow is not too
great.
4.2.3.3 Considerations
Often seines are the only means of sampling larger amphibian larvae that live
in open-water regions of ponds (e.g. large Ambystomatid salamander larvae). A
major disadvantage of seines is that they are hard to handle in vegetation-choked
ponds and lakes often frequented by larval amphibians. In such an environment,
it can take over half an hour to sort through the vegetation and muck gathered
in a 5-min seine haul. During this sorting process, smaller amphibian larvae can
be hard to detect (and seine mesh is often coarse enough to enable their escape)
meaning that seines are most often used for large larvae. As with dip-netting,
seines are typically used to estimate catch per unit effort, although it is possible to
estimate number captured per unit area in some conditions (Shaffer et al. 1994).
4.2.4.2 Application
Litterbags have been used in stream environments where debris collects, and this
technique targets species known to utilize leaf packs. They are easy to deploy,
and quick and inexpensive to construct. They also manage to exploit an attribute
of species that makes them otherwise hard to sample. In the absence of litterbag
sampling, larvae of many stream-dwelling species are difficult to collect. They are
also able to capture more secretive or uncommon species that might be missed in a
dipnet survey. While litterbag sampling can produce species presence and relative
density data for a stream reach, it does not likely provide accurate estimates of abso-
lute population size or density, since it is unclear what exact area of the stream is
being sampled with each bag (Chalmers and Droege 2002; Waldron et al. 2003).
4.2.4.3 Considerations
Litterbags are a highly specialized sampling tool. They are useful only for species
that dwell in streams and use leaf packs. But if such a species is being targeted, the
advantages of litterbags are substantial. Litterbag size is an important consider-
ation, as medium and large bags can capture more individuals and species; how-
ever, the size of the focal stream may only accommodate a smaller bag (Waldron
et al. 2003; Talley and Crisman 2007). Additionally, samples can be biased if
potential competitors or predatory individuals colonize the bag. Researchers
can discourage predatory (usually larger) adults and species by using finer mesh
or submerging bags in deeper water, leading to a preferential capture of larvae
(Waldron et al. 2003). Finally, the utility of litterbags may vary seasonally, as the
abundance of natural leaf pack habitats can vary throughout the year, depending
on the surrounding habitat cover.
64 | Amphibian ecology and conservation
4.2.5 Trapping
4.2.5.1 Description
Traps can be an effective means to capture amphibian larvae, requiring the
researcher to simply deploy the traps and check them after a time period sufficient
to have captured resident larvae. Most traps used by amphibian researchers are
of a funnel design, which channels larvae into a large holding section that can be
accessed by the researcher to recover captured animals. Commercially available
wire minnow traps have been used in many amphibian studies (e.g. Fronzuto and
Verrell 2000; Ghioca and Smith 2007). Home-made funnel traps using plastic
bottles (e.g. 2-L plastic drinks bottles) have also been used successfully (Calef
1973; Richter 1995). Collapsible traps made of fine nylon mesh and available
commercially (Promar, Gardena, CA, USA; Figure 4.1c), have capture rates equal
to or better than traditional wire minnow traps (Adams et al. 1997; C. Pearl,
personal communication). The finer mesh of these collapsible traps allows for the
retention of much smaller larvae, and the compact size and weight make them
suitable for backcountry work. Pyramid-shaped crayfish traps (Johnson and
Barichivich 2004) are an alternative to minnow and collapsible mesh traps that
can be particularly effective when it is important for part of the trap to extend
above the water surface. In all cases, traps are deployed by dropping each in a
predetermined area of the pond with a line attached and tied to a tree or float to
make locating and retrieving it easier.
4.2.5.2 Application
Traps can be effective in capturing amphibian larvae present within many aquatic
habitats with, perhaps, the exception of fast-moving water. Trapping is particu-
larly suited to detection of species presence, and to estimate catch per unit effort
(a metric of relative abundance). Trapping is sometimes the only suitable method
in deep water, steep-sided pond basins, or frozen ponds where wading in to con-
duct sampling is not feasible. Additionally, it may be easier to sample habitats
with structurally complex bottoms or vegetation-choked areas using traps, where
seining and dip-netting may be difficult. Lastly, funnel trapping will often cap-
ture rare, secretive, or more nocturnal species not detected using other methods
conducted in a short time period and during the daytime.
4.2.5.3 Considerations
Whereas traps can be left overnight, this practice requires a sampling location to
be revisited within a short time period (usually 12–24 h) to avoid trap mortality,
especially of non-target species or life stages. It is especially important, when traps
4 Larval sampling | 65
are to be left for extended periods, to keep part of the trap above water to allow
access to water surface for trapped animals. Secondly, wire mesh size in commer-
cial minnow traps (around 6 6 mm) may be too large to effectively trap smaller
larvae, in which case sealing the trap with window screening composed of finer
mesh or the use of nylon collapsible traps can address this issue. Also, any trapping
methods and resulting data are based on an assumption of equal capture prob-
ability among individuals and populations. This can be violated if the presence
of conspecifics, competitors, or predators in the cage alters capture probabilities.
There could also be community-level biases if populations being sampled differ
in species composition. For instance, predators present in one habitat but not the
other may alter the behaviour of a target species and subsequent capture probabil-
ities. There can be other specific sampling biases in trapping certain species and
in some habitat types (e.g. playa wetlands; Ghioca and Smith 2007). Weather,
larval size and developmental stage, and resource availability can all affect capture
rates using traps (Adams et al. 1997). Some researchers use baits when trapping
amphibian larvae. However, for at least some species, it appears that baiting traps
does not increase trap effectiveness (Adams et al. 1997).
4.2.6 Mark–recapture
4.2.6.1 Description
Mark–recapture techniques are commonly used in studies of amphibian adult
populations, but can also be useful for the larval stage as well. However, rapid
growth often accompanied by a dramatic shift in body design can render marking
methods (often developed for juvenile and adults; see Chapter 8) unsuitable for
larval amphibians. Successful marking techniques for larvae include temporary
injectible organic dyes (Seale and Borass 1974) and externally staining dyes, which
typically stain amphibian larvae for less than 24 h (Pfennig 1999; Jung et al. 2002;
Harris et al. 2003), but can also slow growth rates (Travis 1981). More permanent,
yet onerous, marking techniques using paint sprays, stains, dimethyl sulfoxide,
and Super Glue (Ireland 1989) have largely been replaced by more convenient and
robust visible implant elastomer (VIE) tagging (Figure 4.1d). Passive integrated
transponder (PIT) tags may have limited use in all but the largest larval species due
to tag size (down to about 8 mm) and surgery required to implant, although the
development of smaller “injectable” PIT tags, applied using a hypodermic needle,
may expand the potential for this technique (Biomark, Boise, ID, USA).
VIE tagging appears to hold the most promise for amphibians, balancing the
ease of marking and longevity of the actual mark. Elastomer marks consist of a
silicone-based polymer material that is injected subcutaneously and cures into a
66 | Amphibian ecology and conservation
pliable and biologically inert solid (Northwest Marine Technology, Shaw Island,
WA, USA). Whereas some elastomer colors are visible to the naked eye when pre-
sent under translucent skin, fluorescent colors are often used and easily detected
using an ultraviolet light source (portable lights are available for field purposes).
Lowe (2003) used VIE to mark larvae of the spring salamander (Gyrinophilus
porphyriticus) and has indicated that marks can still be seen 10 years after mark-
ing (W. Lowe, personal communication). Fading of the elastomer does not
appear to be a common problem, although marks can migrate from the point
of injection or be lost altogether. Grant (2008) found that wood frog (Rana syl-
vatica) tadpoles can retain marks through metamorphosis and that larger marks
( 2 mm) were more likely to migrate in two larval stream salamander species.
Additionally, it was indicated that stream salamanders could be marked without
anesthesia, while wood frog tadpoles required anesthesia and also had poorer
mark retention (E. Grant, personal communication).
4.2.6.2 Application
Mark–recapture studies can be conducted in just about any habitat type, assuming
that the same method of capture is used for each sampling period. Assuming that
a sufficient proportion of the population is originally marked, mark–recapture
techniques can provide robust estimates of absolute population size, especially
when combined with robust capture–recapture estimation models (Chapter 24).
Jung et al. (2002) found that mark–recapture methods provided the most accurate
estimates of population size for two species of tadpoles in desert pool habitats.
4.2.6.3 Considerations
Regardless of technique, small amphibian larvae are relatively difficult to mark.
The VIE technique will be more useful for species with larger larvae, or at least
with individuals farther along in development. Gyrinophilus porphyriticus larvae
as small as 2 cm in total length have been marked successfully, as well as R. syl-
vatica individuals down to 2.5 cm in total length. Additionally, any substantial
loss of tags within a marked cohort can seriously bias population size estimates.
Consider this when deciding which technique to use and for what exact purpose.
4.3.2 Electroshocking
Electroshocking, developed initially to sample fish, is used primarily within
streams and rivers. However, it can stun and facilitate capture of amphibian lar-
vae as well (Brown and May 2007). It appears to be most effective for lentic species
found in slow-moving parts of rivers (Shaffer et al. 1994). Many stream-dwelling
amphibian species use retreats or bury themselves in the substrate in ways that
prevent their detection and capture even if stunned during electroshocking.
4.4 Conclusions
As in many aspects of field biology, the techniques used for sampling amphib-
ian larvae are often passed without criticism or comment from one generation
of researchers to the next. In many cases, there has been little effort to ask why
one technique should be used as opposed to its alternatives or how a given
technique may be most effectively applied. The many techniques outlined in
this chapter are connected through the references listed below to an enormous
cumulative effort to understand the most effective and efficient means to esti-
mate the presence and density of larvae. Most of the techniques require little
equipment, and that equipment is typically relatively inexpensive. Neither are
the techniques difficult to master. Collectively, this means that there is little
68 | Amphibian ecology and conservation
reason not to try multiple techniques and to calibrate and understand the con-
sequences of altering the timing and intensity of sampling. The modest effort
to do so will greatly increase the reliability of the information gathered and, in
all probability, lead to unforeseen insights into the biology of the species being
studied.
4.5 Acknowledgments
We thank S. Cortwright, M. McPeek, E. Werner, and H. Wilbur for teaching us
about larval amphibian sampling. M. Adams, E. Grant, B. Hossack, W. Lowe,
and C. Pearl provided helpful insight and details into the techniques they use.
4.6 References
Adams, M.J., Richter, K.O., and Leonard, W.P. (1997). Surveying and monitoring
amphibians using aquatic funnel traps. In D.H. Olson, W.P. Leonard, and R. Bury
(eds), Sampling Amphibians in Lentic Habitats: Methods and Approaches for the Pacific
Northwest: Northwest Fauna Number 4, pp. 47–54. Society for Northwestern Vertebrate
Biology, Olympia, WA.
Brown, L.R. and May, J.T. (2007). Aquatic vertebrate assemblages of the Upper Clear
Creek Watershed, California. Western North American Naturalist, 67, 439–51.
Calef, G.W. (1973). Natural mortality of tadpoles in a population of Rana aurora. Ecology,
54, 741–58.
Chalmers, R.J. and Droege. S. (2002). Leaf litter bags as an index to populations of north-
ern two-lined salamanders (Eurycea bislineata). Wildlife Society Bulletin, 30, 71–4.
Duellman, W.E. and Trueb, L. (1986). Biology of Amphibians. John Hopkins University
Press, Baltimore, MD.
Freidenburg, L.K. (2003). Spatial Ecology of the Wood Frog (Rana sylvatica). PhD
Dissertation, University of Connecticut, Storrs, CT.
Fronzuto, J. and Verrell, P. (2000) Sampling aquatic salamanders: tests of the efficiency
of two funnel traps. Journal of Herpetology, 34, 146–7.
Ghioca, D.M. and Smith, L.M. (2007). Biases in trapping larval amphibians in playa
wetlands. Journal of Wildlife Management, 71, 991–5.
Grant, E.H.C. (2008). Visual implant elastomer mark retention through metamorphosis
in amphibian larvae. Journal of Wildlife Management, 72, 1247–52.
Gunzburger, M.S. (2007). Evaluation of seven aquatic sampling methods for amphibians
and other aquatic fauna. Applied Herpetology, 4, 47–63.
Harris, R.N, Alford, R.A., and Wilbur, H.M. (1988). Density and phenology of
Notophthalmus viridescens dorsalis in a natural pond. Herpetologica, 44, 234–42.
Harris, R.N., Vess, T.J., Hammond, J.I., and Lindermuth, C.J. (2003). Context-
dependent kin discrimination in larval four-toed salamanders Hemidactylium scutatum
(Caudata: Plethodontidae). Herpetologica, 59, 164–77.
4 Larval sampling | 69
Travis, J. (1981). The effect of staining on the growth of Hyla gratiosa tadpoles. Copeia,
1981, 193–6.
Waldron, J.L., Dodd, Jr., C.K., and Corser, J.D. (2003). Leaf litterbags: factors affecting
capture of stream-dwelling salamanders. Applied Herpetology, 1, 23–36.
Werner, E.E., Skelly, D.K., Relyea, R.A., and Yurewicz, K.L. (2007). Amphibian species
richness across environmental gradients. Oikos, 116, 1697–1721.
5
Dietary assessments of larval amphibians
Matt R. Whiles and Ronald Altig
5.1 Background
Diet analyses of larval amphibians provide information on foraging patterns,
nutritional requirements, and trophic interactions in aquatic food webs, infor-
mation that is critical for successful conservation and management. Numerous
amphibian species around the globe are declining or presumed extinct (Stuart
et al. 2004). Thus, there is an urgent need for detailed information on larval
diets so that we can understand whether food-related factors are limiting wild
populations and facilitate successful rearing of species in captivity.
Some species of all three orders of living amphibians have free-living, feeding,
larval forms. Larval caecilians and salamanders, both of which occur in a variety
of lentic and lotic habitats, are all carnivores that eat small aquatic organisms
by suction-feeding, and their diets and trophic status are thus relatively easily
assessed. Anuran larvae (tadpoles) occur in almost every conceivable type of
freshwater habitat and show great diversity in feeding modes. As such, morpho-
logical diversity of the oral structures, which glean materials from substrata,
and the buccopharyngeal apparatus, which selectively captures food particles,
of tadpoles is large.
Whereas a few tadpoles are macrocarnivores (e.g. Lepidobatrachus,
Leptodactylidae), most either rasp materials associated with substrata such as
biofilms, periphyton, and deposited organic particles (e.g. ranids and many
hylids), harvest naturally suspended particles in the water column (e.g. micro-
hylids, pipids, and rhinophrynids), capture particles in the surface film (i.e.
neustonic tadpoles with umbelliform oral structures), or consume conspecific
or heterospecific eggs. Most of these feeding modes need further study (Altig
et al. 2007). Although many tadpoles are classified as general herbivores or
detritivores, they likely obtain a fair amount of energy and nutrients from more
72 | Amphibian ecology and conservation
specimens. In cases where fatty acid or isotope analyses are also planned, speci-
mens may need to be frozen rather than placed in fluid preservatives. Freezing
leads to distortion of some materials, and removal of guts and processing or pres-
ervation should take place immediately after specimens are thawed.
is very informative. Estimates of caloric contents for many types of prey are
available (e.g. Cummins and Wuychek 1971; Rodgers and Qadri 1977), but
these estimates tend to be quite variable. Depending on the level of precision
desired, analyses such as bomb calorimetry should be performed on actual prey
items or other materials from the guts or specimens collected from the same
habitats at the same time. As with DM and AFDM, size/caloric-content rela-
tionships can be developed.
While collection and identification of gut contents from salamanders and
caecilians is relatively straightforward, tadpoles present a challenge because
their long guts contain mixtures of materials of different digestibility that are at
different stages of digestion. In this case, the most recently ingested materials
from the foregut should be the foci of analyses. This is generally accomplished
by removing a section of the foregut that is a consistent proportion of the total
gut length (e.g. the anterior quarter of the gut) and preparing it for examination
under a microscope (for identification of individual food particles) or for ana-
lyses of caloric content, fatty acid profiles, or isotopic composition (see methods
for fatty acids and isotopes below). Obviously, method of collection and preser-
vation can limit the types of analyses that can be performed on gut contents.
For identification of foregut contents, materials are generally rinsed from
the gut, slide-mounted, and examined under a microscope. One rather simple
and effective method is to place materials in glycerin and gently homogenize
them. If needed, water can be added and the materials sonicated to further
break up contents. Depending on the abundance of particles, either the entire
sample or a subsample is filtered with low vacuum on to a 0.45 μm gridded
membrane filter to obtain a reasonable dispersion of particles across the filter
(e.g. 10–20 particles/grid). Filters should be allowed to dry at ambient tem-
perature for 2–3 h to remove excess water, and then two ot three drops of Type
A immersion oil, which clears the filter, are added. After approximately 24 h
the filters will clear and can then be placed on a slide and covered with a cover-
slip, which is sealed with enamel nail polish. Slide-mounted materials are gen-
erally subsampled (e.g. some fi xed number of fields of view examined, a fi xed
number of grids examined, or all particles along transects are identified and
measured). Measurements of individual particles can be made using methods
ranging from an ocular micrometer to sophisticated image analysis systems.
Data on identified food types are generally expressed as area or converted to
mass, or, in the case of algae, biovolume (Lowe and LaLiberte 2006; Steinman
et al. 2006). There are many variations for mounting and analyzing gut con-
tents (for recent examples using various taxonomic groups, see Evans-White et al.
2003; Ranvestel et al. 2004; Rosi-Marshall and Wallace 2002). Regardless of
5 Dietary assessments of larvae | 77
modifications, the key points are to avoid damaging materials beyond recogni-
tion and to obtain an evenly dispersed, representative sample.
Masses of materials in guts can be estimated with methods ranging from
simply weighing materials to using size/mass relationships (e.g. body length/
mass relationships for prey; see above section on salamander and caecilian guts)
either developed by the investigator or gleaned from the literature. For direct
weighing, DM or AFDM estimates are best. Dry mass estimates are generally
developed by drying materials at 55–60°C to constant mass. For AFDM, dried
materials are weighed, then ashed in a furnace (≈500°C) for 1–2 h, and weighed
again; mass of remaining ash is subtracted from the original dry mass to obtain
AFDM. In either case, samples should be allowed to cool in a dessicator before
weighing so that moisture is not absorbed in the cooling process. Actual drying
and ashing times will vary with sample size. Small amounts of materials can
be processed in this manner on small pieces of heavy duty aluminum foil or
glass fiber filters. Development of wet mass/DM or AFDM relationships can
streamline procedures and decrease the number of samples that are destroyed.
Likewise, if caloric equivalents are derived through calorimetry, similar rela-
tionships can be developed.
Gut content data expressed as mass or area (or in some cases biovolume for
algae (see Lowe and LaLiberte 2006) can be combined with estimates of assimi-
lation efficiencies for more accurate assessments of trophic status. Procedures
vary, but a simple approach is to express food types as percent of total gut con-
tents and multiply each by its respective assimilation efficiency. Resultant values
for the food types are then summed and the percentage contribution of each to
assimilated materials can be calculated. If caloric content is estimated, indices of
prey or food-type importance can be calculated using basically the same basic
procedure; masses of each food type in the gut are multiplied by their corre-
sponding caloric density and percentage caloric contribution of each food type is
the total number of calories attributable to a given food type divided by the total
caloric content in the gut. Regester et al. (2008) provide examples of these types
of analyses and related procedures applied to pond-breeding ambystomatid com-
munities. Examples using various freshwater omnivores can be found in Bowen
et al. (1995) and Evans-White et al. (2003). Regester et al. (2008) and Skelly and
Golon (2003) provide examples of methods for estimating assimilation efficien-
cies of amphibian larvae.
In cases where algae are obviously consumed, analyses of photosynthetic pig-
ments may be performed on gut contents to estimate algal biomass. For this pro-
cedure, specimens and/or their gut contents must be frozen immediately upon
collection and kept dark and frozen prior to analyses. Chlorophyll a, which can
78 | Amphibian ecology and conservation
be used to estimate algal biomass, can be extracted from samples of gut contents
with an organic solvent and concentration is then estimated with spectropho-
tometry, fluorometry, or high-performance liquid chromatography. Steinman
et al. (2006) provide a review of methods to estimate algal biomass via photo-
synthetic pigment analyses. Although informative to a degree, pigment analyses
alone cannot account for the digestibility of algal materials, and some tadpoles
will re-ingest feces containing undigested algal materials. The ability of tad-
poles to digest algal materials can be assessed through microscopic examination
of cells in feces (e.g. whether full chloroplasts are present) or culturing algae
from feces can identify the quantity and types of materials that are still viable
after passing through guts (Peterson and Boulton 1999). Simple counts of prey
items or food types in guts may not provide reliable information on trophic sta-
tus, but can be valuable for assessing feeding preferences and interactions with
other consumers. A variety of indices, including Horn’s index and Morisita’s
index, can be applied to gut-content data to quantify diet overlap among species
(see Brower et al. 1998; Krebs 2001; Chapter 18). Likewise, if data on the abun-
dance of prey and/or food type in the environment are also available, feeding
selectivity can be assessed and quantified using a variety of available indices.
Relatively simple and commonly used indices include Ivlev’s index of electivity
(Ivlev 1961) and the selectivity index (Chesson 1983).
can provide valuable information on assimilatory diets and methods for their
analyses are standardized. Stable-isotope analyses focus on ratios of heavier
(e.g. 15N and 14C) and lighter isotopes (14N and 13C). Ratios are expressed as ,
which is parts per thousand (‰) deviation from a standard, calculated as 13C
or 15N [(Rsample Rstandard)/Rstandard] 1000 where R 13C/12C or 15N/14N.
Common standards are Pee Dee belemnite limestone for C and atmospheric N
(‰ value set to 0). A positive (N) or less negative (C) value indicates the sample
is enriched with the heavy isotope.
Carbon isotope signatures of consumers are generally similar to those of their
food source or sources (i.e. you are what you eat), which can have 13C val-
ues ranging from slightly higher than atmospheric CO2 (−8‰) to −40‰ for a
variety of reasons that influence degree of fractionation (see Peterson and Fry
1987; Fry 2006; Hershey et al. 2006). In particular, in streams, water velocity
can influence carbon isotopic compositions of foods such as algae (Finlay et al.
1999). Differences in C3 and C4 photosynthetic pathways result in differences
among some plants and their detritus. Likewise, there are often measurable dif-
ferences in carbon isotope ratios between terrestrial and aquatic autotrophs and
their detritus.
Nitrogen is more often used to assess trophic position, as 15N shows a fairly
consistent change with each trophic step (≈3.4‰ increase in 15N with each
trophic step up from primary consumer to predator; Vander Zanden and
Rasmussen 1999; Post 2002). This fairly predictable fractionation results
from the tendency for organisms to excrete more of the lighter nitrogen isotope
than the heavy one. However, fractionation can vary and there is evidence that
the increase in 15N with each trophic step is considerably less than 3.4‰ in
some tropical stream food webs (Kilham and Pringle 2000). In general, basal
resources such as periphyton and detritus in freshwater habitats will have 15N
values ranging from near 0 to 3‰, primary consumers range from 3 to 6‰,
and predators from 6 to 12‰, but values can vary considerably for a variety
of reasons (Fry 2006). Whiles et al. (2006) and Verburg et al. (2007) provide
examples of isotopic analyses of stream food webs with abundant and diverse
tadpole assemblages.
Most studies examine naturally occurring isotopic ratios (natural abundance
studies) in consumer tissues and food resources, but additions of enriched
materials (e.g. tracer studies using 15N enriched ammonium or nitrate or 13C
enriched acetate) are also used, particularly when natural isotopic ratios among
food types are similar. Tracer studies can be informative for assessing diets and
trophic interactions, as well as examining roles of consumers in biogeochem-
ical cycling, because materials (e.g. N or C) are followed from basal resources
80 | Amphibian ecology and conservation
through top consumers. Stable-isotope tracer additions have been used to exam-
ine carbon cycling through consumers in streams (Hall 1995), nitrogen cycling
in stream food webs (Hall et al. 1998), and movement of nutrients from streams
to riparian habitats via consumers (Sanzone et al. 2003). Tracer studies could
be particularly informative for assessing roles of amphibians in freshwater eco-
system processes, as well as material and energy exchanges between aquatic and
terrestrial systems.
5.5.2 Procedures
Stable-isotope analyses can be performed on entire organisms or, in the case of
larger individuals, muscle tissue (generally 1–2 g wet weight). While muscle tissue
is the usual focus for larger animals, blood and other tissues such as liver or skin
are sometimes examined because they have different incorporation and turn-
over rates, and can thus be used to assess temporal patterns of food availability
and assimilation (Dalerum and Angerbjorn 2005). For example, while isotopic
analyses of muscle tissues may reflect assimilation over a period of months, liver
or blood plasma, which turnover much more rapidly, can reflect a period of days
or weeks. This isotopic memory is an important consideration, as diets of some
amphibians can change considerably through development and with changing
resource availability.
Muscle tissues from large individuals can be sampled non-lethally using ster-
ile biopsy punches, which are available from many surgical supply companies. If
biopsies are collected, wounds should be treated with an appropriate antibiotic
before releasing the specimen. Live amphibians and prey items that are to be proc-
essed in their entirety should be allowed to clear their guts as much as possible
by placing them in filtered water from the same habitat for approximately 12 h.
Alternatively, gut tracts can be removed before processing, but this can be a tedi-
ous process for small specimens. The best methods for preservation of samples for
isotopic analyses are freezing or drying at low temperature (55–60°C to constant
mass). In the field, samples should be placed on ice or in liquid nitrogen and then
transferred to a freezer (−80°C) or drying oven when available. Dried samples
should be stored in a desiccant such as drierite or silica. Cross-contamination
must be avoided throughout sampling and processing of all materials.
Comprehensive analyses of food webs require sampling of all potential food
items that are available for larval amphibians or their prey. Sampling of prey,
algae, biofilms, detritus, and other potential resources simply requires collection
of representative samples. These samples can be collected using standard proced-
ures (e.g. Hauer and Lamberti 2006), but care must be taken to avoid cross-
contamination. Sampling of fine particles from substrates or the water column
5 Dietary assessments of larvae | 81
can be collected and processed on glass-fiber filters. As with any other sampling,
replicates should be collected to account for natural variability. In flowing waters,
areas of different velocities should be sampled, as current velocity can alter the
carbon isotope compositions of periphyton (Finlay et al. 1999).
Prior to isotopic analyses, all samples are dried to constant weight at 55–60°C,
ground into a fine powder, weighed, and packed into tin capsules. Samples of
particulates on glass-fiber filter can either be carefully scraped from the filters,
or a portion of the filter, with sampled material on it, can be packed into the
tin capsule. While muscle tissue is the usual focus for larger animals, for most
prey items (e.g. aquatic invertebrates) the entire organism is processed, or, in
the case of very small prey, individuals are combined. Samples for N isotope
analysis should contain at least approximately 100 μg of DM of N; samples
for C isotope analysis should contain approximately 2 mg of DM of C. Sample
sizes for analysis of both C and N isotopes vary with C:N ratio of the sample,
with larger sample sizes needed (e.g. 10–60 mg) for materials with high C:N
ratio (e.g. sediments with low %N). Samples of basal resources and sediments
from calcareous regions should be processed to remove carbonates before ana-
lysis. Carbonates can easily be removed by acidification via exposure to HCl
vapors (Yamamuro and Kayanne 1995). Isotopic analyses are performed using
a continuous-flow isotope-ratio mass spectrometer.
Analyses and interpretation of stable-isotope data range from simply com-
paring values of resources and consumers to assess relative contributions, to
quantifying trophic status (Vander Zanden et al. 2000; Post 2002) or niche
breadth (Layman et al. 2007). Depending on the nature of the system, stable-
isotope analyses will sometimes reveal clear relationships among consumers
and resources. In other cases, environmental complexities and high degrees
of omnivory can confound results. In cases where multiple food sources are
likely contributing to the isotopic composition of a consumer, as would be the
case with most tadpoles, mixing models can be used to assess relative contribu-
tions of different foods. Simple two-source mixing models can be used to esti-
mate relative contributions of two resources (e.g. periphyton and detritus; see
Hershey et al. 2006), whereas more complex models are necessary when more
than two food sources are likely.
terms of their importance for growth and development. Given their nutritional
value, that they are conserved, and that various resources and consumers have
unique combinations of fatty acids, fatty acid profiles can be good indicators
of diet and trophic status. Analyses of fatty acid profiles have been successfully
used to assess food web structure in marine, lake, and stream systems (Muller-
Navarra et al. 2000; Stübing et al. 2003; Sushchik et al. 2003; as well as aquatic–
terrestrial food web linkages, Koussoroplis et al. 2008). In some cases, fatty acid
analyses can provide detailed taxonomic information, including specific types
of algae assimilated (e.g. Napolitano et al. 1997).
Samples of muscle tissues (large amphibians), whole organisms (small
amphibians, prey items), and basal resources (algae, detritus) can be collected
in the same manner as described above for isotopic analyses, except that all
materials should be frozen at −80°C as quickly as possible. Moisture content
of materials is quantified with freeze drying so that lipids are not altered or
volatilized. Lipids are extracted from homogenized materials using chloroform/
methanol solvent extraction (Bligh and Dyer 1957). Analyses involve separation
of polar and non-polar lipids with a solid-phase extraction system, conversion
into fatty acid methyl esters, and separation using a gas chromatograph with a
flame ionization or mass spectrometer detector.
Discriminant function analysis and similar approaches can be used to assess
trophic interactions among sampled resources and consumers (Budge et al.
2002). As with isotopic analyses, resolution will vary among systems, and in
situations where trophic interactions are complex multiple approaches (e.g. fatty
acid analyses combined with isotopic analyses and/or gut content analyses) will
obviously produce more robust results.
5.6 Summary
Understanding the food habits of larval amphibians is actually the first step in
understanding the interactions of amphibians and the communities in which
they live and ultimately their ecological significance (Altig et al. 2007). Given the
current status of many amphibian species and populations, and the likelihood
that past and ongoing declines have ecosystem-level consequences (Ranvestel
et al. 2004; Whiles et al. 2006), there is urgent need for detailed, quantitative
information on the myriad ecological roles of larvae. The methods reviewed
herein provide the basic tools for assessing diet, nutritional ecology, and trophic
interactions. Choosing the proper technique ultimately depends on the research
question and resource limitations. Because each approach has inherent limita-
tions, we recommend combined approaches.
5 Dietary assessments of larvae | 83
5.7 References
Altig, R., Whiles, M. R., and Taylor, C. L. (2007). What do tadpoles really eat? Assessing
the trophic status of an understudied and imperiled group of consumers in freshwater
habitats. Freshwater Biology, 52, 386–95.
Benke, A. C., Huryn, A. D., Smock, L. A., and Wallace, J. B. (1999). Length-mass rela-
tionships for freshwater macroinvertebrates in North America with particular reference
to the southeastern United States. Journal of the North American Benthological Society
18, 308–43.
Bligh, E.G. and Dyer, W. J. (1959). A rapid method of total lipid extraction and purifica-
tion. Canadian Journal of Biochemistry and Physiology 37: 911–917.
Bottrell, H. H., Duncan A., Gliwicz, Z. M., Grygierek, E., Herzig A., Hillbricht-Ilkowska
A., Kurasawa H., Larsson P., and Weglenska, T. (1976). A review of some problems in
zooplankton production studies. Norwegian Journal of Zoology, 24, 419–56.
Bowen, S. H., Lutz, E. V., and Ahlgren, M. O. (1995). Dietary-protein and energy as
determinants of food quality—trophic strategies compared. Ecology, 76, 899–907.
Brower, J. E., Zar, J. H., and von Ende, C. N. (1998). Field and Laboratory Methods for
General Ecology. McGraw-Hill, Boston, MA.
Budge, S. M., Iverson, S. J., Bowen, W. D., and Ackman, R. G. (2002). Among- and
within-species variability in fatty acid signatures of marine fish and invertebrates on the
Scotian Shelf, Georges Bank, and southern Gulf of St. Lawrence. Canadian Journal of
Fisheries and Aquatic Sciences, 59, 886–98.
Chesson, J. (1983). The estimation and analysis of preference and its relationship to for-
aging models. Ecology, 64, 1297–1304.
Culver, D. A., Boucherle, M. M., Bean, D. J., and Fletcher, J. W. (1985). Biomass of
freshwater crustacean zooplankton from length-weight regressions. Canadian Journal
of Fisheries and Aquatic Sciences, 42, 1380–90.
Cummins, K. W. and Wuycheck, J. C. (1971). Caloric equivalents for investigations
in ecological energetics. Mitteilungen Internationale Vereinigung für Theoretische und
angewandte Limnologie, 18, 1–158.
Dalerum, F. and Angerbjorn, A. (2005). Resolving temporal variation in vertebrate diets
using naturally occurring stable isotopes. Oecologia, 144, 647–58.
Duellman, W. E. and Trueb, L. (1986). Biology of Amphibians. McGraw-Hill, New York.
Dumont, H. J., Van de Velde, I., and Dumont, S. (1975). The dry weight estimate of
biomass in a selection of Cladocera, Copepoda, and Rotifera from the plankton, per-
iphyton and benthos of continental waters. Oecologia, 19, 75–97.
Dünker, N., Wake, M. H., and Olson, W. M. (2000). Embryonic and larval develop-
ment in the caecilian Icthyophis kohtaoensis: a staging table. Journal of Morphology, 243,
3–34.
Evans-White, M. A., Dodds, W. K., and Whiles, M. R. (2003). Ecosystem significance
of crayfishes and central stonerollers in a tallgrass prairie stream: functional differences
between co-occurring omnivores. Journal of the North American Benthological Society,
22, 423–41.
Finlay, J. C., Power, M. E., and Cabana, G. (1999). Effects of water velocity on algal car-
bon isotope ratios: implications for river food web studies. Limnology and Oceanography,
44, 1198–1203.
84 | Amphibian ecology and conservation
Post, D. M. (2002). Using stable isotopes to estimate trophic position: models, methods,
and assumptions. Ecology, 83, 703–18.
Ranvestel, A. W., Lips, K. R., Pringle, C. M., Whiles, M. R., and Bixby, R. J. (2004).
Neotropical tadpoles influence stream benthos: evidence for the ecological conse-
quences of decline in amphibian populations. Freshwater Biology, 49, 274–85.
Regester, K. J., Whiles, M. R., and Lips, K. R. (2008). Variation in the trophic basis of
production and energy flow associated with emergence of larval salamander assem-
blages (Ambystomatidae) among forest ponds. Freshwater Biology 53, 1754–67.
Rodgers, D. W. and Qadri, S. U. (1977). Seasonal variations in calorific values of some lit-
toral benthic invertebrates of the Ottawa River, Ontario. Canadian Journal of Zoology,
55, 881–4.
Rogers, L. E., Hinds W. T., and Buschom, R. L. (1976). A general weight vs. length rela-
tionship for insects. Annals of the Entomological Society of America, 69, 387–9.
Rosi-Marshall, E. J. and Wallace, J. B. (2002). Invertebrate food webs along a stream
resource gradient. Freshwater Biology, 47, 129–41.
Sanzone, D. M., Meyer, J. L., Marti, E., Gardiner, E. P., Tank, J. L., and Grimm, N. B.
(2003). Carbon and nitrogen transfer from a desert stream to riparian predators.
Oecologia, 134, 238–50.
Skelly, D. K. and Golon, J. (2003). Assimilation of natural benthic substrates by two spe-
cies of tadpoles. Herpetologica 59, 37–42.
Smith, D. G. (2001). Pennak’s Freshwater Invertebrates of the United States, Porifera to
Crustacea. John Wiley and Sons, New York.
Solé, M., Beckmann, O., Pelz, B., Kwet, A., and Engels, W. (2005). Stomach-flushing for
diet analysis in anurans: an improved protocol evaluated in a case study in Araucaria
forests, southern Brazil. Studies on Neotropical Fauna and Environment, 40, 23–8.
Solomon, C. T., Flecker, A. S., and Taylor, B. W. (2004). Testing the role of sediment-
mediated interactions between tadpoles and armored catfish in a neotropical stream.
Copeia, 2004, 610–16.
Steidl, R. J., Hayes, J. P., and Schauber, E. (1997). Statistical power analysis in wildlife
research. Journal of Wildlife Management, 61, 270–9.
Steinman, A. D., Lamberti, G. A., and Leavitt, P. R. (2006). Biomass and pigments of
benthic algae. In F. R. Hauer and G. A. Lamberti (eds), Methods in Stream Ecology,
pp 352–79. Academic Press, San Diego, CA.
Stuart, S. N., Chanson, J. S., Cox, N. A., Young, B. E., Rodrigues, A. S. L., Fischman, D. L.,
and Waller, R. W. (2004). Status and trends of amphibian declines and extinctions
worldwide. Science, 306, 1783–6.
Stübing, D., Hagen, W., and Schmidt, K. (2003). On the use of lipid biomarkers in mar-
ine food web analyses: an experimental case study on the Antarctic krill, Euphausia
superba. Limnology and Oceanography, 48, 1685–1700.
Sushchik, N. N., Gladyshev, M. I., Moskvichova, A. V., Makhutova, O. N., and
Kalachova, G. S. (2003). Comparison of fatty acid composition in major lipid classes
of the dominant benthic invertebrates of the Yenisei River. Comparative Biochemistry
and Physiology, 134B, 111–22.
Thorpe, J. H. and Covich, A. P. (2001). Ecology and Classification of North American
Freshwater Invertebrates. Academic Press, San Diego, CA.
86 | Amphibian ecology and conservation
Vander Zanden, M. J. and Rasmussen, J. B. (1999). Primary consumer 15N and 13C and
the trophic position of aquatic consumers. Ecology, 80, 1395–1404.
Vander Zanden, M. J., Shuter, B. J., Lester, N. P., and Rasmussen, J. B. (2000). Within-
and among-population variation in the trophic position of a pelagic predator, lake
trout (Salvelinus namaycush). Canadian Journal of Fisheries and Aquatic Sciences, 57,
725–31.
Verburg, P., Kilham, S. S., Pringle, C. M., Lips, K. R., and Drake, D. L. (2007). A sta-
ble isotope study of a neotropical stream food web prior to the extirpation of its large
amphibian community. Journal of Tropical Ecology, 23, 643–51.
Whiles, M. R., Lips, K. R., Pringle, C. M., Kilham, S. S., Bixby, R. J., Brenes, R., Connelly, S.,
Colon-Gaud, J. C., Hunte-Brown, M., Huryn, A. D. et al. (2006). The effects of
amphibian population declines on the structure and function of Neotropical stream
ecosystems. Frontiers in Ecology and Environment, 4, 27–34.
Yamamuro, M. and Kayanne, H. (1995). Rapid direct determination of organic-carbon
and nitrogen in carbonate-bearing sediments with a Yanaco Mt-5 Chn Analyzer.
Limnology and Oceanography, 40, 1001–5.
6
Aquatic mesocosms
Raymond D. Semlitsch and Michelle D. Boone
6.1 Introduction
We have often told our students that any experiment can be easily criticized for
lack of realism because all experiments, whether in the laboratory or field, by
design are contrived by humans and are not natural. However, in spite of such
criticism the more challenging issue is coming up with new venues or innova-
tive designs to answer important ecological or conservation questions and to
balance trade-offs. As amphibian ecology has become more experimental, we
have been confronted with numerous trade-offs concerning the choice of venue,
design, realism, and replication. Each choice is often coupled with benefits as
well as limitations. As in many rapidly developing fields, criticism is an expected
and healthy process that pushes investigators to think about their choices and
develop solutions. The use of mesocosms in studies of aquatic ecology across
an array of taxonomic groups has been central in that debate (e.g. Kimball and
Levin, 1985; Carpenter 1996; Schindler 1998). Within the field of amphibian
biology, a number of reviews also have stimulated a healthy discussion on the use
of experiments (Hairston 1989a, 1989b), and in particular, use of mesocosms
(e.g. Jaeger and Walls 1989; Rowe and Dunson 1994; Skelly and Kiesecker
2001; Skelly 2002).
The purpose of our chapter is to provide background on the use of aquatic
mesocosms in ecological and conservation studies of amphibians. We describe
the history, various types, set-ups, and designs and present selected case studies
to emphasize their versatility. We also provide examples and comments on limi-
tations of the technique and strengths of inference that come with the results.
Our goal is to provide readers with an effective technique to answer a range of
ecological and conservation questions.
88 | Amphibian ecology and conservation
Full-Sun Treatment
High-Shade Treatment
Low-Shade Treatment
Fig. 6.2 Examples of different cattle tanks and pools used to rear amphibian larvae.
90 | Amphibian ecology and conservation
that you typically cannot alter in the laboratory. Secondly, if you have evaluated
a mechanism within or between organism(s), or developed a theoretical model,
then a logical next step is to see if your predictions hold in a more complex and nat-
ural environment; mesocosm studies are a logical next step. Thirdly, if you want to
evaluate single factors in isolation, a mesocosm may be ideal. Fourthly, manipula-
tions in mesocosms can help you eliminate less important factors and find system
drivers, so that you can design a better, more efficient large-scale field study.
If you need a short-term answer to a basic experimental question, then a labora-
tory experiment can offer an expedient answer: can trematodes infect limb buds?
What concentration of an insecticide is lethal to tadpoles or zooplankton? Do
overwintered bullfrog tadpoles prey upon recently hatched tadpoles? If your ques-
tion is site- or habitat-specific or requires a vast area to encompass the entire ecosys-
tem, then field experiments may be the best way to go: can ponds in city parks and
golf courses support native fauna? What effects do forest-management practices
have on amphibian communities? Studies in mesocosms provide a useful surro-
gate for the field and allow us to evaluate single and multiple factors that may be
important at a larger scale, which allows us to better design larger field studies and
provides insight into the system: do changes in food webs from eutrophication
cause increases in number of trematodes and in the number of infected amphib-
ians? Does a sublethal pesticide used on golf courses cause changes in aquatic
community dynamics? Although mesocosms are often referred to as a “black box,”
suggesting that the mechanisms underlying the outcome are not clear, additional
studies can be done to discern mechanisms by behavioral observations, periodic
sampling of covariates (e.g. temperature, primary production, dissolved carbon),
or with a paired laboratory study (e.g. prey palatability).
Mesocosms are useful in addressing a variety of questions in ecology, evolu-
tion, behavior, and conservation (see case studies below). We surveyed the litera-
ture from 2004 to 2007 through the Ecological Society of America (including
the journals Ecology, Ecological Applications, Frontiers in Ecology and Evolution,
and Ecological Monographs), Oecologia, Evolution, Animal Behavior, and Behavior
by searching the terms “amphibian” or “frog” or “salamander” and then select-
ing all studies that used mesocosms. A total of 30 articles using mesocosms or
enclosures were found only in the journals Oecologia or of the Ecological Society
of America. The studies addressed a range of questions in ecology (e.g. species
interactions, population biology, community structure), evolution (e.g. pheno-
typic plasticity), behavior (e.g. oviposition selection), and conservation (e.g.
effects of pathogens, habitat fragmentation, pesticides). Whereas mesocosms
have been historically used for basic questions of ecology and evolution, they are
becoming more frequently used in the field of conservation and have enormous
92 | Amphibian ecology and conservation
(a)
14
Mesocosms
12 Enclosures
10
Number of studies
0
<120 200–450 750–1300 >1500
Volume (L)
(b)
12
10
Number of studies
0
0 0.1–0.3 0.875–1.8
Leaf litter (kg)
Fig. 6.3 Abundance of mesocosm studies conducted: (a) using different volumes and
(b) with different amounts of leaf litter (see text for methods and results).
6 Aquatic mesocosms | 93
purchased in galvanized steel (which must be sealed with an epoxy paint or lined)
and polyethylene plastic (Behlen or Rubbermaid produce tanks in blue or black).
The polyethylene ponds are most commonly used today because they are long-
lasting (10 years), are not prone to rusting, and do not require sealing as metal tanks
do. Wading pools are useful and cheap but short-lived (1–2 years) and usually made
of PVC/vinyl, which can leach phthalates, chemicals that are associated with endo-
crine disruption. Water tests found that Behlen polyethylene tanks did not leach
estrogenic compounds (S.I. Storrs, personal communication). However, wading
pools can be especially useful to address short-term behavioral questions related to
factors that may influence oviposition (Vonesh and Buck 2007) or plasticity in tad-
pole morphological development associated with predators and competition which
may occur in a matter of weeks (Relyea 2004; Relyea and Auld 2005).
Field caging studies enable us to study a discrete location or habitat while
moving toward greater realism. Field cages can range in size from small to large.
In our survey, most of the cages were relatively small (2.4–120 L; Garcia et al.
2004; Griffis-Kyle and Ritchie 2007; Ireland et al. 2007) or large (pond size or
sections; Scott 1990; Loman 2004; Boone et al. 2004). Aquatic field cages are
often made of fine screen mesh, such as fiberglass window screen, mosquito net-
ting, or plastic meshes, which are exposed to the larger matrix allowing for water
exchange and influx of aquatic life smaller than the mesh size. Alternatively,
field cages can also be made of polyethylene plastics that will prevent the flow
of materials into the enclosure after initial set-up and also prevent things like
pesticides or predators from moving into or out of the enclosure. These field
cages whether made of permeable or impermeable material can be attached to
untreated lumber into a box or rectangle to be placed in a pond (Figure 6.2). If
the screen, mesh, or fabric has rigidity, it is also possible to use it to construct
circular cages to hold animals (Figure 6.2). Alternatively, “bags” can be sewn or
fabricated that can be supported with an external structure like multiple stakes,
drinking water pipe, and/or floatation devices (Figure 6.2).
The advantage of field cages is that you can study sites of particular interest
and ask questions in the relevant environment giving your study greater realism
and the ability to manipulate factors of interest. In contrast, field cages often do
not represent independent observations (unlike wading pools or tanks) given
that there is the potential for exchange among enclosures. For example, if a
researcher was examining how a fish predator may influence growth and devel-
opment of tadpoles, the presence of fish in the pond or the presence/absence of
fish as a treatment could influence all the enclosures if there was water exchange.
However, these potential problems can be eliminated or minimized with choice
of mesh and some design considerations for the cages.
6 Aquatic mesocosms | 95
the type of natural system you are attempting to represent (e.g. forested pond,
grassy pond, or marsh) is the best guide to selecting the appropriate nutrient
base to add.
It is important to establish the primary components of an aquatic food web in
mesocosm tanks in order for them to be self-sustaining. A complex community
of both phyto- and zooplankton is critical to help balance nutrient exchange,
bacterial growth, and nitrogenous waste production as well as serve as a food
source for amphibian larvae. Temporary wetlands are often an ideal source of
inoculum for mesocosms because there will be fewer insect predators and no
fish. All inoculum added to mesocosms should be strained to eliminate detri-
mental insect or fish larvae, and eggs or larvae of non-target amphibian species.
Several additions over time and from varying natural ponds will help ensure
adequate diversity and complexity.
the latter case would constitute “pseudoreplicates” (Hurlbert 1984) and would
increase the probability of making a type I error or determining there is a sig-
nificant difference related to treatment when there is not. However, if you are
interested in how overwintered bullfrog tadpoles may influence the amphibian
community, then manipulating bullfrog abundance in replicate cages in a single
pond would be acceptable. Replication in field cage studies requires replicating
at the level of your question and factor of influence (e.g. golf course), which
will likely be at separate locations. Replication is also necessary within ponds to
determine variation among specific cage locations (e.g. aspect, slope, or shading
along a shoreline). Split-plot or nested designs are especially useful in studies
using cages in ponds and may allow the testing of hypotheses not testable with
less sophisticated designs (Gunzburger 2007).
The number of replicates you should use is also an important consideration.
Typically there is a trade-off between the number of replicates and the total
size of the study related to logistical and financial constraints. However, the
more divergent your replicates, the greater the variation within your treatments,
and the less your ability to detect significant differences between treatments.
Therefore, consideration of the level of variation inherent in your replicates will
influence the number of replicates that you need. While laboratory studies fre-
quently have five or more replicates because of the ease of replication, mesocosm
tank studies will often be limited to three to five replicates (Boone and James
2005), which is typically sufficient to detect significant differences in behavior,
morphology, development, and survival. You can estimate a priori how many
replicates you should use by determining your desired statistical power (1−;
a power of 0.8 or greater is desirable) based on your criteria and the effect
size (size of the difference). Using a statistical program like SAS (Proc Power)
or using online calculators (e.g. www.math.yorku.ca/SCS/Online/power/) you
can determine the ideal number of replicates to use based on the system you are
studying.
The random assignment of treatments is one of the most critical steps of all
true experiments and should be applied to mesocosm studies. Randomizing
your treatments is extremely important to avoid confounding variation
among replicate tanks or cages with variation among treatments. You can use
a random number chart or a random number generator in SAS (Proc Plan).
Treatments may be randomly assigned across all tanks or cages (completely
random design), or they may be blocked so that experimental units are grouped
along a known gradient that may cause variation (e.g. shading) not relevant to
the experimental question and which can be removed statistically (randomized
complete block).
98 | Amphibian ecology and conservation
grow fast enough to metamorphose in fast-drying tanks and suffered high pre-
dation by newts. Bufo were very sensitive to high density and produced few met-
amorphs without newt predators. This effect was reversed by newts selectively
preying on Scaphiopus and Rana tadpoles, especially in rapidly drying tanks.
Pseudacris performed poorly in all slow-drying tanks but showed moderate suc-
cess in high-density tanks where newts had removed most competitors. The
overall conclusion of this study was that all three factors interacted with each
other and with species’ life history to determine community structure. Further,
it suggested that competition dominates simple habitats of short duration and
predation becomes increasingly important in ponds with longer hydroperiods
that permit the establishment of diverse predators. Both of the studies described
above were influential in focusing attention on the complex interactions of pre-
dation, competition, and pond drying in aquatic communities rather than his-
torical single-factor arguments.
6.7.3 Ecotoxicology
The field of toxicology historically has taken a reductionist approach by using
single species and chemical testing in the laboratory. However, there is growing
effort to develop multispecies, community, and ecosystem-level experimental
approaches that better mimic real-world problems (Kimball and Levin 1985;
Rowe and Dunson 1994; Boone and James 2005; Semlitsch and Bridges 2005).
A model study by Boone et al. (2007) examined the role that chemical con-
tamination, competition, and predation play singly and in combination in an
aquatic amphibian community. They established replicate aquatic communities
in 64 polyethylene cattle tanks (round, 1.85 m diameter, 0.6 cm height, 1480 L
volume) by adding 1000 L of tap water, 1.0 kg of leaf litter, plankton from nat-
ural ponds, 60 Bufo americanus tadpoles, 30 R. sphenocephala tadpoles, and 10
Ambystoma maculatum salamander larvae to each. Four factors were manipulated
and replicated four times in a fully factorial design: no or two bluegill sunfish,
no or six overwintered bullfrog tadpoles, 0 or 2.5 mg/L carbaryl (an insecticide),
and 0 or 10 mg/L ammonium nitrate fertilizer. The results showed that bluegills
had the largest impact on the community by eliminating B. americanus and A.
maculatum and reducing the abundance of R. sphenocephala. Chemical contam-
inants had the second strongest effect on the community with the insecticide
reducing A. maculatum abundance by 50% and increasing the mass of anurans
(R. sphenocephala and B. americanus) at metamorphosis; the fertilizer positively
influenced time and mass at metamorphosis for both anuran and salamander
larvae. Presence of overwintered bullfrog tadpoles reduced mass and increased
time to metamorphosis of the anurans. Although both bluegill and overwin-
tered bullfrog tadpoles had negative effects on the amphibian community, they
performed better in the presence of one another and in contaminated ponds.
These results indicate that predicting complex interactions from single-factor
effects may not be straightforward. Further, the research supports the hypoth-
esis that combinations of factors can contribute to population declines.
and (4) a control. The five pools were placed at three distances from the nat-
ural breeding pond in each treatment (see Figure 1 in Hocking and Semlitsch
2007): one pool at 10 m, two pools at 64 m, and two pools at 115 m from
a natural breeding pond. The four forest treatments were positioned around
three replicate natural breeding ponds for a total of 60 wading pools. Pools
filled naturally with rainwater by April. Following the first oviposition event,
all pools were checked every 24–48 h for treefrog eggs. Eggs were removed
and counted. The results indicated that treefrogs laid significantly more eggs in
the clear-cuts (low-CWD 77 185 eggs; high-CWD 51 990 eggs) than in either
the partial (13 553 eggs) or the control (14 068 eggs) treatments. Further, the
interaction of forest isolation treatment showed that oviposition in both clear-
cut treatments decreased with increasing isolation whereas oviposition was the
same at all levels of isolation in the partial and control treatments. Hocking and
Semlitsch (2007) suggested that the strong preference for open-canopy pools
likely benefits the larval stage because of higher-quality food and warmer water
temperatures but that isolation of breeding pools beyond 50–100 m can inhibit
oviposition by female treefrogs. They also add that the possible benefits to one
stage (tadpoles) might be diminished by the risk of desiccation to the terrestrial
juvenile stage after metamorphosis. Thus, mesocosms can be used to address
land-use questions beyond the boundary of aquatic ecosystems.
6.8 Conclusion
Although we are both advocates for the use of mesocosms in amphibian ecology
and conservation, we also agree that a pluralistic approach to any question is
likely necessary for complete understanding. The use of laboratory experiments
to understand mechanisms as well as field studies to anchor mesocosm results
back to the real world are critical. The use of cattle watering tanks, wading
pools, and cages is a powerful technique in any researcher’s toolbox. The choice
of mesocosm should reflect knowledge of the natural history of the species and
system being studied. If used properly, at the correct scale for the questions one
is asking, many questions can be clearly answered and results can have strong
inferences back to natural systems. However, we caution readers that neither
type of mesocosm described here can address all ecological and conservation
questions alone. One has only to read Schindler (1998) to realize that most
large-scale ecosystem processes cannot be replicated in cattle tanks or cages.
But, for many population- and community-level studies of amphibians, meso-
cosms offer an excellent approach.
102 | Amphibian ecology and conservation
6.9 References
Boone, M. D. and Bridges-Britton, C. M. (2006). Examining multiple sublethal con-
taminants on the gray treefrog, Hyla versicolor: effects of an insecticide, herbicide, and
fertilizer. Environmental Contamination and Chemistry, 25, 3261–5.
Boone, M. D. and James, S. M. (2003). Interactions of an insecticide, herbicide, and
natural stressors in amphibian community mesocosms. Ecological Applications, 13,
829–41.
Boone, M. D. and James, S. M. (2005). Use of aquatic and terrestrial mesocosms in eco-
toxicology. Applied Herpetology, 2, 231–57.
Boone, M. D., Semlitsch, R. D., Fairchild, J. F., and Rothermel, B. B. (2004). Effects of
an insecticide on amphibians in large-scale experimental ponds. Ecological Applications,
14, 685–91.
Boone, M. D., Semlitsch, R. D., Little, E. E., and Doyle, M. C. (2007). Multiple stressors
in amphibian communities: interactive effects of chemical contamination, bullfrog
tadpoles, and bluegill sunfish. Ecological Applications, 17, 291–301.
Boone, M. D., Semlitsch, R. D., and Mosby, C. (2008). Suitability of golf course ponds
for amphibian metamorphosis when bullfrogs are removed. Conservation Biology, 22,
172–9.
Brockelman, W. Y. (1969). An analysis of density effects and predation in Bufo americanus
tadpoles. Ecology, 50, 632–43.
Carpenter, S. R. (1996). Microcosm experiments have limited relevance for community
and ecosystem ecology. Ecology, 77, 677–80.
Clausen, J., Keck, D. D., and Hiesey, W. M. (1947). Heredity of geographically and eco-
logically isolated races. American Naturalist, 81, 114–33.
Diamond, J. (1986). Overview: laboratory experiments, field experiments, and natural
experiments. In J. Diamond and T. Case (eds), Community Ecology, pp. 3–22. Harper
and Row Publishers, New York.
Drake, J. A., Huxel, G. R., and Hewitt, C. L. (1996). Microcosms as models for generat-
ing and testing community theory. Ecology, 77, 670–7.
Garcia, T. S., Stacy, J., and Sih, A. (2004). Larval salamander response to UV radiation
and predation risk: color change and microhabitat use. Ecological Applications, 14,
1055–64.
Gascon, C. (1992). Aquatic predators and tadpole prey in central Amazonia: field data
and experimental manipulations. Ecology, 73, 971–80.
Gascon, C. and Travis, J. (1992). Does the spatial scale of experimentation matter? A test
with tadpoles and dragonflies. Ecology, 73, 2237–43.
Griffis-Kyle, K. L. and Ritchie, M. E. (2007). Amphibian survival, growth, and devel-
opment in response to mineral nitrogen exposure and predator cues in the field: An
experimental approach. Oecologia, 152, 633–42.
Gunzburger, M. S. (2007). Habitat segregation in two sister taxa of hylid treefrogs.
Herpetologica, 63, 301–10.
Hairston, Sr, N. G. (1989a). Hard choices in ecological experimentation. Herpetologica,
45, 119–22.
Hairston, Sr, N. G. (1989b). Ecological Experiments. Cambridge University Press, Cambridge.
6 Aquatic mesocosms | 103
7.1 Introduction
Most amphibian species spend portions or even their entire life cycle in water.
Whether their habitats are streams, wetlands, vernal ponds, farm ponds, or
larger bodies of water, amphibians are directly affected by several natural and
anthropogenic chemical and physical characteristics. Dissolved oxygen, tem-
perature, pH, salinity and water conductivity, organic carbons, and pollutants
are important factors of their habitats that can affect survival, growth, mat-
uration, and physical development. In addition to direct effects, these char-
acteristics interact with other factors such as predators, prey, parasites, and
competitors to affect populations. Just as one example, the incidence of infest-
ation of tadpoles by Ribeioria ondatrae, a trematode parasite that causes limb
malformations, has been linked to the nutrient status of ponds (Johnson and
Chase 2004). Field studies of amphibians should at least include a description
of naturally occurring environmental conditions. In some cases, this descrip-
tion should also include analysis of environmental contaminants. This chapter
discusses chemical and physical factors that are most important to aquatic life
stages of amphibians and directs the reader to reviews for more complicated
issues dealing with water quality.
Because of space limitations, the descriptions of these factors and the methods
used in analyzing them are abbreviated. More complete descriptions of physio-
logical requirements can be found in Feder and Bruggren (1992) and McDiarmid
and Altig (1999). I recommend US Geological Survey (USGS) (2008a) and
Eaton et al. (2005) for standardized protocols of water-quality analysis.
These organisms take in oxygen through moist membranes such as gill epithe-
lium, dermis, or other structures. Aside from for rare exceptions, only aquatic
life stages of amphibians potentially face oxygen problems; atmospheric oxy-
gen is sufficient for terrestrial forms. Some species of amphibian, such as tailed
frogs (Ascaphus spp.), live in swift-moving lotic environments where water is
constantly mixed with air and oxygen concentrations are usually near satur-
ation. Others, however, inhabit quiescent, warm water, lentic environments
where oxygen concentrations can fluctuate or persist at low concentrations.
Under laboratory conditions, oxygen concentrations below 4 mg/L are deemed
stressful to amphibian larvae and other aquatic organisms (ASTM 1988) and
prolonged exposure to such hypoxic conditions in the field can be considered
adverse. However, there are differences among species and animals can accli-
mate over time to low oxygen concentrations.
Many amphibians have anatomical features, behaviors, and physiological
processes that allow survival under temporary hypoxic conditions. For example,
members of the families Ranidae and Ambystomatidae assimilate oxygen
through dermal uptake, gills, and lungs. Under normoxic conditions (non-
stressful oxygen concentrations), approximately 70% of oxygen uptake by these
larvae is through the dermis, 20% via the gills, and 10% through the lungs
(Gatten et al. 1984). In contrast, Bufo spp. and Ascaphus spp. do not inflate their
lungs until just before metamorphosis (Feder 1984) and Plethondontids lack
lungs entirely.
In hypoxic conditions, larvae with lungs often swim to the surface and gulp
air into the lungs and pulmonary uptake of oxygen becomes more important.
Surfacing, however, incurs metabolic costs and exposes tadpoles to predators such
as turtles (Feder 1983). However, in an experiment using aquariums, McIntyre
and McCollum (2000) found that the alternative behavior of not coming to the
surface could also incur risks. Young, pre-lunged bullfrog (Rana catesbeiana)
tadpoles experienced a higher degree of predation in high-oxygen tanks than in
low-oxygen ones because predacious tiger salamanders (Ambystoma tigrinum)
spent a greater proportion of time searching for tadpoles at the bottom of the
tanks under high-oxygen conditions than when the dissolved oxygen was low.
Hypoxia can induce similar physiological responses in amphibians as in other
vertebrates including changes in blood pH, build-up of lactate in muscles, leth-
argy, and, under severe conditions, mortality (Ultsch et al. 1999). Chronic and
intermittent hypoxia can reduce growth and developmental rates and delay
hatching of salamander embryos (Valls and Mills 2007).
Several factors affect the concentration and measurement of dissolved oxygen.
Oxygen concentrations can vary widely through the course of a day, especially
7 Water-quality criteria for amphibians | 107
15
12
Dissolved oxygen (mg/L)
0
12:00 4 8 12:00 4 8 12:00
am am am
Time of day
example, YSI Instruments, Orion, Accumet, and Oakton are a few of the com-
panies that manufacture portable field meters. These almost always come with
temperature sensors and may be bundled with conductivity and pH sensors.
Colorimeter kits (e.g. LaMotte, Orion, Chemetrics) are available and tend to be
less expensive than electronic meters but require reagents for every test and may
be less sensitive. When collecting water for oxygen analysis it is important to fill
plastic or glass bottles to the very top without adding air and to keep the bottle
cool until analysis. Ideally, oxygen should be analyzed on site to obtain the most
accurate values.
7.3 Temperature
Because amphibians are poikilotherms they have limited ability to regulate
their body temperature and are greatly affected by the temperature of their sur-
rounding environment. Water temperature, therefore, is extremely important in
affecting metabolic rates, other physiological processes, and behavior.
As a class, amphibians demonstrate a wide tolerance to temperature regi-
mens: some species are able to withstand “temperature changes of 30°C on a
daily basis and up to about 35°C on a seasonal basis” (Rome et al. 1992, p. 183).
Wood frogs (Rana sylvatica) have the most northern distribution of any North
American species (Conant and Collins 1998) and live above the Arctic Circle.
Couch’s spadefoot (Scaphiophus couchii) inhabits very arid, hot desert areas in
the west where temperatures may exceed 45°C. Some species are eurythermic;
for example, American bullfrogs range from central Mexico to northern Maine
and Nova Scotia and tiger salamanders can be found from southern Texas into
Alberta and Saskatchewan. Given time to acclimate, adult and larval amphib-
ians in temperate and northern climes can survive near and even subzero tem-
peratures by becoming dormant and greatly reducing their metabolic rates
(Voituron et al. 2003). Those that inhabit dry, hot regions often burrow into the
ground, enmesh themselves in cocoons, and estivate (Pinder et al. 1992). Some
species, such as northern leopard frogs (Rana pipiens), can shunt water from vital
organs to prevent tissue damage from ice crystal formation and concentrate glu-
cose in organs to serve as a type of anti-freeze (Tattersall and Ultsch 2008).
In general, between 10 and 40°C, each 10° increase in ambient tempera-
ture increases metabolism by 1.4–2.4 times (Rome et al. 1992). Higher meta-
bolic rates require greater oxygen; however, oxygen concentrations decrease
as water temperatures increase (see below). At temperatures near 0°C most
amphibians are very sluggish. At high temperatures, say above 30–35°C for
some species but as low as 25–30°C for less tolerant animals, thermal stress can
7 Water-quality criteria for amphibians | 109
result in reduced mobility, abnormally high heart rates, and eventually death.
Subtle interactions between temperature and other stressors can also occur. For
example, Anderson et al. (2001) demonstrated complex relationships between
predators and temperature in Pacific treefrogs (Pseudacris regilla). At 25.7°C
tadpoles grew faster and exceeded the size limitation of their insect predators
more quickly than tadpoles raised at 9.9°C. However, tadpoles of similar size
encountered greater predation rates at the warmer temperature. Temperature
can have a strong influence on life history progress (Camp and Marshall 2000)
and microhabitat use (Crawford and Semlitsch 2008) in both aquatic and ter-
restrial amphibians.
There is a direct relationship between water temperature and dissolved oxygen
concentrations. At sea level, where the partial pressure of oxygen, PO, is highest,
the concentration of dissolved oxygen is around 14 ppm at 0°C and declines as
water temperatures increase (Figure 7.2). The same general relationship exists
at higher elevations except that PO saturation concentrations decline. Because
oxygen demand increases as water increases in temperature, physiological stress
can be very high in warm, low oxygenated waters.
Measuring water temperature is straightforward with any digital or analog
thermometer. Many electronic meters used for other water chemistry parameters
such as pH, conductivity, or dissolved oxygen come equipped with thermom-
eters. Some thermometers automatically record minimum and maximum tem-
peratures between readings so that a range can be ascertained. More sophisticated
systems include data loggers that keep a continuous record of temperatures.
15
Dissolved oxygen (mg/L)
13
11
5
0 5 10 15 20 25 30
Temperature (°C)
Fig. 7.2 Relationship between water temperature and dissolved oxygen concentration
at sea level.
110 | Amphibian ecology and conservation
7.4 pH
pH, the negative log of the hydrogen ion concentration in water, is another key
characteristic in describing amphibian habitats. The pH scale ranges from 0 to 14
which corresponds to a solution of 1 M H (100) to 1 M OH with all H bound
with oxygen as a hydroxide. A pH of 7 defines neutral solutions, those with lower
pH values are acidic, and those with higher values are alkaline. For ecological
purposes, pH ranges of 6.0–7.5 are generally considered to be circumneutral or
within a range that should present no harm to most aquatic organisms. A corre-
sponding term, alkalinity, is the ability of water to buffer changes in pH.
A key factor in determining pH is the type of bedrock found within the
watershed. Sedimentary rock such as sandstone, limestone, and dolomite are
high in carbonates and bicarbonates that bind with hydrogen ions and form
natural buffers. Alkalinity is typically high in these waters. Watersheds with a
bedrock of igneous rock such as granite or basalt typically have low alkalinity
due to the absence of carbonates.
Another natural factor affecting pH is the concentration of organic acids
including tannic, fulvic, and humic acids. These organic compounds are weak
acids in that they have low dissociation constants compared to mineral acids.
Organic acids can be important in maintaining a low pH in isolated water bod-
ies with high organic matter such as fens and bogs, but they also serve to some
degree as buffers against anthropogenic sources of acidity.
Two major sources of anthropogenic acidity are acid deposition and acid
mine drainage. Acid deposition comes from the production of nitrous oxides
(NOx) and sulfates (SO42) during the combustion of fossil fuels. These mol-
ecules combine with hydrogen in the air and are deposited on to land and water
either with rain or attached to dust and organic particles that precipitate in
dry form. Once in water the hydrogen and anions dissociate and free hydrogen
ions increase acidity. Acid deposition can be of major concern in wetlands and
streams that lie downwind of major industrial centers or cities and in water-
sheds with low alkalinity. The US Environmental Protection Agency (USEPA)
(2008) is a good reference for current aspects of acid rain and deposition. Acid
mine drainage is most prevalent in unreclaimed mined areas. Inversion of soil
during mining places layers high in sulfides towards the surface and oxidation
of these soils results in acidic seeps and leaching. The USGS (2008b) provides a
good review of acid mine drainage.
Many review articles have been written on the adverse effects of acid depos-
ition and anthropogenic acidity on amphibians (e.g. Sparling 1995; Rowe
and Freda 2000). Variation in sensitivity to low pH occurs at the life stages,
7 Water-quality criteria for amphibians | 111
Colored DOCs (tannic, humic, and fulvic acids) can be measured indir-
ectly by comparing the color of water to a set of standards available in kit form.
Turbidity, which is affected by the total amount of inorganic and organic sus-
pended particles, can be measured with a turibidimeter that measures light
penetration through a standardized vial in units called nephelometric turbidity
units (NTUs).
7.7 Pollutants
The number of types of pollutants or contaminants that can affect amphibian
populations is enormous and beyond the scope of this brief chapter. Sparling
(2003) surveyed the potential effects of contaminants on amphibians and
Sparling et al. (2000) provided an extensive review of contaminant effects on
amphibians and reptiles. Here I list a brief summary of contaminant classes and
their significance.
7.7.2 Pesticides
Pesticides are chemicals that are used to control plant and animal species con-
sidered to be noxious or unwanted by humans. The types of pesticides can be
distinguished by intended target and by chemical class, which usually corres-
ponds to primary mode of action. By target the primary classes are insecticides,
fungicides, herbicides, and rodenticides.
Very roughly and with exceptions, insecticides are most acutely toxic to
amphibians and other aquatic organisms of these classes. However, insecti-
cides, fungicides, and herbicides exert a wide variety of sublethal or chronic
effects that affect individual and population health. Rodenticides are not of
much concern to amphibians, simply because of how they are used. There is a
large and growing number of chemical families used as pesticides (see Cowman
and Mazanti 2000 for a more complete review). Of greatest concern due to
7 Water-quality criteria for amphibians | 115
7.7.3 Metals
Metals and metalloids such as zinc, copper, lead, chromium, arsenic, cadmium,
mercury, selenium, and others are naturally occurring elements but are also
released through many different industrial processes at concentrations that can
be toxic to amphibians. Linder and Grillitsch (2000) reviewed the ecotoxicol-
ogy of metals on amphibians and reptiles and showed that effects can range
from direct mortality to a host of sublethal and potentially debilitating effects.
Toxicity is affected by pH in that metal solubility increases with acidity and sol-
uble forms of metals are more bioavailable than non-soluble forms.
Ion-specific probes are available for lead, silver, and copper, but the method
for measuring is more complicated than inserting the probe directly into a water
sample and should be done under laboratory conditions. Fortunately, metals are
very stable, and once samples have been acidified to pH
2 they can be stored
for several months before analysis. Colorimetric methods exist for iron, man-
ganese, aluminum, molybdenum, and copper, but most metals are analyzed
through atomic absorption spectrophotometry or ion-coupled spectrophotom-
etry (ICP) in a laboratory.
(e.g. DDT, toxaphene, dieldrin) in this category. All of these chemicals are
organically based and some have chlorine or other halogens incorporated into
the molecular structure. PAHs include benzene, toluene, benzo[a]pyrene, and
hundreds of other molecules that form through natural processes of decompos-
ition, but also come from combustion of fossil fuels, processing of petroleum,
and other sources. PCBs were used as lubricating fluids and in electrical trans-
formers until the 1970s when their use in North America was banned. Dioxins
and furans are formed during forest fires but also through the combustion
of fossil fuels and are strong carcinogens. Similarly, many of the chlorinated
hydrocarbon pesticides have been banned because they are extremely persist-
ent in the environment and cause cancer, genotoxicity, endocrine disruption,
and other harmful effects to humans and wildlife. Nevertheless, because they
are extremely persistent, they can still be found throughout the world. At most
environmentally realistic concentrations these chemicals produce a variety of
sublethal effects including genotoxicity, inhibition of growth and develop-
ment, endocrine disruption, skeletal defects and malformations, and reduced
hatching success (reviewed by Sparling 2000). Sophisticated laboratory methods
including HPLC or gas chromatography are required for the analysis of these
chemicals.
7.7.5 Pharmaceuticals
Ecotoxicologists are just becoming aware of a myriad of chemicals that pass
through our waste-treatment plants and enter natural bodies of waters essen-
tially intact. These include prescription and non-prescription medicines, anti-
biotics, caffeine, and other products that pass through human bodies or are
disposed of into toilets and down drains. These chemicals can exert numerous
effects such as endocrine disruption and many physiological and behavioral
changes that have yet to be defined. Because our awareness of these substances is
so new, methods for determining their concentrations in field-collected matrices
are often not available. Others can be determined through gas chromatography/
mass spectrophotometry.
methods of measuring these factors and the choice of method should be based
on desired accuracy and precision as well as economics.
7.9 References
Anderson, M. T., Kiesecker, J. M., Chivers, D. P., and Blaustein, A. R. (2001). The direct
and indirect effects of temperature on a predator-prey relationship. Canadian Journal
of Zoology, 79, 1834–41.
Andren, C., Henrikson, L. Olsson, M., and Nilson, G. (1989). Effects of pH and alu-
minium on embryonic and early larval stages of Swedish brown frogs Rana arvalis, R.
temporaria and R. dalmatina. Holoarctic Ecology, 11, 127–35.
ASTM (1988). Standard Practice for Conducting Acute Toxicity Tests with Fishes,
Macroinvertebrates, and Amphibians. ASTM, West Conshohocken, PA.
Balinsky, J. B. (1981). Adaptation of nitrogen metabolism to hyperosmotic environment
in Amphibia. Journal of Experimental Zoology, 215, 335–50.
Banks, M. S., Crocker, J., Connery, B., and Amirbahman, A. (2007). Mercury bio-
accumulation in green frog (Rana clamitans) from Acadia National Park, Maine, USA.
Environmental Toxicology and Chemistry, 26, 118–25.
Brodkin, M., Vatnick, I., Simon, M., Hopey, H., Butler-Holston, K., and Leonard, M.
(2003). Effects of acid stress in adult Rana pipiens. Journal of Experimental Zoology,
298A, 16–22.
Calfee, R. D., Bridges, C. M., and Little, E. E. (2006). Sensitivity of two salamander
(Ambystoma) species to ultraviolet radiation. Journal of Herpetology, 40, 35–42.
Camp, C. D. and Marshall, J. L. (2000). The role of thermal environment in determining
the life history of a terrestrial salamander. Canadian Journal of Zoology, 78, 1702–11.
Carpenter, S. R., Cole, J. J., Hodgon, J. R., Kitchell, J. F., Pace, M. L., Bade, D.,
Cottingham, K. L., Essington, T. E., Houser, J. N., and Schindler, D. E. (2001).
Trophic cascades, nutrients, and lake productivity: whole-lake experiments. Ecological
Monographs, 71, 163–86.
Conant R. and Collins, J. T. (1998). Reptiles and Amphibians. East/Central North America.
Peterson Field Guides. Houghton-Mifflin, Boston, MA.
Cowman, D. F. and Mazanti, L. E. (2000). Ecotoxicology of “New Generation” pesticides
to amphibians. In D. W. Sparling, G. Linder, and C. A. Bishop (eds), Ecotoxicology of
Amphibians and Reptiles, pp. 233–68. SETAC Press, Pensacola, FL.
Crawford, J. A. and Semlitsch, R. D. (2008). Abiotic factors influencing abundance and
microhabitat use of stream salamanders in southern Appalachian forests. Forest Ecology
and Management, 255, 1841–7.
Davidson, C., Shaffer, H. B., and Jennings, M. R. (2001). Declines of the California red-
legged frog: climate, UV-B, habitat, and pesticides hypotheses. Ecological Applications,
11, 464–79.
Diamond, S. A., Trenham, P. C., Adams, M. J., Hossack, B. R., Knapp, R. A., Stark, S. L.,
Bradford, D., Corn, C. S., Czarnowski, K., Brooks, P. D. et al. (2005). Estimated ultra-
violet radiation doses in wetlands in six national parks. Ecosystems, 8, 462–77.
Dunson, W. A. and Connell, J. (1982). Specific inhibition of hatching in amphibian
embryos by low pH. Journal of Herpetology, 16, 314–16.
118 | Amphibian ecology and conservation
Eaton, A. D., Clesceri, L. S., Rice, E. W., and Greenberg, A. E. (2005). Standard Methods
for the Examination of Water and Waste Water, 21st edn. American Public Health
Association Press, Washington DC.
Feder, M. E. (1983). The relation of air breathing and locomotion to predation on tad-
poles, Rana berlandieri, by turtles. Physiological Zoology, 56, 522–31.
Feder, M. E. (1984). Consequences of aerial respiration for amphibian larvae. In
R. S. Seymour (ed.), Respiration and Metabolism of Embryonic Vertebrates, pp.
71–86. Junk, Dordrecht.
Feder, M. E. and Burggren, W. W. (1992). Environmental Physiology of the Amphibians.
University of Chicago Press, Chicago, IL.
Freda, J. (1990). Effects of acidification on amphibians. In J. P. Baker (ed.), State of Science
and State of Technology, SOS/T report no. 13, Biological Effects of Changes in Surface
Water Acid-Base Chemistry, pp. 2-185–2-202. National Acid Precipitation Assessment
Program, Washington DC.
Freda, J. (1991). The effects of aluminum and other metals on amphibians. Environmental
Pollution, 71, 305–28.
Freda, J. and Dunson, W. A. (1985). The influence of external cation concentration on
the hatching of amphibian embryos in water of low pH. Canadian Journal of Zoology,
63, 2649–56.
Freda, J., Cavdek, V., and McDonald, D. G. (1989). Role of organic complexation in the
toxicity of aluminum to Rana pipiens embryos and Bufo americanus tadpoles. Canadian
Journal of Fisheries and Aquatic Sciences, 47, 217–24.
Gatten, Jr, R. E., Caldwell, J. P., and Stockard, M. E. (1984). Anaerobic metabolism dur-
ing intense swimming by anural larvae. Herpetologica, 40, 164–9.
Gomez-Mestre, I., Tejedo, M. R., and Estepa, J. (2004). Developmental alterations and
osmoregulatory physiology of a larval anuran under osmotic stress. Physiological and
Biochemical Zoology, 77, 267–74.
Gosner, K. L. and Black, J. H. (1957). The effects of acidity on the development and
hatching of New Jersey frogs. Ecology, 38, 256–62.
Horne, M. T. and Dunson, W. A. (1995a). The interactive effects of low pH, toxic metals,
and DOC on a simulated temporary pond community. Environmental Pollution, 89,
155–61.
Horne, M. T. and Dunson, W. A. (1995b). Effects of low pH, metals, and water hardness on
larval amphibians. Archives of Environmental Contamination and Toxicology, 29, 500–5.
Johnson, P. T. J. and Chase, J. M. (2004). Parasites in the food web: linking amphibian
malformations and aquatic eutrophication. Ecology Letters, 7, 521–6.
Karraker, N. E., Gibbs, J. P., and Vonesh, J. R. (2008). Impacts of road deicing salt on the
demography of vernal pool-breeding amphibians. Ecological Applications, 18, 724–34.
Linder, G. and Grillitsch, B. (2000). Ecotoxicology of metals. In D. W. Sparling,
G. Linder, and C. A. Bishop (eds), Ecotoxicology of Amphibians and Reptiles,
pp. 325–408. SETAC Press, Pensacola, FL.
McDiarmid, R. W. and Altig, R. (1999). Tadpoles: The Biology of Anuran Larvae.
University of Chicago Press, Chicago, IL.
McIntyre, P. B. and McCollum, S. A. (2000). Responses of bullfrog tadpoles to hypoxia
and predators. Oecologia, 125, 301–8.
7 Water-quality criteria for amphibians | 119
Ortiz-Santaliestra, M. E., Marco, A., Fernandez, M. J., and Lizana, M. (2006). Influence
of developmental stage on sensitivity to ammonium nitrate of aquatic stages of amphib-
ians. Environmental Toxicology and Chemistry, 25, 105–11.
Pinder, A. W., Storey, K. B., and Ultsch, G. R. (1992). Estivation and hibernation. In
M. E. Feder and W. W. Burggren (eds), Environmental Physiology of the Amphibians,
pp. 250–74. University of Chicago Press, Chicago, IL.
Relyea, R. A. and Mills, N. (2001). Predator-induced stress makes the pesticide carbaryl
more deadly to gray treefrog tadpoles (Hyla versicolor). Proceedings of the National
Academy of Sciences USA, 98, 2491–6.
Rome, L. C., Stevens, E. D., and John-Alder, H. B. (1992). The influence of temperature
and thermal acclimation on physiological function. In M. E. Feder and W. W. Burggren
(eds), Environmental Physiology of the Amphibians, pp. 183–205. University of Chicago
Press, Chicago, IL.
Rowe, C. L. and Freda, J. (2000). Effects of acidification on amphibians at multiple levels
of organization. In D. W. Sparling, G. Linder, and C. A. Bishop (eds), Ecotoxicology of
Amphibians and Reptiles, pp. 545–72. SETAC Press, Pensacola, FL.
Sadinski, W. J. and Dunson, W. A. (1992). A multilevel study of effects of low pH on
amphibians of temporary ponds. Journal of Herpetology, 26, 413–22.
Shoemaker, V. H., Hillman, S. S., Hillyard, S. D., Jackson, D. C., McClanahan, L. L.,
Withers, P. C., and Wygoda, M. L. (1992). Exchange of water, ions, and respiratory
gases in terrestrial amphibians. In M. E. Feder and W. W. Burggren (eds), Environmental
Physiology of the Amphibians, pp. 125–50. University of Chicago Press, Chicago, IL.
Skei, J. K. and Dolmen, D. (2006). Effects of pH, aluminum, and soft water on larvae of the
amphibians Bufo bufo and Triturus vulgaris. Canadian Journal of Zoology, 84, 1668–77.
Sparling, D. W. (1995). Acidic deposition: a review of biological effects. In D. Hoffman,
B. A. Rattner, G. A. Burton, and J. Cairns Jr (eds), Handbook of Ecotoxicology,
pp. 301–29. Lewis Publishers, Boca Raton, FL.
Sparling, D. W. (2000). Ecotoxicology of organic contaminants to amphibians. In
D. W. Sparling, G. Linder, and C. A. Bishop (eds), Ecotoxicology of Amphibians and
Reptiles, pp. 461–94. SETAC Press, Pensacola, FL.
Sparling, D. W. (2003). A review of the role of contaminants in amphibian declines.
In D. Hoffman, B. A. Rattner Jr, and J. Cairns (eds), Handbook of Ecotoxicology,
pp. 1099–1128. Lewis Publishers, Boca Raton, FL.
Sparling, D. W. and Lowe, T. P. (1996). Environmental hazards of aluminum to plants, inver-
tebrates, fish, and wildlife. Reviews Environmental Contamination Toxicology, 145, 1–127.
Sparling, D. W., Linder, G., and Bishop, C. A. (2000). Ecotoxicology of Amphibians and
Reptiles. SETAC Press, Pensacola, FL.
Sparling, D. W., Fellers, G. M., and McConnell, L. L. (2001). Pesticides and amphibian
population declines in California, USA. Environmental Toxicology and Chemistry, 20,
1591–5.
Tattersall, G. J. and Ultsch, G. R. (2008). Physiological ecology of aquatic overwintering
in ranid frogs. Biological Review, 83, 119–40.
Ultsch, G. R., Bradford, D. F., and Freda, J. (1999). Physiology: coping with the environ-
ment. In R. W. McDiarmid and R. Altig (eds), Tadpoles. The Biology of Anuran Larvae,
pp. 189–214. University of Chicago Press, Chicago, IL.
120 | Amphibian ecology and conservation
8.1 Introduction
In this chapter, I review marking and measuring techniques for post-metamorphic
amphibians; larvae are covered in Chapter 4 in this volume. Measuring the size
of individuals is important for determining age, categorizing life history stages
(juvenile or adult), and calculating growth rates (larval measurements are covered
in Chapter 3). More field and behavioral studies requiring marking are now being
conducted than in the past, and often investigators need very specific marking
techniques and non-invasive ways of identifying individuals. Additionally, ethical
issues relative to certain techniques, such as radioactive tags and mutilation pro-
cedures, are now more commonly considered as, for example, when considering
toe-clipping. Some older techniques are used much less often than they used to
be. Major advancements in biotelemetry (see Chapter 11) and passive integrated
transponder (PIT) tags have added needed flexibility to the choice of marking
methods.
Recent reviews of marking and identification techniques include Baker and
Gent (1998) and Ferner (2007) for amphibians and reptiles, and Donnelly et al.
(1994) for amphibians. Additional references are found in a number of papers
published in a special edition of Mertensiella (Henle and Veith 1997).
Some criteria for ideal marks or tags are as follows (after Ferner 2007):
• they should not affect the survivorship or behavior of the organism;
• they should allow the animal to be as free from stress or pain as possible;
• they should identify the animal as a particular individual, if desirable;
• they should last indefinitely, or at least the duration of the study;
• they should be easily read and/or observable;
124 | Amphibian ecology and conservation
Table 8.1 Sources of materials for marking amphibian species (as reviewed in Ferner
2007 unless otherwise indicated).
Technique item Source of materials
described for use with one group of amphibians might be adapted for a dif-
ferent group. Therefore, investigators might benefit by considering all of the
techniques available, regardless of how originally used. Further, there may be a
need for standardizing marking techniques in comparative studies, or those that
might be carried on by other researchers in the future.
In the following text, I emphasize the use of marking techniques for adult
amphibians which have proved most successful. When available, information
on the advantages and disadvantages of each technique has been given. This
chapter provides information adequate for considering techniques without
consulting the original source. However, once a technique has been tentatively
selected, it is advisable to consult the primary literature if time and facilities
permit. The use of complex techniques, such as telemetry or PIT tags, requires
careful review of the original sources. Sources for materials used with various
techniques are given in Table 8.1, although availability and contact information
are likely to change constantly.
8.2 Toe-clipping
8.2.1 Anurans
Although several toe-clipping methods are proposed in the literature, that of
Martof (1953) is most widely used. He cut toes of Rana clamitans with scissors,
trying to avoid damage to webbing between them. Little blood loss or loss of
swimming power was observed and no regeneration of the digits was found.
This system assigns serial numbers to the digits of the feet. No more than two
toes were removed from any one foot. The left hind foot indicated units, and
for those larger than five, combinations of two digits (the fifth and one other)
were excised. For example, clipping the fifth and third toes designated tens and
the front feet denoted hundreds. With this arrangement, 6399 individuals can
be marked in series; the greatest number can be indicated by removing the two
outer digits on each foot. The only concern Martof expressed about this pro-
cedure was that confusion might arise where only a single toe on a single hand
foot was removed with recapture of a frog with an injured foot. He suggested
alleviating this problem with a special designation, such as a zero marking by
removal of the second and fourth toes on the other hind foot. Waichman (1992)
also proposed a numbering system for toe-clipping that letters the four limbs
(A through D) and numbers the toes. This scheme makes all 959 combinations
available using two and three toes, but not removing more than two toes for each
limb (see Table 8.2).
Table 8.2 Alphanumeric code for toe-clipping amphibians and reptiles from Waichman (1992)
One toe Two toes A2A5 A3D2 A5B5 B1D4 B4C1
A2B1 A3D3 A5C1 B1D5 B4C2
A2B2 A3D4 A5C2 B2B3 B4C3
A1 A1A2 A2B3 A3D5 A5C3 B2B3 B4C4
A2 A1A3 A2B4 A4A5 A5C4 B2B4 B4C5
A3 A1A4 A2B5 A4B1 A5C5 B2B5 B4D1
A4 A1A5 A2C1 A4B2 A5D1 B2D1 B4D2
A5 A1B1 A2C2 A4B3 A5D2 B2D2 B4D3
B1 A1B2 A2C3 A4B4 A5D3 B2D3 B4D4
B2 A1B3 A2C4 A4B5 A5D3 B2D4 B4D5
B3 A1B4 A2C5 A4C1 A5D4 B2D5 B5C1
B4 A1B5 A2D1 A4C2 A5D5 B3B4 B5C2
B5 A1C1 A2D2 A4C3 B1B2 B3B5 B5C2
C1 A1C2 A2D3 A4C4 B1B3 B3C1 B5C3
C2 A1C3 A2D4 A4C5 B1B4 B3C2 B5C4
C3 A1C4 A2D5 A4D1 B1B5 B3C3 B5C5
C4 A1C5 A3A4 A4D2 B1C1 B3C4 B5D1
C5 A1D1 A3A5 A4D3 B1C2 B3C5 B5D2
D1 A1D2 A3C1 A4D4 B1C3 B3D1 B5D3
D2 A1D3 A3C2 A4D5 B1C4 B3D2 B5D4
D3 A1D4 A3C3 A5B1 B1C5 B3D3 B5D5
D4 A1D5 A3C4 A5B2 B1D1 B3D4 C1C2
D5 A1A3 A3C5 A5B3 B1D2 B3D5 C1C3
A1A4 A3D1 A5B4 B1D3 B4B5 C1C4
C1C5 C5D3 A1A2C3 A1A4B2 A1A5D1 A2A4B5 A2A5D4
C1D1 C5D4 A1A2C4 A1A4B3 A1A5D2 A2A4C1 A2A5D5
C1D2 C5D5 A1A2C5 A1A4B4 A1A5D3 A2A4C2 A3A4B1
C1D3 D1D2 A1A2D1 A1A4B5 A1A5D4 A2A4C3 A3A4B2
C1D4 D1D3 A1A2D2 A1A4C1 A1A5D5 A2A4C4 A3A4B3
C1D5 D1D4 A1A2D3 A1A4C2 A2A3B1 A2A4C5 A3A4B4
C2C3 D1D5 A1A2D4 A1A4C3 A2A3B2 A2A4D1 A3A4B5
C2C4 D2D3 A1A2D5 A1A4C4 A2A3B3 A2A4D2 A3A4C1
C2C5 D2D4 A1A3B1 A1A4C5 A2A3B4 A2A4D3 A3A4C2
C3C4 D2D5 A1A3B2 A1A4D1 A2A3B5 A2A4D4 A3A4C3
C3C5 D3D4 A1A3B3 A1A4D2 A2A3C1 A2A4D5 A3A4C4
C3D1 D3D5 A1A3B4 A1A4D3 A2A3C2 A2A5B1 A3A4C5
C3D2 D4D5 A1A3B5 A1A4D4 A2A3C3 A2A5B2 A3A4D1
C3D3 A1A3C1 A1A4D5 A2A3C4 A2A5B3 A3A4D2
C3D4 Three toes A1A3C2 A1A5B1 A2A3C5 A2A5B4 A3A4D3
C3D5 A1A3C3 A1A5B2 A2A3D1 A2A5B5 A3A4D4
C4C5 A1A3C4 A1A5B3 A2A3D2 A2A5C1 A3A4D5
C4D1 A1A2B1 A1A3C5 A1A5B4 A2A3D3 A2A5C2 A3A5B1
C4D2 A1A2B2 A1A3D1 A1A5B5 A2A3D4 A2A5C3 A3A5B2
C4D3 A1A2B3 A1A3D2 A1A5C1 A2A3D5 A2A5C4 A3A5B3
C4D4 A1A2B4 A1A3D3 A1A5C2 A2A4B1 A2A5C5 A3A5B4
C4D5 A1A2B5 A1A3D4 A1A5C3 A2A4B2 A2A5D1 A3A5B5
C5D1 A1A2C1 A1A3D5 A1A5C4 A2A4B3 A2A5D2 A3A5C1
C5D2 A1A2C2 A1A4B1 A1A5C5 A2A4B4 A2A5D3 and so on
Codes A5 and B5 are available only for animals with five foretoes.
128 | Amphibian ecology and conservation
Regardless of the numbering system used, the potential for toe regeneration
must be considered (see Table 8.3). In addition, the thumbs of the forefeet are
required by males during amplexus, and they should not be clipped (for example,
Briggs and Storm 1970). Another potential problem is understanding how toe-
clipping may stress an animal and affect resource acquisition. Daugherty (1976)
indicated a problem with weight loss in toe-clipped Rana pipiens. The most ser-
ious criticism of toe-clipping anurans was raised by Clarke (1972), who noted
that the probability of recapturing Bufo fowleri decreased as the number of digits
excised increased; researchers today do not clip as many toes as Clarke, however.
Halliday (1995) reported that studies of Atelopus elegans, Atelopus carbonerensis,
Bufo marinus, Bufo granulosus, and Bufo bufo provided little or no evidence for
any impact on survival after toe-clipping. An assessment and concern about use
of toe-clipping of Hyla labialis was provided by Lüddecke and Amezquita (1999).
Fewer than 1% of toe-clipped Rana pretiosa showed any sign of infection, and no
mortality was found (Reaser and Dexter 1996). They stressed that hygienic, ster-
ile techniques must be used, and that impacts should be carefully monitored in
Table 8.3 Digit-regeneration occurrence and times for selected toe-clipped amphibian
species
Species Time interval for regeneration
Anurans
Bufo hemiophrys Several months but recognizable
Hyla regilla Only slight after 1 year
Hyperolius viridiflavus ferniquei Complete in newly metamorphosed
Rana catesbeiana None in 3-year study
Rana erythraea Slow or nonexistent over 4 years
Caudates
Ambystoma opacum Some in short period
Taricha spp. Several years
Taricha granulosa Indefinitely, if kept below or at 10C
Triturus cristatus Slow compared to tail clips
Triturus vulgaris Up to 10 months; little winter regeneration
(Griffiths 1984)
Plethodon cinereus 7 months in laboratory
Plethodon glutinosus At least 2 years
Plethodon wehrlei 50% after 100 days; regenerated digits
identifiable by lack of pigmentation
Cryptobranchus alleganiensis 1 year
Batrochoseps spp. Slow; regenerated toes recognizable
After review by Ferner (2007) unless otherwise indicated.
8 Working with post-metamorphic amphibians | 129
8.2.2 Salamanders
Toe-clipping has been the common method of marking salamanders in recent
years, and in most cases something similar to Martof ’s (1953) system of mark-
ing anurans is used. Another scheme (as above) was provided by Waichman
(1992) (Table 8.2). The potential of toe regeneration needs to be considered in
salamanders, but whether or not it is a deterrent to using toe-clipping depends
upon the species and study objectives. Table 8.3 summarizes reports in the lit-
erature concerning the regeneration of digits in salamanders.
Some salamanders, such as juvenile (
2.5 cm snout–vent length, or SVL)
dusky salamanders (Desmognathus fuscus), have toes too small for precise clip-
ping. In such cases, animals have been marked successfully by clipping small
pieces of the tail at either right angles to the longitudinal axis of the body or in
the transverse plane at different angles (Orser and Shine 1972). This technique
allowed groups of marked juveniles to be distinguished for at least 1 month
before regeneration became a problem.
Mutilation techniques, such as toe-clipping, may use some form of local or
general anesthesia. Peterman and Semlitsch (2006) stressed the importance of
properly buffering solutions of the anesthetic MS-222 in plethodontid salaman-
ders, and the effect of various concentrations on the time to become anesthe-
tized and recover. They found induction time usually decreased, and the time
for recovery increased, with increasing concentrations of MS-222. This study
should be consulted before using this anesthetic.
Davis and Ovaska (2001) found that toe-clipping in Plethodon vehiculum
resulted in less weight gain than animals marked with fluorescent tags or used
as controls. These authors used unique combinations of three clipped toes per
salamander, removing no more than one toe per foot. The number of ambigu-
ous marks in toe-clipped individuals increased dramatically in frequency com-
pared with the fluorescent tagged animals after 50 days, but the majority of both
retained useful marks throughout the 87-week field study.
8.3 Branding
8.3.1 Anurans
Kaplan (1958) described a branding technique for frogs involving the incorp-
oration of India ink into scarified skin of the venter. Numbers were etched in
the skin with a hypodermic needle and then filled with India ink. Erythema
disappeared quickly, no infection was observed, and the numerals remained
distinguishable for more than 3 months. Kaplan also described an electric tat-
tooing technique as an improvement upon the scarification method. Higgins
India Ink™ was found most effective if mixed with a drop of glycerin to aid its
spreading into the skin. Surplus ink was wiped away, leaving a clear, permanent
mark. Some initial inflammation was observed in the frogs, but was only rarely
prolonged. Additional branding techniques are provided in Table 8.4.
Table 8.4 Branding techniques used for marking amphibian species (as reviewed in
Ferner 2007)
Species Description of branding technique
8.3.2 Salamanders
Salamander branding techniques involve heat or freeze brands, or the application
of pigments or dyes into scarified skin. Taber et al. (1975) branded Cryptobranchus
alleganiensis with heated 1.5-cm-high numerical brands of thin nickel-chrome
wire. These remained clear through a 2-year study, but some faded and required
rebranding. Bull et al. (1983) tested freeze-branding application times while mark-
ing Ambystoma macrodactylum. Letter brands of 5 mm high made with silver-tipped
copper rods were immersed for 5 min in liquid nitrogen. A 0.75-s application pro-
duced the best results, with maximum clarity occurring at 3 months. Woolley
(1962) used a black Carter’s felt ink pencil to mark salamanders. Dilute acetic acid
or ammonium hydroxide was used to remove mucus on the salamander’s tail before
application. These marks lasted at least a month.
Taylor and Deegan (1982) injected a fluorescent pigment into red eft
(Notophthalmus viridescens) dermis using a compressed air spray gun with minimal
mortality. After 39 days, individuals began succumbing to water mold (Saprolegnia)
infections, but these could not be linked to the use of fluorescent branding.
Nishikawa and Service (1988) described a technique using dry fluorescent
dust applied with pressurized air on Plethodon jordoni and Plethodon glutino-
sus. The technique required four to 10 colours of inert fluorescent pigments,
canisters, a spray gun with 6.35 mm nozzle, a hose, a single-stage regulator,
and pressurized air. Animals were placed on a dry enamel pan, and the spray
gun was held 1 cm away from the skin surface. Marks were applied using lower
pressures (25 psi) for smaller animals and greater pressures (40 psi) for larger
animals, depending on the size of the fluorescent particles used. Three to four
marks could be applied to each animal on either side of the body (cranial and
caudal to the forelimb, mid-body, and cranial and caudal to the hind limb) for
a maximum of 10 possible locations. Nishikawa and Service (1988) reported
higher recapture rates than for any salamander toe-clipping studies performed
to that time.
8.4.1 Anurans
Kaplan (1958) used aluminum toe bands to tag frogs (“butt-end” bird band,
#1242, size 2½). These numbered, cylindrical bands were placed around a toe, and
the two ends were pressed together with pliers. The bands were tightened so as not
132 | Amphibian ecology and conservation
to restrict circulation, but they pierced the webbing of the foot. These remained
fixed indefinitely and caused no apparent hindrance in the movement of the frog.
A variety of tags and bands used for anurans is summarized in Table 8.5.
McAllister et al. (2004) recommended against the knee tags used by Watson
et al. (2003) in R. pretiosa due to skin and muscle lacerations in 33% of the
recaptured animals. Emlen (1968) reported no differences in behavior, mor-
tality, emigration rates, or weight loss between tagged and untagged American
bullfrogs (Rana catesbeiana). Due to soiling and staining, seasonal replacement
of waistbands was necessary. However, Robertson (1984) reported a negative
impact using Emlen’s (1968) waistband design on 10 species of Australian hylids
and leptodactylids.
Windmiller (1996) suspended yarn tags in a plastic bag filled with fluorescent
pigment which was pressed into the yarn, and attached to the dorsal integument
of juvenile ranid frogs. The pigment trails were followed at night by illuminat-
ing them with a longwave (366 nm) ultraviolet lamp (Blak-Ray ® model ML-49)
while wearing safety/ultraviolet-enhancing glasses. Buchan et al. (2005) used
soft VIAlpha Numeric Tags (US $1.00 per tag) and an injector (US$120) to
insert the tag. Buchan et al. (2005) found high survivorship, retention, and
readability of these tags in both the laboratory and field.
Table 8.5 Tagging and banding techniques used for marking amphibian species (as
reviewed in Ferner 2007 unless otherwise specified)
Species Description of tagging/banding technique
8.4.2 Salamanders
Woolley (1973) tagged Eurycea lucifuga and Eurycea longicauda with a sub-
cutaneous injection of two parts Liquitex acrylic polymer to one part distilled
water. This mixture was injected into the lateral proximal caudal region with
a 22-gauge hypodermic needle, leaving a mark 7–10 mm in diameter. Woolley
observed no adverse effects with this procedure and found slight fading in very
few individuals. The author suggested that a series of acrylic colours be used to
differentiate individuals.
Subcutaneous injections of fluorescent, elastomer dyes were used in Plethodon
vehiculum by Davis and Ovaska (2001). Red, orange, or yellow dyes (VIEs)
were injected into six locations in various combinations to create 816 unique
marks. The elastomer was mixed with a hardener following the manufacturer’s
instructions, and 0.1 mL was placed in a 0.3 mL syringe. The authors recom-
mend fluorescent marking or pattern mapping over toe-clipping in studies of
plethodontid salamanders.
Bailey (2004) evaluated VIE marking in Eurycea bislineata by documenting
mark retention, salamander survivorship, and growth rates. No adverse effects
and no loss of marks were found over the 11-month study. Misreading marks
in the field by observers was found to be a factor in some cases, so a training
period for investigators is recommended to minimize observer bias. Rittenhouse
et al. (2006) used the same technique with powdered fluorescent pigments with
newly metamorphosed Ambystoma maculatum, as described for Rana sylvatica
(see Table 8.4). No temporal effects of powder-dusting were detected.
8.4.3 Caecilians
Gower et al. (2006) reviewed the use of alphanumeric fluorescent tags (VIAlpha
tags) on the Indian caecilian Gegeneophis ramaswamii. They found that making
an incision with a scalpel blade before using the injector increased the efficiency
of tagging, and that anesthesia was needed to quiet specimens. Equipment ster-
ilization between uses minimizes the risk of spreading pathogens.
Table 8.6 Trailing devices used for marking amphibian species (as reviewed in Ferner
2007 unless otherwise indicated)
Species Description of trailing technique
difficult to attach, and they may snag on the substrate and vegetation. Still, they
are useful in certain applications. Table 8.6 reviews these techniques.
8.6.1 Anurans
While more commonly used in salamanders, the patterns of anurans may be
unique enough to make individual recognition possible. Wengert and Gabrial
(2006) used photographs of chin-spot patterns in Rana mucosa and found a high
success rate in reidentification over a 3-month study; changes in spot pattern
over time were observed so regular recaptures might minimize misidentifica-
tions using this technique. The effectiveness of using photographic identifica-
tion on long-term studies of adult anurans has yet to be determined.
Pointing out concerns in using artificial or invasive marks with the protected
Leiopelma archeyi (Archey’s frog) in New Zealand, Bradfield (2004) developed a
highly accurate photographic identification technique using natural markings.
This study is a model on how to develop and test a photographic pattern mapping
8 Working with post-metamorphic amphibians | 135
technique. A series of six digital photographs were taken from dorsal, ventral,
cranial (facial), caudal, right lateral, and left lateral views. Photographs using a
flash were most successful in discerning usable characteristics of the various black
markings. A filing system for the photos used subgroupings which substantially
reduced the amount of time needed to identify recaptures. Intra- and interobserver
consistency of identification was tested and found to be high. Overall, 99.2% of
identifications were successful once recaptures were assigned to their subgroups.
8.6.2 Salamanders
Naturally occurring variation in morphology was used for Triturus cristatus and
Triturus vulgaris by photographing belly patterns (Hagstrom 1973). With this
technique, one must be sure that the patterns are both recognizably different
and constant through time. Ontogenetic belly pattern variation of T. cristatus
was documented by Arntzen and Teunis (1993). Healy (1974) found the vari-
ation in dorsal spot patterns of Notophthalmus viridescens useful in identifying
individuals. This technique requires recapture and handling which may not be
consistent with the animals’ behavior.
Digital photographs of dorsal spot patterns were tested for their effectiveness
in the identification of Eurycea bislineata by Bailey (2004). Observers were given
color printouts of the dorsal view of each animal with the SVL written below
the image for size reference. The observers compared these photographs to test
animals and were timed as to how long it took them to identify each animal.
Photo-identification rates were very high (about 94% correct), but were poten-
tially lower if observers were also given un-marked (photographed) individuals
in the attempted comparisons. Use of Adobe Photoshop® was effective in com-
paring qualities of integument patterns of vertebrates in digital photographs
and may be useful in matching images (L. Rifai, personal communication).
Sweeney et al. (1994) describe the use of computer analysis for the belly patterns
of T. cristatus. Computer applications for animal pattern analysis can now be
found on websites such as www.conservationresearch.co.uk/.
Loafman (1991) photographed the natural variation of spot patterns in adult
Ambystoma maculatum to successfully identify about 97% of the recaptures.
Adult A. opacum were photographed in a box and successfully reidentified
using their distinctive bar pattern over a 1-year period; using head patterns
alone, 80% of the adults could be distinguished from one another (Doody
1995).
The use of spot patterns for identification of A. maculatum is reviewed by
Grant and Nanjappa (2006) with special attention to possible errors in this
technique. They quantified error rates in specific mapping techniques, modified
136 | Amphibian ecology and conservation
Table 8.7 Use of PIT tags for marking amphibian species (as reviewed in Ferner 2007
unless otherwise indicated)
Species Description of technique
Anurans
Bufo bufo Needle injection into pinched dorsum, massaged
caudally to base of spine
Limnodynastes peronii and Inserted with needle behind front limb along side of
Litoria aurea body
Rana catesbeiana, Rana clamitans, Needle injected into pinched ventral side
and Rana pipiens
Rana pretiosa Inserted through 2 mm incision on dorsum 1–2 cm
caudal to eyes, then massaged back along spine
Rana sylvatica Inserted through 2 m incision above the scapula;
“PIT pack” used for detection (Blomquist et al. 2008)
Rana temporaria Needle injection into pinched dorsum, massaged
caudally to base of spine
Caudates
Ambystoma opacum Inserted into body cavity at 3 mm incision 5 mm
cranial to hind limb
Taricha torosa Needle insertion into abdomen
8 Working with post-metamorphic amphibians | 137
8.7.1 Anurans
All anurans in the Camper and Dixon (1988) project had the PIT tags implanted
intra-abdominally using the implanter syringe with the needle dipped in 70%
ethanol before implantation. Wounds were cleaned after injection with 70% etha-
nol and sealed with Krazy Glue®. Synthetic liquid bandages can also be used.
Ireland et al. (2003) found that the tag migrated on its own to the caudal
lymph space between the hind legs in all but 5% of the frogs they studied. A
field scanner was used for identification and the authors believed the technique
to be the best available for marking medium-to-large-sized anurans. The pri-
mary drawback is the cost of the AVID microchips (≈US$8.00) and scanner
(≈$1000). McAllister et al. (2004) inserted 12-mm tags (125 kHz model) on
frogs 42 mm and massaged them back along the body to the base of the spine
to minimize the chance of their being lost through the incision prior to healing
(about 2 weeks).
Pyke (2005) provides a good review on the use of PIT tags and reports on the
marking of 3000 individuals of nine species. The tags were supplied by Trovan
and are “individually-packaged needles inside hermetically sealed packages.”
The resulting small wound was sealed with Vetbond® (n-butyl cyanoacrylate
adhesive) and fewer than 1% of the animals exhibited any sort of distress call
during the procedure. As with previous studies, Pyke (2005) found little imme-
diate impact or effect on reproduction and long-term survival.
8.7.2 Salamanders
Ott and Scott (1999) anesthetized salamanders (see Table 8.6) for about 10
min with 2-phenoxy ethanol (30 drops/500 ml distilled H2O) and sealed the
incision with New Skin® liquid bandage. Recovery from anesthesia was in con-
tainers of distilled H2O with the head above water for a 3–4-h period until the
animals appeared active when prodded gently. Where no adverse impacts of PIT
tags were reported, Ott and Scott (1999) pointed out the disadvantage of cost
and marking time.
8.7.3 Caecilians
PIT tags probably hold the most promise for marking caecilians, but it is only
appropriate for larger specimens and requires perfecting the injection technique.
Inasmuch as the long salamanders Siren and Amphiuma have been marked suc-
cessfully over a period of more than 5 years (C. K. Dodd Jr, personal commu-
nication), PIT tagging holds promise for studies of caecilians. Since caecilians
may be less likely to be recaptured than some other amphibians, the cost of lost
tags should be considered.
138 | Amphibian ecology and conservation
8.9 Recommendations
After extensive review of the literature and considering the criteria for selecting
a marking technique listed in the Introduction to this chapter, the following are
recommended.
8.10 References
Arntzen, J. W. and Teunis, S. F. M. (1993). A six year study on the population dynamics of
the crested newt (Triturus cristatus) following the colonization of a newly created pond.
Herpetological Journal, 3, 99–110.
Bailey, L. L. (2004). Evaluating elastomer marking and photo identification methods for
terrestrial salamanders: marking effects and observer bias. Herpetological Review, 35,
38–41.
Baker, J. and Gent, T. (1998). Marking and recognition of animals. In T. Gent and
S. Gibson (eds), Herpetofauna Worker’s Manual, pp. 45–54. Joint Nature Conservation
Committee, Peterborough.
Blomquist, S. M., Zydlewski, J. D., and Hunter, M. L. (2008). Efficacy of PIT tags for
tracking the terrestrial anurans Rana pipiens and Rana sylvatica. Herpetological Review,
39, 174–9.
Bradfield, K. S. (2004). Photographic Identification of Individual Archey’s Frogs, Leiopelma
archeyi, from Natural Markings. DOC Science Internal Series 191. New Zealand
Department of Conservation, Wellington.
Briggs, J. L. and Storm, R. L. (1970). Growth and population structure of the cascade
frog, Rana cascadae Slater. Herpetologica, 26, 283–300.
Buchan, A., Sun, L., and Wagner, R. S. (2005). Using alpha numeric fluorescent tags for
individual identification of amphibians. Herpetological Review, 36, 43–44.
Bull, E. L., Wallace, R., and Bennett, D. H. (1983). Freeze-branding: a long term mark-
ing technique on long-toed salamanders. Herpetological Review, 14, 81–2.
Camper, J. D. and Dixon, J. R. (1988). Evaluation of a microchip marking system for
amphibians and reptiles. Texas Parks and Wildlife Department, Research Publication,
7100–59, 1–22.
Clarke, R. D. (1972). The effect of toe clipping on survival in Fowler’s toad (Bufo wood-
housei fowleri). Copeia, 1972, 182–5.
Daugherty, C. H. (1976). Freeze branding as a technique for marking anurans. Copeia,
1976, 836–8.
Davis, T. M. and Ovaska, K. (2001). Individual recognition of amphibians: effects of
toe clipping and fluorescent tagging on the salamander Plethodon vehiculum. Journal of
Herpetology, 35, 217–25.
Donnelly, M. A., Guyer, C., Juterbock, J. E., and Alford, R. A. (1994). Techniques for mark-
ing amphibians. In W. R. Heyer, M. A. Donnelly, R. W. McDiarmid, L. C. Hayek, and
M. S. Foster (eds), Measuring and Monitoring Biological Diversity, Standard Methods for
Amphibians, appendix 2, pp. 277–284. Smithsonian Institution Press, Washington DC.
Doody, J. S. (1995). A photographic mark-recapture method for patterned amphibians.
Herpetological Review, 26, 19–21.
Elmberg, J. (1989). Knee-tagging: a new marking technique for anurans. Amphibia-
Reptilia, 10, 101–4.
Emlen, S. T. (1968). A technique for marking anuran amphibians for behavioral studies.
Herpetologica, 24, 172–3.
Ferner, J. W. (2007). A Review of Marking and Individual Recognition Techniques for
Amphibians and Reptiles. Herpetological Circular No. 35. Society for the Study of
Amphibians and Reptiles, Salt Lake City, UT.
140 | Amphibian ecology and conservation
Gower, D. J., Oommen, O. V., and Wilkinson, M. (2006). Marking amphibians with
alpha numeric fluorescent tags: caecilians lead the way. Herpetological Review, 37, 302.
Grant, E. H. C. and Nanjappa, P. (2006). Addressing error in identification of Ambystoma
maculatum (Spotted Salamanders) using spot patterns. Herpetological Review, 37, 57–60.
Griffiths, R. A. (1984). Seasonal behaviour and intrahabitat movements in an urban
population of smooth newts Triturus vulgaris (Amphibian: Salamandridae). Journal
of Zoology, 203, 241–51.
Hagstrom, T. (1973). Identification of newt specimens (Urodela, Triturus) by recording
the belly pattern and a description of photographic equipment for such registrations.
British Journal of Herpetology, 4, 321–6.
Halliday, T. (1995). More on toe-clipping. Froglog, 12, 2–3.
Healy, W. R. (1974). Population consequences of alternative life histories in Notophthalmus
v. viridescens. Copeia, 1974, 221–9.
Henle, K. and Veith, M. (eds) (1997). Naturschutzrelevante Methoden der
Feldherpetologie. Mertensiella, 7.
Ireland, D., Osbourne, N., and Berrill, M. (2003). Marking medium to large sized
anurans with Passive Integrated Transponder (PIT) tags. Herpetological Review, 34,
218–20.
Kaplan, H. M. (1958). Marking and banding frogs and turtles. Herpetologica, 14,
131–2.
Kuhn, J. (1994). Methoden der Anuren-Marierung für Frielandstudien: Übersicht-
Knie- Ringetiketten-Erfahrunger mit der Phalangenamputation. Zeitscrift für
Feldherpetologie, 1, 177–92.
Loafman, P. (1991). Identifying individual spotted salamanders by spot pattern.
Herpetological Review, 22, 91–2.
Lüddecke, H. and Amezquita, A. (1999). Assessment of disc clipping on the survival and
behavior of the Andean frog Hyla labialis. Copeia, 1999, 824–30.
Martof, B. S. (1953). Territoriality in the green frog, Rana clamitans. Ecology, 34,
165–74.
McAllister, K. R., Watson, J. W., Risenhoover, K., and McBride, T. (2004). Marking
and radiotelemetry of Oregon spotted frogs (Rana pretiosa). Northwestern Naturalist,
85, 20–5.
McCarthy, M. A. and Parris, K. M. (2004). Clarifying the effect of toe clipping on frogs
with Bayesian statistics. Journal of Applied Ecology, 41, 780–6.
Meyer, F. and Grosse, W.-R. (1997). Populationsökologische Studien an Amphibien
mit Hilfe der fotografischen Individualerkennung: Übersicht zur Methodik und
Anwendung bei der Kreuzkröte (Bufo calamita). Mertensiella, 7, 79–92.
Nishikawa, K. C. and Service, P. M. (1988). A fluorescent marking technique for individ-
ual recognition of terrestrial salamanders. Journal of Herpetology, 22, 351–3.
Orser, P. N. and Shine, D. J. (1972). Effects of urbanization on the salamander
Desmognathus fuscus fuscus. Ecology, 53, 1148–54.
Ott, J. A. and Scott, D. E. (1999). Effects of toe-clipping and PIT-tagging on growth and
survival in metamorphic Ambystoma opacum. Journal of Herpetology, 33, 344–8.
Parris, K. M. and McCarthy, M. A. (2001). Identifying effects of toe clipping on anuran
return rates: the importance of statistical power. Amphibia-Reptilia, 22, 275–89.
8 Working with post-metamorphic amphibians | 141
North America literature (e.g. Paton and Crouch 2002) and in the European
literature as “spawn clumps” (Griffith and Raper 1994), “batches” (Barton and
Rafinski 2006), “clutches” (Ficetola et al. 2006), or “egg masses” (Sofianidou
and Kyriakopoulou-Sklavounou 1983; Waringer-Löschenkohl 1991; Hartel
2008).
Nest counts can monitor populations in a similar manner to egg-mass counts.
Both types of count estimate annual breeding effort of populations and provide
an index of population size. In addition, both counts can test for effects of cli-
mate on population-level breeding effort (Corser and Dodd 2004), habitat rela-
tionships (Chalmers and Loftin 2006), population presence, and life-history
information (e.g. relationships between egg size and clutch size, and between
egg size and embryonic survival). Comparisons of egg size and clutch size among
populations can investigate maternal investment patterns as adaptations to local
environmental conditions. For nest counts, behavioural information can often
be gained, such as frequency of solitary and communal nesting and patterns of
nest attendance or brooding.
In the following sections we describe characteristics that make a species suit-
able for egg mass or nest counts. We also discuss factors that need to be consid-
ered when designing studies to collect and analyze these types of survey data.
Both egg mass and nest counts can be used to study habitat use over a variety of
spatial scales, thus are good candidates for landscape and metapopulation stud-
ies designed to manage and conserve at-risk species.
can search sites completely. For example, Egan and Paton (2004) found that
most pond-breeding amphibian sites were less than 0.05 ha in southern New
England, USA. Oviposition sites selected by some tropical species are extremely
small bodies of water, such as treeholes, aerial plants, or water-filled seedpods
(Wells 2007). However, although their oviposition habitat has been described,
to our knowledge no one has used egg-mass counts to monitor tropical amphib-
ian populations. Searching for egg masses of species that breed in large lakes is
probably not feasible due to low detection probabilities, and thus other monitor-
ing strategies should be employed. For stream-breeding species, egg masses can
be detected (e.g. Ascaphidae), although stream monitoring programs typically
search for juveniles and adults (Adams and Bury 2002).
Some anurans lay eggs in foamy nests on the water’s surface or in water-filled
basins (e.g. members of the subfamily Leptodactylinae; Wells 2007), which can
be highly visible on the surface of shallow, temporary pools. Also, rhacophorid
and hyperoliid frogs build foam nests in trees and other substrates. Thus, one
possible technique to monitor their populations might be to attempt to count
foam nests, but given the difficulties of finding the nests of most foam-nesting
species, this is probably not a cost-effective technique for most species.
There are many arboreal tropical species that deposit eggs on leaves, branches,
or tree trunks above standing water (see summary in Wells 2007). Neotropical
hylids oviposit on upper surfaces of leaves and develop rapidly, where tadpoles
fall into the water below to complete metamorphosis. Phyllomedusine hylid
frogs from Central and South America deposit eggs on leaves and other sub-
strates. Although monitoring schemes have not incorporated egg mass or nest
counts of these arboreal breeding species, it is something that tropical biologists
could consider.
Table 9.1 Examples of studies where anuran populations were assessed by counting
egg masses for more than 2 years.
Species Country Habitat Years Source
Table 9.2 Examples of studies where urodele populations were assessed by counting
egg masses.
Species Country Habitat Years Source
extensive surveys of nests and attending females in the same populations begin-
ning in the early 1970s and then again in the 1990s. These data were critical in
showing that at least three large populations had severely declined, assuming
that each pond surveyed represented a separate population or deme.
We know of no studies that have used anuran nest counts to track changes
in population size. However, nest sites of Eleutherodactylus coqui have been
studied extensively in Puerto Rico (Stewart and Pough 1983; Townsend 1989;
Townsend and Stewart 1994). Counting nests may track E. coqui population
trends because researchers were able to confidently check all suitable potential
nest sites. For E. coqui, preferred nest sites are elevated, enclosed, and often occur
in curled dead leaves of Cecropia peltata and fronds of the palm Prestoea mon-
tana. In an experimental study, adding artificial nest sites increased population
density, including the number of clutches, indicating that preferred nest sites
are limited (Stewart and Pough 1983). One advantage of diurnal nest-count
surveys is that they can be conducted under most weather conditions, whereas
frog-call surveys in many areas are best done at night under rainy, moist, or
humid conditions.
150 | Amphibian ecology and conservation
Fig. 9.2 Egg masses of spotted salamanders (Ambystoma maculatum; upper panel)
and wood frogs (Rana sylvatica; lower panel) are commonly found in seasonal pools
in eastern North America. Photographs by Scott Egan.
9 Egg mass and nest counts | 151
a longer survey window. For example, in Rhode Island, R. sylvatica eggs may
take over 20 days to hatch in cold ponds (Crouch and Paton 2000). In contrast,
R. catesbeiana eggs that are deposited in June may hatch in under 3 days.
A. maculatum egg masses in four seasonal ponds. Crouch and Paton (2000) esti-
mated about 12% variation by two independent observers surveying R. sylvatica
egg masses. Egan (2001) found that independent double-observer counts of A.
maculatum egg masses varied by 25%, while R. sylvatica egg-mass counts varied
by 11%. All these species had relatively high detection probabilities (84–100%,
Grant et al. 2005). Based on work by Grant et al. (2005) and Williams et al.
(2002), studies using multiple observers need to incorporate an observer param-
eter when modeling trend estimates.
although Harris et al. (1995) found the females could be disturbed daily in meso-
cosms without causing desertion. Some species, such as the anuran Cophixalus
ornatus, appear to treat humans as predators and will lunge toward human fin-
gers and not desert (Felton et al. 2006). Preliminary studies should be conducted
to assess the effects of nest discovery and degree of nest disturbance. Quickly
checking for nest presence will be less likely to cause desertion than temporar-
ily removing and measuring the attendant and counting embryos. In many spe-
cies, desertion by the nest attendant is associated with reduced embryonic success
(Harris et al. 1995). Therefore, if the objective is to monitor or conserve popula-
tion size, then it is important to not cause desertion.
about to occur. Care must be taken at this stage because movement of late-stage
embryos can induce hatching. Estimates of hatching or embryonic success typ-
ically take several visits to nest sites.
where N represents the population size (number of the egg masses) at time t.
Density dependence can be tested by regressing
N (growth rate) against the
number of egg masses (see also Hartel 2008 for R. dalmatina and Meyer et al.
1998 for R. temporaria).
If one is interested in annual changes in site occupancy (i.e. the proportion
of ponds with breeding populations), then recent developments by MacKenzie
et al. (2006) could be of interest. This is relevant for species with lower detection
probabilities, particularly for species or areas where it is difficult to determine
total population size based on total egg mass or nest counts.
9.16 Summary
Egg-mass counts and nest counts are both viable techniques that have been used
successfully by herpetologists to monitor population trends of amphibian spe-
cies. Egg mass and nest counts are particularly powerful because one can model
annual breeding effort or true population sizes, rather than crude indices of popu-
lation sizes that are provided by other techniques such as calling surveys. In add-
ition, for conservation biologists interested in assessing factors that might affect
populations, egg counts and nest counts are useful tools. In particular, biologists
have been using egg counts recently to assess the effects of habitat structure at
a variety of spatial scales on estimates of annual reproductive effort (e.g. Egan
2001; Egan and Paton 2004; Skidds et al. 2007). Biologists working in the tropics,
158 | Amphibian ecology and conservation
where monitoring information is critical given recent disease events there, might
be particularly interested in incorporating egg counts or nest counts in monitoring
schemes as a supplement to existing calling surveys and visual encounter surveys.
9.17 References
Adams, M. H. and Bury, R. B. (2002). The endemic headwater stream amphibians of
the American Northwest: associations with environmental gradients in a large forested
preserve. Global Ecology and Biogeography, 11, 169–78.
Adams, M. H., Bury, R. B., and Swarts, S. A. (1998). Amphibians of the Fort Lewis
Military Reservation, Washington: Sampling techniques and community patterns.
Northwestern Naturalist, 79, 12–18.
Alford, R. A. and Richards, S. J. (1999). Global amphibian declines: a problem in applied
ecology. Annual Review of Ecology and Systematics, 30, 133–56.
Andreone, F., Cadle, J. E., Cox, N., Glaw, F., Nussbaum, R. A., Raxworthy, C. J., Stuart,
S. N., Vallan, D., and Vences, M. (2005). Species review of amphibian extinction
risks in Madagascar: conclusions from the global amphibian assessment. Conservation
Biology, 19, 1790–1802.
Bailey, L. L., Simons, T. R., and Pollock, K. H. (2004). Estimating site occupancy
and species detection probability parameters for terrestrial salamanders. Ecological
Applications, 14, 692–702.
Banning, J. L., Weddle, A. L., Wahl III, G. W., Simon, M. A., Lauer, A., Walters, R. L.,
and Harris, R. N. (2008). Antifungal skin bacteria, embryonic survival, and communal
nesting in four-toed salamanders, Hemidactylium scutatum. Oecologia, 156, 423–9.
Barton, K. and Rafinski, J. (2006). Co-occurrence of Agile Frog (Rana dalmatina Fitz. in
Bonaparte) with Common Frog (Rana temporaria L.) in breeding sites in southeastern
Poland. Polish Journal of Ecology, 54, 151–7.
Bickford, D. (2004). Differential parental care behaviors of arboreal and terrestrial micro-
hylid frogs from Papua New Guinea. Behavioral Ecology and Sociobiology, 55, 402–9.
Breitenbach, G. L. (1982). The frequency of joint nesting and solitary brooding in the
salamander, Hemidactylium scutatum. Journal of Herpetology, 16, 341–6.
Brodman, R. (1995). Annual variation in breeding success of two syntopic species of
Ambystoma salamanders. Journal of Herpetology, 29, 111–13.
Buckley, J. and Beebee, T. J. C. (2004). Monitoring the conservation status of an endan-
gered amphibian: the natterjack toad Bufo calamita in Britain. Animal Conservation,
7, 221–8.
Bury, R. B. and Adams, M. J. (2000). Inventory and monitoring of amphibians in North
Cascades and Olympic National Parks, 1995–1998. Final Report. Forest and Rangeland
Ecosystem Science Center, Corvallis, OR.
Calhoun, A. J. K., Walls, T. E., Stockwell, S. S., and McCollough, M. C. (2003).
Evaluating vernal pools as a bases for conservation strategies: a Maine case study.
Wetlands, 23, 70–81.
Chalmers, R. J. and Loftin, C. S. (2006). Wetland and microhabitat use by nesting four-
toed salamanders in Maine. Journal of Herpetology, 40, 478–85.
9 Egg mass and nest counts | 159
Cooke, A. S. (1985). The deposition and fate of spawn clumps of the common frog
Rana temporaria at a site in Cambridgeshire, 1971–1983. Biological Conservation, 32,
165–87.
Corn, P. S. and Livo, L. J. (1989). Leopard frog and wood frog reproduction in Colorado
and Wyoming. Northwestern Naturalist, 70, 1–9.
Corser, J. D. (2001). Decline of disjunct green salamander (Aneides aeneus) populations
in the southern Appalachians. Biological Conservation, 97, 119–26.
Corser, J. D. and Dodd Jr, C. K. (2004). Fluctuations in a metapopulation of nesting four-
toed salamanders, Hemidactylium scutatum, in the Great Smoky Mountains National
Park, USA, 1999–2003. Natural Areas Journal, 24, 135–40.
Crochet, P.-A., Chaline, O., Cheylan, M., and Guillaume, C. P. (2004). No evidence
of general decline in an amphibian community of Southern France. Biological
Conservation, 119, 297–304.
Crouch, W. B. and Paton, P. W. C. (2000). Using egg mass counts to monitor wood frog
populations. Wildlife Society Bulletin, 28, 895–901.
Crouch, W. B. and Paton, P. W. C. (2002). Assessing the use of call surveys to monitor
breeding anurans in Rhode Island. Journal of Herpetology, 36, 185–92.
DeAlmeida Prado, C. P., Uetanabaro, M., and Lopes, F. S. (2000). Reproductive strat-
egies of Leptodactylus chaquensis and L. podicipinus in the Pantanal, Brazil. Journal of
Herpetology, 34, 135–9.
Dodd Jr, C. K. (2003). Monitoring amphibians in Great Smoky Mountains National Park.
Circular Number 1258. US Geological Survey, Tallahassee, FL.
Doody, J. S. and Young, J. E. (1995). Temporal variation in reproduction and clutch mor-
tality of leopard frogs (Rana utricularia) in South Mississippi. Journal of Herpetology,
29, 614–16.
Driscoll, D. A. (1998). Counts of calling males as estimates of population size in
the endangered frogs Geocrina alba and G. vitellina. Journal of Herpetology, 32,
475–81.
Egan, R. S. (2001). Within-pond and Landscape-level Factors Influencing the Breeding
Effort of Rana sylvatica and Ambystoma maculatum. MSc thesis, University of Rhode
Island, Kingston, RI.
Egan, R. S. and Paton, P. W. C. (2004). Within-pond parameters affecting oviposition by
wood frogs and spotted salamanders. Wetlands, 24, 1–13.
Elmberg, J. (1990). Long-term survival, length of breeding season, and operational sex
ratio in a boreal population of common frogs, Rana temporaria L. Canadian Journal of
Zoology, 68, 121–7.
Felton, A., Alford, R. A., Felton, A. M., and Schwarzkopf, L. (2006). Multiple mate
choice criteria and the importance of age for male mating success in the microhylid
frog, Cophixalus ornatus. Behavioral Ecology and Sociobiology, 59, 786–95.
Ficetola G. F., Valota, M., and de Bernardi, F. (2006). Temporal variability of spawning
site selection in the frog Rana dalmatina: consequence for habitat management. Animal
Biodiversity and Conservation, 29, 157–63.
Freda, J., Sadinski, W. J., and Dunson, W. A. (1991). Long term monitoring of amphib-
ian populations with respect to the effects of acidic deposition. Water, Air, and Soil
Pollution, 55, 445–62.
160 | Amphibian ecology and conservation
Kusano, T., Sakai, A., and Hatanaka, S. (2005). Natural egg mortality and clutch size
of the Japanese treefrog Rhacophorus arboreus (Amphibia: Rachophoridae). Current
Herpetology, 25, 79–84.
Lips, K. R., Reaser, J. K., Young, B. E., and Ibáñez, R. (2001). Amphibian monitoring in
Latin America: a protocol manual. Society for the Study of Amphibians and Reptiles,
Herpetological Circular No. 30, 1–115.
Loman, J. and Andersson, G. (2007). Monitoring brown frogs Rana arvalis and Rana
temporaria in 120 south Swedish ponds 1989–2005. Mixed trends in different habitats.
Biological Conservation, 35, 46–56.
MacKenzie, D. I., Nichols, J. D., Royle, J. A., Pollock, K. H., Bailey, L. L., and Hines, J. E.
(2006). Occupancy Estimation and Modeling: Inferring Patterns and Dynamics of Species
Occurrence. Elsevier Press, Amsterdam.
Marsh, D. M. (2001). Fluctuations in amphibian populations: a meta-analysis. Biological
Conservation, 101, 327–35.
Martin, A. A. and Cooper, A. K. 1972. The ecology of terrestrial anuran eggs, genus
Crinia (Leptodactylidae). Copeia, 1972, 163–9.
Mazerolle, M. J., Bailey, L. L., Kendall, W. L., Royle, J. A., Converse, S. J., and Nichols, J. D.
(2007). Making great leaps forward: accounting for detectability in herpetological field
studies. Herpetologica, 41, 672–89.
Meyer, A. H., Schmidt, B. R., and Grossenbacher, K. (1998). Analysis of three amphib-
ian populations with quarter-century long time-series. Proceedings of the Royal Society of
London Series B Biological Sciences 265, 523–8.
Muths, E., Corn, P. S., Pessier, A. P., and Green, D. E. (2003). Evidence for disease-
related amphibian decline in Colorado. Biological Conservation, 110, 357–65.
Neckel-Oliveira, S. and Wachlevski, M. (2004). Predation on the arboreal eggs of three
species of Phyllomedusa in Central Amazônia. Journal of Herpetology, 38, 244–8.
Nichols J. D., Hines, J. E., Sauer, J. R., Fallon, F., Fallon, J., and Heglund, P. J. (2000).
A double-observer approach for estimating detection probability and abundance from
avian point counts. Auk, 117, 393–408.
Parmelee, J. R., Knutson, M. G., and Lyon, J. E. (2002). A Field Guide to Amphibian
Larvae and Eggs of Minnesota, Wisconsin, and Iowa. Information and Technology
Report USGS/BRD/ITR-2002-0004. US Geological Survey, Washington DC.
Paton, P. W. C. and Crouch, W. B. (2002). Using the phenology of pond-breeding
amphibians to develop conservation strategies. Conservation Biology, 16, 194–204.
Pechmann, J. H. K., Scott, D. E., Gibbons, J. W., and Semlitsch, R. D. (1989). Influence
of wetland hydroperiod on diversity and abundance of metamorphosing juvenile
amphibians. Wetland Ecology and Management, 1, 3–11.
Peterson, M. G. (2000). Nest, but not egg, fidelity in a terrestrial salamander. Ethology,
106, 781–94.
Petranka, J. W. (1998). Salamanders of the United States and Canada. Smithsonian
Institution Press, Washington DC.
Petranka, J. W. and Holbrook, C. T. (2006). Wetland restoration for amphibians: should
local sites be designed to support metapopulations or patchy populations? Restoration
Ecology, 14, 404–11.
Petranka, J. W., Smith, C. K., and Scott, A. F. (2004). Identifying the minimal demographic
unit for monitoring pond-breeding amphibians. Ecological Applications, 14, 1065–78.
162 | Amphibian ecology and conservation
Regester, K. J. and Woosley, L. B. (2005). Marking salamander egg masses with Visible
Fluorescent Elastomer: retention time and effect on embryonic development. American
Midland Naturalist, 152, 52–60.
Richards, S. J. and Alford, R. A. (1992). Nest construction by an Australian rainforest
frog of the Litoria lesueuri complex (Anura: Hylidae). Copeia, 1992, 1120–3.
Richter, S. C. (2000). Larval caddisfly predation on the eggs and embryos of Rana capito
and Rana sphenocephala. Journal of Herpetology, 34, 590–3.
Richter, S. C., Young, J. E., Johnson, G. N., and Seigel, R. A. (2003). Stochastic variation
in reproductive success of a rare frog, Rana sevosa: implications for conservation and for
monitoring amphibian populations. Biological Conservation, 111, 171–7.
Roberts, W. E. (1994). Explosive breeding aggregations and parachuting in a neotropical
frog, Agalychis saltator (Hylidae). Journal of Herpetology, 28, 193–9.
Rödel, M. O. and Ernst, R. (2004). Measuring and monitoring amphibian diversity in
tropical forests. I. An evaluation of methods with recommendations for standardiza-
tion. Ecotropica, 10, 1–14.
Seale, D. (1982). Physical factors influencing oviposition by the wood frog, Rana sylvatica,
in Pennsylvania. Copeia, 1982, 627–35.
Sherman, C. K. and Morton, M. L. (1993). Population declines of Yosemite Toads in the
eastern Sierra Nevada of California. Journal of Herpetology, 27, 186–98.
Skidds, D. E., Golet, F. C., Paton, P. W. C., and Mitchell, J. C. (2007). Habitat correlates
of reproductive effort in wood frogs and spotted salamanders in an urbanizing water-
shed. Journal of Herpetology, 41, 439–50.
Sofianidou, T. S. and Kyriakopoulou-Sklavounou, P. (1983). Studies on the biology of
the frog, Rana dalmatina, Bonap. during the breeding season in Greece. Amphibia-
Reptilia, 4, 125–36.
Stewart, M. M. and Pough, F. H. (1983). Population density of tropical forest frogs: rela-
tion to retreat sites. Science, 221, 570–2.
Storfer, A. (2003). Amphibian declines: future directions. Diversity and Distributions, 9,
151–63.
Summers, K., McKeon, C. S., and Heying, H. (2006). The evolution of parental care and
egg size: a comparative analysis in frogs. Proceedings of the Royal Society of London Series
B Biological Sciences 273, 687–92.
Thurley, T. and Bell, B. D. (1994). Habitat distribution and predation on a western popu-
lation of terrestrial Leiopelma (Anura: Leiopelmatidae) in the northern King Country,
New Zealand. New Zealand Journal of Zoology, 21, 431–6.
Townsend, D. S. (1989). The consequences of microhabitat choice for male reproductive
success in a tropical frog (Eleutherodacylus coqui). Herpetologica, 45, 451–8.
Townsend, D. S. and Stewart, M. M. (1994). Reproductive ecology of the Puerto Rican
frog Eleutherodactylus coqui. Journal of Herpetology, 28, 34–40.
University of Canberra (2003). Survey Standards for Australian Frogs. Final Report.
Applied Ecology Research Group, Canberra.
Veith, M., Lötters, S., Andreone, F., and Rödel, M.-O. (2004). Measuring and monitor-
ing amphibian diversity in tropical forests. II. Estimating species richness from stand-
ardized transect censing. Ecotropica, 10, 85–99.
Vonesh, J. (2000). Dipteran predation on the arboreal eggs of four Hyperolius frog species
in western Uganda. Copeia, 2000, 560–6.
9 Egg mass and nest counts | 163
Wahl III, G. W., Harris, R. N., and Nelms, T. (2008). Nest site selection and embryonic
survival in four-toed salamanders, Hemidactylium scutatum (Caudata: Plethodontidae).
Herpetologica, 64, 12–19.
Waringer-Löschenkohl, A. (1991). Breeding ecology of Rana dalmatina in lower Austria:
a 7-year study. Alytes, 9, 121–34.
Weir, L. A., Royle, J. A., Nanjappa, P., and Jung, R. E. (2005). Modeling anuran detection
and site occupancy on North American Amphibian Monitoring Program (NAAMP)
routes in Maryland. Journal of Herpetology, 39, 627–39.
Wells, K. D. (2007). The Ecology and Behaviour of Amphibians. University of Chicago
Press, Chicago, IL.
Werner, E. E., Yurewicz, K. L., Skelly, D. K., and Relyea, R. A. (2007). Turnover in
an amphibian metacommunity: the role of local and regional factors. Oikos, 116,
1713–25.
Williams, L. R. (2005). Restoration of ponds in a landscape and changes in Common
Frog (Rana temporaria) populations, 1983–2005. Herpetological Bulletin, 94, 22–9.
Williams, B. K., Nichols, J. D., and Conroy, M. J. (2002). Analysis and Management of
Animal Populations: Modeling, Estimation and Decision Making. Academic Press, San
Diego, CA.
Windmiller, B. S. (1996). The Pond, the Forest, and the City: Spotted Salamander Ecology
and Conservation in a Human-dominated Landscape. PhD dissertation, Tufts University,
Medford, MA.
Zheng, Y. and Fu, J. (2006). Making a doughnut-shaped egg mass: oviposition behaviour
of Vibrissaphora boringiae (Anura: Megophryidae). Amphibia-Reptilia, 28, 309–11.
164 | Amphibian ecology and conservation
Appendix 9.1 Oviposition strategies of anuran families of the world, with comments on
whether egg counts could be used monitor their populations (modified from Wells 2007).
Family Distribution Deposition site Eggs could be counted?
10.1 Introduction
Dietary information is essential to understand amphibian life history, popu-
lation fluctuations, and the impact of habitat modification, and to develop
conservation strategies (Anderson 1991). In ecology, trophic relationships
represent a functional connection between different taxa and are a key
subject in many aut- and synecological studies. Studies on diets are essen-
tial for assessments of energy flow and food webs in ecological communi-
ties. Variables involved in an animal’s diet are seldom obvious and include
behavioral, physiological, and morphological features of both predator and
prey (Simon and Toft 1991). Therefore, dietary information from locally
imperilled species is a necessary component in designing management and
conservation programs.
Traditionally, amphibians are described as generalist predators feeding
mainly on arthropods, but mollusks, annelids, and even small vertebrates
are also frequently consumed. Amphibians are able to distinguish between
different prey types allowing for different degrees of feeding specialization
(Freed 1982). Dietary specialization is often associated with morphological,
physiological, and behavioral characteristics that facilitate location, identifi-
cation, capture, ingestion, and digestion of prey items. Feeding dynamics can
vary considerably between species, but differences between the three major
amphibian lineages and between species feeding on land and in water become
obvious. Therefore, the feeding mechanisms and dietary specialties of the dif-
ferent groups are discussed separately (for amphibian larvae see Chapter 5).
in the diet analysis. Stomach flush should be used more widely and stomach
dissection should be thrown away ultimately.” Since prey size, composition,
and diversity are often correlated with the sex and size of the predator (Lima
and Magnusson 2000), it is important to note the snout–vent length and the
gender of each specimen.
Fig. 10.1 Complete flushing set consisting of two containers, a spatula, forceps, two
syringes, an infusion tube (for small frogs) or a rubber or PVC hose, small airtight vials,
and 70% ethanol solution.
10 Dietary assessments of adult amphibians | 171
can be directly washed with 70% ethanol from the gauze into a small vial. Frogs
should be released where they were captured as soon as possible after stomach
flushing to facilitate a return to normal foraging activity and behavior.
groups. Some authors categorize the prey items into ecological guilds, such as
‘rather active’ or ‘sedentary’ taxa, or flying, ground-dwelling, or burrowing spe-
cies. These methods facilitate assessments of foraging strategies in the absence of
behavioral information, since the quantity and volume of prey items belonging
to different guilds allow researchers to draw conclusions about the feeding strat-
egy of the amphibian predator.
Even if researchers manage to capture amphibians shortly after they have
started foraging, and even if the flushing procedure is applied shortly thereafter,
researchers frequently find that parts of the stomach contents have already been
digested. Soft tissues generally are digested more rapidly than strongly chitinized
parts, thus making prey items break into fragments. Hirai and Matsui (2001)
noticed that certain specific parts of prey remain intact even after digestion has
started. Examples are the heads of ants or wasps, abdomen and elytra of beetles,
thoraxes of flies, moths, beetles, or Orthopterans (grasshoppers, crickets, katy-
dids), forewings of flies and bees, and the saltatorial legs of Orthopterans. Hirai
and Matsui (2001) calculated regression formulae between body parts and total
body length, which can be used as a rough estimate of prey size. To be complete,
however, information on the size of undamaged items is required for a thorough
assessment of prey use.
In most dietary studies, information about the number, frequency, and vol-
ume and/or mass of the prey items of each taxonomic or ecological group is
presented. These data allow direct comparison between different prey groups.
The mass of prey items can be assessed with an analytic balance. Here, storage
in formalin is superior to storage in ethanol since the later dehydrates the prey
items. Furthermore, since stomach contents are commonly wet, it is important
to remove excess moisture from the prey item before weighing to avoid bias.
The best way would be to store the prey items in a drying oven at approximately
50°C until a constant dry mass is achieved. Information about prey number,
frequency, and mass are relatively easy to obtain, as there are a number of differ-
ent techniques used to measure the volume of prey items. Volume and mass of
prey are directly correlated with the amount of energy they can provide; thus,
this information is most appropriate in combination with the identification of
important prey groups. Most researchers attempt to reconstitute volumes from
linear measures using formulae for geometric shapes; body appendices of prey
items should be ignored when measurements are taken. The most commonly
used formula is the one for ellipsoid bodies (Colli and Zamboni 1999):
2
4 L ⎛ W ⎞
V ⎜ ⎟
3 2⎝ 2 ⎠
174 | Amphibian ecology and conservation
The volume of prey items also can be measured directly by introducing them
into a measuring cylinder graduated in intervals of 0.1 ml. Water is added from a
pipette graduated in intervals of 0.01 ml until the item is completely covered and
the water has reached the next highest graduation in the cylinder. The volume
can then be taken as the water level in the cylinder minus the amount added
with the pipette (Magnusson et al. 2003; Cuello et al. 2006).
traps (Brower et al. 1998) in addition to pitfall traps and combining the results.
These traps sample sedentary taxa in addition to mobile fauna. Both methods
underestimate flying prey items.
In grassy or herbaceous habitats, the composition and abundance of prey
items can be estimated along transects using linear sweeps with insect nets.
Each sweep sample should be transferred into a Berlese–Tullgren trap for at
least 30 min to separate prey items from litter (leaves, seeds, sticks). Properly
conducted, this method can provide a comprehensive assessment of inverte-
brate availability. It should be noted that highly mobile flying insects such as
Odonates (dragonflies) might be underrepresented. Furthermore, the time and
temperature at which the sweeps are conducted can have a strong influence on
the result depending on the activity periods of the insects.
Arboreal habitat can be sampled most easily using fogging techniques.
Applying insecticides through fogging is widely used in pest control, and a wide
range of commercial fogging machines is available. Fumigated insects falling
from tree canopies are collected on plastic sheets laid out on the ground or by
plastic trays or aluminum funnels (for a short review, see Adis et al. 1998).
Insecticide fogging has become established as a standard technique since the
1980s for assessing arthropod diversity, especially in tropical forest canopies
(Adis et al. 1998). A variety of “knockdown” insecticides have been tested, but
natural pyrethrum is now established as a standard as it allows for collection of
live specimens and its biodegradability facilitates application in ecologically sen-
sitive habitats. Adis et al. (1998) suggested recommendations to standardize fog-
ging procedures. When applying fogging procedures, researchers must consider
that the entire vertical habitat that is sampled may overestimate the prey avail-
able to amphibians, which inhabit only parts of the vertical habitat structure.
Aquatic habitats can be sampled using a fine mashed net, such as the D-frame
kick net or surber stream-bottom sampler. The D-frame kick net is either cone-
or bag-shaped for the capture of organisms and commonly used by lying the
spine of the net firmly on the ground and hauling it a specific number of times
along a specific distance. This type of net is easy to transport and can be used in
a variety of habitat types. However, the D-frame kick net must have a defined or
delimited area that is sampled/kicked in order to compare results obtained from
different sites or samples. Cao et al. (2005) investigated the performance and
comparability between surber and D-frame kick-net samples and found that
subsamples with the same number of individuals were highly and consistently
comparable between sampling devices. Rocky stream habitats are best sampled
by two persons. One collector should place a fine mashed net so that its bottom
lies on the ground. Another collector should turn over as many rocks as possible
176 | Amphibian ecology and conservation
upstream from the net for 1 min. Wiping off clinging insects from rocks and
branches may enhance collection quantities and species diversity.
Once relative abundances of the prey items in the habitat and in the stomach
contents are identified, an electivity index can be calculated, as proposed by
Jacobs (1974):
Rk Pk
D
(Rk Pk ) (2Rk Pk )
whereby Rk is the proportion of prey category k in stomach contents, and Pk is the
proportion of prey category k in the environment. D varies from 1 (complete
selection of preference for prey category k) through 0 (prey category k is taken in
the same proportion as found in the environment) to 1 (prey category k is pre-
sent in the environment but absent in the diet). When interpreting the results,
biases caused by sample methods need to be considered.
B′
∑ p k
100R
where pk is the occurrence frequency of kth prey, is the standard deviation of
the occurrence frequencies, and R is the number of prey species exploited by the
guild.
In order to compare the niche breath between multiple species or changes in
the niche breadth within different seasons, the Shannon–Weaver diversity index
H (Weaver and Shannon 1949; see Chapter 18) can be used:
H ′ ∑ pi In Pi
i
where pi is the relative abundance of each prey category, calculated as the pro-
portion of prey items of a given category to the total number of prey items.
10 Dietary assessments of adult amphibians | 177
∑P P ij ik
O jk n
i
n
∑P ∑P
i
ij
2
i
2
ik
Here, Ojk is niche overlap and pij and pik represent the proportions of the ith
resource used by the jth and kth species.
over long periods of time. Isotope signatures—that is, ratios of heavier (e.g.
15
N and 14C) and lighter (14N and 13C) isotopes—of consumers are generally
similar to those of their food and can thus provide insights into food-web struc-
ture. Fatty acids are essential energy storage lipids in consumers that cannot
be readily synthesized and must be obtained from their food. Differences in
fatty acid composition among available food types are reflected in fatty acid
compositions in consumers; therefore, profiles can be good indicators of diets.
For both techniques, tissue samples (or in the case of small individuals or food
items, the whole organism) are simply collected and preserved in a manner that
does not alter isotopic or fatty acid composition, depending on the analyses to
be performed, and then sent to an analytical laboratory for analyses (for a more
detailed description of these techniques see Chapter 5).
sizes are an important issue for dietary analyses. Considering that diet may vary
between the sexes, adequate samples from both genders should be obtained.
When working with stomach contents, researchers should be aware that they
will probably not get a picture of what prey items the focal species feeds upon,
but rather a picture of the items that are hard to digest. Frogs are capable of
digesting prey rapidly: Blanchard’s cricket frogs (Acris crepitans blanchardi) fed
dyed Drosophila (fruit flies) and killed 7 h later had no trace of prey remain-
ing in their stomachs (Johnson and Christiansen 1976). In another study with
A. crepitans immediately preserved frogs had a mean of seven prey items per
stomach, whereas frogs preserved 6 h after capture revealed 3.6 prey items per
stomach (Caldwell 1996). In a preliminary study with 20 Physalaemus lisei,
frogs were fed three different prey categories (Oligochaeta, Curculionid beetles,
and Aranea) at the same time, and four frogs were stomach-flushed every 4 h.
Oligochaeta (worms) could not be detected after 4 h, whereas the pronotum of
spiders was still undigested after 8 h and the beetle elytra were still present after
12 h (M. Solé, unpublished results). In conclusion, researchers will underesti-
mate the prey categories that basically consist of soft tissue (worms, mollusks,
and fly larvae) and overestimate those with strongly keratinized bodies (ants and
beetles). One way in which biologists can mediate this problem is to carefully
choose the sampling period according to the feeding period of each species.
10.5 Conclusions
Dietary information is pivotal for successful development of conservation strat-
egies on species level and the understanding of ecosystem function. Unfortunately,
this kind of information is not available for the vast majority of taxa and is often
incomplete. Seasonal variations, ontogenetic shifts, and relationships between
site-specific prey availability, presence of potential competitors, and diet compos-
ition are commonly not addressed. Future studies should focus on these issues.
In the light of global amphibian declines an Amphibian Conservation Action
Plan was formulated by the International Union for Conservation of Nature to
prevent further biodiversity loss and captive breeding programmes are being
developed for the most threatened species (Gascon et al. 2007). Such approaches
depend crucially on autecological knowledge such as information on the diet of
the target species. Here, every kind of information can be valuable.
10.6 Acknowledgments
We are grateful to Ken Dodd, Marian Griffey, and Matt Whiles for helpful com-
ments on the manuscript and to Klaus Riede for fruitful discussions concerning
quantitative insect sampling methods and help with the literature. The work of
DR was funded by the Graduiertenförderung des Landes Nordrhein-Westfalen.
10.7 References
Adis, J., Basset, Y., Floren, A., Hammond, P. M., and Linsenmair, K. E. (1998). Canopy fog-
ging on an overstorey tree- recommendations for standardization. Ecotropica, 4, 93–7.
10 Dietary assessments of adult amphibians | 181
Freed, A. N. (1982). A treefrog’s menu: selection for an evening’s meal. Oecologia, 53,
20–6.
Gascon, C., Collins, J. P., Moore, R. D., Church, D. R., McKay, J. E., and Mendelson III,
J. R. (2007). Amphibian Conservation Action Plan - Proceedings: IUCN/SSC Amphibian
Conservation Summit 2005. IUCN, Gland.
Gladfelter, W. B. and Johnson, W. S. (1983). Feeding niche separation in a guild of trop-
ical reef fish (Holocentridae). Ecology, 64, 552–63.
Hart, R. K., Calver, M. C., and Dickman, C. R. (2002). The index of relative import-
ance: an alternative approach to reducing bias in descriptive studies of animal diets.
Wildlife Research, 29, 415–21.
Hirai, T. and Matsui, M. (2001). Attempts to estimate the original size of partly digested
prey recovered from stomachs of Japanese anurans. Herpetological Review, 32, 14–16.
Houston, W. W. K. (1973). The food of the common frog, Rana temporaria, on high
moorland in northern England. Journal of Zoology, 171, 153–65.
Hulbert, S. H. (1978). The measurement of niche overlap and some relatives. Ecology,
59, 67–77.
Hurtubia, J. (1973). Trophic diversity measurement in sympatric predatory species.
Journal of Ecology, 549, 885–90.
Jacobs, J. (1974). Quantitative measurement of food selection: a modification of the for-
age ration and Ivlev’s electivity index. Oecologia, 14, 413–17.
Johnson, B. K. and Christiansen, J. L. (1976). The food and food habits of Blanchard’s
cricket frog, Acris crepitans blanchardi (Amphibia, Anura, Hylidae), in Iowa. Journal of
Herpetology, 10, 63–74.
Joly, P. (1987). Le regime alimentaire des amphibiens: methods d’etude. Alytes, 6, 11–17.
Juncá, F. A. and Eterovick, P. C. (2007). Feeding ecology of two sympatric species of
Aromobatidae, Allobates marchesianus and Anomaloglossus stepheni, in Central Amazon.
Journal of Herpetology, 41, 301–8.
Kupfer, A., Nabhitabhata, J., and Himstedt, W. (2005). From water into soil: trophic
ecology of a caecilian amphibian (Genus Ichthyophis). Acta Oecologica, 28, 95–105.
Legler, J. M. (1977). Stomach flushing: a technique for chelonian dietary studies.
Herpetologica, 33, 281–4.
Legler, J. M. and Sullivan, L. J. (1979). The application of stomach-flushing to lizards and
anurans. Herpetologica, 35, 107–10.
Lima, A. P. (1998). The effects of size on the diets of six sympatric species of postmeta-
morphic litter anurans in central Amazonia. Journal of Herpetology, 32, 392–9.
Lima, A. P. and Magnusson, W. E. (2000). Does foraging activity change with ontogeny?
An assessment for six sympatric species of postmetamorphic litter anurans in Central
Amazonia. Journal of Herpetology, 34, 192–200.
MacNally, R. (1983). Trophic relationships of two sympatric species of Ranidella (Anura).
Herpetologica, 39, 130–40.
Magnusson, W. E., Lima, A. P., Silva, W. A., and Araújo, M. C. (2003). Use of geomet-
ric forms to estimate volume of invertebrates in ecological studies of dietary overlap.
Copeia, 2003, 13–19.
Mahan, R. D. and Johnson, J. R. (2007). Diet of the gray treefrog (Hyla versicolor) in rela-
tion to foraging site location. Journal of Herpetology, 41, 16–23.
10 Dietary assessments of adult amphibians | 183
Maneyro, R. and da Rosa, I. (2004). Temporal and spatial changes in the diet of Hyla
pulchella (Anura, Hylidae) in southern Uruguay. Phyllomedusa, 3, 101–13.
Matori, R. and Aun, L. (1994). Aspects of the ecology of a population of Tropidurus spinu-
losus. Amphibia-Reptilia, 15, 317–26.
May, R. M. and MacArthur, R. H. (1972). Niche overlap as a function of environmental
variability. Proceedings of the National Academy of Sciences USA, 69, 1109–13.
Miranda, T., Ebner, M., Solé, M., and Kwet, A. (2006). Spatial, seasonal and intrapopu-
lational variation in the diet of Pseudis cardosoi (Anura: Hylidae) from the Araucária
Plateau of Rio Grande do Sul, Brazil. South American Journal of Herpetology, 1, 121–30.
O’Reilly, J. C. (2000). Feeding in caecilians. In K. Schwenk (ed.), Feeding, Form, Function
and Evolution in Tetrapod Vertebrates. Academic Press, New York.
Pianka, E. R. (1973). The structure of lizard communities. Annual Review of Ecology and
Systematics, 4, 53–74.
Pianka, E. R. (1974). Niche overlap and diffuse competition. Proceedings of the National
Academy of Sciences USA, 71, 2141–5.
Pianka, E. R., Oliphant, M. S., and Iverson, Z. L. (1971). Food habits of albacore bluefin,
tuna and bonito in Califorina waters. California Department of Fish and Game Bulletin,
152, 1–350.
Rocha, C. F. D., Van Sluys, M., Alves, M. A. S., Bergallo, H. G., and Vrcibradic, D. (2000).
Activity of leaf-litter frogs: when should frogs be sampled? Journal of Herpetology, 34,
285–7.
Schoener, T. W. (1967). The ecological significance of sexual dimorphism in size in the
lizard Anolis conspersus. Science, 155, 474–7.
Schoener, T. W. (1974). The compression hypothesis and temporal resource partitioning.
Proceedings of the National Academy of Sciences USA, 71, 4169–72.
Silva, H. R. and Britto-Pereira, M. C. (2006). How much fruit do fruit-eating frogs
eat? An investigation on the diet of Xenohyla truncata (Lissamphibia: Anura: Hylidae).
Journal of Zoology, 270, 692–8.
Simon, M. P. and Toft, C. A. (1991). Diet specialization in small vertebrates: mite-eating
in frogs. Oikos, 61, 263–78.
Simpson, E. H. (1949). Measurement of diversity. Nature, 163, 688.
Sokol, O. M. (1969). Feeding in the pipid frog Hymenochirus boettergi (Tornier).
Herpetologica, 25, 9–24.
Solé, M. and Pelz, B. (2007). Do male tree frogs feed during the breeding season? Stomach
flushing of five syntopic hylid species in Rio Grande do Sul, Brazil. Journal of Natural
History, 41, 2757–63.
Solé, M., Beckmann, O., Pelz, B., Kwet, A., and Engels, W. (2005). Stomach-flushing for
diet analysis in anurans: an improved protocol evaluated in a case study in Araucaria
forests, southern Brazil. Studies on Neotropical Fauna and Environment, 40, 23–8.
Toft, C. A. (1980a). Feeding ecology of thirteen syntopic species of anurans in a seasonal
tropical environment. Oecologia, 45, 131–41.
Toft, C. A. (1980b). Seasonal variation in populations of Panamanian litter frogs and
their prey: a comparison of wetter and drier sites. Oecologia, 47, 34–8.
Toft, C. A. (1981). Feeding ecology of Panamanian litter anurans: patterns in diet and
foraging mode. Journal of Herpetology, 15, 139–44.
184 | Amphibian ecology and conservation
11.1 Introduction
There is no greater need for details on the year-round movements and refuges of
amphibians in their natural habitats than today, mainly due to the pressure to
develop lands for human habitation, agriculture, and commerce, and to define
better wetland/buffer zones to protect amphibian communities. Unfortunately,
there is a dearth of information on habitat use for most amphibians due to their
relatively small size and secretive nature. Efforts to track individuals using radio-
isotopes began decades ago (Madison and Shoop 1970), and other techniques
have followed, including fluorescent powder, spool tracking, radiotelemetry,
and harmonic radar, but for most amphibians details on habitat use through
the annual cycle are still lacking. The recent surge in radiotelemetry use is pro-
viding new insights into habitat use and movements, but only a few studies have
addressed whether the data might reflect transmitter transport, attachment,
implantation, or the periodic disturbance of the investigator.
Favorable outcomes in the use of radiotelemetry require having a clear pur-
pose for the study, obtaining reliable equipment and knowing how to use it,
choosing the transmitter and external/internal tagging technique that best
fit the need and the animal chosen for study, having the knowledge and skills
necessary to use the attachment or implant techniques, anticipating the poten-
tial indirect and direct impacts of the radiotransmitter on the individual during
and after tracking, and planning in advance how to analyze the anticipated data
set. In this chapter, we discuss our experiences with amphibians to reduce dupli-
cation of effort and trauma to the animals studied, and to accelerate technical
advances. We do not report all radiotelemetry studies, nor do we repeat the con-
tent of earlier reviews (see Heyer et al. 1994), but focus instead on more recent
studies and models for future application of radiotelemetry procedure.
186 | Amphibian ecology and conservation
11.2 Equipment
The basic radiotelemetry equipment for hand-tracking amphibians includes
radiotransmitters, a portable radio receiver, an antenna, and a cable. There are
many international vendors of telemetry equipment but no single authority that
has direct experience with all products and applications. Our advice is to rely
on the recommendation of experienced users having the broadest biological and
technical experience, and then choose the most reliable equipment available.
11.2.2 Transmitters
The types and sizes of transmitters, transmitter antennas and batteries vary,
but the options grow fewer with smaller animals. In general, the weight of the
transmitter/battery/attachment device should not exceed 10% of body weight
on land (see Heyer et al. 1994), it should be neutrally buoyant in water, and
the detection range should be at least 30–50 m line-of-sight. The smaller the
11 Movement patterns and radiotelemetry | 187
transmitter/battery unit required to remain under the 10% weight rule, the
more one has to compromise signal distance, lifetime, or pulse rate. There
are three trade-offs in selecting a transmitter/battery combination at a given
weight: (1) the greater the pulse rate (from about 50 beats/min to continuous),
the more quickly and reliably you can determine signal direction, but the
shorter the life of the transmitter/battery unit, (2) the larger (more powerful)
the transmitter, the greater the detection distance, but the shorter the life of
the transmitter/battery unit, and (3) the longer the transmitter antenna, the
greater the detection distance, but the greater the impact on the animal. One
should begin by determining the weight of the completed unit (including the
waistband if used), and then choose the transmitter type/battery/attachment
option that best serves the purpose of the study. In addition, in choosing trans-
mitter frequencies one should avoid CB frequencies to ensure that there is no
interference at the field site, especially near major highways or commercial
properties. Finally, one may want to choose a vendor (e.g. Holohil Systems,
Carp, Ontario, Canada) that will refurbish transmitters with spent batteries
for a significant cost reduction.
11.3.1 Surgery
Transmitters are typically implanted into the coelomic cavity. Some studies have
implanted transmitters under the skin (Blais 1996; Lemckert and Brassil 2003),
but sloughing of the transmitter through the skin (without edema or infection)
can occur when using this technique (Blais 1996). Lumps on the body also pro-
mote abrasion/injury and suture failure as the animal moves through restrictive
environments (Weick et al. 2005). We do not recommend subcutaneous trans-
mitter implants for amphibians.
The following surgical procedure with minor variations has been used suc-
cessfully (Colberg et al. 1997; Madison and Farrand 1998; Faccio 2003; Crook
and Whiteman 2006; Rittenhouse and Semlitsch 2006; Cecala et al. 2007). For
coelomic implants, all instruments and transmitters should be sterilized, such as
in 95% ethanol followed by distilled water rinse and drying. For anesthetiza-
tion, we recommend near total submersion in MS-222 (3-aminobenzoic acid
ester methanosulfate salt) diluted with distilled water to a 0.25% solution and
buffered to pH 7.0 with sodium bicarbonate (this concentration is specifically
for adult Ambystoma tigrinum). The concentration should be adjusted to animal
size, species, and habitat condition (see Lowe 2004; Crook and Whiteman 2006;
Peterman and Semlitsch 2006), so that it takes about 10 min or more for the loss of
the “righting” response or reaction to the poke of a probe. More rapid anesthesia,
190 | Amphibian ecology and conservation
prior to release, check for transmitter functioning one last time, and tune the
receiver to the strongest signal output of the transmitter. Tuning on an additional
channel is also suggested (Baldwin et al. 2006). Re- or multi-programming the
receiver to an exact frequency setting improves the chance of re-locating a lost
signal. We also suggest that the investigator stay nearby after release and obtain
positions every 30–60 min for several hours to determine the initial movement
headings, which are generally straight and can often be used to find a lost sig-
nal the next day. Searching for a lost signal without any idea which direction to
search can be time-consuming and frustrating.
for multiple position checks, especially following rain when animals are most
likely to move. Removing an animal from a subterranean retreat should follow
a circular excavation procedure to avoid injuring the animal and to enhance the
description of the refuge (Madison and Farrand 1998).
Several factors can influence signal direction and detectability. Animals near
or under metal objects, boulders, bluffs, small ridges and depressions, utility
poles, or fences, as well as underground or underwater, will be more difficult to
locate because of signal deflection and dampening. Madison and Farrand (1998)
experienced about a 1 m distance reduction for each centimeter of transmitter
depth, but this dampening can vary with different equipment. The reduction
for water depth is much less than for soil. If an arboreal animal is being tracked,
or if the observer and/or animal is on higher ground, the normal signal range
will be increased and the animal will sound closer than it actually is. An obser-
ver looking for a lost signal should raise the antenna high overhead and, where
possible, move to higher ground to improve detection distance.
#10 #3
= Summer areas
Shrubs
= Over-wintering
sites
Pond
Woods = Foraging
forays
#9
= Migratory
movements
Swamp
80 m
(a)
(c) (b)
Fig. 11.2 Tracks for three radio-implanted northern green frogs (Rana clamitans),
Broome County, NY, USA, showing vegetation types, habitat photographs, and
major movement/refuge details. Wetland margins were structurally complex,
consisting of Leersia, Polygonum, Sparganium, Scirpus, and Carex. The swamp had an
11 Movement patterns and radiotelemetry | 195
CA, USA). The latter permits detailed maps of space use to be overlaid with vege-
tation and soil maps for comprehensive studies of habitat use.
A common mistake in quantifying space use is rushing to label and quan-
tify the “home range” for an amphibian, which should include all behaviors
relating to reproduction and survival (Burt 1943), or calculating the average or
maximum distance moved from a breeding pond without qualifying that the
estimates are based on a small subset of individuals over a few weeks or months.
Lacking for the most part is information on year-round movements in struc-
turally different habitats. Results are seldom scaled or weighted according to
the duration of a telemetry record, for example averaging a one-week distance
record with a 4-month distance record can underestimate the potential for a
species to move long distances (Madison and Farrand 1998). The potential for
individuals to occupy widely spaced and diverse habitats for different purposes
through the annual cycle must be kept in mind, and a good but preliminary
descriptive model for these movements is that for green frogs (Figure 11.2), as
described in Lamoureux and Madison (1999) and Lamoureux et al. (2002).
More quantitative indices of multi-seasonal space use also occur (e.g. Spieler and
Linsenmair 1998; Watson et al. 2003; Baldwin et al. 2006).
The most common home-range representation is the minimum convex
polygon, and while this method may give fairly accurate short-term “daily”
VEGETATION W E
Pinus rigida-Quercus (alba, Velutina)-Vaccinium palladium
Quercus coccinea-Vaccinium palladium/Gaylussacia baccata S
Pinus rigida /Quercus (coccinea, alba)-Vaccinium palladium
Grass 6
Road 5
0 20 40 80 m
Water
(a)
1 (c)
(b)
3
2
11.6.2 Movements
Little is known about the full extent of habitat use in free-ranging amphib-
ians other than that obtained using radiotelemetry, and so it is difficult to test
whether transmitters might be affecting movements and habitat use. In field
enclosures, Rana sylvatica and Rana pipiens with and without belted transmit-
ters with 15 cm antennas showed some species-specific differences in behavior
in response to simulated predation threat, prompting some caution in the use of
the attached units studied (Blomquist and Hunter 2007). Studies of a rainforest
frog (Litoria lesueuri) in laboratory enclosures showed no difference in the dis-
tance or frequency of movement before and after transmitter attachment, except
for a slight reduction the first day after attachment (Rowley and Alford 2007).
In nature, casual observations on non-implanted A. maculatum salamanders
during tracking of implanted individuals revealed comparable movements,
occurrence, and mortality (Madison 1997). Cross-study comparisons show
that wood frogs monitored by drift fences (Vasconcelos and Calhoun 2004)
emigrated comparable distances to implanted individuals (Baldwin et al. 2006;
Rittenhouse and Semlitsch 2007). In addition, implanted tiger salamanders
showed courtship movements in a breeding pond similar to those observed for
non-implanted salamanders in captivity (Madison 1998). The above studies,
although quite incomplete, suggest no major effects of implanted or attached
transmitters on short-term movements.
11.6.3 Reproduction
Reproduction is seldom verified for amphibians carrying transmitters, although
there are exceptions. Six gravid female frogs (Buergeria buergeri) were given
waist-band transmitters (with a 3 cm whip antenna) and monitored along a
stream for up to 8 days during successful mating and oviposition (Fukuyama
et al. 1988). In implant studies, Johnson (2006) recorded a female Hyla versicolor
ovipositing a year after carrying an implant and undergoing two implant surger-
ies. Tiger salamanders showed normal breeding activity soon after implantation
(Madison 1998), and an implanted male spotted salamander returned 190 m to
the same location at a breeding pond after carrying an implant in upland habitat
for a year (Madison 1997). Further studies on possible long-term reproductive
effects are obviously needed.
11 Movement patterns and radiotelemetry | 199
11.7 Conclusions
In order to fully understand amphibian movement patterns and habitat use,
radiotelemetry is an indispensable tool, despite some shortcomings in meth-
odology. Waist bands for external transmitters with short whip antennas (less
than ≈3 cm), or antennas built into the waist band rather than trailing behind
the animal, are tentatively recommended for short-term tracking of anurans
under conditions where the animals can be inspected periodically and the waist
bands removed at the end of study (e.g. Fukuyama et al. 1988). However, long-
term tracking using waist-band transmitters is not recommended because of the
increased likelihood of skin trauma and getting snagged or trapped by objects in
post-breeding and over-wintering refuges. External transmitters and antennas
sutured to amphibians are not recommended.
Transmitters with antennas coiled in the potting material are recommended
for short-term studies as gastric pills (e.g. Schabetsberger et al. 2004), or for
extended studies as coelomic implants in both anurans (e.g. Lamoureux and
Madison 1999) and salamanders (Madison 1997; Madison and Farrand 1998;
Faccio 2003; Rittenhouse and Semlitsch 2006; McDonough and Paton 2007).
Especially needed, however, is information on potential weight effects on fitness
from the long-term transport of transmitter implants.
The future challenge in the use of radiotelemetry in amphibians is threefold:
to minimize duplication in method development and unnecessary trauma
200 | Amphibian ecology and conservation
11.8 References
Baldwin, R. F., Calhoun, A. J. K., and deMaynadier, P. G. (2006). Conservation planning
for amphibian species with complex habitat requirements: A case study using move-
ments and habitat selection of the Wood Frog Rana sylvatica. Journal of Herpetology,
40, 442–53.
Blais, D. P. (1996). Movement, Home Range, and Other Aspects of the Biology of the Eastern
Hellbender (Cryptobranchus alleganiensis alleganiensus): a Radiotelemetric Study.
Master’s thesis, State University of New York at Binghamton, Binghamton, NY.
Blomquist, S. M. and Hunter Jr, M. L. (2007). Externally attached radio-transmitters
have limited effects on the antipredator behavior and vagility of Rana pipiens and Rana
sylvatica. Journal of Herpetology, 41, 430–8.
Burt, W. H. (1943). Territoriality and home range concepts as applied to mammals.
Journal of Mammalogy, 24, 346–52.
Cecala, K. K., Price, S. J., and Dorcas, M. E. (2007). A comparison of the effectiveness
of recommended doses of MS-222 (tricaine methanesulfonate) and Orajel (R) (benzo-
caine) for amphibian anesthesia. Herpetological Review, 38, 63–6.
Colberg, M. E., Denardo, D. F., Rojek, N. A., and Miller, J. W. (1997). Surgical proced-
ure for radio transmitter implantation into aquatic, larval salamanders. Herpetological
Review, 28, 77–8.
Crook, A. C. and Whiteman, H. H. (2006). An evaluation of MS-222 and benzocaine as
anesthetics for metamorphic and paedomorphic tiger salamanders (Ambystoma tigri-
num nebulosum). American Midland Naturalist, 155, 417–21.
Faccio, S. D. (2003). Postbreeding emigration and habitat use by Jefferson and spotted
salamanders in Vermont. Journal of Herpetology, 37, 479–89.
Fukuyama, K., Kusano, T., and Nakane, M. (1988). A radio-tracking study of the behav-
iour of females of the frog Buergeria buergeri (Rhacophoridae, Amphibia) in a breeding
stream in Japan. Japanese Journal of Herpetology, 12, 102–7.
Goldberg, C. S., Goode, M. J., Schwalbe, C. R., and Jarchow, J. L. (2002). External
and implanted methods of radiotransmitter attachment to a terrestrial anuran
(Eleutherodactylus augusti). Herpetological Review, 33, 191–4.
Gray, M. J., Miller, D. L., and Smith, L. M. (2005). Coelomic response and signal range
of implant transmitters in Bufo cognatus. Herpetological Review, 36, 285–8.
Heyer, W. R., Donnelly, M. A., McDiarmid, R. W., Hayek, L.-A. C., and Foster, M. S.
(eds) (1994). Measuring and Monitoring Diversity: Standard Methods for Amphibians,
Smithsonian Institution Press, Washington DC.
Indermaur, L.B., Schmidt, B.R., and Tockner, K. (2008). Effect of transmitter mass and
tracking duration on body mass of two anuran species. Amphibia-Reptilia, 29, 263–9.
11 Movement patterns and radiotelemetry | 201
Given that the use of enclosures is often expensive and time-consuming, why
should a researcher use them? Enclosures allow researchers to expose animals to
experimental treatments that test explicit hypotheses and to recapture animals
to measure response variables. More importantly, field experiments maximize
realism relative to laboratory or mesocosm studies and provide the strong-
est inferences as to whether the factors under investigation operate in nature
(Hairston 1989; Morin 1998; Boone and James 2005).
When designing enclosure experiments, researchers must balance trade-offs
in experimental design because it is impossible to simultaneously maximize
realism, precision, and generality (Hairston 1989). Precision is maximized by
creating replicates of experimental treatments that are as similar as possible to
one another. Generality, which is the ability to extrapolate to a broad area or
to a number of conditions, is maximized by placing replicate enclosures in as
many locations or conditions as possible. In general, experimental designs that
maximize generality compromise precision (and visa versa), and designs that
maximize realism also compromise precision.
Optimally balancing these trade-offs depends on the questions that are being
addressed. For example, a researcher who wishes to address how density affects
the growth of juvenile amphibians may elect to maximize precision and reduce
experimental error by placing enclosures at one site and in close proximity to one
another (Altwegg 2003). In contrast, a researcher who wishes to determine the
rate at which air-borne agricultural pesticides are accumulating in amphibians
in a nearby mountain range may want to establish test sites at numerous loca-
tions to maximize generality.
In the sections that follow we provide examples of how terrestrial enclosures
have been used to address questions about the ecology and conservation biology
of amphibians. We also provide information on cage designs, add cautionary
notes about experimental artifacts, and discuss trade-offs that should be consid-
ered when designing experiments.
Fig. 12.1 Small cages can be used to confine amphibians to specific microhabitats
during short-term experiments.
12 Field enclosures and terrestrial cages | 207
Measuring habitat variables for use as covariates in later analyses can be useful
in understanding the mechanisms underlying this within-treatment variabil-
ity (e.g. aspect and surface temperature; Harper 2007). Densities of predators,
competitors, and prey may also be important explanatory variables.
Research questions focused on habitat type have important implications for
enclosure design and construction. Because the treatment includes the surround-
ing habitat as well as the conditions within the enclosure, realism may be improved
by using mesh rather than solid materials for the enclosure walls. Mesh allows
movement of air, water, nutrients, and small organisms in and out of enclosures
and creates conditions within the enclosure that are more typical of the habitat type
as a whole. Large enclosures ( 200 m2) are more likely to incorporate most micro-
habitats available within a larger habitat type and to provide the range of choices
that would be available to free-roaming individuals, thereby increasing realism.
Large enclosures are also likely to support a more representative community includ-
ing competitors, predators, and small mammals that dig burrows used by some
amphibians (Regosin et al. 2003b). Conversely, if the research question focuses
on microhabitats, small enclosures (Figure 12.1) can purposefully limit choice to
determine the effects of specific microhabitats and their associated organisms on
amphibian performance (Maerz et al. 2005; Rothermel and Luhring 2005).
Price and Shields Plethodon cinereus and 0.26 52 Interspecific 3.8–7.71 8 weeks Growth
(2002) Plethodon glutinosus interactions
Rothermel and Ambystoma maculatum, 125 8 Juvenile Variable (total 2 months Rate of movement and
Semlitsch (2002) Ambystoma texanum, emigration n 179) distance moved
Bufo americanus
Rothermel and Ambystoma maculatum, 0.025 8 Dehydration 401 22–25 h Change in mass
Semlitsch (2002) Ambystoma texanum,
Bufo americanus
Altwegg (2003) Rana lessonae 9 12 Density 2.8–8.3 1 year Growth and survival
dependence
Altwegg and Reyer Rana lessonae and 9 12 Carryover effects 2.8–8.3 1 year Growth and survival
(2003) Rana esculenta
Regosin et al. Rana sylvatica 272 17 Over-wintering Not stocked 9 months Sex ratio, density,
(2003a) ecology distance from pond
Marsh and Beckman Plethodon cinereus 0.81 24 Road effects 2.5 2 months Detectibility and
(2004) surface activity
Moseley et al. Ambystoma 20.88 and 9 and 18 Microhabitat use Not reported 8 months Movement and habitat
(2004) talpoideum 9 selection
Regosin et al. Ambystoma 3.8 10 Burrow-occupancy 0.26–0.531 8 days Probability of burrow
(2004) maculatum patterns occupancy
Williams and Plethodon cinereus 1.18 124 Detection 0.85–1.71 2 months Proportion detected
Berkson (2004) probabilities
Yurewicz and Plethodon cinereus 0.81 28 Cost of 1.21 2 years Egg survival, female
Wilbur (2004) reproduction growth, production of
ova
Boone (2005) Rana sphenocephala, 2 48 Contaminant 3–5 7–12 months Survival and growth
Rana blairi, Rana exposure
clamitans,
Bufo woodhousii
Maerz et al. Rana clamitans 0.06 38 Foraging success 171 38 h Change in mass
(2005)
Rothermel and Ambystoma 0.025 48 Forestry practices 401 72 h Water loss and
Luhring (2005) talpoideum survival
Croshaw and Scott Ambystoma opacum 112–223 4 Nest-site selection 10.76 1 month Number of nests per
(2006) area
Greenlees et al. Bufo marinus 2.88 60 Effects on native Standard Not reported Invertebrate
(2006) invertebrates biomass abundance
Rothermel and Ambystoma maculatum 18 12 Forest 1.3 2 years Survival to maturity
Semlitsch (2006) and Ambystoma opacum fragmentation
Todd and Bufo terrestris 16 8 Forestry practices 1.8 2 months Growth and survival
Rothermel (2006)
Harper (2007) Rana sylvatica and 9 64 Forestry practices 2 3 months Survival
Bufo americanus
Table 12.1 Continued
Reference Species Enclosure Number of Research Stocking Duration of Response metrics
size (m2) enclosures focus density experiment
(per/m2)
choosing species that have relatively small home ranges, as is often the case for
direct-developing species (Price and Shields 2002).
Large enclosures maximize realism, allowing increased movement and access
to a greater range of microhabitats and interacting organisms. One important
trade-off is that the probability of recapturing individuals decreases as enclosure
size increases. On the other hand, large enclosures can hold more individuals
and provide more precise estimates of mean responses.
Large enclosures can also allow densities to be at or near “natural levels”.
Unfortunately, estimates of natural densities for terrestrial life history stages
are not available for most amphibian species, and density can have significant
effects on growth and survival (Figure 12.3). It may be worthwhile to estimate
the natural densities of study species if estimates are unavailable. Terrestrial
densities can be extremely high following metamorphosis, but quickly decrease
thereafter due to mortality and migration (Cohen and Alford 1993). Therefore,
short-term experiments using recently metamorphosed individuals can use high
Fig. 12.3 The density of animals in enclosures can have profound effects on
individual growth rates, probabilities of survival, and age at sexual maturity. These
American toads were the same size at metamorphosis and were raised in similar
enclosures, but at densities of 1/m2 (left) and 10/m2 (right). This photo was taken 3
months after the animals were introduced to enclosures.
216 | Amphibian ecology and conservation
densities without sacrificing realism (Beck and Congdon 1999), but experiments
that are longer term or use adults may require larger enclosures to approximate
natural densities. Because amphibians occur at higher densities in high-quality
habitat (Patrick et al. 2008), ensuring high habitat quality within enclosures
(e.g. sufficient moisture, cover, and prey) can improve realism when enclosures
are small and densities are relatively high.
even the least expensive silt fencing will survive at least a full year under most
conditions. After the walls are secured, soil is backfilled into the trench and
compressed so that a portion of the wall is underground. If pitfall traps are used
to recapture animals, they can be installed before the trench is filled. During
construction, soil, leaf litter, and vegetation are often disturbed. To maximize
realism, the time between the building of the enclosures and the introduction of
animals should be sufficient to allow accumulation of leaf litter and invertebrate
prey, and the regrowth of vegetation.
Animals are most likely to attempt to escape from the corners of the enclosures,
by climbing up the posts, or through the overlapping ends of the material used for
the walls. Rounded corners can be more difficult to climb, and placing the support
posts 10cm or more below the top of the wall can also reduce escapes. Strips of add-
itional material across the top edge of the walls can serve as baffles, both to keep
enclosed animals in and to prevent animals from natural populations from enter-
ing (Figure 12.4). Quick detection and repair of damaged or worn enclosures is
also necessary to prevent escapes. Enclosures with lids may be necessary for arbor-
eal species (Beard et al. 2002), and can be useful for other species to exclude preda-
tors, provide shade, and reduce escapes (Harper and Semlitsch 2007). However,
and Congdon 1999). There are many techniques that can be used to identify
individuals, including toe-clipping, visible implant elastomer, passive integra-
tive transponder tags (PIT tags), and photographing unique dorsal patterns
(Chapter 8). Group marks may be necessary when it is impractical to mark all
individuals. In general, the marking technique chosen should strive to minimize
the increased probability of mortality and stress to the animals while maximiz-
ing the probability of successful identification.
early and often to ensure that useful data are gathered even if natural disas-
ters or unexpectedly high mortality occur before the end of the experiment.
Frequent censuses also improve estimates of capture probability and survival
(Chapter 25). However, it is important to consider that handling animals too
frequently can affect survival and induce unnecessary stress. The optimal fre-
quency of censuses for estimating capture probability and survival also depends
on the statistical methods that will be used. Some methods require equal time
between sampling intervals or a minimum number of sample periods. Other
statistical methods rely on multiple consecutive samples spaced at regular inter-
vals (e.g. 3 days of sampling every 2 weeks). These methods are discussed in
greater detail in Chapter 25. It is important to become familiar with these stat-
istical methods prior to gathering the data to ensure that the frequency and
methods used in the censuses maximize statistical power.
survival (Altwegg 2003; Chapter 24). However, when capture probabilities are
low, there is a large amount of error associated with these estimates. Methods
for incorporating errors from mark–recapture estimates into other statistical
analyses of multiple experimental populations, such as analysis of variance, have
received limited attention. Survival estimates can be improved by using an end-
point associated with breeding migrations, such as first reproduction, because
migrations increase the likelihood of capture (Pechmann 1995; Regosin et al.
2003b).
Growth is usually measured as change in mass or length over time. Some
researchers prefer measuring snout–vent length, often with additional measure-
ments including head width or tibia length (Altwegg and Reyer 2003), because
mass is prone to short-term fluctuations.
Reproductive maturity can be determined by dissection, candling, or exter-
nal signs of maturity such as nuptial pads or a swollen cloaca (Pechmann 1995;
Rothermel and Semlitsch 2006; Harper and Semlitsch 2007). Age at repro-
ductive maturity or the proportion of individuals that have reached maturity at
the conclusion of the experiment are demographically important response met-
rics. Although fecundity is a component of individual fitness and population
growth, clutch size has not typically been used as a response metric for enclosure
experiments (but see Yurewicz and Wilbur 2004).
Moseley et al. 2004) and distance moved (Rothermel and Semlitsch 2002).
Studies of habitat selection typically quantify the proportion of individuals that
select a particular habitat type (Southerland 1986; Moseley et al. 2004; Patrick
et al. 2008). Regosin et al. (2004) used probability of burrow occupancy as a
response metric.
12.9 References
Altwegg, R. (2003). Multistage density dependence in an amphibian. Oecologia, 136,
46–50.
Altwegg, R. and Reyer, H.-U. (2003). Patterns of natural selection on size at metamor-
phosis in water frogs. Evolution, 57, 872–82.
Beard, K. H., Vogt, K. A., and Kulmatiski, A. (2002). Top-down effects of a terrestrial
frog on forest nutrient dynamics. Oecologia, 133, 583–93.
Beck, C. W. and Congdon, J. D. (1999). Effects of individual variation in age and size
at metamorphosis on growth and survivorship of southern toad (Bufo terrestris) meta-
morphs. Canadian Journal of Zoology, 77, 944–51.
224 | Amphibian ecology and conservation
Jaeger, R.G. (1971). Competitive exclusion as a factor influencing the distributions of two
species of terrestrial salamanders. Ecology, 52, 632–7.
Laposata, M. M. and Dunson, W. A. (1999). Enclosure design for ecotoxicological stud-
ies of terrestrial salamanders. Herpetological Review, 30, 28–30.
Maerz, J. C., Blossey, B., and Nuzzo, V. (2005). Green frogs show reduced foraging
success in habitats invaded by Japanese knotweed. Biodiversity and Conservation, 14,
2901–11.
Marsh, D. M. and Beckman, N. G. (2004). Effects of forest roads on the abundance and
activity of terrestrial salamanders. Ecological Applications, 14, 1882–91.
Morin, P. J. (1998). Realism, precision, and generality in experimental tests of ecological
theory. In W. J. Resetarits and J. Bernardo (eds), Issues and Perspectives in Experimental
Ecology, pp. 50–70. Oxford University Press, Oxford.
Moseley, K. R., Castleberry, S. B., and Ford, W. M. (2004). Coarse woody debris and pine
litter manipulation effects on movement and microhabitat use of Ambystoma talpoi-
deum in a Pinus taeda stand. Forest Ecology and Management, 191, 387–96.
Parris, M. J. (2001). Hybridization in leopard frogs (Rana pipiens complex): terrestrial
performance of newly metamorphosed hybrid and parental genotypes in field enclos-
ures. Canadian Journal of Zoology, 79, 1552–8.
Patrick, D. A., Harper, E. B., Hunter Jr, M. L., and Calhoun, A. J. K. (2008). Terrestrial
habitat selection and strong density-dependent mortality in recently metamorphosed
amphibians. Ecology, 89, 2563–74.
Pearson, P. G. (1955). Population ecology of the spadefoot toad, Scaphiopus h. holbrooki
(Harlan). Ecological Monographs, 25, 233–67.
Pechmann, J. H. K. (1994). Population Regulation in Complex Life Cycles: Aquatic and
Terrestrial Density-dependence in Pond-breeding Amphibians. PhD dissertation, Duke
University, Durham, NC.
Pechmann, J. H. K. (1995). Use of large field enclosures to study the terrestrial ecology of
pond-breeding amphibians. Herpetologica, 51, 434–50.
Petranka, J. W. and Sih, A. (1986). Environmental instability, competition and density-
dependent growth and survivorship of a stream-dwelling salamander. Ecology, 67,
729–36.
Price, J. E. and Shields, J. A.S. (2002). Size-dependent interactions between two terrestrial
amphibians, Plethodon cinereus and Plethodon glutinosus. Herpetologica, 58, 141–55.
Regosin, J. V., Windmiller, B. S., and Reed, J. M. (2003a). Terrestrial habitat use and win-
ter densities of the Wood Frog (Rana sylvatica). Journal of Herpetology, 37, 390–4.
Regosin, J. V., Windmiller, B. S., and Reed, J. M. (2003b). Influence of abundance of
small-mammal burrows and conspecifics on the density and distribution of spot-
ted sala manders (Ambystoma maculatum) in terrestrial habitats. Canadian Journal of
Zoology, 81, 596–605.
Regosin, J. V., Windmiller, B. S., and Reed, J. M. (2004). Effects of conspecifics on the
burrow occupancy behaviour of spotted salamanders. Copeia, 2004, 152–8.
Regosin, J. V., Windmiller, B. S., Homan, R. N., and Reed, J. M. (2005). Variation
in terrestrial habitat use by four poolbreeding amphibian species. Journal of Wildlife
Management, 69, 1481–93.
226 | Amphibian ecology and conservation
13.1 Introduction
Many of the simplest yet most highly productive sampling methods in her-
petological field research use some type of trap or attraction device to increase
capture rates or target secretive species. These techniques fall into two general
categories: those that actually trap animals, accumulating captures on their own
over time (passive traps) and those that attract animals but require an observer
to actively capture them at the moment of the census (active traps). The most
popular examples of these two trap categories are drift fences (generally with
pitfall and/or funnel traps) and coverboards, respectively. Both of these methods
are inherently simple concepts, and their description and explanation need not
be made complex or complicated. Both techniques are usually best modified by
the investigator who can use common sense to focus on the needs of a particular
project that involves capturing animals in a field situation. However, we provide
a general discussion of some of the fundamental issues that investigators who
use these techniques must face, with particular emphasis on how choice of cap-
ture method and sampling design influence interpretation of capture data.
most importantly, passive traps are the most effective ways to sample many rare or
secretive amphibian species, many of which are of conservation concern.
A huge variety of passive amphibian traps have been developed for nearly any
imaginable habitat or situation; however, nearly all are variants on two basic trap
types, the funnel trap and the pitfall trap. Funnel traps consist of a tapering funnel-
shaped entrance that guides animals into a larger holding chamber (Figure 13.1).
Once within the chamber, animals are unable to find their way back out the small
entrance hole, becoming trapped. Pitfall traps work on a similar principle, con-
sisting of some type of container that is sunk into the ground, with the rim level
with the surface (Figure 13.2). Animals that fall into pitfalls are unable to climb
out, becoming trapped. First described for use with herpetofauna by Gibbons and
Bennett (1974) and Gibbons and Semlitsch (1981), drift fences are vertical bar-
riers that intercept the intended trajectory of amphibians moving from one loca-
tion to another. The fence typically guides animals toward a pitfall bucket, funnel
trap, or other capture device (Figure 13.2). In most terrestrial and some aquatic
situations, drift fences dramatically increase amphibian capture rates (Friend et al.
1989) and, under some circumstances, have been responsible for more amphibian
captures per day, month, year, or decade than any other method used in field stud-
ies of amphibians (Pechmann et al. 1991; Gibbons et al. 2006). Although drift
Fig. 13.1 Several varieties of funnel trap are commonly used to sample amphibians
in terrestrial and aquatic habitats. Front: soft-drink bottle funnel trap; back (left
to right): plywood and hardware cloth box trap, steel “Gee” minnow trap, plastic
minnow trap, and collapsible nylon trap. Photograph by John Willson, Savannah River
Ecology Laboratory.
13 Drift fences, coverboards, and other traps | 231
fences have been used most extensively in the southeastern USA, they have proven
effective for many amphibian species worldwide (e.g. Gittins 1983; Friend 1984;
Bury and Corn 1987; Friend et al. 1989; Jehle et al. 1995; Weddeling et al. 2004).
However, drift fences were relatively ineffective for sampling anurans in forests of
Queensland, Australia (Parris et al. 1999) and recorded lower numbers of anuran
species than automated recording devices or nocturnal line-transect surveys in
Taiwan (Hsu et al. 1985).
Metal stakes
Drainage holes
Fig. 13.2 Schematic of a terrestrial drift fence with large pitfall traps. Figure reprinted
from Gibbons and Semlitsch (1981).
232 | Amphibian ecology and conservation
(a)
Small
wetland or
Large wetland
(b) Habitat 1
Habitat 2
from any obvious breeding wetland, hibernaculum, or other habitat focal point.
For such applications, drift fences are often constructed in cross- or X-shaped
arrays with a central trap and traps placed along each section of the fence (see
Corn 1994). Because the goal of this type of study is generally to test hypoth-
eses about amphibian abundance between areas, each array makes a convenient
sampling unit for statistical comparisons. In this case, care should be taken to
ensure that arrays are comparable (same length of fencing, number of traps, etc.)
and that arrays are located randomly or systematically across the treatments or
habitats.
Finally, passive traps can be used for quantitative sampling of aquatic
amphibians that are not easily captured by other methods. For this applica-
tion, a variety of aquatic funnel traps can be effective, with the size of the
trap being dictated by the size of the target species. For small species or life
stages, small funnel traps made from plastic soft-drinks bottles would suffice
(e.g. Griffiths 1985; Willson and Dorcas 2003), while targeting larger spe-
cies such at the giant salamanders, Siren and Amphiuma, would require lar-
ger traps such as commercially available minnow or crawfi sh traps (Johnson
and Barichivich 2004; Willson et al. 2005). In most cases, aquatic traps
should be set in water shallow enough to allow captured animals access to
air and care should be taken to monitor fluctuations in water level that could
submerge traps. Specific microhabitats where traps are set will vary by spe-
cies, but heavily vegetated shallow areas are often preferable. Capture rates
may also be increased by setting aquatic traps along natural barriers such as
submerged logs or the shoreline or by the use of aquatic drift fences to direct
amphibians into traps (Enge 1997b; Willson and Dorcas 2004; Palis et al.
2007). Finally, as in the previous example, traps may be set in any number
of spatial confi gurations. Often a simple linear transect along a shoreline
is sufficient. However, if the goal of the study is to compare captures stat-
istically across different treatments (habitats, wetlands, etc.) standardized
arrays of traps can be set that will serve as the sampling units in statistical
comparisons.
Passive traps restrain captured animals, so frequent (generally at least daily)
monitoring of traps is necessary to avoid mortality of captured animals. During
hot or dry periods it is often advisable to provide access to moisture (e.g. a water
bowl or damp sponge) within traps to avoid desiccation of captured animals.
Finally, natural amphibian predators such as mid-sized mammals, birds, and
large snakes often learn to target drift fences, and predator-control measures
(raised covers for pitfalls, live-trapping and removal, or wide-width steel mesh
trap covers, etc.) may be necessary to avoid undue predation.
13 Drift fences, coverboards, and other traps | 235
boards generally need not be large, but at least one study reported a positive
correlation between the size of coverboards and the number of salamanders
captured per board (Moore 2005). Moreover, larger or thicker boards may be
preferable in warm or dry habitats as they generally hold moisture better than
smaller boards. One study conducted in southern Georgia, USA, noted lower
amphibian captures under coverboards than natural-cover objects, presumably
resulting from warmer and more variable temperatures under boards (Houze
and Chandler 2002). As some time is often necessary for suitable microhabi-
tats (e.g. rotten leaf litter, burrows) to develop under refugia, the investiga-
tor should consider allowing boards to “weather” for several weeks or months
before amphibian censuses are initiated.
Construction of other active traps varies, but the general goal is to create
microhabitat conditions that attract target amphibian species. For example,
sections of PVC pipe may be inserted vertically into the ground or affixed to
tree trunks to provide arboreal refugia for hylid treefrogs (Moulton et al. 1996;
Boughton et al. 2000). Johnson (2005) described several modifications to the
so-called hylid tube technique, maximizing standardization of microhabitat
within tubes and increasing ease of census and frog capture.
13 Drift fences, coverboards, and other traps | 239
As uses of coverboards and other active traps vary, so will the designs for their
placement. In general active traps should be placed in habitats that are favorable
for amphibians, including well-shaded areas with abundant moisture. For spe-
cies that breed in aquatic habitats, breeding sites may be the most appropriate
places to maximize captures. For example, an obvious location to deploy PVC
tubes for hylid frogs would be around the periphery of wetland breeding sites.
The spatial distribution of active traps also depends on the goals of the study.
For simple amphibian inventories (documenting species presence) or collecting
individuals for use in the laboratory, placing devices haphazardly in the most
optimal habitats is the most cost-effective method. For studies where statistical
comparisons are to be made, replicated sampling units must be designated. In
general, because captures per individual trap are low, arrays of several boards or
other traps are usually designated as the sampling unit. An array can consist of
any arbitrary number of traps, generally arranged in a systematic pattern (e.g.
a grid or transect). Replicate arrays are then placed systematically or randomly
within treatments. Ideally, a power analysis can be used to determine the num-
ber of arrays (sample size) that is needed to obtain sufficient statistical power for
the analysis.
For example, a researcher might wish to compare salamander abundance
across three forest types within a relatively small geographical area. Having
determined that salamanders are generally fairly common (say, an average of
one salamander found per five coverboards checked), an array of 10 cover-
boards, spaced 5 m apart in a linear transect, might be determined as the sam-
pling unit. Geographic Information Systems (GIS) technology could then be
used to generate randomized locations for five arrays to be placed within each
of the three habitat types. Thus, the total number of boards to be used would
be 150 (10 boards/array 5 arrays per habitat treatment 3 habitats) and the
sample size would be five per treatment.
Unlike passive traps, active traps can be monitored on nearly any tem-
poral schedule, and the timing of censuses will reflect the question of inter-
est. Generally, as animals may remain within the same refuge for extended
periods, allowing some time (e.g. a few days or a week) between censuses may
minimize repeated counts of the same individual animal. Alternatively, ani-
mals may avoid boards that are disturbed too frequently; one study noted that
salamander captures under coverboards that were checked daily were reduced
compared to boards that were censused on longer time intervals (1 or 3 weeks;
Marsh and Goicochea 2003). Similarly, when designing a monitoring scheme,
care should be taken to make samples as repeatable as possible. This typic-
ally means that environmental conditions should be as comparable as possible
240 | Amphibian ecology and conservation
between samples, and many researchers set up environmental criteria for deter-
mining census times based on the biology of the study animal. For example,
coverboard arrays might be checked once-weekly at 7–9 am on days without
precipitation or only on nights with a temperature greater than 15°C and at
least 1 cm of rain.
premise is violated. Perhaps the most obvious case where the equal catchability
assumption is violated is when comparing capture rates among species, as noted
for the drift-fence method. It is often the case, regardless of the sampling method,
that some species are highly catchable, while others are seldom encountered.
Thus, it is generally unreasonable to assume that simply because one species is
captured more frequently than another, that it is truly more abundant. Likewise,
differences in catchability across time are critical to consider in temporal com-
parisons and the potential for activity patterns to influence catchability must
be considered when interpreting capture rates. For example, in warm climates,
salamanders often retreat deep underground during the summer and are seldom
captured using any method. In this case, low capture rates of salamanders in the
summer are best explained by seasonal differences in catchability rather than by
changes in actual abundance within the landscape. A growing body of literature
uses mark–recapture or occupancy modeling to incorporate detection probabil-
ity (catchability) in interpretations of amphibian abundance data (Chapter 24;
Mazerolle et al. 2007). Ideally, any researcher wishing to use abundance indices
as indicators of population size or density should consider using these methods to
test the equal catchability assumption.
A final consideration when using active traps is the potential for use of these
methods to actually improve the overall quality of the habitat for target spe-
cies. For example, if one of the factors limiting population size in an amphibian
species is availability of suitable refugia, adding boards, hylid tubes, or other
artificial refugia to the habitat may permit an increase in population density.
Although such a situation would probably be favorable in studies designed to
conserve at-risk amphibian species, it is worth remembering that population
densities estimated in areas with abundant artificial refugia may not necessarily
be representative of those in unaltered habitats. To our knowledge the poten-
tial for artificial refugia to improve habitat quality for amphibians has not been
addressed experimentally and warrants future investigation.
13.4 References
Adams, M. J., Richter, K. O., and Leonard, W. P. (1997). Surveying and monitoring
amphibians using aquatic funnel traps. In D. H. Olson, W. P. Leonard, and R. B. Bury
(eds), Sampling Amphibians in Lentic Habitats, Northwest Fauna, no. 4, pp. 47–54.
Society for Northwestern Vertebrate Biology, Olympia, WA.
Aresco, M. J. (2005). Mitigation measures to reduce highway mortality of turtles and other
herpetofauna at a north Florida lake. Journal of Wildlife Management, 69, 549–60.
Arntzen, J. W., Oldham, R. S., and Latham, D. M. (1995). Cost effective drift fences for
toads and newts. Amphibia-Reptilia, 16, 137–45.
242 | Amphibian ecology and conservation
Boughton, R. G., Staiger, J., and Franz, R. (2000). Use of PVC pipe refugia as a sampling
technique for hylid treefrogs. American Midland Naturalist, 144, 168–77.
Buech, R. R. and Egeland, L. M. (2002). Efficacy of three funnel traps for capturing
amphibian larvae in seasonal forest ponds. Herpetological Review, 33, 182–5.
Bury, R. B. and Corn, P. S. (1987). Evaluation of pitfall trapping in northwestern forests:
trap arrays with drift fences. Journal of Wildlife Management, 51, 112–19.
Corn, P. S. (1994). Straight-line drift fences and pitfall traps. In W. R. Heyer,
M. A. Donnelly, R. W. McDiarmid, and L. C. Hayek (eds), Measuring and Monitoring
Biological Diversity. Standard Methods for Amphibians, pp. 109–17. Smithsonian
Institution Press, Washington DC.
Degraaf, R. M. and Yamasaki, M. (1992). A nondestructive technique to monitor the
relative abundance of terrestrial salamanders. Wildlife Society Bulletin, 20, 260–4.
Dodd, Jr, C. K. (1991). Drift fence-associated sampling bias of amphibians at a Florida
sandhills temporary pond. Journal of Herpetology, 25, 296–301.
Dodd, Jr, C. K. and Scott, D. E. (1994). Drift fences encircling breeding sites. In
W. R. Heyer, M. A. Donnelly, R. W. McDiarmid, and L. C. Hayek (eds), Measuring
and Monitoring Biological Diversity. Standard Methods for Amphibians, pp. 125–30.
Smithsonian Institution Press, Washington DC.
Enge, K. M. (1997a). Use of silt fencing and funnel traps for drift fences. Herpetological
Review, 28, 30–1.
Enge, K. M. (1997b). A Standardized Protocol for Drift-fence Surveys. Technical report
no. 14. Florida Game and Fresh Water Fish Commission, Tallahassee, FL.
Enge, K. M. (2001). The pitfalls of pitfall traps. Journal of Herpetology, 35, 467–78.
Fellers, G. M. and Drost, C. A. (1994). Sampling with artificial cover. In W. R. Heyer,
M. A. Donnelly, R. W. McDiarmid, and L. C. Hayek (eds), Measuring and Monitoring
Biological Diversity. Standard Methods for Amphibians, pp. 146–50. Smithsonian
Institution Press, Washington DC.
Friend, G. R. (1984). Relative efficiency of two pitfall-drift fence systems for sampling
small vertebrates. Australian Zoologist, 21, 423–34.
Friend, G. R., Smith, G. T., Mitchell, D. S., and Dickman, C. R. (1989). Influence of pit-
fall and drift fence design on capture rates of small vertebrates in semi-arid habitats of
western Australia. Australian Wildlife Research, 16, 1–10.
Gascon, C. (1994). Sampling with artificial pools. In W. R. Heyer, M. A. Donnelly,
R. W. McDiarmid, and L. C. Hayek (eds), Measuring and Monitoring Biological
Diversity. Standard Methods for Amphibians, pp. 144–6. Smithsonian Institution Press,
Washington DC.
Gibbons, J. W. and Bennett, D. H. (1974). Determination of anuran terrestrial activity
patterns by a drift fence method. Copeia 1974, 236–43.
Gibbons, J. W. and Semlitsch, R. D. (1981). Terrestrial drift fences with pitfall traps: an
effective technique for quantitative sampling of animal populations. Brimleyana, 7,
1–16.
Gibbons, J. W., Winne, C. T., Scott, D. E., Willson, J. D., Glaudas, X., Andrews, K. M.,
Todd, B. D., Fedewa, L. A., Wilkinson, L., Tsaliagos, R. N. et al. (2006). Remarkable
amphibian biomass and abundance in an isolated wetland: Implications for wetland
conservation. Conservation Biology, 20, 1457–65.
13 Drift fences, coverboards, and other traps | 243
Gittins, S. P. (1983). The breeding migration of the common toad (Bufo bufo) to a pond
in mid-Wales. Journal of Zoology, London, 199, 555–62.
Grant, B. W., Tucker, A. D., Lovich, J. E., Mills, A. M., Dixon, P. M., and Gibbons, J. W.
(1992). The use of coverboards in estimating patterns of reptile and amphibian bio-
diversity. In D. R. McCullough, and R. H. Barrett (eds), Wildlife 2001, pp. 379–403.
Elsevier Science, London.
Greenberg, C. H., Neary, D. G., and Harris, L. D. (1994). A comparison of herpetofaunal
sampling effectiveness of pitfall, single-ended, and double-ended funnel traps used
with drift fences. Journal of Herpetology, 28, 319–24.
Griffiths, R. A. (1985). A simple funnel trap for studying newt populations and an evalu-
ation of trap behavior in smooth and palmate newts, Triturus vulgaris and Triturus
helveticus. Herpetological Journal, 1, 5–10.
Heyer, W. R., Donnelly, M. A., McDiarmid, R. W., and Hayek, L. C. (eds) (1994).
Measuring and Monitoring Biological Diversity. Standard Methods for Amphibians,
Smithsonian Institution Press, Washington DC.
Houze, C. M. and Chandler, C. R. (2002). Evaluation of coverboards for sampling terres-
trial salamanders in south Georgia. Journal of Herpetology, 36, 75–81.
Hsu, M. Y., Kam, Y. C., and Fellers, G. M. (2005). Effectiveness of amphibian monitor-
ing techniques in a Taiwanese subtropical forest. Herpetological Journal, 15, 73–9.
Jehle, R., Hodl, W., and Thonke, A. (1995). Structure and dynamics of central European
amphibian populations: a comparison between Triturus dobrogicus (Amphibia,
Urodela) and Pleobates fuscus (Amphibia, Anura). Australian Journal of Ecology, 20,
362–6.
Johnson, J. R. (2005). A novel arboreal pipe-trap designed to capture the gray treefrog
(Hyla versicolor). Herpetological Review, 36, 274–7.
Johnson, S. A. (2003). Orientation and migration distances of a pond-breeding salaman-
der (Notophthalmus perstriatus, Salamandridae). Alytes, 21, 3–22.
Johnson, S. A. and Barichivich, W. J. (2004). A simple technique for trapping Siren lac-
ertina, Amphiuma means, and other aquatic vertebrates. Journal of Freshwater Ecology,
19, 263–9.
Luhring, T. M. and Young, C. A. (2006). Innovative techniques for sampling stream-
inhabiting salamanders. Herpetological Review, 37, 181–3.
Marsh, D. M. and Goicochea, M. A. (2003). Monitoring terrestrial salamanders: biases
caused by intense sampling and choice of cover objects. Journal of Herpetology, 37,
460–6.
Mazerolle, M. J., Bailey, L. L., Kendall, W. L., Royle, J. A., Converse, S. J., and Nichols,
J. D. (2007). Making great leaps in herpetology: accounting for detectability in field
studies. Journal of Herpetology, 41, 672–89.
Mitchell, J. C., Erdle, S. Y., and Pagels, J. F. (1993). Evaluation of capture techniques for
amphibian, reptile, and small mammal communities in saturated forested wetlands.
Wetlands, 13, 130–6.
Moore, J. D. (2005). Use of native wood as a new coverboard type for monitoring red-
backed salamanders. Herpetological Review, 36, 268–71.
Moulton, C. A., Fleming, W. J., and Nerney, B. R. (1996). The use of PVC pipes to cap-
ture hylid frogs. Herpetological Review, 27, 186–7.
244 | Amphibian ecology and conservation
Palis, J. G., Adams, S. M., and Peterson, M. J. (2007). Evaluation of two types of
commercially-made aquatic funnel traps for capturing ranid frogs. Herpetological
Review, 38, 166–7.
Parris, K. M., Norton, T. W., and Cunningham, R. B. (1999). A comparison of tech-
niques for sampling amphibians in the forests of south-east Queensland, Australia.
Herpetologica, 55, 271–83.
Pauley, T. K. and Little, M. (1998). A new technique to monitor larval and juvenile sala-
manders in stream habitats. Banisteria, 12, 32–6.
Pechmann, J. H. K., Scott, D. E., Semlitsch, R. D., Caldwell, J. P., Vitt, L. J., and
Gibbons, J. W. (1991). Declining amphibian populations: the problem of separating
human impacts from natural fluctuations. Science, 253, 892–5.
Phillips, C. A. and Sexton, O. J. (1989). Orientation and sexual differences during
breeding migrations of the spotted salamander, Ambystoma maculatum. Copeia, 1989,
17–22.
Pittman, S. E., Jendrek, A. L., Price, S. J., and Dorcas, M. E. (2008). Habitat selection
and site fidelity of Cope’s gray treefrog (Hyla chrysoscelis) at the aquatic-terrestrial eco-
tone. Journal of Herpetology, 42, 378–85.
Resetarits, W. J. and Wilbur, H. M. (1991). Calling site choice by Hyla chrysoscelis: effects
of predators, competitors, and oviposition sites. Ecology, 72, 778–86.
Rice, A. N., Rice, K. G., Waddle, J. H., and Mazzotti, F. J. (2006). A portable non-
invasive trapping array for sampling amphibians and reptiles. Herpetological Review,
37, 429–30.
Ryan, T. J., Philippi, T., Leiden, Y. A., Dorcas, M. E., Wigley, T. B., and Gibbons, J. W.
(2002). Monitoring herpetofauna in a managed forest landscape: effects of habitat
types and census techniques. Forest Ecology and Management, 167, 83–90.
Semlitsch, R. D. (1985). Analysis of climatic factors influencing migrations of the sala-
mander Ambystoma talpoideum. Copeia, 1985, 477–89.
Semlitsch, R. D. and Pechmann, J. H.K. (1985). Diel patterns of migratory activity for
several species of pond-breeding salamanders. Copeia, 1985, 86–91.
Shaffer, H. B., Alford, R. A., Woodward, B. D., Richards, S. J., Altig, R. G., and Gascon, C.
(1994). Quantitative sampling of amphibian larvae. In W. R. Heyer, M. A. Donnelly,
R. W. McDiarmid, and L. C. Hayek (eds), Measuring and Monitoring Biological
Diversity. Standard Methods for Amphibians, pp. 130–41. Smithsonian Institution
Press, Washington DC.
Stevens, C. E. and Paszkowski, C. A. (2005). A comparison of two pitfall trap designs in
sampling boreal anurans. Herpetological Review, 36, 147–9.
Todd, B. D. and Winne, C. T. (2006). Ontogenetic and interspecific variation in timing
of movement and responses to climatic factors during migrations by pond-breeding
amphibians. Canadian Journal of Zoology, 84, 715–22.
Todd, B. D., Winne, C. T., Willson, J. D., and Gibbons, J. W. (2007). Getting the drift:
examining the effects of timing, trap type, and taxon on herpetofaunal drift fence sur-
veys. American Midland Naturalist, 158, 292–305.
Vogt, R. C. and Hine, R. L. (1982). Evaluation of techniques for assessment of amphibian
and reptile populations in Wisconsin. In N. J. Scott, Jr (ed.). Herpetological Communities,
pp. 201–17. Wildlife Research Report 13. United States Fish and Wildlife Service,
Washington DC.
13 Drift fences, coverboards, and other traps | 245
Waldron, J. L., Dodd, Jr, C. K., and Corser, J. D. (2003). Leaf litterbags: factors affecting
capture of stream-dwelling salamanders. Applied Herpetology, 1, 23–6.
Weddeling, K., Hachtel, M., Sander, U., and Tarkhnishvili, D. (2004). Bias in estimation
of newt population size: a field study at five ponds using drift fences, pitfalls, and funnel
traps. Herpetological Journal, 14, 1–7.
Willson, J. D. and Dorcas, M. E. (2003). Quantitative sampling of stream salamanders:
comparison of dipnetting and funnel trapping techniques. Herpetological Review, 34,
128–30.
Willson, J. D. and Dorcas, M. E. (2004). A comparison of aquatic drift fences with
traditional funnel trapping as a quantitative method for sampling amphibians.
Herpetological Review, 35, 148–50.
Willson, J. D., Winne, C. T., and Fedewa, L. A. (2005). Unveiling escape and capture
rates of aquatic snakes and salamanders (Siren spp. and Amphiuma means) in commer-
cial funnel traps. Journal of Freshwater Ecology, 20, 397–403.
This page intentionally left blank
14
Area-based surveys
David M. Marsh and Lillian M.B. Haywood
Table 14.1 Summary of a literature review of 89 studies that used area-based sampling.
Studies are divided by whether the primary sampling units were plots/quadrats, transects,
transects nested within plots, or quadrats nested within plots or transects. Summary data
show the percentage of studies of each type that were: (1) terrestrial, aquatic, or both;
(2) tropical or temperate; (3) single-species or multi-species studies, (4) targeting frogs,
salamanders, caecilians, or multiple groups; and (5) conducted during daytime, at night,
or both.
Approach N Terrestrial/ Tropical/ Single/ Frogs/salamanders/ Day/night/
aquatic/both temperate multiple caecilians/multiple both
species
on our own experience using plots and transects to sample terrestrial salaman-
ders, stream salamanders, and neotropical frogs.
soil underneath cover objects is wetter than the surrounding leaf litter. Cover-
object surveys, like leaf-litter surveys, can disturb the microhabitats of interest,
and Smith and Petranka (2000) recommend re-sampling plots once annually.
Habitat data collected during cover surveys usually include herb cover, canopy
cover, soil moisture, and soil temperature, as well as data on the number, type,
and decay classes of cover objects searched. Salamander counts can be divided
by the number of cover objects on a plot to correct for differences in search
effort (e.g. Marsh and Beckman 2004); however, whether or not this yields
more accurate estimates of relative abundance has not been tested.
The ability to resample salamanders more frequently without degrading cover
objects is one big advantage for artificial cover objects (ACOs; see Chapter 13)
over natural-cover surveys. However, counts from grids of ACOs may be only
weakly correlated with counts from natural-cover surveys (Hyde and Simons
2001), and ACOs may under-represent smaller size classes of salamanders
(Marsh and Goicochea 2003). Thus, natural-cover surveys are recommended
over ACOs for relative abundance and diversity estimation until ACOs have
been more thoroughly validated.
Temperature and moisture are often positively associated with frog calling activ-
ity (Townsend and Stewart 1994; Duellman 1995), although very heavy rains
can actually reduce activity. Practicing transects in a variety of weather con-
ditions can allow one to determine the best times for carrying out nocturnal
transects. Habitat data measured in conjunction with nocturnal transects often
include herb cover, canopy cover, tree stems, and tree diameter at breast height.
14.4 Modifications
The previous examples highlight some of the basic uses of plot and transect
surveys. However, several modifications to these approaches have also been pro-
posed. We cover two in detail: distance sampling (a modification of transect
surveys) and adaptive cluster sampling (a modification of quadrat surveys).
is small and round or rectangular, plots are far more practical than very short
transects. Also, plots may be useful when many observers are available for sam-
pling. Observers can cover a rectangular plot in parallel, making these more con-
venient to sample than several distinct transects. Finally, plots should generally be
selected when one is sampling amphibians in conjunction with mark–recapture
analysis. Even if amphibians move only a few meters between captures, they will
tend to wander off of transects and small quadrats, leading to low detection rates
and difficulty with estimating population parameters (Schaub et al. 2004).
These suggestions are valid for all animals, but for amphibians there are
additional considerations. Because amphibian activity is highly dependent on
weather conditions, amphibian counts can vary by orders of magnitude from
day to day or even from hour to hour. There are two main approaches to deal-
ing with this kind of variation. The first is to restrict surveys to periods of time
when amphibians are most likely to be observed. This requires flexibility with
scheduling surveys but can dramatically reduce the amount of survey effort
needed. It also requires some initial surveys to get a sense of amphibian activity
patterns. The second approach is to carry out surveys in a variety of condi-
tions, but to collect data on weather-related variables that may be associated
with activity levels (e.g. rainfall within the past week, rainfall at the time of
the survey, temperature, humidity). These variables can then be included as
covariates in any analysis of amphibian counts, thereby controlling some of the
background variation.
In conjunction with either of these two approaches, surveys can be paired or
grouped in a way that ensures that background variation will not confound a
hypothesis being tested. For example, in a comparison of forest to pasture, each
forest plot could be paired with a pasture plot so that one of each (or two or three
of each) is surveyed on the same date. The pairing of these surveys ensures that
any day-to-day variation affects both habitats equally. Additionally, by using
survey date as a random effect in the statistical analysis, one can remove most of
this variation from the variable of interest.
Pasture
Pasture
Forest
Forest
Pasture Pasture
Forest
Forest
Fig. 14.1 Four possible designs for area-based sampling to compare a single forest
area to surrounding pasture. Equal numbers of samples would be randomly located
within pasture but these are not shown. (a) Five plots of 200 m2. (b) Ten transects
of 50 m 2 m. (c) Forty 5 m 5 m quadrats. (d) Four 5 m 5 m quadrats nested
within each of ten 200 m2 plots.
(a) Transects
Pasture
Forest
Forest
Pasture
(b) Quadrats
Pasture
Forest
Forest
Pasture
Fig. 14.2 Two possible designs for area-based sampling to compare five forest
patches to surrounding pasture. (a) Two transects of 50 m 2 m within each forest
patch. (b) Eight 5 m 5 m quadrats within each forest patch.
14 Area-based surveys | 259
14.9 References
Allmon, W. D. (1991). A plot study of forest floor litter frogs, central Amazon, Brazil.
Journal of Tropical Ecology, 7, 503–22.
14 Area-based surveys | 261
Bailey, L. L., Simons, T. R., and Pollock, K. H. (2004). Spatial and temporal variation
in detection probability of Plethodon salamanders using the robust capture-recapture
design. Journal of Wildlife Management, 68, 14–24.
Barr, G. E. and Babbitt, K. J. (2002). Effects of biotic and abiotic factors on the distribu-
tion and abundance of larval two-lined salamanders (Eurycea bislineata) across spatial
scales. Oecologia, 133, 176–85.
Cochran, W. G. (1977). Sampling Techniques. Wiley, New York.
Dodd, Jr, C. K. (1990). Line transect estimation of Red Hills salamander burrow density
using a Fourier series. Copeia, 1990, 555–7.
Duellman, W. E. (1995). Temporal fluctuations in abundances of anuran amphibians in
a seasonal Amazonian rainforest. Journal of Herpetology, 29, 13–21.
Fogarty, J. H. and Vilella, F. J. (2001). Evaluating methodologies to survey
Eleutherodactylus frogs in montane forests of Puerto Rico. Wildlife Society Bulletin,
29, 948–55.
Funk, W. C., Almeida-Reinoso, D., Nogales-Sornosa, F., and Bustamante, M. R. (2003).
Monitoring population trends of Eleutherodactylus frogs. Journal of Herpetology, 37,
245–56.
Gibbs, J. P., Droege, S., and Eagle, P. (1998). Monitoring populations of plants and animals.
Bioscience, 48, 935–40.
Gower, D. J., Loader, S. P., Moncrieff, C. B., and Wilkinson, M. (2004). Niche separ-
ation and comparative abundance of Boulengerula boulengeri and Scolecomorphus vittatus
(Amphibia: Gymnophiona) in an East Usambara forest, Tanzania. African Journal of
Herpetology, 53, 183–90.
Gregory, R. D., Gibbons, D. W., and Donald, P. F. (2004). Bird census and survey
techniques. In W. J. Sutherland, I. Newton, and R. E. Green (eds), Bird Ecology and
Conservation: a Handbook of Techniques, pp. 17–55. Oxford University Press, Oxford.
Hairston, Sr, N. G. (1987). Community Ecology and Salamander Guilds. Cambridge
University Press, New York.
Hyde E. J. and Simons, T. R. (2001). Sampling plethodontid salamanders: sources of
variability. Journal of Wildlife Management, 65, 624–32.
Inger, R. F. (1980). Densities of floor–dwelling frogs and lizards in lowland forests of SE
Asia and Central America. American Naturalist, 115, 761–70.
Jaeger, R. and Inger, R. F. (1994). Standard techniques for inventory and monitoring:
quadrat sampling. In W. R. Heyer, M. A. Donnely, R. W. McDiarmid, L. C. Hayek,
and M. S. Foster (ed.), Measuring and Monitoring Biological Diversity. Standard Methods
for Amphibians, pp. 97–102. Smithsonian Institution Press, Washington DC.
Lehtinen, R. M., Ramanamanjato, J. B., and Raveloarison, J. G. (2003). Edge effects and
extinction proneness in a herpetofauna from Madagascar. Biodiversity and Conservation,
12, 1357–70.
Marsh, D. M. and Pearman, P. B. (1997). Effects of habitat fragmentation on the abun-
dance of two species of Leptodactylid frogs in an Andean montane forest. Conservation
Biology, 11, 1323–8.
Marsh, D. M. and Goicochea, M. A. (2003). Monitoring terrestrial salamanders: biases
due to frequent sampling and choice of cover objects. Journal of Herpetology, 37,
460–6.
262 | Amphibian ecology and conservation
Marsh, D. M. and Beckman, N. G. (2004). Effects of forest roads on the abundance and
activity of terrestrial salamanders in the Southern Appalachians. Ecological Applications,
14, 1882–91.
Measey, G. J. (2006). Surveying biodiversity of soil herpetofauna: towards a standard
quantitative methodology. European Journal of Soil Biology, 42, S103–10.
Measey, G. J., Gower, D. J., Oommen, O. V., and Wilkinson, M. (2003). Quantitative
surveying of endogeic limbless vertebrates – a case study of Gegeneophis ramaswamii
(Amphibia: Gymnophiona: Caeciliidae) in southern India. Applied Soil Ecology, 23,
43–53.
Noon, B. R., Ishwar, N. M., Vasudevan, K. (2006). Efficiency of adaptive cluster and
random sampling in detecting terrestrial herpetofauna in a tropical rainforest. Wildlife
Society Bulletin, 34, 59–68.
Pearman, P. B. (1997). Correlates of amphibian diversity in an altered landscape of
Amazonian Ecuador. Conservation Biology, 11, 1211–25.
R Development Core Team (2005). R: a Language and Environment for Statistical
Computing, Reference Index Version 2.2.1. www.R-project.org. R Foundation for
Statistical Computing, Vienna.
Rocha, C. F. D., Van Sluys, M., Alves, M. A.S., Bergallo, H. G., and Vrcibradic, D. (2000).
Activity of leaf-litter frogs: when should frogs be sampled? Journal of Herpetology, 34,
285–7.
Rocha, C. F. D., Van Sluys, M., Alves, M. A. S., Bergallo, H. G., and Vrcibradic, D.
(2001). Estimates of forest floor litter frog communities: a comparison of two methods.
Austral Ecology, 26, 14–21.
Schaub, M., Gimenez, O., Schmidt, B. R., and Pradel, R. (2004). Estimating survival
and temporary emigration in the multistate capture–recapture framework. Ecology,
85, 2107–13.
Smith, C. K. and Petranka, J. W. (2000). Monitoring terrestrial salamanders: repeatabil-
ity and validity of area-constrained cover object searches. Journal of Herpetology, 34,
547–57.
Thomas, L., Laake, J. L., Strindberg, S., Marques, F. F. C., Buckland, S. T., Borchers,
D. L., Anderson, D. R., Burnham, K. P., Hedley, S. L., Pollard, J. H. et al. (2006).
Distance 5.0. Release 2. www.ruwpa.st-and.ac.uk/distance/. Research Unit for Wildlife
Population Assessment, University of St Andrews, St Andrews.
Thompson, S. K. and Seber, G. A. F. (1996). Adaptive Sampling. John Wiley and Sons,
New York.
Toft, C. A. (1980). Feeding ecology of thirteen syntopic species of anurans in a seasonal
tropical environment. Oecologia, 45, 131–41.
Townsend, D. S. and Stewart, M. M. (1994). Reproductive ecology of the Puerto Rican
frog Eleutherodactylus coqui. Journal of Herpetology, 28, 34–40.
Vonesh, J. R. (2001). Patterns of richness and abundance in a tropical African leaf-litter
herpetofauna. Biotropica, 33, 502–10.
Woolbright, L. L. (1991). The impact of Hurricane Hugo on forest frogs in Puerto Rico.
Biotropica, 23, 462–7.
15
Rapid assessments of amphibian diversity
James R. Vonesh, Joseph C. Mitchell, Kim Howell, and
Andrew J. Crawford
information. The best candidate sites are those that are only marginally sur-
veyed and are highly threatened (Sayre and Roca 2000). It is in these cases that
data resulting from an RA, though typically limited in scope, can have import-
ant implications for decision-making.
For example, the Atewa Range Forest Reserve of Ghana contains some of the
last remaining upland evergreen forest in the country and is home to endemic
plant and insect species (McCullough et al. 2007). In 2006, Conservation
International initiated a Rapid Assessment Program of amphibians (and other
taxa) within the Atewa Range Forest Reserve in response to proposed mineral
exploration in the reserve. Sampling sites were selected in areas suspected to sup-
port high biodiversity that also had large mineral deposits (McCullough et al.
2007). Field surveys were conducted over just 18 days at the beginning of the
rainy season. Nearly one-third of the 32 amphibian species observed were listed
as threatened on the IUCN Red List. One species was considered Critically
Endangered (Kouamé et al. 2007). Based in part on the amphibian assessment
results, the authors recommended that Atewa Range Forest Reserve be elevated
to national park status and that mineral exploitation be prohibited (Kouamé
et al. 2007; McCullough et al. 2007).
15.2 Planning an RA
When an RA is indicated, the next steps involve planning and preparation.
Because RAs may involve government and non-government participants from
multiple regions, are often conducted in remote, logistically challenging field
localities, and investigate unknown or poorly described faunas, substantial
advanced planning is necessary for success.
15.2.4 Permits
Conducting an RA may require permits from local, regional, national, and inter-
national governing agencies. Most national governments, and many state and
local governments, require permits to study amphibians within their boundaries
and public lands. Many protected areas (e.g. national parks) require additional
permission. National agencies and international treaties (e.g. Convention on
International Trade in Endangered Species of Wild Fauna and Flora (CITES),
U.S. Fish and Wildlife Service, Convention on Biological Diversity) also regu-
late export and import of specimens and tissue samples. There is considerable
variation among regions in permit fees and the times required to process and
receive permits, ranging from weeks to well over a year. Thus, the permitting
process should be initiated well in advance of the planned fieldwork.
other data sets, and images or video or audio clips. Data management is the process
that helps ensure that diverse data sets can be efficiently collected, integrated and
analyzed, and archived so that they can be easily retrieved in the future (Gotelli
and Ellison 2004). Planning for good data management begins early in project
development by making a table of the types of data to be produced in the field;
for example habitat variables (Chapter 17), conditions, counts, specimens, and
other media such as photos, audio (Chapter 16), or video. This information can
be used to design protocols (e.g. field forms, labels, code numbers) for collection
and management of disparate types of data. Once data are being generated it is
important to review data in the field, and verify that team members are using iden-
tical protocols. Data should be transcribed regularly and transferred to a database
as soon as possible. If a portable computer is brought to the field, raw data should
still be maintained on paper in case of damage to the computer. Widely used
software for data management include Microsoft Access and the non-proprietary
OpenOffice BASE. Protocols need to be established for reviewing, cleaning, and
backing up data on a regular basis. Museum specimen data should be stored in a
manner compliant with the Distributed Generic Information Retrieval (DiGIR)
protocols, such as the database management software, Specify (http://www.
specifysoftware.org).
sampling. Finally, GIS (Chapter 19) coupled with ecological niche modeling pro-
vide a new tool for remotely identifying potential habitat for species of concern or
biodiversity hotspots in general (e.g. Raxworthy et al. 2003; Pawar et al. 2007).
Development and accessibility of new mapping technologies over the past two
decades have had important impacts on the way RAs are conducted. Spatial tech-
nologies, including GIS, remote sensing, and GPS are commonly used in defin-
ing project scope and establishing sampling localities. The Rapid Ecological
Assessment approach developed by Conservation International places strong
emphasis on using satellite and aerial imagery to classify landscapes of interest
into vegetation or land-use cover categories (Sayre et al. 2000). Sampling sites are
spread across these vegetation categories to establish the framework within which
field sampling is conducted and to facilitate linkages between fine-scale sampling
by field teams with landscape-level assessment of biodiversity (Nagendra and
Gadgil 1999; Sayre et al. 2000). Sayre et al. (2000) provide a detailed example of
how natural-colour and colour-infrared satellite and aerial fly-over imagery were
used to develop an initial landscape characterization of Parque Nacional del Este,
Dominican Republic, which was subsequently used to identify specific sampling
sites and generate detailed site maps. While the utility of remote sensing imagery
when planning an RA is readily apparent, costs can be considerable. Commercial
satellite imagery may cost US$3000–5000 per scene and an aerial photo acqui-
sition mission may cost between $20 000 and 120 000 (Sayre et al. 2000). For
projects with limited resources, Google Earth™ maps integrate data from satel-
lite imagery, aerial photography, and a GIS 3D globe to provide resolution of at
least 15 m for most terrestrial areas.
We draw a distinction between the ways sampling sites are determined in most
RA studies compared to traditional ecological inventory methods. Traditional
ecological inventories emphasized objective field sampling based on rand-
omized selection of replicated samples and substantial effort may be given to
issues such as defining sampling coverage and determining detection probabil-
ities (e.g. Buckland et al. 1993; 2001, 2004; Heyer et al. 1994; Williams et al.
2002). This emphasis on randomization, replication, and estimating detection
functions greatly increases the kinds of inferences that can be made. However,
for many RAs, sampling may be opportunistic, or determined by issues such as
logistics, access, efficiency, and a priori perceptions of areas likely to be high-
est in diversity or under the greatest threat. Although replicate units may be
sampled in some cases, the resulting data may not be rigorously applied to all
questions of interest to the ecologist or biogeographer (Sayre 2000). As a result,
the results from many RAs are best viewed as qualitative or semi-quantitative
(e.g. Kouamé et al. 2007).
270 | Amphibian ecology and conservation
(e.g. animals are being processed). The number of animals observed can then
be expressed in terms of animals observed per area (or distance) per person
searching, or per unit time per person. VES is an effective technique for build-
ing species lists rapidly (Crump and Scott, 1994; Rodda et al. 2007), requires
little equipment (e.g. minimally a headlamp and field notebook), and can
be implemented in a variety of habitat types. For these reasons it is perhaps the
most widely employed sampling method in RA of amphibian diversity, so we
provide a brief overview of the method here. For additional information see
Crump and Scott (1994), Corn and Bury (1990), Campbell and Christman
(1982), Rödel and Ernst (2004), and Rueda-Almonacid et al. (2006).
There is considerable variation among past studies in how VES is conducted.
In some instances (e.g. Mitchell 2006; Kouamé et al. 2007) experienced her-
petologists selectively search areas and microhabitats determined most likely to
yield amphibians. This approach has the benefit of potentially yielding more
animals and species per effort than randomized sampling approaches. However,
it is most subject to variation in the skill levels of the field personnel and is lim-
ited in its ability to generalize about relative abundances and habitat associations
because (micro-)habitats are searched in a biased manner. Alternatively, an area
may be searched via VES using a randomized-walk design (Crump and Scott
1994). In this case, prior to going to the field site the researcher generates a ran-
dom sequence of compass headings as well as a random distance to be searched
along each heading. A starting heading and distance is selected at random and
field personnel simply work through the list of headings and distances, search-
ing for animals, for a specified time. Since the path to be sampled is determined
randomly, this approach has the advantage that replicated random walks from
different areas could be compared statistically. However, it is important to appre-
ciate that differences in the observed number of animals among sites may be as
much due to differences in detection among sites as much as differences in true
abundance. If the sampling design requires greater effort per area, VES can be
conducted using a quadrat design (Crump and Scott 1994). In this case, quad-
rats of a known area are established and each is then sampled systematically by
searching parallel transects across the plot (e.g. Hairston 1980; Aichinger 1987;
Donnelly 1989). Crump and Scott (1994) recommend plot sizes of 10 m 10 m
or 25 m 25 m, depending upon amphibian densities. To statistically compare
areas of interest (e.g. different habitats) plots must be replicated and randomly
located within areas to be compared (e.g. using GPS coordinates or a trail grid
system). Finally, transect designs are often used in coordination with VES for
sampling across habitat gradients that may affect amphibian diversity and abun-
dance (e.g. moisture, elevation). Although these considerations may determine
272 | Amphibian ecology and conservation
transect direction, distances to be sampled and starting points are best deter-
mined a priori and multiple randomly located replicate transects will be needed
if areas are to be compared statistically.
It is also important to determine and carefully define the intensity of field
sampling in advance. Crump and Scott (1994) suggest three levels of sampling
intensity for sampling amphibians. The least intensive surveys are counts of
animals that are active on the surface (Hairston 1980; Vonesh 2000). This
approach is minimally invasive and thus is suitable for search areas with sensi-
tive faunas and also requires the least amount of time, increasing the amount
of area that can be covered. Intermediate-level searches count exposed ani-
mals but also turn over surface cover objects (note that cover objects must
be returned to their original position to minimize disturbance). This level
of VES can yield species often overlooked by low-intensity VES because of
their secretive and semi-fossorial life histories (e.g. many salamanders and
caecilians). When the most complete inventory possible is desired, all pos-
sible microhabitats are searched; surface objects are turned, decaying logs are
torn apart, the leaf litter is systematically raked, epiphytes and tree holes are
searched (Crump and Scott 1994), and aquatic habitats are searched for adults
and larvae (Chapter 4). Such intensity requires considerable labor per unit
area and causes the greatest disturbance to the habitat but may be more effect-
ive at sampling some rare species.
lucid reviews of these issues and specifically focus on their application to her-
petological studies. Given these limitations, VES is best viewed as a qualitative
or semi-quantitative approach and should be used only when qualitative results
are acceptable (e.g. threatened species were observed; Kouamé et al. 2007). VES
may be a suitable sampling approach given the objectives of many RA projects.
However, when resources and objectives allow, other more powerful study designs
(e.g. distance sampling methods; Thomas et al. 2002; Funk et al. 2003; Fogarty
and Viletta 2001; Viletta and Fogarty 2005) should be considered.
15.6 Summary
Rapid assessment is an approach that can be useful for conservation planning
because it can provide an efficient preliminary characterization of the amphibian
fauna. However, because the methods are often qualitative or semi-quantitative,
the RA approach is of most value when the fauna of the study areas is unknown
and when even qualitative information is needed urgently to help decision-
making. When resources, time, and project objectives allow, researchers should
consider more quantitative approaches.
276 | Amphibian ecology and conservation
15.7 References
Aguirre, A. A. and Lampo, M. (2006). Protocolo de bioseguridad y cuarentena para preve-
nir la transmisión de enfermedades en anfibios. In A. Angulo, J. V. Rueda-Almonacid,
J. V. Rodríguez-Mahecha, and E. La Marca (eds), Técnicas de Inventario y Monitoreo
para los Anfibios de la Región Tropical Andina, pp. 73–92. Conservación Internacional,
Bogotá.
Aichinger, M. (1987). Annual activity patterns of anurans of anurans in a seasonal
Neotropical environment. Oecologia, 71, 583–92.
AmphibiaWeb (2008). Information on Amphibian Biology and Conservation. AmphibiaWeb,
Berkeley, CA. http://amphibiaweb.org.
Bailey, L. L., Simons, T. R., and Pollock, K. H. (2004). Estimating site occupancy
and species detection probability parameters for terrestrial salamanders. Ecological
Applications, 14, 692–702.
Buckland, S. T., Anderson, D. R., Burnham, K. P., and Laake, J. L. (1993). Distance
Sampling: Estimating Abundance of Biological Populations. Chapman and Hall,
London.
Buckland, S. T., Anderson, D. R., Burnham, K. P., Laake, J. L., Borchers, D. L., and
Thomas, L. (2001). Introduction to Distance Sampling: Estimating Abundance of
Biological Populations. Oxford University Press, New York.
Buckland, S. T., Anderson, D. R., Burnham, K. P., Laake, J. L., Borchers, D., and
Thomas, L. (2004). Advanced Distance Sampling: Estimating Abundance of Biological
Populations. Oxford University Press, Oxford.
Campbell, H. W. and Christman, S. P. (1982). Field techniques for herpetofaunal com-
munity analysis. In N. J. Scott, Jr (ed.), Herpetofaunal Communities, pp. 193–200.
Wildlife Research Report 13. U.S. Department of the Interior, Fish and Wildlife
Service, Washington DC.
Chao, A., Chazdon, R. L., Colwell, R. K., and Shen, T.-J. (2005). A new statistical
approach for assessing similarity of species composition with incidence and abundance
data. Ecology Letters, 8, 148–59.
Corn, P. S. and Bury, R. B. (1990). Sampling Methods for Terrestrial Amphibians and Reptiles.
General Technical Report, PNW-GTR-256. U.S. Forest Service, Fort Collins, CO.
Crump, M. L. and Scott, Jr, N. J. (1994). Visual encounter surveys. In W. R. Heyer,
M. A. Donnelly, R. W. McDiarmid, L. A. C. Hayek, and M. S. Foster (eds), Measuring
and Monitoring Biological Diversity, Standard Methods for Amphibians, pp. 84–92.
Smithsonian Institution Press, Washington DC.
Donnelly, M. A. (1989). Demographic effects of reproductive resource supplementation
in a territorial frog, Dendrobates pumilio. Ecological Monographs, 59, 207–21.
Fellers, G. M. and Freel, K. L. (1995). A standardized protocol for surveying aquatic amphib-
ians. Technical Report NPS/WRUC/NRTR-95-01. U.S. National Park Service,
Davis, CA.
Ficetola, G. F., Miaud, C., Pompanon, F., and Taberlet, P. (2008). Species detection using
environmental DNA from water samples. Biology Letters, 4, 423–5.
Fogarty, J. H. and Vilella, F. J. (2001). Evaluating methodologies to survey Eleutherodactylus
frogs in montane forests of Puerto Rico. Wildlife Society Bulletin, 29, 948–55.
15 Rapid assessments of amphibian diversity | 277
Fouquet, A., Gilles, A., Vences, M., Marty, C., Blanc, M., and Gemmell, N. J. (2007).
Underestimation of species richness in Neotropical frogs revealed by mtDNA analyses.
PLoS ONE, 2, e1109.
Funk, W., Almeida-Reinoso, D., Nogales-Sornosa, F., and Bustamante, M. (2003).
Monitoring population trends of Eleutherodactylus frogs. Journal of Herpetology, 37,
245–56.
Goldberg, C. S., Kaplan, M. E., and Schwalbe, C. R. (2003). From the frog’s mouth:
buccal swabs for collection of DNA from amphibians. Herpetological Review, 34,
220–1.
Gonser, R. A. and Collura, R. V. (1996). Waste not, want not: toe-clips as a source of
DNA. Journal of Herpetology, 30, 445–7.
Gotelli, N. J. and Colwell, R. K. (2001). Quantifying biodiversity: procedures and pitfalls
in the measurement and comparison of species richness. Ecology Letters, 4, 379–91.
Gotelli. N. J. and Ellison, A. M. (2004). Chapter 8: Managing and curating data. A
Primer of Ecological Statistics, pp. 207–36. Sinauer Associates, Sunderland, MA.
Hairston, N. G. (1980). The experimental test of an analysis of field distributions:
Competition in terrestrial salamanders. Ecology, 61, 817–26.
Hayek, L.-A. (1994). Research design for quantitative amphibian studies. In W. R. Heyer,
M. A. Donnelly, R. W. McDiarmid, L. A. C. Hayek, and M. S. Foster (eds), Measuring
and Monitoring Biological Diversity, Standard Methods for Amphibians, pp. 21–38.
Smithsonian Institution Press, Washington DC.
Heyer, W. R., Donnelly, M. A., McDiarmid, R. W., Hayek, L. A. C., and Foster, M. S.
(eds) (1994). Measuring and Monitoring Biological Diversity, Standard Methods for
Amphibians. Smithsonian Institution Press, Washington DC.
Howell, K. M. (2002). Amphibians and reptiles: Herptiles. In G. Davies, (ed.), African
Forest Biodiversity, pp. 17–44. Earthwatch, London.
IUCN, Conservation International, and NatureServe (2006). Global Amphibian
Assessment. http://globalamphibians.org.
Jacobs, J. F. and Heyer, W. R. (1994). Collecting tissue for biochemical analysis. In
W. R. Heyer, M. A. Donnelly, R. W. McDiarmid, L. A. C. Hayek, and M. S. Foster
(eds), Measuring and Monitoring Biological Diversity, Standard Methods for Amphibians,
pp. 103–7. Smithsonian Institution Press, Washington DC.
Kerr, J. T., Sugar, A., and Packer, L. (2000). Indicator taxa, rapid biodiversity assessment,
and nestedness in an endangered ecosystem. Conservation Biology, 14, 1726–34.
Kouamé, N. G., Bonteng, C. O., and Rödel, M.-O. (2007). A rapid survey of the amphib-
ians from Atewa Range Forest Reserve, Eastern Region, Ghana. In J. McCullough,
L. E. Alonso, P. Naskrescki, H. E. Wright, and Y. Osei-Owusu (eds), A Rapid Biological
Assessement of the Atewa Range Forest Reserve, Eastern Ghana. RAP Bulletin of Biological
Assessment 47, pp. 76–83. Conservation International, University of Chicago Press,
Chicago, IL.
Lips, K. R., Reaser, J. K., Young, B. E., and Ibáñez, R. (2001). Amphibian monitoring in
Latin America: a protocol manual/Monitoreo de Anfibios en América Latina: Manual
de protocolos. SSAR Herpetological Circular no. 30.
Magnusson, W. E. and Mourão, G. (2004). Statistics without Math. Sinauer Associates,
Sunderland, MA.
278 | Amphibian ecology and conservation
Roelants, K., Gower, D. J., Wilkinson, M., Loader, S. P., Biju, S. D., Guillaume, K.,
Moriau, L., and Bossuyt, F. (2007). Global patterns of diversification in the history
of modern amphibians. Proceedings of the National Academy of Sciences USA, 104,
887–92.
Rueda-Almonacid, J. V., Castro, F., and Cortez, C. (2006). Técnicas para el inventario
y muestreo de anfibios: Una compilación. In A. Angulo, J. V. Rueda-Almonacid,
J. V. Rodríguez-Mahecha, and E. La Marca (eds), Técnicas de Inventario y
Monitoreo para los Anfibios de la Región Tropical Andina, pp. 135–71. Conservación
Internacional, Bogotá.
Sayre, R. (2000). Overview: Rapid Ecological Assessment after ten years. In R. Sayre,
E. Roca, G. Sedaghatkish, B. Young, S. Keel, R. L. Roca, and S. Sheppard (eds), Nature
in Focus Rapid Ecological Assessment, pp. 1–18. Island Press, Washington DC.
Sayre, R. and Roca E. (2000). Careful planning: a key to success. In R. Sayre, E. Roca,
G. Sedaghatkish, B. Young, S. Keel, R. L. Roca, and S. Sheppard (eds), Nature in Focus
Rapid Ecological Assessment, pp. 33–44. Island Press, Washington DC.
Sayre, R., Roca, E., Sedaghatkish, G., Young, B., Keel, S., Roca, R. L., and
Sheppard S. (eds) (2000). Nature in Focus: Rapid Ecological Assessment. Island
Press, Washington DC.
Schmidt, B. (2004). Declining amphibian populations: the pitfalls of count data in the
study of diversity, distributions, dynamics, and demography. Herpetological Journal,
14, 167–74.
Schmidt, B. (2005). Monitoring the distribution of pond breeding amphibians when spe-
cies are detected imperfectly. Aquatic Conservation: Marine and Freshwater Ecosystems,
15, 681–92.
Seutin, G., White, B. N., and Boag, P. T. (1991). Preservation of avian blood and tissue
samples for DNA analyses. Canadian Journal of Zoology, 69, 82–90.
Smith, M. A., Poyarkov, Jr, N. A., and Hebert, P. D. N. (2008). CO1 DNA barcod-
ing amphibians: take the chance, meet the challenge. Molecular Ecology Resources, 8,
235–46.
Thomas, L., Buckland, S. T., Burnham, K. P., Anderson, D. R., Laake, J. L., Borchers, D. L.,
and Strindberg, S. (2002). Distance sampling. In A. H. El-Shaarawi and W. W. Piegorsch
(eds), Encyclopedia of Environmetrics, vol. 1, pp. 544–52. John Wiley & Sons, Chichester.
Timm, R. M. (1994). The mammal fauna. In L. A. McDade, K. S. Bawa, H. A. Hespenheide,
and G. S. Hartshorn (eds), LaSelva: Ecology and Natural History of a Neotropical Rainforest,
pp. 229–37. University of Chicago Press, Chicago, IL.
Vences, M., Thomas, M., Bonett, R. M., and Vieites, D. R. (2005). Deciphering
amphibian diversity through DNA barcoding: chances and challenges. Philosophical
Transactions of the Royal Society of London Series B Biological Sciences, 360, 1859–68.
Vilella, F. and Fogarty, J. (2005). Diversity and abundance of forest frogs (Anura:
Leptodactylidae) before and after Hurricane Georges in the Cordillera Central of
Puerto Rico. Caribbean Journal of Science, 41,157–62.
Vonesh, J. R. (2000). Dipteran predation on the arboreal eggs of four Hyperolius frog spe-
cies in western Uganda. Copeia, 2000, 560–6.
Vonesh, J. R. (2001a). Patterns of richness and abundance in a tropical African leaf-litter
herpetofauna. Biotropica, 33, 502–10.
280 | Amphibian ecology and conservation
Vonesh, J. R. (2001b). Natural history and biogeography of the amphibians and reptiles
of Kibale National Park, Uganda. Contemporary Herpetology. http://www.nhm.ac.uk/
hosted_sites/ch/.
Whiles, M. R., Lips, K. R., Pringle, C. M., Kilham, S. S., Bixby, R. J., Brenes, R.,
Connelly, S., Colon-Gaud, J. C., Hunte-Brown, M., Huryn, A. D. et al. (2006). The
effects of amphibian population declines on the structure and function of Neotropical
stream ecosystems. Frontiers in Ecology and the Environment, 4, 27–34.
Williams, B. K., Nichols, J. D., and Conroy, M. J. (2002). Analysis and Management of
Animal Populations. Academic Press, New York.
Young, B., Sedaghatkish, G., and Roca, R. (2000). Fauna surveys. In R. Sayre, E. Roca,
G. Sedaghatkish, B. Young, S. Keel, R. L. Roca, and S. Sheppard (eds), Nature in Focus
Rapid Ecological Assessment. pp. 93–117. Island Press,Washington DC.
16
Auditory monitoring of anuran populations
Michael E. Dorcas, Steven J. Price, Susan C. Walls, and
William J. Barichivich
16.1 Introduction
Because anurans rely on vocalization for most communication, detection of
species-specific calls provides relatively efficient mechanisms for studying and
evaluating the status of anuran populations. Consequently, most amphibian
monitoring programs focus on anurans and use call detection as their sole or pri-
mary monitoring technique. The overall goal of these programs is to determine
and monitor the status of populations over time. Some monitoring programs
may actually attempt some form of quantification of anuran population size or
density, but this is often difficult when using only calling data. In this chapter we
describe approaches for monitoring anuran populations based solely on auditory
techniques. These approaches include manual calling surveys (MCS), automated
recording systems (ARS), or some combination of the two. MCS can be used by
researchers to monitor multiple amphibian populations, often over large spatial
scales. ARS can be used by researchers interested in intensively monitoring popu-
lations of anurans at a single or a few locations. In this chapter, we describe both
approaches and the potential costs and benefits of each. We also explain how
data resulting from ARS can be used to optimize manual survey protocol and to
interpret data resulting from MCS. Our goal is to provide an overview of these
techniques and questions that can be addressed using them.
16.2 MCS
Generally, MCS simply involve observers listening to the vocalizations of male
frogs and recording all species detected for the duration of the survey or for a
given area. In many surveys, observers also score abundance of each species
282 | Amphibian ecology and conservation
as well as some European (e.g. Anthony 2002; Pellet and Schmidt 2005; Schmidt
2005; Scott et al. 2008), and Central American countries (e.g. Kaiser 2008).
researchers typically allow 1 min between arriving at each site and beginning the
chorus survey. All anuran species heard in a specified time frame (usually 5 min,
but see section 16.2.4) are recorded; each calling anuran species is given a score,
known as an amphibian calling index (ACI), ranging from 1 to 3, where 1 means
distinct calls of individuals that can be counted and have no overlapping calls,
2 means calls of individuals that can be distinguished but have some overlap-
ping calls, and 3 means a full chorus, with calls of individuals indistinguishable.
Generally, MCS should not be conducted during heavy rain, high wind, or other
inclement weather that could affect the detection of calling anurans. At each site,
air temperature, relative humidity, barometric pressure, wind speed, and other
variables are often measured immediately after recording call data to be used as
sample covariates in occupancy analysis. These covariates can be used to help
explain variation in site occupancy or detection probabilites.
In regions where anurans vocalize in a somewhat predictable manner, many
MCS, such as NAAMP, recommend sampling each stop along a roadside route
during a specified sampling period, such as during early spring, late spring, and
summer. The NAAMP protocol recommends a single survey, based on conveni-
ence to volunteer observers, despite studies that have shown that significant vari-
ation in calling behavior does occur within the specified sampling periods (e.g.
Todd et al. 2003; Gooch et al. 2006; Kirlin et al. 2006). For example, Gooch
et al. (2006) conducted three MCS during the NAAMP-specified “summer”
sampling period in the western Piedmont region of North Carolina, USA, and
found that detection probabilities increased for some species (e.g. Acris crepitans,
Rana catesbeiana) and decreased for other species (e.g. Rana clamitans) from
survey 1 to survey 3 (Figure 16.1). Variation in calling behavior with prescribed
sampling periods may cause the observer to not detect a species if single sampling
occasions are employed. At least two surveys are required to calculate detection
probabilities (preferably a minimum of three; J. Nichols, personal communi-
cation) during a sampling period (see MacKenzie et al. 2002 and Table 6.1 of
MacKenzie et al. 2006 for more details).
1.0 Survey 1
0.9 Survey 2
Survey 3
0.8
Detection probability (p)
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
Acris crepitans Hyla chrysoscelis Rana catesbeiana Rana clamitans
in species-specific calling behaviors and abiotic and biotic conditions can lead to
the inference of absence despite the presence of a species. Aspects of survey proto-
col and observer bias can also influence detection probabilities. Fortunately,
recent advances in statistical techniques have allowed for calculation of species-
specific detection probabilities (e.g. program PRESENCE; MacKenzie et al.
2002, 2006; Chapter 24), which can greatly aid in inferring population status
and potentially long-term population trends.
Inter- and intraspecific variation in anuran calling behavior may affect detec-
tion probability and should be considered when conducting a MCS or evaluat-
ing MCS data. Anurans exhibit a vast array of acoustic properties (Duellman
and Trueb 1986) which influence probability of detection by observers. Some
species have calls that can carry long distances (e.g. 1 km), whereas calls of
other species cannot be detected until the observer is 100 m or less from the
breeding site. Many species, such as R. catesbeiana, may call sporadically every
few minutes. Other species (e.g. Pseudacris crucifer) call more continuously. In
species-rich communities, louder, higher-pitch calls of one species may interfere
with the detection of other, quieter species (Droege and Eagle 2005) or inhibit
calling in another sympatric species (Littlejohn and Martin 1969). In general,
MCS are best suited for regions where all species vocalize during a somewhat
286 | Amphibian ecology and conservation
studies have been conducted on the effects of noise on anuran calling behavior.
Weir et al. (2005) found that only three of 10 species decreased calling behavior
due to increased traffic. Regardless, anthropogenic noise can affect the obser-
ver’s ability to detect frogs and should be recorded during MCS. Additionally,
anthropogenic noise can also reduce a species proclivity to call, thus lowering its
detectability (Sun and Narins 2005; C. Steelman, personal communication)
Several studies have investigated the effects of survey length on anuran detec-
tion. The majority of MCS range from 3 to 10 min per stop. Pierce and Gutzwiller
(2004) found that 15 min was required to detect 90% of all species known to be
present, whereas Shirose et al. (1997) found that 3 min surveys were adequate to
detect most species. Gooch et al. (2006) found that 94% of summer-breeding
anurans in the North Carolina Piedmont region were detected within the first
5 min of the MCS and detection probabilities were slightly higher as observers
spent longer listening (3 min compared to 10 min). However, these and other
investigations highlight that some species may go undetected even if surveys are
extended up to an hour. The length of survey should be determined based on
the specific objectives of the MCS; however, for detecting long-term population
trends for most species, 3–5 min appears to be adequate.
Variation among observers can also substantially influence the quality of
MCS data. Some large-scale MCS (e.g. NAAMP) rely heavily on volunteers,
who may vary in experience and/or hearing ability. As a result, observers may
not detect all species present, include species not present, or incorrectly identify
vocalizations, resulting in flawed assessments of anuran populations (Lotz and
Allen 2007). Weir et al. (2005) found that volunteer experience may influence
species detection. However, studies by Genet and Sargent (2003), Shirose et al.
(1997), and Lotz and Allen (2007) suggest that even relatively inexperienced
volunteers were reliable in their abilities to determine species, yet estimating
abundance categories often differed among observers. Pierce and Gutzwiller
(2007) showed 79% agreement among nine observers conducting MCS in cen-
tral Texas, USA, and stressed the importance of accounting for interobserver
variation in data analysis. NAAMP currently requires all volunteers to pass an
online anuran call test (available at www.pwrc.usgs.gov/frogquiz/) prior to con-
ducting MCS. In general, training of volunteers will likely increase the prob-
ability of ensuring quality data.
the ACI (section 16.2.3). These types of abundance data are of limited utility in
assessing population densities. However, Nelson and Graves (2004) compared
population estimates (via mark–recapture) of male R. clamitans with ACI col-
lected at the same sites and found abundance to be correlated positively with
ACI. In contrast, Corn et al. (2000) found that another measure of call fre-
quency failed to reflect “a relatively large population” of Bufo woodhousii and
only weakly distinguished among different-sized populations of Pseudacris
maculata. The need to be able to use indices of relative abundance, such as the
calling index, in occupancy models has been conceptualized and is currently an
active area of research (Royle and Nichols 2003; Dorazio 2007).
Because call surveys are based upon the vocalizations of adult male anurans,
they do not provide complete information on population structure; that is,
non-calling females and subadults are not assessed by this method (Stevens and
Paszkowski 2004). Similarly, this method is not useful for other non-calling
amphibians, such as salamanders and caecilians. The males of many species of
anurans will vocalize in contexts not necessarily related to breeding; moreover,
the MCS does not consider the presence of egg masses, tadpoles, or metamorph-
osing juveniles in the population. Thus, the MCS provides no information
about whether there was successful reproduction at a given site. Depending on
the questions of interest in a given study, the MCS is therefore most useful
when used in conjunction with other survey methods, such as visual encoun-
ter surveys and the use of traps or dipnets to assess the larval component of the
population.
16.3 ARS
In addition to MCS, automated systems (ARS) can be used to detect anuran
vocalizations and can be useful in monitoring many anuran populations
(Peterson and Dorcas 1994). Typically, ARS (or so-called frogloggers) are used
to collect data intensively at a single or a few locations, whereas MCS provide
more superficial data but for a larger number of sites. ARS can be used to survey
for anuran species in places difficult to access for MCS and can be left in the field
for extended periods of time, thus increasing the probability of detecting a given
species. ARS may be the only practical way to reliably detect species that have
very short or unpredictable breeding seasons, such as R. capito. ARS minimize
disturbance to calling anurans and provide a permanent sampling record that
can be evaluated by multiple experts if required (Mohr and Dorcas 1999; Todd
et al. 2003). When combined with information on environmental variation,
data from ARS can be incorporated into models that can be used to optimize
16 Auditory monitoring of anurans | 289
monitoring programs based on MCS (Bridges and Dorcas 2000; Oseen and
Wassersug 2002).
been used for this purpose. Timers generally work in one of two ways: controlling
the function of the recorder or interrupting the power supply to the recorder. In
addition to time-based triggers, environmental triggers can be used to activate an
ARS. For example, to detect explosive breeders like Spea or Scaphiopus, a tipping-
bucket rain gauge with a reed valve could be used so a rainfall event would trip
an ARS into service.
The choice of microphones is a crucial decision, as recordings will only be as
good as the microphone, regardless of the recording device. At least nine types
of microphone are available but condenser or dynamic varieties are those most
commonly used. Unlike dynamic microphones, condenser microphones require
a power supply. Condenser microphone power sources are often small (e.g. a
single AA battery) and may not be sufficient if the ARS is deployed for extended
periods of time without maintenance. This issue can be addressed by using the
timer/controller to control the microphone as well as the recorder. Another con-
sideration in selecting a microphone is the directional or acceptance cone. Cones
can range from a 360° circle around the microphone (omnidirectional) to just a
few degrees in front of the microphone (unidirectional or shotgun). To capture
the vocalizations of species that call while partially or completely submerged
(e.g. Rana sevosa or Rana subaquavocalis), a hydrophone would be more appro-
priate to use than a microphone (Platz 1993).
The final considerations of building an ARS are power supply and pro-
tection. Rechargeable batteries generally provide the best option for most
researchers. Cost, power, size, and weight should be considered when selecting
batteries. In environments with sufficient sunlight, rechargeable batteries can
be supplemented by a solar panel and can generally operate without interrup-
tion. With the exception of the microphone, all components of an ARS should
be firmly mounted inside a protective case. The case should provide adequate
environmental protection and be large enough to hold all the components,
yet small enough to be reasonably portable for ease of field deployment. For
most field deployments, waterproof boxes (e.g. Otter or Pelican brands) work
well, but less costly options exist. These include surplus ammunition cans,
polyethylene coolers, plastic tool boxes, and plastic watertight marine boxes.
Consideration should be given to the environment in which the ARS will be
deployed. In areas frequented by people, the ARS can be locked, hidden, or
even buried. Animals, such as raccoons and bears, may damage equipment
and thus more rugged cases might be needed in some situations (Corn et al.
2000). Microphones should also be protected by a windscreen to reduce wind
noise, along with a cover to shield the windscreen and microphone from the
environment.
16 Auditory monitoring of anurans | 291
3.0 3.0
Hyla cinerea Rana clamitans
Mean calling intensity
2.5 2.5
2.0 2.0
1.5 1.5
1.0 1.0
0.5 0.5
0.0 0.0
0 2 4 6 8 10 12 14 16 18 20 22 0 2 4 6 8 10 12 14 16 18 20 22
3.0 3.0
Hyla gratiosa Rana sphenocephala
Mean calling intensity
2.5 2.5
2.0 2.0
1.5 1.5
1.0 1.0
0.5 0.5
0.0 0.0
0 2 4 6 8 10 12 14 16 18 20 22 0 2 4 6 8 10 12 14 16 18 20 22
Hour of day Hour of day
Fig. 16.2 Daily calling pattern of Hyla cinerea, Hyla gratiosa, Rana clamitans, and
Rana sphenocephala recorded using an ARS at Carolina Bay near Aiken, SC, USA.
Mean calling activity was calculated by averaging the recorded calling activity levels
for each 30 min time recording period over all days of the study (from 16 June to
12 July 1997). Error bars denote 1 standard deviation. Note that calling in both
species of treefrog peaked during the time period when manual calling surveys are
recommended (dusk to midnight), but peak calling in Rana peaked well after midnight
and R. sphenocephala called almost exclusively after midnight, and would thus likely be
missed on most calling surveys. Adapted from Bridges and Dorcas (2000).
292 | Amphibian ecology and conservation
one must be careful not to assume that anurans call primarily during times trad-
itionally recognized as peak calling periods (Bridges and Dorcas 2000).
In some cases, researchers use ARS to attempt to detect a species that is
either rare or has very unpredictable calling patterns. Anurans such as Spea and
Scaphiopus that generally only call under certain conditions (i.e. during or after
heavy rains) may be particularly hard to detect using MCS and ARS may offer
the best opportunity to detect their populations. For some extremely rare species
(e.g. R. sevosa) detecting every known population is important for proper man-
agement and thus, increasing detectability of populations using ARS can play a
vital role in conservation efforts.
Because ARS can collect data at regular and precise intervals, mathematical
models can be developed that allow prediction of when and under what condi-
tions species are likely to call, thus providing data that can be used to optimize
MCS. Typically, data collected simultaneously with anuran calling data, such as
temperature, precipitation, wind speed, and time of day, are used as independ-
ent variables in a logistic regression to predict the optimal conditions to conduct
MCS (i.e. the times when detectability is maximized; Oseen and Wassersug
2002). Anurans may respond differently to environmental variables across their
ranges, and thus development of models should be done for particular regions
as needed. Such models were developed for three species of winter-breeding
anurans (Pseudacris crucifer, Pseudacris feriarum, and Rana sphenocephala) in the
western Piedmont region of North Carolina (Steelman and Dorcas, in press).
Models showed that for P. crucifer day of year, time, precipitation, and water
temperature positively influenced calling and air temperature negatively influ-
enced calling; for P. feriarum time, precipitation, air temperature, and water
temperature positively influenced calling, and day of year negatively influenced
calling; for R. sphenocephala day of year, time, precipitation, and air temperature
positively influenced calling, and higher water temperature negatively influ-
enced calling. The models described for the winter-breeding species above were
tested using previously collected data from MCS along with spot measurements
of environmental data (Kirlin et al. 2006). Models accurately predicted whether
a species was calling approximately 70% of the time.
In addition to using models to predict the best times and conditions to manu-
ally sample anurans, ARS can be used to interpret data collected previously,
assuming sufficient environmental data are collected at the time of the surveys.
Fortunately, nearly all existing MCS programs require collection of at least some
environmental data. To do this, an investigator would insert the spot measure-
ments of environmental variables collected by volunteers at the time the surveys
were conducted into the model equation for each species of interest to generate a
294 | Amphibian ecology and conservation
likelihood of calling (e.g. ranging from 0 to 1). Calling likelihoods can then be
used to interpret previously collected calling survey results. If a species was not
heard at a particular location but the model indicates a high likelihood of calling
if it was present (e.g. 0.9), then the investigator can have a higher confidence in
concluding that the species was not present, rather than it simply not vocalizing
and being undetectable.
16.4 Conclusions
Auditory monitoring of anuran populations, either by MCS or by ARS, is a use-
ful tool for assessing the status and trends of populations, as well as the responses
of populations and communities to environmental and anthropogenic change.
The use of auditory monitoring has increased tremendously since the realiza-
tion that amphibian populations are undergoing global population declines,
and has become a standard approach to monitoring in many state, national,
and international programs. Auditory monitoring has limitations in that (1) its
utility in estimating abundance is restricted, (2) this method provides little
information about the complete structure of a population (i.e. the status of
non-calling adult females and subadults) or about whether a given popula-
tion has experienced successful reproduction, and (3) this method cannot be
used on non-calling amphibians, such as salamanders and caecilians. The two
means by which auditory monitoring may be conducted (manual or automated)
differ in their relative costs and benefits, and the appropriate approach for a
16 Auditory monitoring of anurans | 295
given study ultimately depends on the study objectives and resources available.
Nevertheless, MCS can provide vital information on changes in occupancy
states of various sites, which can be overlaid with environmental data to assess
potential causes of change in occupancy for a given species or community. Data
from ARS can be used to optimize the effectiveness of MCS or interpret data
based on MCS. As such, auditory monitoring has great utility in assessing the
extent of declines of anuran amphibians, a topic of heightened concern in ecol-
ogy and conservation biology.
16.5 Acknowledgments
We thank P. Stephen Corn, Erin Muths, and Charlotte K. Steelman for com-
ments that improved the manuscript. Manuscript preparation was partially sup-
ported by the Department of Biology at Davidson College, Duke Power, and
National Science Foundation grant (DEB-0347326) to M.E.D and the U.S.
Geological Survey’s Amphibian Research and Monitoring Initiative to SCW
and WJB. The use of trade or product names does not imply endorsement by
US government agencies.
16.6 References
Acevedo, M. A. and Villanueva-Rivera, L. J. (2006). Using automated digital recording
systems as effective tools for the monitoring of birds and amphibians. Wildlife Society
Bulletin, 34, 211–14.
Agranat, I. D. (2007). Automatic Detection of Cerulean Warblers Using Autonomous
Recording Units and Song Scope Bioacoustics Software. www.wildlifeacoustics.com/
Anthony, B. P. (2002). Results of the first batrachian survey in Europe using road call
counts. Alytes, 20, 55–66.
Barichivich, W. J. (2003). Appendix IV: Guidelines for building and operating remote
field recorders. In C. K. Dodd, Jr, Monitoring Amphibians in Great Smoky Mountains
National Park, pp. 87–96. U.S. Geological Survey Circular no. 1258. U.S. Geological
Survey, Tallahassee, FL.
Blair, W. F. (1961). Calling and spawning seasons in a mixed population of anurans.
Ecology, 42, 99–110.
Bowers, D. G., Anderson, D. E., and Euliss, Jr, N. H. (1998). Anurans as indicator of
wetland condition in the Prairie Pothole Region of North Dakota: an environmen-
tal monitoring and assessment program pilot project. In M. Lannoo (ed.), Status and
Conservation of Midwestern Amphibians, pp. 369–78. University of Iowa Press, Iowa
City, IO.
Bridges, A. S. and Dorcas, M. E. (2000). Temporal variation in anuran calling behavior:
implications for surveys and monitoring programs. Copeia, 2000, 587–92.
Corn, P. S. and Muths, E. (2002). Variable breeding phenology affects the exposure of
amphibian embryos to ultraviolet radiation. Ecology, 83, 2958–63.
296 | Amphibian ecology and conservation
Corn, P. S., Muths, E., and Iko, W. M. (2000). A comparison in Colorado of three methods
to monitor breeding amphibians. Northwestern Naturalist, 81, 22–30.
de Solla, S. R., Shirose, L. J., Fernie, K. J., Barrett, G. C., Brousseau, C. S., and Bishop, C. A.
(2005). Effect of sampling effort and species detectability on volunteer based anuran moni-
toring programs. Biological Conservation, 121, 585–94.
Dorazio, R. M. (2007). On the choice of statistical models for estimating occurrence and
extinction from animal surveys. Ecology, 88, 2773–82.
Dorcas, M. E. and Foltz, K. D. (1991). Environmental effects on anuran advertisement
calling. American Zoologist, 31, 3111A.
Droege, S. and Eagle, P. (2005). Evaluating calling surveys. In M. Lannoo (ed.), Amphibian
Declines: The Conservation Status of United States Species, pp. 314–19. University of
California Press, Berkeley, CA.
Duellman, W. E. and Trueb, L. (1986). Biology of Amphibians. McGraw-Hill Publishing
Company, New York.
Genet, K. S. and Sargent, L. G. ( 2003). Evaluation of methods and data quality from a
volunteer-based amphibian call survey. Wildlife Society Bulletin, 31, 703–14.
Gooch, M. M., Heupel, A. H., Price, S. J., and Dorcas, M. E. (2006). The effects of survey
protocol on detection probabilities and site occupancy estimates of summer breeding
anurans. Applied Herpetology, 3, 129–42.
Inkley, D. B. (2006). Final report assessment of utility of Frogwatch USA data
1998–2005. www.nwf.org/frogwatchUSA/FinalRptFW-ScienceAssess.pdf.
National Wildlife Federation for U.S. Geological Survey, Washington DC.
Johnson, D. H. and Batie, R. D. (2001). Surveys of calling amphibians in North Dakota.
The Prairie Naturalist, 33, 227–47.
Kaiser, K. (2008). Evaluation of a long-term amphibian monitoring protocol in Central
America. Journal of Herpetology, 42, 104–10.
Kirlin, M., Gooch, M. M., Price, S. J., and Dorcas, M. E. (2006). Predictors of winter
anuran calling activity in the North Carolina Piedmont. Journal of the North Carolina
Academy of Science, 122, 10–18.
Knutson, M. G., Sauer, J. R., Olsen, D. A., Mossman, M. J., Hemesath, L. M., and
Lannoo, M. J. (1999). Effects of landscape composition and wetland fragmentation
on frog and toad abundance and species richness in Iowa and Wisconsin, U.S.A.
Conservation Biology, 13, 1437–46.
Littlejohn, M. J. and Martin, M. M. (1969). Acoustic interaction between two species of
leptodactylid frog. Animal Behaviour, 17, 785–91.
Lotz, A. and Allen, C. R. (2007). Observer bias in anuran call surveys. Journal of Wildlife
Management, 71, 675–9.
MacKenzie, D. L., Nichols, J. D., Lachman, G. B., Droege, S., Royle, J. A., and Langtimm,
C. A. (2002). Estimating site occupancy rates when detection probabilities are less than
one. Ecology, 83, 2248–55.
MacKenzie, D. I., Nichols, J. D., Royle, J. A., Pollock, K. H., Bailey, L. L., and Hines, J. E.
(2006). Occupancy Estimation and Modeling: Inferring Patterns and Dynamics of Species
Occurrence. Academic Press, San Diego, CA.
Marsh, D. M. and Trenham, P. C. (2008). Current trends in plant and animal population
monitoring. Conservation Biology, 22, 647–55.
16 Auditory monitoring of anurans | 297
Martof, B. S. (1953). Territoriality in the green frog, Rana clamitans. Ecology, 34, 165–74.
Mohr, J. R. and Dorcas, M. E. (1999). A comparison of anuran calling patterns at two
Carolina bays in South Carolina. Journal of the Elisha Mitchell Scientific Society, 115,
63–70.
Nelson, G. L. and Graves, B. M. (2004). Anuran population monitoring: comparison
of the North American Amphibian Monitoring Program’s calling index with Mark-
Recapture estimates for Rana clamitans. Journal of Herpetology, 38, 355–9.
Oseen, K. L. and Wassersug, R. J. (2002). Environmental factors influencing calling in
sympatric anurans. Oecologia, 133, 616–25.
Pellet, J. and Schmidt, B. R. (2005). Monitoring distributions using call surveys: esti-
mating site occupancy, detection probabilities and inferring absence. Biological
Conservation, 123, 27–35.
Penman, T. D., Lemckert, F. L., and Mahony, M. J. (2005). A cost-benefit analysis of
automated call recorders. Applied Herpetology, 2, 389–400.
Peterson, C. R. and Dorcas, M. E. (1992). The use of automated data acquisition techniques
in monitoring amphibian and reptile populations. In D. McCullough and R. Barrett
(eds), Wildlife 2001: Populations, pp. 369–78. Elsevier Applied Science, London.
Peterson, C. R. and Dorcas, M. E. (1994). Automated data acquisition. In W. Heyer,
R. McDairmid, M. Donnelly, and L. Hayek (eds), Measuring and Monitoring Biological
Diversity. Standard Methods for Amphibians, pp. 47–57. Smithsonian Institution Press,
Washington DC.
Pierce, B. A. and Gutzwiller, K. J. (2004). Auditory sampling of frogs: detection effi-
ciency in relation to survey duration. Journal of Herpetology, 38, 495–500.
Pierce, B. A. and Gutzwiller, K. J. (2007). Interobserver variation in frog call surveys.
Journal of Herpetology, 41, 424–9.
Platz, J. E. (1993). Rana subaquavocalis, a remarkable new species of Leopard Frog
(Rana pipiens complex) from southeastern Arizona that calls under water. Journal of
Herpetology, 27, 154–62.
Royle, J. A. and Nichols, J. D. (2003). Estimating abundance from repeated presence-
absence data or point counts. Ecology, 84, 777–90.
Saenz, D., Fitzgerald, L. A., Baum, K. A., and Conner, R. N. (2006). Abiotic correl-
ates of anuran calling phenology: the importance of rain, temperature and season.
Herpetological Monographs, 20, 64–82.
Schmidt, B. R. (2005). Monitoring the distribution of pond-breeding amphibians
when species are detected imperfectly. Aquatic Conservation: Marine and Freshwater
Ecosystems, 15, 681–92.
Scott, W. A., Pithart, D., and Adamson, J. K. (2008). Long-term United Kingdom
trends in the breeding phenology of the common frog, Rana temporaria. Journal of
Herpetology, 42, 89–96.
Shirose, L. J., Bishop, C. A., Green, D. M., MacDonald, C. J., Brooks, R. J., and Helferty,
N. J. (1997). Validation tests of an amphibian call count survey technique in Ontario,
Canada. Herpetologica, 53, 312–20.
Steelman, C. K. and Dorcas, M. E. (in press). Anuran calling survey optimization:
developing and testing predictive models of anuran calling activity. Journal of
Herpetology.
298 | Amphibian ecology and conservation
Stevens, C. E. and Paszakowski, C. A. (2004). Using chorus-size ranks from call sur-
veys to estimate reproductive activity of the wood frog (Rana sylvatica). Journal of
Herpetology, 38, 404–10.
Sun, J. W. C. and Narins, P. M. (2005). Anthropogenic sounds differentially affect
amphibian call rate. Biological Conservation, 121, 419–27.
Todd, M. J., Cocklin, R. R., and Dorcas, M. E. (2003). Temporal and spatial variation in
anuran calling activity in the western Piedmont of North Carolina. Journal of the North
Carolina Academy of Science, 119, 103–10.
Villanueva-Rivera, L. J. (2007). Digital recorders increase detection of Eleutherodactylus
frogs. Herpetological Review, 38, 59–63.
Walker, S. J. (2002). Frog Census 2001: Community monitoring of water quality and habitat
condition in South Australia using frogs as indicators. Environment Protection Authority,
Adelaide.
Weeber, R. C. and Vallitanios, M. (2000). The Marsh Monitoring Program 1995–1999:
Monitoring Great Lakes wetlands and their amphibian and bird inhabitants. www.
bsc-eoc.org/mmpreport.html. Bird Studies Canada with the U.S. Environmental
Protection Agency, Ontario.
Weir, L. A. and Mossman, M. J. (2005). North American amphibian monitoring pro-
gram (NAAMP). In M. Lannoo (ed.), Amphibian Declines: The Conservation Status of
United States Species, pp. 307–13. University of California Press, Berkeley, CA.
Weir, L. A., Royle, J. A., Nanjappa, P., and Jung, R. E. (2005). Modeling anuran detection
and site occupancy on North American Amphibian Monitoring Program (NAAMP)
routes in Maryland. Journal of Herpetology, 39, 627–39.
Woolbright, L. L. (1985). Patterns of nocturnal movement and calling by the tropical
frog, Eleutherodactylus coqui. Herpetologica, 41, 1–9.
Wright, A. H. and Wright, A. A. (1949). Handbook of Frogs and Toads of the United States
and Canada. 3rd edn. Comstock Publishing Associates, Ithaca, NY.
17
Measuring habitat
Kimberly J. Babbitt, Jessica S. Veysey, and George W. Tanner
17.1 Introduction
Understanding wildlife–habitat relations is fundamental to sound management
and conservation and provides important basic information on species’ ecology.
Obtaining accurate information on amphibian habitat associations can be chal-
lenging as many species have complex life cycles requiring very different habitat
types during different life-history stages, and many species are highly cryptic
during much of the year. However, numerous amphibian species are experien-
cing significant threats from habitat loss and degradation. Better information
on habitat requirements and habitat use is necessary to understand how these
changes affect habitat suitability for different species and to improve manage-
ment targeted at enhancing amphibian habitat.
Use-availability designs are the most common approaches for examining habi-
tat selection (Thomas and Taylor 1990, 2006; Garshelis 2000). These approaches
compare the proportional use of each habitat type by an organism to the relative
area of each habitat. Use-availability studies can be classified into four general
designs (Thomas and Taylor 1990, 2006). In design 1, use is determined for all
known individuals in a population but the individuals are not identified. Habitat
is considered equally available to all individuals. This design is an assessment of
habitat use at the population level and does not allow examination of variation
in habitat selection among individuals. This design would be appropriate for
studies where individuals are documented using visual encounter surveys or area-
constrained searches but where no individuals are marked.
Design 2 is used when individuals are uniquely marked and, as in design 1,
habitat availability is considered equivalent for all individuals. In design 3, use
is determined for individually marked organisms but, unlike designs 1 and 2,
available habitat is estimated for each individual animal. This design is appro-
priate when, for example, individual home ranges have been estimated and only
areas within each home range are deemed available. Design 4 examines use of
habitat by uniquely marked individuals at multiple time periods (e.g. each time
an animal is located) and pairs measurements of used habitat with measure-
ments of available habitat taken at each interval. This design is particularly use-
ful for examination of small-scale habitat (i.e. microhabitat).
Use-availability studies generally focus on habitat selection based on broad
habitat types, or multiple habitat parameters that are analyzed individually
(Garshelis 2000). An alternative approach to use-availability studies is the site-
attribute approach. This approach differs from use-availability studies in that
the focus is on measuring habitat features that potentially influence use in areas
where an individual has been documented and comparing those values to meas-
urements made on the same features in areas that are not used. Differences
between used and non-used areas are then assumed to reveal which features are
responsible for selection for (or against) habitat use. Thus, the focus is not on the
amount of use in one habitat compared to another, but rather whether the site
was used or not and what specific features appear to determine use. Often, site-
attribute studies are used to examine habitat variables at biologically important
sites (e.g. breeding sites).
we study, it is certain that some working assumptions will go into any determin-
ation of availability. At the broadest level, areas beyond the study area are not
examined, and so by default are unavailable. If we delineate the home range of
an organism we can define unavailable habitat as that outside the home-range
boundary. However, within a home range, intraspecific interactions may mean
that a portion of the home range is not just unused but unavailable. In contrast,
some habitat may go unused during much of the year but yet be both available
and critical (e.g. over-wintering sites, breeding sites). Knowledge of the organism’s
natural history can help to address some questions about what is unavailable and
what is unused, but cautious interpretation is always warranted because improper
designation of habitat availability can bias results.
wind speed, and handheld versions are very inexpensive. Wind speed can also be
estimated using a Beaufort scale of winds, which provides descriptive measures
of wind effects that are given a number from 1 to 12. Wind direction can be
determined at the same time as wind velocity by taking a compass reading in the
direction of wind flow.
All the weather parameters mentioned can be measured using relatively inex-
pensive equipment or electronic data devices that require more money but less
personnel time. Further, meteorological stations are maintained throughout
many areas of the world and may provide free access to data. Availability may
differ with geographic location, but researchers should check into availability
before purchasing expensive equipment. These stations are an excellent source
of climate information, but the utility of station data depends on the scale of
the research project. If the purpose of an investigation is to understand the role
of local habitat relations then weather-station data will likely be insufficient for
examining among-site differences.
Morphometry, the habitat’s basic shape and structure, limits the community
types possible in a habitat, through its influence on water temperature, current
strength, sediment and vegetation patterns, water-holding capacity, substrate,
and weathering rates. Morphometric variables include contours, perimeter,
length, width, area, depth, slope, and surficial geology. For streams, morph-
ometry can be assessed for the wetted or dry channel, or for individual habi-
tat reaches. Channel slope is particularly important for streams because slope
affects dispersal potential and flow velocity.
Many morphometric variables can be mapped using global positioning sys-
tem (GPS) or professional survey equipment. Perimeter, length, and width can
also be assessed with pacing or a measuring tape. Wet contours and water depth
can be charted with an echo sounder, weighted line, or meter stick. Area can be
calculated from length and width; or from maps, aerial photographs, or GIS lay-
ers. The reliability of map-derived area measurements increases with basin size
(Skidds et al. 2007). Slope can be measured with a clinometer. Surficial geology
can be determined from geologic maps or field-verified from test or gravel pits.
Hydroperiod, or the timing and duration that a habitat holds water during a
given year, strongly influences community composition (Wellborn et al. 1996;
Babbitt et al. 2003). For example, in ephemeral waterbodies, invertebrates are
often the dominant predators and amphibians must have a short larval stage
to complete metamorphosis before the habitat dries. Conversely, in perman-
ent waterbodies, amphibians can over-winter as larvae and fish are typically
major predators. Hydroperiod can be categorized for lentic habitats as tempor-
ary, semi-permanent, or permanent; and for lotic habitats as intermittent or
perennial. Such categorizations can be made by examining vegetation, or water
level and substrate, towards the end of the dry season. Detailed data can be
obtained by recording changes in the extent of inundation (Skidds et al. 2007),
or by tracking water depth and noting the drying date. Hydroperiod can vary
widely from year to year in seasonal waterbodies, so hydroperiod data should be
collected over several years.
Hydrologic connectivity (i.e. the extent to which aquatic environments are
linked) influences both hydroperiod and the inputs to, and outputs from, a
habitat. Connectivity may be observed by inspection for inflows and outflows
during the rainy season, or detected from contour maps or historic aerial photo-
graphs. Monitoring wells or a hydrologic analysis may be necessary to establish
groundwater connectivity.
For streams, flow intensity (e.g. velocity, variability) also strongly influences
habitat suitability. Flow affects dispersal, predator distribution, scouring, and
sedimentation, and the ability of amphibians, and their sperm (i.e. for external
17 Measuring habitat | 307
fertilization) and food (i.e. invertebrates and algae), to maintain channel pos-
ition. Velocity is measured with a flow meter, a pitot tube, or by calculating the
time required for a buoyant object or dye to travel a specified distance. Discharge,
or flow volume per unit time, is equal to the stream’s cross-sectional area times
the velocity. Variability can be described by flow variance per unit time, or by the
return interval of specific events (e.g. 100-year floods). Flow timing determines
desiccation and scouring risks, and can be understood by examining velocity and
discharge records.
Water temperature impacts amphibians at multiple ecological scales. For
instance, temperature affects decomposition and photosynthesis, which influ-
ence food availability and thus amphibian abundance. Amphibian development
rates also vary with water temperature. Temperature measurement is discussed
in section 17.7. Drainage area, elevation and aspect, and canopy closure all affect
water temperature. Drainage area and elevation can be determined from topo-
graphic maps. Elevation can also be measured with an altimeter. Canopy clos-
ure is discussed in section 17.11.
Aquatic substrates can be organic or inorganic. Organic substrates provide
over-wintering habitat and are integral to aquatic food webs. Inorganic sub-
strates (e.g. cobbles) provide oviposition sites and refuges for amphibians and
their invertebrate and algal food. Substrate particle size is a critical stream-
habitat variable, as particle size affects size of interstitial hiding places which
affects susceptibility to predation (Barr and Babbitt 2002). Small particles (i.e.
density of randomly dispersed trees. Variants of this method, either the angle-
order (Morista 1957) or the corrected-point-distance (Batcheler 1973), are rec-
ommended for use when non-random distributions of individuals occur.
Tree height, up to 10 m, can be measured using a telescoping stage stadia rod
elevated from the base of the tree to the highest point of the canopy crown. A
manual or automated clinometer can be used to measure the height of any tree
for which the base of the stem and top of the canopy (apex) can be seen. Stadia
rods and clinometers also can be used to measure height to base of canopy and
thus compute canopy thickness.
Tree canopy cover is defined as the vertical projection of the total areal extent
of the canopy upon the ground and is typically expressed as m2/ha or more com-
monly as a percentage. Changes in canopy cover can alter suitability of aquatic
sites (e.g. via effects on temperature and hydroperiod), resulting in changes in
amphibian community composition (Skelly et al. 1999). Light interception and
irradiance at a point on the ground, however, are influenced by both the vertical
and angular aspects of the canopy (Nuttle 1987). Several ground-based methods
have been developed to measure tree canopy cover including, but not limited to,
line-interception, moosehorn, spherical densitometer, and hemispherical pho-
tography (Fiala et al. 2006). The line-intercept method employs multiple line
transects (each 15–20 m long) distributed within a stand. Tree canopy cover is
estimated as the percentage of each line lying beneath the vertical projection of
the canopy disregarding any breaks within the outline of the canopy.
The moosehorn and the spherical densitometer are both sighting devices that
employ mirrors with etched markings. The moosehorn has a flat mirror to reflect
the canopy directly above the sampling location onto the reflective grid. Percent
canopy cover is determined by counting the number of cross-hair intersections
or points intercepting the reflected canopy divided by the total number of points
on the grid. The spherical densitometer uses a convex or concave spherical mir-
ror. Canopy measurements are taken while facing in all four cardinal directions
from a single point. Canopy cover is estimated by averaging the proportion of
the 24 squares etched on the reflective spherical surface that is contacted by can-
opy. A simpler approach to measuring canopy cover is to use a tube and ocularly
estimate percent canopy in broader categories (e.g. 0, 5, 25, 50, 75, 100%).
Horizontal photographs, when taken at a designated height with a hemi-
spherical wide-angle lens, record the extent of the overhead canopy dependent
upon the angle of view of the lens. Photographs taken at bright midday condi-
tions should be avoided (Fiala et al. 2006). Photogrammetric data obtained
using hemispherical lenses are analyzed following Rich (1989). These data are
312 | Amphibian ecology and conservation
percent cover or translated to m2/ha (Baillie et al. 1999). Using quadrats, CWD
volume (m3/ha) can be estimated from the density of downed logs and diameter
measurements at the large and small ends of each log. CWD decomposition stage
may be of interest and be classified according to Maser et al. (1979).
Litter depth and moisture content can greatly influence microhabitat suit-
ability. Litter depth to mineral soil is easily measured using a ruler. To estimate
biomass (g of dry weight/m2) and moisture content (%), litter can be collected
in the field, weighed, and oven dried to a constant weight (Bonham 1989). Dry
weight and moisture content of understory plants can be determined by clip-
ping the vegetation at ground level within quadrats and weighing, drying, and
re-weighing the vegetative samples.
17.13 Conclusion
Characterization of amphibian habitat has lagged behind many other taxa.
Increased research on habitat suitability will provide valuable data that will
assist in management and conservation of amphibian species. Habitat suitabil-
ity is influenced by both the features of the environment outlined in this chapter
17 Measuring habitat | 315
as well as changes that occur at the landscape level. Researchers are encouraged
to design studies that include approaches outlined in this chapter as well as those
detailed in Chapter 20.
17.14 References
Babbitt, K. J., Baber, M. J. and Tarr, T. L. (2003). Patterns of larval amphibian distribution
along a wetlands hydroperiod gradient. Canadian Journal of Zoology, 81, 1539–52.
Baillie, B. R., Cummins, T. L., and Kimberley, M. O. (1999). Measuring woody debris in
the small streams of New Zealand’s pine plantations. New Zealand Journal of Marine
and Freshwater Research, 33, 87–97.
Bandoni de Oliveira, F. and Navas, C. A. (2004). Plant selection and seasonal patterns of
vocal activity in two populations of the bromeligen treefrog Scinax perpusillus (Anura,
Hylidae). Journal of Herpetology, 38, 331–9.
Barr, G. E. and Babbitt, K. J. (2002). Effects of biotic and abiotic factors on distribution
and abundance of larval two-lined salamanders (Eurycea bislineata): changes across
spatial scales. Oecologia, 133, 176–85.
Batcheler, C. L. (1973). Estimating density and dispersion from truncated or unre-
stricted joint point-distance nearest-neighbor distances. Proceeding of the New Zealand
Ecological Society, 20, 131–47.
Bonham, C. D. (1989). Measurements for Terrestrial Vegetation. John Wiley and Sons,
New York.
Brooks, P. D., O’Reilly, C. M., Diamond, S. A., Campbell, D. H., Knapp, R., Bradford, D.,
Corn, P. S., Hossack, B., and Tonnessen, K. (2005). Spatial and temporal variability in
the amount and source of dissolved organic carbon: implications for ultraviolet expos-
ure in amphibian habitats. Ecosystems, 8, 478–87.
Cottam, G. and Curtis, J. T. (1956). The use of distance measures in phytosociological
sampling. Ecology, 37, 451–60.
Davey, B. G. and Conyers, M. K. (1988). Determining the pH of acid soils. Soil Science,
146, 141–50.
Dayton, G. H. and Fitzgerald, L. E. (2006) Habitat suitability models for desert amphib-
ians. Biological Conservation, 132, 40–9.
Fiala, A. C. S., Garman, S. L., and Gray, A. N. (2006). Comparison of five cover estima-
tion techniques in the western Oregon Cascades. Forest Ecology and Management, 232,
188–97.
Garshelis, D. L. (2000). Delusions in habitat evaluation: measuring use, selection,
and importance. In L. Boitani and T. K. Fuller (eds), Research Techniques in Animal
Ecology: Controversies and Consequences, pp. 111–64. Columbia University Press,
New York.
Goslee, S. C. (2006). Behavior of vegetation sampling methods in the presence of spatial
autocorrelation. Plant Ecology, 187, 203–12.
Griffis-Kyle, K. L. (2005). Ontogenic delays in effects of nitrite exposure on tiger salaman-
ders (Ambystoma tigrinum tigrinum) and wood frogs (Rana sylvatica). Environmental
Toxicology and Chemistry, 24, 1523–7.
316 | Amphibian ecology and conservation
Higgins, K. F., Jenkins, K. J., Clambey, G. K., Uresk, D. W., Naugle, D. E., Norland, J. E.,
and Barker, W. T. (2005). Vegetation sampling and measurement. In C. E. Braun
(ed.), Techniques for Wildlife Investigations and Management, 6th edn, pp. 524–54. The
Wildlife Society, Bethesda, MD.
Howe, C. M., Berrill, M., Pauli, B. D., Helbing, C. C., Werry, K., and Veldhoen, N.
(2004). Toxicity of glyphosate-based pesticides to four North American frog species.
Environmental Toxicology and Chemistry, 23, 1928–38.
Irwin, L. L. and Peek, J. M. (1979). Shrub production and biomass trends following five
logging treatments within the cedar-hemlock zone of northern Idaho. Forest Science,
24, 415–26.
Jackson, D. C. (2007). Temperature and hypoxia in ectothermic tetrapods. Journal of
Thermal Biology, 32, 125–33.
Jofre, M. B. and Karasov, W. H. (1999). Direct effect of ammonia on three species of
North American anuran amphibians. Environmental Toxicology and Chemistry, 18,
1806–12.
Johnson, D. H. (1980). Comparison of usage and availability measurements for evaluat-
ing resource preference. Ecology, 61, 65–71.
Levin, S. A. (1992). The problem of pattern and scale in ecology. Ecology, 73, 1943–67.
Luscier, J. D., Thompson, W. L., Wilson, J. M., Gorham, B. E., and Dragut, L. D. (2006).
Using digital photographs and object-based image analysis to estimate percent ground
cover in vegetation plots. Frontiers in Ecology and Environment, 4, 408–13.
Madison, D. M. (1997). The emigration of radio-implanted spotted salamanders,
Ambystoma maculatum. Journal of Herpetology, 31, 542–51.
Manly, B. F. J., McDonald, L. L., Thomas, D. L., McDonald, T. L., and Erickson, W. P.
(2002). Resource Selection by animals: Statistical Design and Analysis for Field Studies,
2nd edn. Kluwer, New York.
Maser, C., Anderson, R. G., Cromack, Jr, K., Williams, J. T., and Martin, R. E. (1979).
Dead and down woody material. In J. W. Thomas (ed.), Wildlife Habitats in Managed
Forests—The Blue Mountains of Oregon and Washington, pp. 78–95. Agricultural
Handbook 553. US Department of Agriculture, Washington DC.
Morista, M. (1957). A new method for the estimation of density by the spacing method
applicable to non-randomly distributed populations. Physiological Ecology, 7, 134–44.
Morrison, M. L., Marcot, B. G., and Mannan, R. W. (2006). Wildlife-habitat Relationships:
Concepts and Applications, 3rd edn. Island Press, Washington DC.
Nuttle, T. (1997). Densiometer bias? Are we measuring the forest or the trees? Wildlife
Society Bulletin, 25, 610–11.
Orton, F., Carr, J. A., and Handy, R. D. (2006). Effects of nitrate and atrazine on lar-
val development and sexual differentiation in the northern leopard frog Rana pipiens.
Environmental Toxicology and Chemistry, 25, 65–71.
Pechmann, J. H. K., Scott, D. E., Semlitsch, R. D., Caldwell, J. P., Vitt, L. T., and
Gibbons, J. W. (1991). Declining amphibian populations: the problem of separating
human impacts from natural fluctuations. Science, 253, 892–5.
Relyea, R. A. and Mills, N. (2001). Predator induced stress makes the pesticide car-
baryl more deadly to gray treefrog tadpoles (Hyla versicolor). Proceedings of the National
Academy of Sciences USA, 98, 2491–6.
17 Measuring habitat | 317
18.1 Introduction
Determining how many species are present in a particular habitat, their abun-
dance, their importance, and how communities are related are critical questions
in ecology and conservation biology. Ecologists use measures of diversity and
similarity to understand community structure and function, particularly in
terms of habitat use, food webs, predator–prey relationships, estimating how
many species can co-exist within a community, energy flow, and nutrient cyc-
ling. Conservation biologists need to estimate diversity to identify areas for pro-
tection and management (Scott et al. 1987; Snodgrass et al. 2000), and to assess
the effects of habitat change through time. Knowledge of these variables also is
very practical for consultants or others doing rapid assessments, such as in the
preparation of environmental impact assessments. Species diversity (richness,
heterogeneity, evenness) is fairly simple to estimate, yet the use of these indices
in amphibian studies is generally less than in many other areas of field research,
and most field amphibian researchers have yet to explore the utility of measures
of similarity.
There are many indices used to express various aspects of diversity and simi-
larity; some of the most common are listed in Table 18.1. Magurran (1988,
2003), Krebs (1999), and Clarke and Warnick (2001) provide good discussions
of some diversity and similarity indices, and these references can serve as a start-
ing point for understanding what the indices do and how they are calculated,
and for where to find more information about them. Whatever index is chosen,
it is important to understand the assumptions surrounding the computation,
the biases of the index, and the way the index should be interpreted. It is very
common for biologists to sacrifice rigor (by violating assumptions) for ease of
computation or precedence of use by other biologists. Some of the easiest indices
to use (for example, Margalef) are also some of the most informative, whereas
322 | Amphibian ecology and conservation
other indices (for example, Shannon) are popular but not very informative by
themselves.
0 0 0 0 0
1 1 1 0.3 1
10 3.3 1.7 1 1
100 10 3.3 2 1
1000 33 5.7 3 1
10 000 100 10 4 1
1
Because the log(0) ∞, 1 is added to the value inasmuch as the log(1y) 0 when y 0.
One way to change weighting would be to transform the data. Table 18.2
shows how data transformation can weight the ‘importance’ of the less abun-
dant or rare species, and thus allow them to have more influence in perceptions
of community patterns. Data transformation also helps in clustering and ordin-
ation methods. Thus, it should be apparent that the objectives of the researcher
become very important in the selection of diversity and similarity indices.
Knowing the limitations and assumptions of the indices helps avoid confusion
and aids in the interpretive process.
A second reason for transforming data is that sampling usually involves large
numbers of zero captures; that is, all species are not caught on each day of sam-
pling. Zeros present complications in data analysis, especially when using para-
metric measures which assume normal distributions. Transformation offers a
means to validate the statistical assumptions of parametric techniques prior to
using them. One must always remember to verify that the data transformation
actually corrects the original issue with the data set. If it does not, transformation
is not warranted. Of course, an alternative approach would be to avoid paramet-
ric statistics, and many of the diversity and similarity indices commonly used
are based on non-parametric techniques.
Species richness does not imply anything about abundance, although the
term frequently is used inaccurately as a synonym of diversity (see below). In
some situations, species richness can be estimated in advance using field guides
or museum specimens, coupled with knowledge of habitat requirements; this
method is termed interpolation. For example, the number of species found
within a forested area of a national park might be predicted from published lit-
erature and museum records. Interpolation can be very inaccurate, depending
on scale. Knowledge of species richness in one watershed might be useful for
predicting species occupancy in an adjacent watershed, but not on a regional
scale. Thus, interpolation does not supplant the necessity of field sampling for
an accurate understanding of richness since myriad factors may influence occu-
pancy at a particular location.
The number of species within an area results from a complex interaction
of resource availability, habitat complexity, biogeography, land-use history,
and phylogenetic history. Habitats with large numbers of amphibian species
include tropical rainforests and areas with diverse topography and, as might be
expected, amphibian species richness is inversely correlated with latitude and,
usually, elevation (Duellman 1999). Even within lowland rainforests, how-
ever, amphibian species richness changes spatially, so sampling considerations
become paramount when estimating how many species are present. An estimate
of species richness is only as accurate as the reliability of the sampling methods
on which the estimate is based.
there are few species with high abundance, accumulation curves have low shoul-
ders and long trajectories to the asymptote. Conversely, areas with large num-
bers of abundant species have steep trajectories and reach asymptotes quickly.
Species diversity is positively correlated with the initial slope of the trajectory of
the accumulation curve.
Both incidence-based (that is, occupancy) and abundance-based (that is,
incorporating diversity) species accumulation curves can be generated. A com-
puter program (such as the Mao Tau of EstimateS, see below) might use 1000
iterations of a data set to predict the expected range of shapes of the accumu-
lation curve. This allows biologists to incorporate confidence intervals (95%)
around the accumulation curve, which provides a better estimate of the actual
numbers of species likely within the area than a simple curve. An example of
species-accumulation curves based on intensive pitfall and other sampling tech-
niques for amphibians is shown in Figure 18.1.
There are few applications of these methods dealing with amphibians. Pineda
and Halffter (2004) used them to verify the completeness of inventories at
both local and regional scales and determined the sampling effort needed for
reaching a plateau. Heyer et al. (1999) tested the utility of museum collections
for conservation decisions, focusing on frogs of the genus Leptodactylus from
Amazonia. The results indicated that the data set was adequate in terms of sam-
pling effort and useful for conservation decisions, at least for amphibians. Dodd
et al. (2007) used species accumulation curves to compare changes in species
presence through time as a consequence of habitat changes. They showed that
long-term changes in habitat management resulted in decreases in species rich-
ness within the amphibian community.
18.3.4 Heterogeneity
The concept of heterogeneity (also sometimes termed ‘species diversity’) pro-
vides a way of expressing both the number of species and a measure of counts or
abundance into a single index. What is more diverse: a community with many
rare species or one with fewer but much more abundant species? What if two
communities have the same number of species, but at different abundances?
In computations, actual (rarely measured) or relative (based on counts) abun-
dances are combined with species richness to measure the heterogeneity of the
community. The result is an index which allows the observer to compare the
heterogeneity of one or more communities (for example, Dodd 1992). An index
by itself tells little; it gains value when it is used comparatively.
Amphibians are sampled as one might for other measures of diversity; that
is, via a variety of techniques that provide capture histories through time.
35
30
Number of amphibian species
25
20
15
10
Historic
Recent
5
0
2 4 6 8 10 12
Number of samples
40
35
Number of reptile species
30
25
20
15
Historic
Recent
10
5
2 4 6 8 10 12
Number of samples
Fig 18.1 Species accumulation curves, with 95% confidence intervals, for amphibians
(top) and reptiles (bottom) sampled at 12 study sites at St. Marks National Wildlife
Refuge, Florida, USA. The historic curve shows relationships in 1977–9, whereas the
recent curve shows the relationships in 2002–5. These species accumulation curves
suggested that fewer species of amphibians were expected in the 2000s than in the
1970s. In the 1970s, the curves predicted that 29 species of amphibians might be
present throughout the sampling sites, but only 19 in the 2000s. The actual values
were 29 (1970s) and 24 (2000s). Reprinted from Dodd et al. (2007).
18 Diversity and similarity | 329
Nˆ C / pˆ
18.4 Similarity
Measures of similarity are used to examine data from a number of sampling
areas in order to compare community similarity, or more correctly, dissimilar-
ity among those sampling areas. The result is a matrix of numbers representing
paired comparisons which can be difficult to interpret without a visual context.
However, these matrix-based comparisons can be fitted into a graphic depiction
of similarity, aligning those communities or sampling areas most similar to one
another into a cluster dendrogram. Cluster dendrograms then can be compared
to see how communities change through time and how variables change from
330 | Amphibian ecology and conservation
18.5 Software
The program EstimateS does most computations required for species accumu-
lation curves and non-parametric analyses of species richness, diversity, and
dominance. A detailed description of the estimators computed can be found in
Colwell and Coddington (1994), Chazdon et al. (1998), and Colwell (2006).
EstimateS (version 8.0) computes randomized species accumulation curves,
statistical estimators of true species richness (S), and a statistical estimator of
20
40
Similarity
60
80
100
WBF HYR SPC EYR NAT SBL SYR CLH LPH PRS BSF UBF
20
40
Similarity
60
80
100
LPH PRS HYR WBF SPC EYR NAT SBL SYR CLH BSF UBF
the true number of species shared between pairs of samples, based on species-by-
sample (or sample-by-species) incidence or abundance matrices. It can be used
to compute an expected species accumulation curve, the Mao Tau (with 95%
confidence limits). The Mao Tau is a sample-based rarefaction curve which pro-
vides a graphic estimate of expected species accumulation (Colwell et al. 2004;
Figure 18.1). The program also can be used to compute both incidence-based
coverage estimates (ICE) and abundance-based coverage estimates (ACE) of
species richness among sampling sites. The derivation and use of these estima-
tors is discussed by Chazdon et al. (1998) and Colwell (2006).
EstimateS further allows computation of Fisher’s , Shannon, and Simpson
diversity indices; the Chao, Jacknife, ICE, ACE and other species richness esti-
mators for abundance and incidence data; modified versions of the Sørensen
and Jaccard similarity indices based on abundances, including the effects of
unseen shared species (Chao et al. 2005); and classic Jaccard, Bray–Curtis, and
Morisita–Horn (both incidence-based and abundance-based) similarity esti-
mators. EstimateS can be downloaded free of charge from http://viceroy.eeb.
uconn.edu/estimates.
Ecological software to accompany Kreb’s Ecological Methodology is available
from Exeter Software (Setauket, New York, USA; www.exetersoftware.com/
cat/ecometh/ecomethodology.html). Version 6.1.3 sells for US$150 (as of May
2008). The software covers only the topics and analyses discussed in the book,
which include various measures of richness, heterogeneity, and equitability, in
addition to a wide range of other ecological analyses. These include binary coef-
ficients, Euclidean distance coefficients, Bray–Curtis metric, Canberra met-
ric, percentage similarity, Morisita’s Index of Similarity, Horn’s Index, Species
Richness (Rarefaction method, Jackknife method for counts), logarithmic ser-
ies, log-normal distribution, Simpson’s Index of Diversity, Shannon–Wiener
measure, Brillouin’s Index of Diversity, and evenness measures. The program
comes with a manual describing the analyses, and includes some information on
the use and theory behind the various indices.
Primer 6 for Windows is another powerful tool for analysis of ecological data,
including both diversity and dominance indices. Information on the program
can be obtained from the website (www.primer-e.com/). Primer 6 allows uni-
variate, graphical, and multivariate analyses of species, abundance, biomass,
and physio-chemical data. The program facilitates grouping data into clusters,
allows identification of species that are responsible for discrimination among
two sample clusters, and graphs species abundance distributions through ordin-
ation and multidimensional scaling plots. As of May 2008 the cost is US$500
for a single-user license for academic research and is available from Primer-E
334 | Amphibian ecology and conservation
18.6 Summary
Indices of diversity and similarity offer a means of critically examining cur-
rent ecological patterns and changes in community composition, especially
when estimates of site occupation across a sufficient number of habitats are not
available. In particular, the Bray–Curtis Similarity Index has proven useful in
assessing the effects of habitat changes on herpetofauna and other taxa during
monitoring programs (Pawar et al. 2004; Pieterson et al. 2006; Dodd et al.
2007), comparing stream and forest faunas (Parris and McCarthy 1999; Huang
and Hou 2004), measuring the success of restoration efforts (Ruiz-Jean and
Aide 2005), prioritizing areas for conservation (Seymour et al. 2001), assess-
ing dietary differences (Whitfield and Donnelly 2006), and in analyses of geo-
graphic differences in diversity patterns (Urbina-Cardona and Londroño-M
2003; Menegon and Salvidio 2005; Smith-Vaniz et al. 2006).
An index-based approach offers insights into potential, if not definitive causes
of community change. Once potential causes are identified, research can be
designed to test hypotheses related to changes in species composition and rela-
tive abundance. Community ecology, conservation, and monitoring programs
should incorporate an evaluation of habitat variables into data-collection proto-
cols, and researchers must be aware of the potential importance of stochastic
or periodic environmental disturbances, such as storms and flooding, when
interpreting species presence/not detected and abundance data. Counting indi-
vidual animals or determining the percentage of site occupancy neglects much
information needed to understand changes in community composition through
time. The use of diversity and similarity indices offers further insight into the
comparative structure of amphibian communities.
18.7 References
Beals, E. W. (1984). Bray-Curtis ordination: an effective strategy for analysis of multi-
variate ecological data. Advances in Ecological Research, 14, 1–56.
Bloom, S. A. (1981). Similarity indices in community studies: potential pitfalls. Marine
Ecology Progress Series, 5, 125–8.
18 Diversity and similarity | 335
Boulinier, T., Nichols, J. D., Sauer, J. R., Hines, J. E., and Pollock, K. H. (1998).
Estimating species richness: the importance of heterogeneity in species detectability.
Ecology, 79, 1018–28.
Brose, U., Martinez, N. D., and Williams, R. J. (2003). Estimating species richness: sen-
sitivity to sample coverage and insensitivity to spatial patterns. Ecology, 84, 2364–77.
Cam, E., Nichols, J. D., Sauer, J. R., and Hines, J. E. (2002). On the estimation of species
richness based on the accumulation of previously unrecorded species. Ecography, 25,
102–8.
Chao, A., Chazdon, R. L., Colwell, R. K., and Shen, T.-J. (2005). A new statistical
approach for assessing similarity of species composition with incidence and abundance
data. Ecology Letters, 8, 148–59.
Chazdon, R. L., Colwell, R. K., Denslow, J. S., and Guariguata, M. R. (1998). Statistical
methods for estimating species richness of woody regeneration in primary and second-
ary rainforests of northeastern Costa Rica. In F. Dallmeier and J. A. Comisky (eds),
Forest Biodiversity Research, Monitoring and Modelling: Conceptual Background and Old
World Case Studies, pp. 285–309. Smithsonian Institution/Man and Biosphere No. 20.
Smithsonian Institution, Washington DC.
Clarke, K. R. and Warnick, R. M. (2001). Change in Marine Communities: an approach to
statistical analysis and interpretation, 2nd edn. Primer-E, Plymouth.
Clarke, K. R. and Gorley, R. N. (2006). PRIMER v6: user manual/tutorial. Primer-E,
Plymouth.
Clarke, K. R., Somerfield, P. J., and Chapman, M. G. (2006). On resemblance measures
for ecological studies, including taxonomic dissimilarities and a zero-adjusted Bray-
Curtis coefficient for denuded assemblages. Journal of Experimental Marine Biology and
Ecology, 330, 55–80.
Colwell, R. K. (2006). EstimateS 8.0 user’s guide. http://viceroy.eeb.uconn.edu/
EstimateSPages/EstSUsersGuide/EstimateSUsersGuide.htm.
Colwell, R. K. and Coddington, J. A. (1994). Estimating terrestrial biodiversity through
extrapolation. Philosophical Transactions of the Royal Society London Series B Biological
Sciences 345, 101–18.
Colwell, R. K., Mao, C. X., and Chang, J. (2004). Interpolating, extrapolating, and com-
paring incidence-based species accumulation curves. Ecology, 85, 2717–27.
de Solla, S. R., Shirose, L. J., Fernie, K. J., Barrett, G. C., Brousseau, C. S., and Bishop, C. A.
(2005). Effect of sampling effort and species detectability on volunteer based anuran moni-
toring programs. Biological Conservation, 121, 585–94.
Dodd, Jr, C. K. (1992). Biological diversity of a temporary pond herpetofauna in north
Florida sandhills. Biodiversity and Conservation, 1, 125–42.
Dodd, Jr, C. K., Barichivich, W. J., Johnson, S. A., and Staiger, J. R. (2007). Changes
in a northwestern Florida Gulf Coast herpetofaunal community over a 28-y period.
American Midland Naturalist, 158, 29–48.
Duellman, W. E. (1999). Global distribution of amphibians: patterns, conservation, and
future challenges. In W. E. Duellman (ed), Patterns of Distribution of Amphibians. A
Global Perspective, pp. 1–30. Johns Hopkins University Press, Baltimore, MD.
Gotelli, N. J. and Colwell, R. K. (2001). Quantifying biodiversity: procedures and pitfalls
in the measurement and comparison of species richness. Ecology Letters, 4, 379–91.
336 | Amphibian ecology and conservation
Heyer, R. W., Coddington, J., Kress, J. W., Acevede, P., Cole, D., Erwin,T. L., Meggers,
B. J., Pogue, M. G., Thorington, R. W., Vari, R. P., Weitzman, M. J., and Weitzman,
S. H. (1999). Amazonic biotic data and conservation decisions. Ciencia e Cultura,
Journal of the Brazilian Association for the Advancement of Science, 51, 372–85.
Huang, C.-Y. and Hou, P.-C. L. (2004). Density and diversity of litter amphibians in a
monsoon forest of southern Taiwan. Zoological Studies, 43, 795–802.
Krebs, C. J. (1999). Ecological Methodology, 2nd edn. Benjamin-Cummings, Menlo
Park, CA.
Magurran, A. E. (1988). Ecological Diversity and its Measurement. Princeton University
Press, Princeton, NJ.
Magurran, A. E. (2003). Measuring Biological Diversity. Blackwell Publishing, Oxford.
May, R. M. (1988). How many species are there on Earth? Science, 241, 1441–9.
Menegon, M. and Salvidio, S. (2005). Amphibian and reptile diversity in the south-
ern Udzungwa Scarp Forest Reserve, south-eastern Tanzania. In B. A. Huber,
B. J. Sinclair, and K.-H. Lampe (eds), African Biodiversity. Molecules, Organisms,
Ecosystems, pp. 205–12. Springer, New York.
Parris, K. M. and McCarthy, M. A. (1999). What influences the structure of frog assem-
blages at forest streams? Australian Journal of Ecology, 24, 495–502.
Pawar, S. S., Rawat, G. S., and Choudhury, B. C. (2004). Recovery of frog and lizard
communities following primary habitat alteration in Mizoram, northeast India. BMC
Ecology, www.biomedcentral.com/1472-6785/4/10.
Pieterson, E. C., Addison, L. M., Agobian, J. N., Brooks-Solveson, B., Cassani, J., and
Everham III, E., M. (2006). Five years of the southwest Florida frog monitoring
network: changes in frog communities as an indicator of landscape change. Florida
Scientist, 69, 117–26.
Pineda, E. and Halffter, G. (2004). Species diversity and habitat fragmentation: frogs in a
tropical montane landscape in Mexico. Biological Conservation, 117, 499–508.
Ruiz-Jean, M. C. and Aide, T. M. (2005). Restoration success: how is it measured?
Restoration Ecology, 13, 569–77.
Scott, J. M., Csuti, B., Jacobi, J. D., and Estes, J. E. (1987). Species richness. A geographic
approach to protecting future biological diversity. BioScience, 37, 782–8.
Seymour, C. L., De Klerk, H. M., Channing, A., and Crowe, T. M. (2001). The bio-
geography of the Anura of sub-equatorial Africa and the prioritisation of areas for their
conservation. Biodiversity and Conservation, 10, 2045–76.
Smith, E. P. and van Belle, G. (1984). Nonparametric estimation of species richness.
Biometrics, 40, 119–29.
Smith-Vaniz, W. F., Jelks, H. L., and Rocha, L. A. (2006). Relevance of cryptic fishes
in biodiversity assessments: a case study at Buck Island Reef National Monument,
St. Croix. Bulletin of Marine Science, 79, 17–48.
Snodgrass, J. W., Komorosski, M. J., Bryan, Jr, L., and Burger, J. (2000). Relationships
among isolated wetland size, hydroperiod, and amphibian species richness: implica-
tions for wetland regulation. Conservation Biology, 14, 414–19.
Soberón, M. J. and Llorente, B. J. (1993). The use of species accumulation functions for
the prediction of species richness. Conservation Biology, 7, 480–8.
Thompson, G. G. and Withers, P. C. (2003). Effect of species richness and relative abun-
dance on the shape of the species accumulation curve. Austral Ecology, 28, 355–60.
18 Diversity and similarity | 337
19.1 Introduction
19.1.1 Relevance of landscape ecology to
amphibian biology and conservation
The field of landscape ecology “deals with the effects of the spatial configuration
of mosaics on a wide variety of ecological phenomena” (Wiens et al. 1993). The
field has emerged from recognition of the need to link ecological processes to
landscape configuration and composition (Turner et al. 2001). Another stimu-
lus has been the revelation that population processes play out over much larger
areas and time frames than previously assumed. Landscape ecology is a rela-
tively young discipline that emerged in Central and Eastern Europe (Naveh
and Lieberman 1994). The field has exploded over the past two decades largely
because of technological advances (Turner et al. 2001). On the data side, a
sudden wealth of spatial data from global positioning systems (GPS), satellite
imagery, and high-resolution aerial photography has materialized, coupled with
the Internet as a medium for its quick distribution. On the analysis side, com-
puter processing speeds have increased exponentially (Moore 1975) and a large
suite of software packages for manipulating and analyzing spatial data (that is,
Geographic Information Systems or GIS) has evolved that capitalize on these
vastly faster processing speeds. The result is a synergism between more data and
faster computers for processing it, thereby creating opportunity for tackling
increasingly complex questions in landscape ecology.
Technological advances have also occurred in a societal context in which envir-
onmental problems have become more and more pervasive. This, in turn, has
forced recognition that conservation actions must be implemented over broad
spatial scales that have not been addressed within the domain of classic ecological
research; that is, conducted on a few hectares at most (Turner et al. 2001). As a
340 | Amphibian ecology and conservation
result, the “landscape perspective” has become widespread in ecology and GIS
has become an essential component of the ecologist’s toolbox.
In the realm of amphibian biology, Douglas Gill’s studies of the metapopula-
tion ecology of red-spotted newts (Gill 1978a, 1978b) were among the first to
catalyze this shift in scale in how we conceptualize population processes. These
studies prompted recognition that the fates of individual, pond-based breeding
populations of amphibians—the focus of nearly all prior research—are often
tightly linked to the fates of other populations nearby as mediated by the flow
of migrants among them. This landscape-as-population rather than pond-
as-population perspective has been the subject of much subsequent research
(Sjögren-Gulve 1994; Pope et al. 2000; Marsh and Trenham 2001; Hels 2002;
Gamble et al. 2007; Richter-Boix et al. 2007; Werner et al. 2007) that has radic-
ally expanded the temporal and spatial scales at which we conceive of planning
approaches for conserving amphibians (Semlitsch 2000; Trenham et al. 2003;
Smith and Green 2005).
Landscape ecology is an increasing focus of amphibian research
(Figure 19.1). The goal of this chapter is to assist amphibian biologists in
0.25
Amphibian(s) and
landscape(s)
0.2
Fraction of total citations
Amphibian(s) and
habitat(s)
0.15
Amphibian(s) and
genetic(s)
0.1
Amphibian(s) and
communit(y, ies)
0.05
Amphibian(s) and
population(s)
0
2002 2003 2004 2005 2006 2007
Year
Fig 19.1 Fraction of amphibian studies published since 1985 featuring keywords
that combine “amphibian(s)” with “landscape(s),” “habitat(s),” “genetic(s),”
“communit(y, ies),” and “population(s),” indicating a recent surge in interest
in landscape-level studies of amphibians relative to that on habitats, genetics,
communities, or populations (Source: Thomson Scientific’s Web of Science; data
plotted for last 5 years only).
19 Landscape ecology and GIS methods | 341
24
17
2 5
5 3
1 6
3
1 1
3
1
1
1
1 2 2
1
2
1
2
3 1
1 500 m
1
2 1
they occupy might imply stability when in fact they are experiencing continuous
decline and are functionally extinct (Löfvenhaft et al. 2004).
19.1.3 GIS
A GIS offers the tools to envision, analyze, process, and model spatial data.
A GIS is essentially computer software capable of storing, manipulating, dis-
playing, and analyzing geographically referenced data. Two-dimensional spa-
tial data can be represented in vector or raster form. Vectors are points, lines,
or polygons representing real-world features (i.e. sampling locations, roads, or
land use, respectively) associated with individual data attributes. In contrast,
rasters are continuous matrices of equally sized, rectangular cells, each usually
containing a single value corresponding to the landscape feature they represent
(e.g. 1 is deciduous forest, 2 is open water, 3 is urban). Adjacent cells of similar
19 Landscape ecology and GIS methods | 343
(a) (c)
(b) (d)
(a) (b)
2 1 0 2 km 2 1 0 2 km
Fig 19.4 Circular neighborhoods used for extracting landscape scale variables
around selected ponds: (a) non-overlapping (independent variables) and (b) highly
overlapping (leading to non-independence of the predictors and pseudo-replication).
348 | Amphibian ecology and conservation
Often, small changes in the spatial pattern of resource distribution can prod-
uce sudden changes in the ecological response (Figure 19.5). These transitions
are termed critical landscape thresholds (With and Crist 1995). Such critical
thresholds have been identified in the response of amphibians to the habitat
100
(a) 30 m
80
60
40 100
m = 0.58
20 (d) 500 m
b = 11.45 80
r 2 = 0.54
0 60
Spotted salamander occurrence (%)
100
40
(b) 100 m
80 m = 1.13
20 b = –9.04
60 r 2 = 0.77
0
40
100
Forest (%) m b r2
(e) 1000 m
20 0–30 2.28 –28.04 0.86 80
30–100 0.30 39.76 0.28
0 60
100
40
(c) 300 m m = 1.25
80
20 b = –12.99
r 2 = 0.88
60
0
40
10 10
20 20
30 30
40 40
50 50
60 60
70 70
80 80
0
–9
0–
–
–
–
–
–
–
–
Forest (%) m b r2
20 Upland forest (%)
0–30 2.97 –33.41 0.96
30–100 0.55 30.86 0.47
0
10 1 0
2 0 20
30 3 0
40 0
50 50
60 60
70 70
80 80
90 90
00
–4
0–
–
–
–
–
–
–
–
–1
alteration by human activities (Gibbs 1998a; Homan et al. 2004; Denoël and
Ficetola 2007). Landscape-composition thresholds (e.g. the percentage of a
land-use type within a certain neighborhood below which the long-term viabil-
ity of the population is compromised), as well as landscape configuration (e.g.
patch isolation as function of the distance between suitable breeding and for-
aging habitat) have been documented (Denoël and Ficetola 2007).
Typically, connectivity measures have not considered the habitat matrix between
breeding sites and simply used “nearest neighbor” measures (Euclidian distances)
(Knutson et al. 1999; Pope et al. 2000), but studies that account for habitat matrix
complexity have started to emerge in recent years (Stevens et al. 2005; Compton
et al. 2007; Figure 19.6).
In metapopulation dynamics theory, connectivity measures are critical for
understanding and describing extinction and recolonization patterns and
processes. Hanski et al. (1994) developed a connectivity index that was used
for understanding amphibian metapopulation processes. This index requires
specific knowledge of site occupancy and patch carrying capacity as well as a
detailed map of the available patches:
n
CONNECT ∑ e
dij
Aj
i =1
where dij is the distance between patch i and patch j, and Aj is the carrying cap-
acity of site j. This so-called Hanski index has provided a useful measure of pond
connectivity for amphibian metapopulations (Schmidt and Pellet 2005).
Fig 19.6 Compton et al. (2007) used a modified kernel estimator (commonly
used in home-range studies) associated with a friction surface to model vernal pool
connectivity for four ambystomatid salamanders at three scales in Massachusetts,
USA. Their approach is different from a simple cost-distance calculation because
they integrated the probability of pools exchanging individuals based on empirical
salamander dispersal data (using the kernel estimator, h 339.6 m) with a
resistance map (expert-based habitat specific values). The figure depicts examples
of the resistant-kernel estimator at three scales in a landscape with a focal pool
(star), five neighboring pools (circles), and two roads: (a) local scale, showing
connectivity to upland habitat from the focal pool; (b) neighborhood scale,
showing the probability of the focal pool receiving dispersing animals from each
neighboring pool; and (c) regional scale, with dark outline indicating pools that are
interconnected by a specified level of dispersal. Darker shading indicates greater
connectivity at each scale (reprinted with permission from Blackwell Publishing).
19 Landscape ecology and GIS methods | 351
that some amphibians perform. The methods for obtaining the data (i.e. pitfall
traps) may impede continuous movements and consequently not reflect habitat
permeability best. Dissimilarities in vagility, habitat selection, and physiology
between species, as well as age-specific differences, demand that such experi-
ments be conducted across multiple species. Also, such experiments are labor-
intensive and time-consuming, and the results cannot easily be translated to
another geographical locality. So-called friction or resistance maps provide a
potential GIS-based solution to modeling permeability realistically. The pro-
cess includes assigning land-use or land-cover classes various permeability values
based on expert opinion or empirical data, and using a GIS to create a least cost
surface expressing connectivity between populations or breeding sites. Ray et al.
(2002) used a land-use map to estimate potential migration zones for common
toads (Bufo bufo) and alpine newts (Triturus alpestris) in Switzerland. Similarly,
Joly et al. (2003) estimated biological connectivity of common toad (B. bufo)
populations in the Rhone floodplain, France, while accounting for potential
road mortality. Compton et al. (2007) used a statewide land-cover data set and a
combination of expert-based and empirical permeability data to model connect-
ivity for ambystomatid salamanders at three different scales in Massachusetts,
USA (Figure 19.6).
be used. Legendre and Fortin (1989), Perry et al. (2002), and Dormann et al.
(2007) review the statistical tools for testing for the presence of spatial auto-
correlation in biological data and provide guidance for using the proper tech-
niques. Practical guidelines for addressing the issue of spatial autocorrelation
for designing of field surveys and data analysis are also provided by Legendre
et al. (2002).
Common means for analyzing spatial patterns of the data include indices
for global spatial autocorrelation (Moran’s I, Geary’s c), indices for spatial clus-
tering of group-level data (Getis–Ord Local G), and spatial autocorrelation
tests (Mantel and partial Mantel tests). Calculation of these indices and tests
can be conducted either using spatial statistics software (GSLIB, Gstat, GS,
VARIOWIN) or as packages implemented in a general statistical framework
with software (R, SAS, SYSTAT). Some GIS packages, such as ArcGIS (ESRI,
Redlands, CA, USA) and IDRISI (Clark Labs, Worcester, MA, USA), also con-
tain geostatistical tools to make these adjustments.
New modeling techniques that account for spatial autocorrelation in the
data, such as autologistic regression (Augustin et al. 1998) and geographic
weighted regression (Fotheringham et al. 2002) have been developed. For
example, Knapp et al. (2003) used autologistic regression to model probability
of pond occupancy in yellow-legged frogs (Rana muscosa) by incorporating an
autocovariate term describing the degree of isolation of ponds. As a result, the
autocovariate isolation term was the most important predictor of occupancy
in yellow-legged frogs. A similar approach was taken by Davidson (2004) for
explaining the relation between amphibian decline and historical pesticide use
in California. Dormann (2007) brings forth further evidence on the import-
ance of accounting for spatial autocorrelation: when modeling the distribution
of organisms including amphibians, reptiles, mammals, birds, plants, insects,
or mites, spatial autocorrelation biased the coefficient estimates, under- or over-
estimating by approximately 25%. Incorporating spatial autocorrelation also
improved model fit.
not be available or updated for large continental areas for many years. Most
importantly, particularly data-deficient regions, such as tropical latitudes, with
the least amphibian research attention, are nevertheless of critical importance
for amphibian conservation.
Lack of a stable “taxonomy” for ecosystem classification also plagues GIS-
based amphibian research. Researchers seeking the latest spatial information
available typically analyze remotely sensed imagery (such as from Landsat,
SPOT, or IKONOS satellite sensors) and implement their own classification
of habitats or land use, often designed to suit the focus organism. This serves
a given study well but reduces the comparability among different studies with
respect to habitat use, availability, and selection by the same organism.
Inconsistent data resolution also creates problems. For example, habitat
changes critical for amphibian population persistence might occur at a finer
scale than satellite imagery can capture. Aerial photos may, however, be suffi-
cient for the task. Whatever the case, misinterpretation due to an inability to
resolve such patterns in some cases but not others may occur because of vari-
ation in coarseness of the spatial data available across studies. Not only spatial,
but also temporal and informational resolution add to the problems of GIS use.
The time lag between the publishing date of spatial data and the study period
imposes obvious restrictions on the ability to model habitat relationships.
“GIS-philia” can also blind some biologists to the inadequacy of their efforts.
GIS represents are thrilling mix of technologies for many amphibian researchers,
and resulting maps often have a visually stunning effect that can distract from
the maps’ limitations. Moreover, no matter how accurate and up-to-date the
spatial data, the animal–habitat correlations that are at the core of GIS analysis
for many amphibian studies are always only part of the puzzle. Sophisticated
analysis and modeling of spatial data require empirical data for model param-
eterization and calibration. Such data are also required for model validation, a
critical step increasingly neglected in many modeling studies. Gathering suffi-
cient field data to confront each model and thereby assess its validity should be
regarded as a requirement for publication of any modeling studies or dissemin-
ation of any guidelines extending from such models. This said, model validation
can be equally or more expensive than the cost of developing the original model,
a cost most researchers (and their funders) are rarely willing to cover.
Landscape thresholds have the potential to become a powerful tool in man-
aging amphibian populations because they offer specific management objectives
(e.g. “retain more than 70% forest within a 2 km radius pond neighborhood
to maintain a healthy population”). However, threshold research is still in its
infancy (Huggett 2005), and the temporal and spatial behaviour of thresholds
356 | Amphibian ecology and conservation
19.5 References
Araújo, M. B., Thuiller, W., and Pearson, R. G. (2006). Climate warming and the decline
of amphibians and reptiles in Europe. Journal of Biogeography, 33, 1712–28.
Augustin, N. H., Mugglestone, M. A., and Buckland, S. T. (1998). The role of simulation
in spatially correlated data. Environmetrics, 9, 175–96.
Cabe, P. R., Page, R. B., Hanlon, T. J., Aldrich, M. E., Connors, L., and Marsh, D. M.
(2007). Fine-scale population differentiation and gene flow in a terrestrial salamander
(Plethodon cinereus) living in continuous habitat. Heredity, 98, 53–60.
Compton, B. W., McGarigal, K., Cushman, S. A., and Gamble, L. R. (2007). A resistant-
kernel model of connectivity for amphibians that breed in vernal pools. Conservation
Biology, 21, 788–99.
19 Landscape ecology and GIS methods | 357
20.1 Introduction
Both public and scientific awareness of the phenomenon of “amphibian declines”
(Blaustein and Wake 1995) have focused the importance of understanding the
physiological underpinnings of amphibian diversity, distribution, and survival
in varied and challenging environments. Contributions to understanding of
ecology have been important goals of physiologists and integrative biologists
interested in the structure, function, and behavior of amphibians (Feder and
Burggren 1992). Such studies are rich in history and possibilities considering
the biological diversity and evolutionary responses of amphibians to varied and
extensive environments. These range from caves and subterranean burrows to
high tree canopies, from forests to scrub and deserts, and from tropical lowlands
to mountain tops. For many species complex life cycles connect developmental
stages with both aquatic and terrestrial environments.
Scientific interest in the functional attributes of frogs began at least as
early as 1671 when Malpighi used frogs as a model for investigating kidney
function. This was followed by research in which amphibians were used for
investigating various aspects of physiology from neuromuscular function to
development and endocrinology. Frogs became familiar animals in college
and high-school anatomy and physiology classes. However, in both scientific
and education endeavors, the “amphibian” was simply a means to understand-
ing physiological principles. Publications by G. Kingsley Noble (1931) and
John Moore (1964) began to focus attention on the physiology of amphibians
as a means to understanding their biology. The development of physiological
ecology as an established discipline in the 1960s evoked parallel emphasis on
364 | Amphibian ecology and conservation
Thermocouples Electronic measurement of temperatures with fine wire probes and sensors; available with sensors <1 mm diameter. Battery
power is available for field portable thermocouple thermometers. See also infrared thermometers.
Thermistors Similar uses and advantages as thermocouples, but with non-linear thermal characteristics that can be unstable over time.
Infrared thermometers Non-contact temperature measurement device that detects emitted infrared energy and converts this into a temperature
reading. Critical considerations include field of view (target size and distance), spectral response, response time, signal
processing, portability, hand-held options, and temperature range. Most infrared thermometers cover environmental and
body temperatures that might be encountered in field studies of amphibians. Infrared thermocouples are small, low-cost
infrared sensors that are self-powered and produce an output that mimics a thermocouple sensor. Infrared imagery may have
varied roles in research with amphibians.
Transmitters Smallest and lightest, 0.35–0.52 g total encapsulated transmitter with antenna and battery, for tracking movement. Crystal
controlled two-stage design, pulsed by a multivibrator. Optional design available with pulse rate determined by temperature.
Several configurations available intended for glue-on applications, but could be modified for implantation depending on
ingenuity of the researcher. Larger units (0.39–3.8 g) can be supplied with helical or subcutaneous antenna. Medical grade
epoxy or wax can be used to protect implanted units.
Passive integrated Many types of implantable transponders are available for individual identification of animals from an implanted chip;
transponders (PITs) location of animals over short range (1–10 m) by use of a diode tag and harmonic radar; or transmission of a temperature
signal by passive acquisition of energy from an external electromagnetic field. No battery is required, and the small size
permits implantation in the body cavity or just beneath the skin. However, usefulness for identification and temperature
measurement is limited to very short distances (generally a few centimeters).
Dataloggers Available with multiple channels for logging multiple inputs in the field, or as mini- or micro-units that can be used in multiple
locations. Measurements include temperature, humidity, pressure, salinity, depth, and voltage. Microloggers can be used to
generate thermal or humidity maps of the environment. Underwater dataloggers measure and record water temperatures to
depths >305 m. Programmable with capability to record maximum/minimum or averaging. Sampling intervals 0.5 s–9 h. Up
to 5-year battery life. Numerous units available from multiple sources with advantages of low-cost, real-time operation,
programmable start time, reusable, and miniature size. Some units may be subject to calibration drift and should be
recalibrated at intervals or before and after use.
iButtons™ Measurement of temperature and/or humidity, recorded from built in sensors into internal memory with user-defined time
interval. Small, light-weight with stainless steel casing. A durable stainless steel package is resistant to dirt, dust, contaminants,
moisture, and shock. Reusable and relatively inexpensive, with lithium battery operation for at least 10 years without
maintenance. Units should be checked for calibration drift and should be recalibrated at intervals or before and after use.
Evaporimeters Various products are available to measure area-specific transepidermal evaporative water loss and related variables. Many of
these instruments are cumbersome, but some are portable with battery operation for field applications. The VapoMeter ®,
manufactured by Delfin Technologies, Kuopio, Finland, is a hand-held easily portable device developed for use with human
skin and can also be used with small animals. Various other products known as dynamic diffusion porometers, leaf porometers,
and leaf hygrometers are sold for use with plants and can measure conductance or resistance directly. Soil hygrometers and
ceramic plates are used to measure water potential of soils.
Note: various products are available from multiple vendors, which can be accessed via the Internet. Further specifications are available online, providing details that
are too numerous to be listed here.
370 | Amphibian ecology and conservation
1.5
0.5
8.6 g
0
5 10 15 20 25 30
Manipulation time (s)
significantly (Figure 20.1) (Navas and Araujo 2000). Ideally, the measurement
should be taken rapidly (within a matter of seconds) upon capture of the animal
while it is held loosely but firmly by means of an insulating material such as dry
cloth, cotton, or glove. Heat transfer can also be minimized if anurans are held
by a hind limb while resting upon the insulation. Typically, investigators have
inserted thermal sensors into the hindgut via entry at the cloaca, with the result
that the core temperature is represented by the rear mass of the body. I have
found it easier and quicker to insert the probe into the stomach via the mouth.
If the tip of the sensor or thermometer is applied gently to the mouth, anurans
usually open the mouth briefly when the probe can be inserted. This method
typically measures the core temperature at a more central location of the body
where the mass and thermal inertia is greatest. Insertion of the probe also is
quicker compared with the time used in locating and probing the much smaller
orifice of the cloaca. Finally, cloacal insertion often causes release of urine from
the urinary bladder, which in turn increases evaporative cooling near the site of
measurement.
Tolerance of broad thermal regimes may be characteristic of many species of
amphibians which might be subjects of investigation. Whatever instruments
20 Physiological ecology | 371
20.2.4 Telemetry
Radiotelemetry offers many advantages over the use of thermometers. Body tem-
peratures can be logged or recorded continuously or at chosen intervals without
disturbance to the animal that is being monitored. The first use of radioteleme-
try in field studies of amphibian body temperature was that of Lillywhite (1970)
who used temperature-sensitive transmitters to record internal body tempera-
tures of bullfrogs at shorelines and shallow water of ponds. Currently, a variety
of transmitter designs, sizes, and other features are available commercially, but
the options are generally not useful for studies involving smaller amphibian
species (
5 g). Where miniaturization is a requirement, dataloggers and
iButtons™ have proven to be useful (Table 20.1). Micro-dataloggers can be
used effectively to continuously monitor temperatures from both animals and
their microhabitats. They can generate a thermal map of the environment, use-
ful for understanding thermal selection behavior and constraints on physiology,
and they can monitor temperatures of real animals or physical models in simi-
lar contexts. Available temperatures can have significant effects on amphibian
activity, metabolism, muscle performance, and developmental rates. In extreme
environments the available microhabitats can quickly reach lethal temperatures,
and species must evolve strategies for avoiding thermal incapacitation or death.
For short-term measurements the unit can be simply fed to the animal (or
physically lodged in the stomach), but longer-term studies require surgical
implantation of the unit within the body cavity. Care must be taken to avoid
infection and to allow adequate recovery from the surgery before an animal is
released into the environment. Surgical implantation also avoids loss of the trans-
mitter, interference with feeding, and postprandial temperature elevation that
are possibly associated with units that are placed within the gut. Implantation
372 | Amphibian ecology and conservation
procedures should incorporate due caution to avoid cutting the intestine and
accidentally implanting units within the gut from which they are subsequently
passed. The user must consider trade-offs between range, life of operation, and
size, where the latter is determined largely by the mass of the batteries. It is rec-
ommended that transmitters or other implanted devices not exceed 3–5% of the
animal’s body mass. The use of passive integrated transponders (PITs) allows
gathering of temperature signals from even small amphibians, but the range is
very short (Table 20.1).
water loss. However, research during the past 40 years has shown that not all
amphibians evaporate water in this manner (reviews in Toledo and Jared 1993;
Lillywhite 2006).
Since the early 1970s a substantial number of anuran species have been
found to occupy very arid habitats and to withstand xeric conditions due to
plastic responses that modulate skin resistance (Shoemaker 1988; Navas et al.
2004). Arboreal species, in particular, exhibit specializations that enable them
to withstand drying conditions including exposure to full sun and low ambient
humidity. Key among these features are uricotelism and low rates of evapor-
ation attributable to superficial lipids that are secreted from specialized cuta-
neous glands and wiped over the outer skin surfaces of the body. The evolution
of uricotelism in xerophilic frogs only makes sense in the context of abilities of
animals to reduce their rates of cutaneous water loss simultaneously (Shoemaker
et al. 1972). Thus, a key adaptation is the ability of anurans to mitigate losses
of body water by means of creating an external lipid barrier to transepidermal
water loss (TEWL) (Lillywhite and Mittal 1999; Lillywhite 2006). Species with
a skin resistance to evaporative water loss exceeding about 10 s/cm have been
referred to as “waterproof” following the early publications of Loveridge (1970),
Shoemaker et al. (1972), and Blaylock et al. (1976). This designation, however,
is not literally accurate.
An alternative strategy for reducing losses of body water is the production of
“cocoons” involving multiple layers of shed epidermis when fossorial species are
subjected to dehydrating conditions while buried in soil (Lillywhite 2006). In
the case of both secreted cocoons and the supraepidermal lipids that are secreted
and wiped by arboreal frogs, subsequent function of the barrier depends on
immobility of the animal for otherwise the structural integrity of the barrier is
disturbed.
Understanding amphibian water balance in these animals requires careful
observations in the field combined with appropriate mechanistic investigations
of structure and function in the laboratory. Most physiological ecologists work-
ing with amphibians appreciate that only a small fraction of extant amphibian
species—clearly less than 2%—have actually been studied in these or related
contexts. Understanding the adaptive variation of skin resistance and its plasticity
is central to the role of water economy in limiting the distribution of amphibians
and their behavioral activities in time and in space. Increasingly, these studies will
become more integrative. Most animals require access to fresh water or a moist
substrate to replenish evaporative losses of body water, and specializations of ven-
tral skin in pelvic and abdominal regions facilitate this process behaviorally and
physiologically (Nagai et al. 1999). Water permeability generally varies directly
374 | Amphibian ecology and conservation
with gaseous permeability (Lillywhite and Mittal 1999), and the skin exchanges
respiratory gases and ions as well as water. Some amphibians rely exclusively on
the skin for respiratory gas exchange. Most studies have focused on these aspects
singly, but more integrative approaches are necessary to understand the mul-
tiple processes important to growth, distribution, and persistence of populations
(Lillywhite et al. 1973; Bartelt 2000; Child et al. 2008).
such measurements can be easily converted to volume flux if changes are meas-
ured over time. If such data are part of experiments, animals are usually fasted
and measured with the bladder empty (‘standard’ mass: Ruibal 1962) so that
defecation and urination do not influence such measurements. Due to the low
rate of metabolism and highly permeable skin characteristic of most amphib-
ians, changes in mass over relatively short intervals (without ingestion or defe-
cation) are assumed to result from exchanges of water insofar as the component
due to metabolic losses of carbon will be negligibly small. Such measurements
can be used to estimate daily water turnover rates and water budgets for amphib-
ians. Due to the generally high water turnover rates of amphibians, isotopic
labeling procedures such as the use of tritiated water (Nagy 1975) cannot be
used without error. Urine flow rates have been used to estimate water turnover
in the aquatic Xenopus laevis (Henderson et al. 1972).
Regional measurements of cutaneous evaporative water loss can be measured
by use of a VapoMeter®, which is a portable, hand-held evaporimeter manufac-
tured by Delfin Technologies, Kuopio, Finland (Table 20.1). These measure-
ments all disturb the animal, but can be carried out in the field. To date, most
information about the evaporative properties of amphibian skin is derived from
laboratory studies (Toledo and Jared 1993; Lillywhite 2006).
20.5 Energetics
Chemical energy is important to amphibians, as the ability to maintain energy
balance is crucial to both survival and reproduction. Energy budgets can be
approached from perspectives of mass balance and are most meaningful when
viewed over broad time scales such as annual cycles. Feeding rates are related
to temperature, water balance, and seasonal and hormonal influences (e.g.
growth hormone). Energy assimilation efficiencies can vary greatly, but most
values average around 80–90%. Rates of metabolism (aerobic energy expend-
iture) are commonly estimated by manometry (pressure/volume changes) or
respirometry (concentration changes of O2 or CO2), but cannot be measured
in free-ranging amphibians by means of doubly labeled water technique due to
characteristically high rates of water turnover (Nagy 1975). Therefore, energy
data incorporated into amphibian energy budgets are based on laboratory esti-
mates of metabolic energy expenditure coupled with time allocations for dif-
ferent activities (e.g. Jorgensen 1988). During periods of restricted resources,
metabolic depression facilitates survival of amphibians by lowering levels of
energy expenditure below normal resting values (generally 20–30% of stand-
ard metabolic rate; Guppy and Withers 1999). Brumation, dormancy, or
376 | Amphibian ecology and conservation
as would the species represented. Various investigators have used models of frogs
cast as agar (from alginate molds; Navas and Araujo 2000), plaster (O’Connor
1989), sponges (Hasegawa et al. 2005), and copper casts or tubes covered with
water-saturated cotton or cloth (Bradford 1984; Bartelt and Peterson 2005).
Care should be taken to ensure that the reflectance of the cloth or model surface
is similar to that of the animals being compared, and that evaporation rates are
not compromised by depletion of water associated with the model. Bartelt and
Peterson (2005) have developed a physical model with a datalogger and water
reservoir, which maintains a wet cloth sleeve over a copper model (Figure 20.2).
To date, physical models have represented either wet- or dry-skinned animals.
Agar replicas of anurans have been used as null models in studies of amphibians
having near-zero resistance to evaporative water loss, and these models have
been shown to exhibit rates of water evaporation and temperatures identical
to those of living frogs which they are intended to mimic (Navas and Araujo
2000). The use of models for investigating amphibians having an intermediate
resistance, or which vary the resistance of skin by wiping of lipids, still present
special challenges.
Copper tube
with cloth sleeve
Soil surface
(to battery pack)
Strain gauge
Plastic container
Fig 20.2 A physical model used to simulate the thermal and evaporative properties
of toads. The model records operative (equilibrial) temperature (Te) and evaporative
water loss (EWL), both tested for accuracy, limitations, and applicability. The model
assumes there is no physiological control over EWL and is designed specifically for
western toads, Anaxyrus ( Bufo) boreas, but can be used or modified to measure Te
and EWL in other species of amphibians. Modified after Bartelt and Peterson (2005)
with permission.
20 Physiological ecology | 379
The use of drugs and chemicals that are introduced into animals or the envir-
onment by investigators must be considered as a potential problem in future
herpetological research in context of unwanted and unknown or inadequately
studied effects on amphibians. Chemicals used as anesthetics, sedatives, chem-
ical markers such as dyes, hormones, and various pharmaceutical agents all have
potential cumulative, long-term effects on populations if used improperly or
without knowledge of potential problems. As an example, the anesthetic tric-
aine methanesulfonate, popularly known as MS-222, has been used widely as an
anesthetic in field studies of amphibians. A review of the literature indicates that
MS-222 was used for decades before field biologists understood its mechanisms
of action, and such uses potentially impact individual animals or populations by
means of increasing stress, impairing sensory perception, and interfering with
abilities of researchers to quantify parasite loads or diagnose bacterial infection
(Byram and Nickerson 2009). During studies of declining populations of hell-
benders (Cryptobranchus alleganiensis), thousands of these animals have been
exposed to MS-222, many of them multiple times. As a consequence, Byram
and Nickerson (2009) recommend that usage of MS-222 should be avoided
when possible. When the anesthetic is used, it should be properly buffered, and
the treated animals should be allowed adequate time and facilities for recovery.
Field biologists need to recognize that the use of this or other anesthetics for
immobilizing animals for marking, etc., can alter the animal’s physiology and
behavior as well as that of associated parasites and microbes. The authors also
suggest that herpetologists and field biologists work as a community to develop
humane and informed field anesthesia protocols. Such a recommendation would
apply also to use of any other chemical in field research to ensure that researchers
do not negatively impact the declining populations they might study.
Finally, it is important to comment briefly on the topics of reproduction and
development. The evolutionary success of amphibians ultimately depends on
their capabilities to reproduce, grow, and in many cases transition between dif-
ferent environments associated with metamorphosis. Amphibian growth and
reproduction is characterized by diversity of reproductive modes, plasticity, and
interplay between internal physiological cycles and external conditions of envir-
onment. The identification of factors in the environment that influence growth
and development will continue to be important, especially with respect to cli-
mate, environmental chemistry, and the quality of food and water. Methods for
measurement of egg properties, body mass, composition, and nutritional and
hormonal status in the field (e.g. Licht et al. 1983) will be important to valid-
ate conclusions based strictly in laboratory studies, which means that access,
timing, sampling effort, and other aspects of field investigation will need to
382 | Amphibian ecology and conservation
20.8 References
Adolph, E. F. (1933). Exchanges of water in the frog. Biological Reviews, 8, 224–40.
Bartelt, P. E. (2000). A Biophysical Analysis of Habitat Selection in Western Toads (Bufo
boreas) in Southeastern Idaho. PhD dissertation, Idaho State University, Pocatello, ID.
Bartelt, P. E. and Peterson, C. R. (2005). Physically modeling operative temperatures and
evaporation rates in amphibians. Journal of Thermal Biology, 30, 93–102.
Bentley, P. J. (1966). Adaptations of amphibian to arid environments. Science, 152,
619–23.
Blaustein, A. R. and Wake, D. B. (1995). The puzzle of declining amphibian populations.
Scientific American, 272, 52–7.
Blaylock, L. A., Ruibal, R., and Plat-Aloia, K. (1976). Skin structure and wiping behavior
of phyllomedusine frogs. Copeia, 1976, 283–95.
Boutilier, R. G., Stiffler, D. F., and Toewes, D. P. (1992). Exchange of respiratory gases, ions,
and water in amphibious and aquatic amphibians. In M. E. Feder and W. W. Burggren
(eds), Environmental Physiology of the Amphibians, pp. 81–124. University of Chicago
Press, Chicago, IL.
Bradford, D. F. (1984). Temperature modulation in a high-elevation amphibian, Rana
muscosa. Copeia, 1984, 966–76.
Brattstrom, B. H. (1962). Thermal control of aggregation behavior in tadpoles.
Herpetologica, 18, 38–46.
20 Physiological ecology | 383
Lajtha, K. and Michener, R. H. (eds) (1994). Stable Isotopes in Ecology and Environmental
Science. Blackwell Scientific Publications, Oxford.
Lauer, A., Simon, M. A., Banning, J. L., Andre, E., Duncan, K., and Harris, R. N. (2007).
Common cutaneous bacterial from the eastern red-backed salamander can inhibit
pathogenic fungi. Copeia, 2007, 630–40.
Lee. A. K. (1968). Water economy of the burrowing frog, Heleiporus eyrei (Grey). Copeia,
1968, 741–5.
Licht, P., McCreery, B. R., Barnes, R., and Pang, R. (1983). Seasonal and stress related
changes in plasma gonadotropins, sex steroids, and corticosterone in the bullfrog, Rana
catesbeiana. General and Comparative Endocrinology, 50, 124–45.
Lillywhite, H. B. (1970). Behavioral temperature regulation in the bullfrog. Copeia,
1970, 158–68.
Lillywhite, H. B. (1975). Physiological correlates of basking in amphibians. Comparative
Biochemistry and Physiology, 52A, 323–30.
Lillywhite, H. B. (2006). Water relations of tetrapod integument. Journal of Experimental
Biology, 209, 202–26.
Lillywhite, H. B. and Mittal, A. K. (1999). Amphibian skin and the aquatic-terrestrial
transition: constraints and compromises. In A. K. Mittal, F. B. Eddy, and J. S. Datta
Munshi (eds), Water/Air Transitions in Biology, pp. 131–44. Oxford and IBH
Publishing, New Delhi.
Lillywhite, H. B. and Navas, C. A. (2006). Animals, energy, and water in extreme envi-
ronments: perspectives from Ithala 2004. Physiological and Biochemical Zoology, 79,
265–73.
Lillywhite, H. B., Licht, P., and Chelgren, P. (1973). The role of behavioral thermoregula-
tion in the growth energetics of the toad, Bufo boreas. Ecology, 54, 375–83.
Lillywhite, H. B., Mittal, A. K., Garg, T. K., and Agrawal, N. (1997). Wiping behav-
ior and its ecophysiological significance in the Indian tree frog Polypedates maculatus.
Copeia, 1997, 88–100.
Lillywhite, H. B., Mittal, A. K., Garg, T. K., and Das, I. (1998). Basking behavior,
sweating and thermal ecology of the Indian tree frog, Polypedates maculatus. Journal of
Herpetology, 32, 169–75.
Lips, K. R., Diffendorfer, J., Mendelson, III, J. R., and Sears, M. W. (2008). Riding the
wave: Reconciling the roles of disease and climate change in amphibian declines. PLOS
Biology, 6, 441–54.
Loveridge, J. P. (1970). Observations on nitrogenous excretion and water relationships of
Chiromantis xerampelina (Amphibia, Anura). Arnoldia (Rhodesia), 5, 1–6.
Magnuson, J. J., Crowder, L. B., and Medvick, P. A. (1979). Temperature as an ecological
resource. American Zoologist, 19, 331–43.
Mayhew, W. (1965). Hibernation in the horned lizard, Phrynosoma m’calli. Comparative
Biochemistry and Physiology, 16, 103–19.
Moore, J. A. (1964). Physiology of the Amphibia. Academic Press, New York.
Nagai, T., Koyama, H., Hoff, K. V. S., and Hillyard, S. D. (1999). Desert toads discrim-
inate salt taste with chemosensory function of the ventral skin. Journal of Comparative
Neurology, 408, 125–36.
20 Physiological ecology | 385
Nagy, K. A. (1975). Water and energy budgets of free-living animals: measurement using
isotopically labeled water. In N. F. Hadley (ed.), Environmental Physiology of Desert
Organisms, pp. 227–45. Dowden, Hutchinson and Ross, Stroudsburg, PA.
Navas, C. A. and Araujo, C. (2000). The use of agar models to study amphibian thermal
ecology. Journal of Herpetology, 32, 330–4.
Navas, C. A., Antoniazzi, M. M., and Jared, C. (2004). A preliminary assessment of
anuran physiological and morphological adaptation to the Caatinga, a Brazilian
semi-arid environment. In S. Morris and A. Vosloo (eds), International Congress
Series, vol. 1275, Animals and Environments, pp. 298–305. Proceedings of the Third
International Conference of Comparative Physiology and Biochemistry. Elsevier,
Cambridge.
Noble, G. K. (1931). The Biology of the Amphibia. McGraw-Hill, New York.
O’Connor, M. P. (1989). Thermoregulation in Anuran Amphibians: Physiology, Biophysics,
and Ecology. PhD dissertation, Colorado State University, Fort Collins, CO.
Overton, E. (1904). Neununddreissig Thesen über die Wasserökonomie der Amphibien
and osmotischen Eigenschaften der Amphibienhaut. Verhandlungen der Physikalisch-
medizineschen Gesselschaft zu Würzburg, 36, 277–95.
Pinder, A. W., Storey, K. B., and Ultsch, G. R. (1992). Estivation and hibernation. In
M. E. Feder and W. W. Burggren (eds), Environmental Physiology of the Amphibians,
pp. 250–76. University of Chicago Press, Chicago, IL.
Pounds, J. A., Fogden, M. P. L., and Campbell, J. H. (1999). Biological response to cli-
mate change on a tropical mountain. Nature, 398, 611–15.
Pounds, J. A., Bustamante, M. R., Coloma, L. A., Consuegra, J. A., Fogden, M. P.L.,
Foster, P. N., La Marca, E., Masters, K. L., Merino-Viteri, A., Puschendorf, R. et al.
(2006). Widespread amphibian extinctions from epidemic disease driven by global
warming. Nature, 439, 161–7.
Ray, C. (1958). Vital limits and rates of desiccation in salamanders. Ecology, 39,
75–83.
Rome, L. C., Stevens, E. D., and John-Alder, H. B. (1992). The influence of tem-
perature and thermal acclimation on physiological function. In M. E. Feder and
W. W. Burggren (eds), Environmental Physiology of the Amphibians, pp. 183–205.
University of Chicago Press, Chicago, IL.
Rowley, J. J. and Alford, A. R. (2007). Non-contact infrared thermometers can accurately
measure amphibian body temperatures. Herpetological Review, 38, 308–11.
Ruibal, R. (1962). The adaptive value of bladder water in the toad, Bufo cognatus.
Physiological Zoology, 35, 218–23.
Seymour, R. S. and Bradford, D. F. (1995). Respiration of amphibian eggs. Physiological
Zoology, 68, 1–25.
Shoemaker, V. H. (1988). Physiological ecology of amphibians in arid environments.
Journal of Arid Environments, 14, 145–53.
Shoemaker, V., Balding, D., Ruibal, R., and McClanahan, Jr, L. (1972). Uricotelism and
low evaporative water loss in a South American frog. Science, 175, 1018–20.
Spotila, J. R. (1972). The role of temperature and water in the ecology and distribution of
some lungless salamanders. Ecological Monographs, 42, 95–125.
386 | Amphibian ecology and conservation
Spotila, J. R., O’Connor, M. P., and Bakken, G. S. (1992). Biophysics of heat and mass
transfer. In M. E. Feder and W. W. Burggren (eds), Environmental Physiology of the
Amphibians, pp. 59–80. University of Chicago Press, Chicago, IL.
Thorson, T. B. (1955). The relationship of water economy to terrestrialism in amphibians.
Ecology, 36, 100–16.
Thorson, T. B. (1964). The partitioning of body water in Amphibia. Physiological Zoology,
37, 395–9.
Toledo, R. C. and Jared, C. (1993). Cutaneous adaptations to water balance in amphib-
ians. Comparative Biochemistry and Physiology, 105A, 593–608.
Tracy, C. R. (1975). Water and energy relations of terrestrial amphibians: insights
from mechanistic modeling. In D. M. Gates and R. M. Schmerl (eds), Perspectives in
Biophysical Ecology, pp. 325–46. Springer-Verlag, New York.
Tracy, C. R. (1976). A model of the dynamic exchange of water and energy between a ter-
restrial amphibian and its environment. Ecological Monographs, 46, 293–326.
Tracy, C. R., Christian, K. A., O’connor M. P., and Tracy, C. R. (1993). Behavioral
thermoregulation by Bufo americanus: the importance of the hydric environment.
Herpetologica, 49, 375–82.
Ultsch, G. R., Bradford, D. F., and Freda, J. (1999). Physiology: coping with the environ-
ment. In R. W. McDiarmid and R. Altig (eds), Tadpoles: The Biology of Anuran Larvae,
pp. 189–214. University of Chicago Press, Chicago, IL.
Warburg, M. R. (1965). Studies on the water economy of some Australian frogs. Australian
Journal of Zoology, 13, 317–30.
Whitford, W. G. and Hutchison, V. H. (1965). Gas exchange in salamanders. Physiological
Zoology, 38, 228–42.
21
Models in field studies of
temperature and moisture
Jodi J. L. Rowley and Ross A. Alford
21.1 Introduction
21.1.1 The importance of understanding the temperature
and moisture environments of amphibians
In ectothermic organisms, variation in body temperature directly affects fac-
tors such as rates of energy acquisition, growth, and reproduction (Shoemaker
and McClanahan 1975; McClanahan 1978). Temperature also exerts a strong
influence on ecological interactions such as predator–prey and host–pathogen
interactions, and changes in temperature can completely reverse the outcome
of such interactions at both the individual and population levels (Elliot et al.
2002; Woodhams et al. 2003). Although the body temperature of terrestrial
ectotherms is broadly correlated with environmental temperature, the actual
body temperatures of many species can differ considerably from macroenviron-
mental temperatures due to species-specific behavior, physiology, morphology,
and microenvironment use. As a result, ectotherms exposed to identical mac-
roenvironmental conditions can experience very different body temperatures
(Kennedy 1997).
The physiology and ecology of amphibians, and their thermal relations,
are also strongly influenced by moisture. The degree of evaporative water loss
(EWL) varies considerably within and among species (Shoemaker and Nagy
1977; Wygoda 1984; Buttemer et al. 1996; Young et al. 2005), and individuals
of many species are able to adjust their rates of water loss over relatively short
periods of time (Withers et al. 1982; Wygoda 1989a; Withers 1995; Tracy et al.
2008). The moisture and thermal environments, and interactions between them,
can constrain the performance of amphibians (Snyder and Hammerson 1993;
388 | Amphibian ecology and conservation
Tracy et al. 1993). However, the actual temperature and humidity experienced
by an amphibian can differ dramatically within a range of available macroen-
vironmental conditions, depending on microenvironment use (Schwarzkopf
and Alford 1996; Seebacher and Alford 2002; Rowley and Alford 2007a). To
understand how the environment is experienced by amphibians in the field,
both the thermal and moisture environments must be characterized as they are
experienced by amphibians.
Understanding the thermal and water relations of amphibians in the field
is important not only for understanding their biology, but also for conserva-
tion. Recent declines in amphibian populations around the world have been
attributed in part to the epidemic disease amphibian chytridiomycosis (Lips
et al. 2006). Changes in thermoregulatory opportunities available to rainforest
frogs may have contributed to their widespread declines in association with the
disease (Pounds et al. 2006). Laboratory experiments have shown that elevated
body temperatures, as can be produced by basking, can cure amphibians of the
disease (Woodhams et al. 2003).
Global warming is also likely to have a large impact on many species, with
macroenvironmental modeling suggesting that the distributions of many
ectothermic species will be dramatically altered (Thomas et al. 2004). This
approach is limited by the fact that the present distributions of many species
may not reflect their fundamental climatic requirements (Parmesan et al.
2005). Understanding those requirements may aid in refining such predictions
(Kennedy 1997; Parmesan et al. 2005).
There are two complementary approaches to collecting field data on tem-
perature and water relations. The first is to sample living animals. This can
involve single measurements from individuals (e.g. Snyder and Hammerson
1993; Navas 1996), or intensive monitoring via mark–recapture or tracking
(e.g. Schwarzkopf and Alford 1996; Seebacher and Alford 2002). Sampling
living animals typically results in a relatively small number of measurements
per unit of investigator effort. In addition, it often requires attaching tags to
animals or repeatedly handling them, both of which may alter normal behavior.
The second approach is to place physical models that mimic the thermal and
hydric properties of animals in the field. Because the thermal and hydric states
of many models can be monitored simultaneously, a large number of microenv-
ironments can be sampled intensively. The best approach is probably a com-
bination of techniques. Living animals provide information on behavior and
microhabitat use, which is used to inform the design of studies using models and
to compare with their results. Models are used to collect large volumes of data
21 Models in temperature and moisture studies | 389
that can provide an integrated picture of how the environment should affect
animals across space and time. Sampling of living animals is covered elsewhere
in this volume; here we focus on the collection and interpretation of data using
physical models.
water loss (a permeable model) and the other (an impermeable model) has
either near-complete resistance or resistance equivalent to the highest known
for the species of interest, if good data on this exist. Although we developed and
have implemented this technique using agar models, it could be adapted to use
with models of other types. If the upper boundary of skin resistance for a spe-
cies is known, the technique we suggest could be modified by designing models
similar to those we use, but with resistance that matches the species of interest.
One way to do this has been suggested by Hasegawa et al. (2005); manipulat-
ing the water content of sponge models can vary their rates of EWL. Another
technique would be to apply an impermeable coating in patches distributed
over the body or concentrated in particular regions with lower EWL, and cali-
brate the resulting models against living animals (Navas et al. 2002).
A pair of permeable and impermeable models should always have equilib-
rium temperatures that indicate the lower and upper limits, respectively, of body
temperature available to a frog at thermal equilibrium in any location. Placing
pairs of models in a wide range of activity and retreat sites known to be used by
a species will provide data that set boundaries on the body temperatures that
can be experienced by that species in the habitat being sampled. If models with
zero and near-complete resistance to EWL are used, this is true even if the range
of skin resistance exhibited by the species is unknown, since the models span
the possible extremes. Using pairs of models ensures that the measured range
encompasses the range available to any individual of a species with variable skin
resistance, regardless of its present state.
similar to that experienced by living frogs due to the highly dynamic role of frog
ventral skin in water exchange. Coating the ventral surface simulates frogs in a
known, realistic situation, and eliminates what would otherwise be an uncon-
trolled source of variation in measurements. To create impermeable models, the
remainder of the surface of half of the models is also coated with PLASTI DIP.
These models match the permeable models in every respect, including spectral
absorbence, which spectrophotometric measurements have shown is not meas-
urably affected by clear PLASTI DIP between 330 and 800 nm. Coated and
uncoated models also cannot be distinguished in infrared images taken through
a B W 094 infrared-pass filter, indicating that their reflectance does not dif-
fer substantially in the range from 800 to 1000 nm. Small thermal datalog-
gers (Thermochron iButtons, Dallas Semicondictor, Dallas, TX, USA; diameter
15 mm, height 6 mm) are embedded in each model to allow regular, repeated
measurements of the models’ core temperature.
Agar Datalogger
Twine
twine, which was pegged at each end of the mold as the agar set, so that the
datalogger was suspended in the centre of the model (Figure 21.1). The free
twine remaining at each end of the models could be used to attach them to
surfaces either by using tape over the twine or by tying the twine around fea-
tures such as branches and roots. We set up three opaque, white, plastic con-
tainers (60 cm 40 cm 40 cm), each with a small water bowl in the center
and a metal fly-screen lid. The containers were housed in a constant tempera-
ture room, which maintained ambient temperature between 19.5 and 21.5°C.
Relative humidity fluctuated between 64 and 96% (mean 74%). A 150 W heat
lamp was provided at one end of each container and was illuminated between
0930 and 2130 h to create a temperature gradient simulating the normal range
of temperatures available to this species. We ran four temporal replicates of the
experiment, creating a total of 12 sets of measurements of frog and model tem-
peratures for comparison. Each replicate ran for 3 days.
At the start of each replicate, models were placed in each container in pairs,
with one permeable and one impermeable model. The pairs were located so
they spanned the range of thermal and light environments available in the con-
tainers. Twelve adult L. caerulea (11 males and one female) were captured near
Townsville, Queensland, Australia. They ranged from 74.1 to 91.8 mm snout–
vent length and 26 to 65 g body mass. Prior to experiments, they were main-
tained in smaller containers at ambient temperature in the constant-temperature
room in which the experiments were carried out. Each frog was used in a single
run of the experiment.
We recorded the body temperatures of frogs five times per day (0900, 1100,
1300, 1500, and 1700 h) over the 3 days of each temporal replicate, producing 15
measurements for each individual. The first time (0900 h) was chosen because
at that time the frogs had not had access to any source of extra heat for almost
12 h, and their body temperatures should have been similar to those that would
be measured during nocturnal readings in the field. Each temporal replicate was
set up at least 60 min before the first temperature reading was taken, allowing
models and frogs to reach a thermal steady state. At each reading, cloacal tem-
peratures were measured using a small, chromel-alumel K-type thermocouple
(diameter ≈1 mm) with the tip coated in plastic, attached to a digital therm-
ometer (Small Pocket Thermometer model 90000, Industrial Automation,
Joondalup, Western Australia). During temperature measurement, each frog
was held by a single leg only. When temperatures were taken, the posture and
location of the individual was noted; this information was later used to deter-
mine whether individuals had changed location, and to select the pair of models
that should represent the upper and lower boundaries of the thermal envelope
394 | Amphibian ecology and conservation
available to each frog at each measurement. All dataloggers and the thermo-
couple were calibrated before the experiment.
All frog body temperatures fell within the broad envelope defined by the
maximum and minimum temperatures of all impermeable and permeable models,
respectively (Figure 21.2a). Temperatures of frogs fell within the narrower enve-
lope defined by the temperatures measured in the same time interval in the
permeable and impermeable models nearest to the frog on 144 of the 180 (80%)
(a) 45
40
Temperature (°C)
35
30
25
20
15
(b)
40
Range of nearest models (°C)
35
30
25
20
15
20 22 24 26 28 30 32
Cloacal temperature (°C)
Fig. 21.3 Two of pairs of models placed in retreat sites used by frogs in the field
very similar rate per unit surface area. We would not recommend allowing models
to fall below 50% of their original mass, as they begin to lose shape, and their
proportional water content, and thus equivalent resistance to EWL, increases at
an accelerating rate. Model size affects rates of heating and cooling, and could
possibly alter equilibrium temperatures by altering the ratio of surface area to
volume. However, our field data (Figure 21.4) suggest that any effects of changes
as permeable models dehydrated on equilibrium temperature were very small.
The distributions of temperature difference between the permeable and imper-
meable models were similar for all levels of desiccation, and the distribution in
the 50–60% desiccated class was more similar to that in the 90–100% class
than it was to those in the 60–70% and 70–80% classes, probably because of
weather conditions at the time of measurement rather than any effect of drying
on model temperature. The outlying extreme differences shown in Figure 21.4
21 Models in temperature and moisture studies | 397
15
10
usually occurred at the same time each day and reflected occasions on which
only the impermeable model was in the sun.
To determine the accuracy with which the models delineated the envelope
of body temperatures available to frogs, we compared each measured frog body
temperature with the minimum and maximum temperatures attained by models
at the nearest 30 min mark (
15 min time difference). We removed 16 frog
body temperatures taken when frogs were found in microhabitats in which
we had not placed models (i.e. under logs, on roads, in mown pastures, or in
exposed sites high in terrestrial vegetation). Of the 145 field body temperatures
remaining (30 taken in the warm/wet season and 115 in the cool/dry season),
121 (83.4%) were within the limits defined by the hottest impermeable model
and the coldest permeable model, and 19 of the remaining 24 were within
±0.5°C of those limits. Only one body temperature was more than 0.9°C out-
side them (1.5°C above; Figure 21.5). Two-thirds of the temperatures that fell
outside the limits were taken during nocturnal activity; in most cases these
frogs were warmer than the models (Figure 21.5), which were all in diurnal
retreat sites. This suggests that the nocturnal activity sites of frogs are warmer
398 | Amphibian ecology and conservation
30
24
21
Night
Day
18
15 18 21 24 27 30
Frog body temperature (°C)
Fig. 21.5 Lines indicating the range of temperatures delineated by the maximum
and minimum temperatures of all permeable and impermeable models at rainforest
field site near Babinda, Queensland, Australia at the times body temperatures were
measured for tracked Litoria lesueuri during the cool/dry and warm/wet seasons.
Models were placed at known diurnal retreat sites of L. lesueuri. The line of equality
is included to ease visualization of where cloacal temperature fell within each pair of
model temperatures.
at night than their diurnal retreat sites are, and that models should be placed
in nocturnal activity sites to more accurately delineate the environment expe-
rienced by frogs.
The physical models produced a relatively accurate outline of the temperature
range into which the body temperatures of thermally equilibrated frogs fell, and
the ways in which the model temperature envelope differed from the frog data
were informative in themselves. This should hold true when this technique is
used for any species of amphibian.
Our field validation is more rigorous than most that have been undertaken
with reptiles (Dzialowski 2005). Many reptile studies evaluate the range of
available temperatures by deliberately placing models to cover the maximum
possible range of environmental temperatures (Dzialowski 2005). Had we done
this, all of our field temperatures would have fallen within the range measured
by models, as our laboratory temperatures did. This is not the main objective
of our technique, which is intended to use models as a means of gaining more
detailed information than is otherwise possible on the range of body tempera-
tures a species is likely to experience, given its known behavior and microhabitat
use. Measured body temperatures that fall outside the range experienced by
21 Models in temperature and moisture studies | 399
models are not a failure of the models. Rather, they indicate that the species
being studied uses a broader range of microenvironments than was provided by
the locations in which models were placed.
55 Black impermeable
Temperature (°C) ± 95% CI
Green impermeable
45
Colorless impermeable
35
25
All permeable
15
10 12 14 16 18 20 22 0 2 4 6 8 10
Time
Fig. 21.6 Temperatures at hourly intervals for permeable and impermeable agar
models of three colors exposed outdoors to ambient summer conditions. Vertical
bars indicate 95% confidence intervals (CI) for the three replicate models of each
color and permeability, lines connect the means for each color and type to aid in
visualization of patterns. Because all permeable models performed almost identically
while well hydrated, we have not attempted to make it possible to distinguish their
lines, although the black models were slightly hotter. 95% Confidence interval bars
were omitted for the last hour because steep slopes of curves made them unreadable.
All were of magnitudes similar to those at hours 10–12 on the previous day.
Wet season
1.1
1.0
Permeable model weight as 0.9
proportion of initial weight 0.8
0.7
0.6
0.5
Dry season
1.1
1.0
0.9
0.8
0.7
0.6
0.5
0 2 4 6 8
Days since model placement
Fig. 21.7 Relative water loss by permeable agar models placed in diurnal retreat
sites of Litoria lesueuri in rainforest near Babinda, Queensland, Australia. There was
substantial cloud cover and rainfall during the winter, nominally dry season, sampling
period (top), while the summer, nominally wet season, sampling period (bottom)
was unusually dry. In both seasons, some retreat sites provided very low rates of
evaporative water loss while some led to higher rates; in the summer, one of the
models reached 50% water loss on day 4, causing data collection to cease.
21.4 References
Bakken, G. S. and Gates, D. M. (1975). Heat-transfer analysis of animals: some implica-
tions for field ecology, physiology, and evolution. In D. M. Gates and R. B. Schmerl
(eds), Perspectives in Biophysical Ecology, pp. 255–90. Springer, New York.
Barbeau, T. R. and Lillywhite, H. B. (2005). Body wiping behaviors associated with
cutaneous lipids in hylid tree frogs of Florida. Journal of Experimental Biology, 208,
2147–56.
Bartelt, P. E. and Peterson, C. R. (2005). Physically modeling operative temperatures and
evaporation rates in amphibians. Journal of Thermal Biology, 30, 93–102.
Bradford, D. F. (1984). Temperature modulation in a high-elevation amphibian, Rana
muscosa. Copeia, 1984, 966–76.
Buttemer, W. A. (1990). Effect of temperature on evaporative water loss of the Australian
tree frogs Litoria caerulea and Litoria chloris. Physiological Zoology, 63, 1043–57.
Buttemer, W. A. and Thomas, C. (2003). Influence of temperature on evaporative water
loss and cutaneous resistance to water vapour diffusion in the orange-thighed frog
(Litoria xanthomera). Australian Journal of Zoology, 51, 111–18.
Buttemer, W. A., van der Wielen, M., Dain, S., and Christy, M. (1996). Cutaneous
properties of the Green and Golden Bell Frog Litoria aurea. Australian Zoologist, 30,
134–8.
Christian, K. A. and Parry, D. (1997). Reduced rates of water loss and chemical properties
of skin secretions of the frogs Litoria caerulea and Cyclorana australis. Australian Journal
of Zoology, 45, 13–20.
Dzialowski, E. M. (2005). Use of operative temperature and standard operative tempera-
ture models in thermal biology. Journal of Thermal Biology, 30, 317–34.
Elliot, S. L., Blanford, S., and Thomas, M. B. (2002). Host-pathogen interactions in
a varying environment: temperature, behavioural fever and fitness. Proceedings of the
Royal Society of London Series B Biological Sciences, 269, 1599–1607.
Hasegawa, M., Suzuki, Y., and Wada, S. (2005). Design and performance of a wet sponge
model for amphibian thermal biology. Current Herpetology, 24, 27–32.
404 | Amphibian ecology and conservation
Kennedy, A.D. (1997). Bridging the gap between general circulation model (GCM) out-
put and biological microenvironments. International Journal of Biometeorology, 40,
119–22.
Kobelt, F. and Linsenmair, K. E. (1992). Adaptations of the reed frog Hyperolius viridifla-
vus (Amphibia: Anura: Hyperoliidae) to its arid environment. Journal of Comparative
Physiology B, 162, 314–26.
Lillywhite, H. B. (2006). Water relations of tetrapod integument. Journal of Experimental
Biology, 209, 202–26.
Lips, K. R., Brem, F., Brenes, R., Reeve, J. D., Alford, R. A., Voyles, J., Carey, C., Livo, L.,
Pessier, A. P., and Collins, J. P. (2006). Emerging infectious disease and the loss of bio-
diversity in a Neotropical amphibian community. Proceedings of the National Academy
of Sciences USA, 103, 3165–70.
McClanahan, L. L. (1978). Skin lipids, water loss, and energy metabolism in a South
American tree frog (Phyllomedusa sauvagei). Physiological Zoology, 51, 179–87.
McClanahan, L. L. and Shoemaker, V. H. (1987). Behavior and thermal relations of the
arboreal frog Phyllomedusa sauvagei. National Geographic Research, 3, 11–21.
Navas, C. A. (1996). Implications of microhabitat selection and patterns of activity on the
thermal ecology of high elevation neotropical anurans. Oecologia, 108, 617–26.
Navas, C. A. and Araujo, C. (2000). The use of agar models to study amphibian thermal
ecology. Journal of Herpetology, 34, 330–4.
Navas, C. A., Jared, C., and Antoniazzi, M. M. (2002). Water economy in the casque-
headed tree-frog Corythomantis greeningi (Hylidae): role of behaviour, skin, and skull
skin co-ossification. Journal of Zoology, 257, 525–32.
O’Connor, M. P. and Tracy, C. R. (1987). Thermal and hydric relations of leopard frogs
in the field. American Zoologist, 27, 118A.
Parmesan, C., Gaines, S., Gonzalez, L., Kaufman, D. M., Kingsolver, J., Peterson, A. T.,
and Sagarin, R. (2005). Empirical perspectives on species borders: from traditional
biogeography to global change. Oikos, 108, 58–75.
Pounds, J.A. , Bustamante, M. R., Coloma, L. A., Consuegra, J. A., Fogden, M. P. L.,
Foster, P. N., La Marca, E., Masters, K. L., Merino-Viteri, A., Puschendorf, R. et al.
(2006). Widespread amphibian extinctions from epidemic disease driven by global
warming. Nature, 439, 161–7.
Rowley, J. J. L. and Alford, R. A. (2007a). Movement patterns and habitat use of rainfor-
est stream frogs in northern Queensland, Australia: implications for extinction vulner-
ability. Wildlife Research, 34, 371–8.
Rowley, J. J. L. and Alford, R. A. (2007b). Non-contact infrared thermometers can accur-
ately measure amphibian body temperatures. Herpetological Review, 38, 308–11.
Rowley, J. J. L. and Alford, R. A. (2007c). Behaviour of Australian rainforest stream
frogs may affect the transmission of chytridiomycosis. Diseases of Aquatic Organisms,
77, 1–9.
Schmuck, R., Kobelt, F., and Linsenmair, K. E. (1988). Adaptations of the reed frog
Hyperolius viridiflavus (Amphibia, Anura, Hyperoliidae) to its arid environment:
V. Iridophores and nitrogen metabolism. Journal of Comparative Physiology B, 158,
537–46.
21 Models in temperature and moisture studies | 405
22.2.2 Sampling
All the genetic methods described below are based on PCR amplification of very
small amounts of DNA. Thus tiny amounts of tissue suffice for analysis, and
sacrifice of individuals is not necessary. There are several ways of obtaining tis-
sue samples for DNA extraction. Small (2–3 mm2) segments of larval tailfins or
tail tips can be excised with a sharp scalpel; terminal digits can be removed with
sharp scissors or buccal swabs can be taken from immature or adult amphibians
(Pidancier et al. 2003). None of these procedures are likely to cause significant
mortality. It is important to ensure that sampling is representative of a popula-
tion. Taking all the samples from one aggregation of tadpoles, for example, may
bias in favour of a few sibships.
Each tissue sample should be preserved immediately in a small volume (≈1 ml)
of at least 70% ethanol, in sealed tubes, and returned to the laboratory. DNA is
stable in alcohol for many months at field or room temperature. Buccal swabs
(e.g. from Epicentre Biotechnologies, Madison, WI, USA) are an exception, and
should be treated as per the manufacturer’s instructions. With future statistical
analysis in mind, it is desirable to sample at least 20 individuals per population.
method, and tend to produce lower yields. However, extraction time is shorter
with a kit and there is less risk of contamination with substances that can
interfere with subsequent PCR amplifications. Some amphibians have high
concentrations of tissue pigments, and these are not always removed by the
phenol/chloroform approach. The Chelex method is attractively cheap and
simple, and contamination problems are rare. However, Chelex extraction pro-
duces denatured (single-stranded), not native, DNA. This makes it unsuitable
for techniques such as amplified fragment length polymorphism (AFLP; see
below) that absolutely require native DNA. FTA cards may also be unsuitable
for AFLP analysis, where DNA needs to be digested with restriction enzymes
prior to PCR amplification.
is the existence of conserved (or universal) primers for PCR reactions that func-
tion with many different species, particularly for the cytochrome b (cytb) gene.
This means that study of a new species can be initiated with minimal lead-in
time for the development of genetic markers. A disadvantage is that due to its
lack of recombination mtDNA can only be considered as a single locus despite
the multiplicity of genes that it bears. Inferences about population structure
from a single locus, particularly one that relies solely on maternal inheritance,
are questionable because they may not accurately reflect events across the gen-
ome as a whole.
polymerase), the same PCR machine, and especially the same amount of DNA
(≈25 ng) in all analyses. Because of reproducibility problems, some high-profile
journals such as Molecular Ecology usually reject population studies based solely
on RAPD analyses. However, RAPDs remain an excellent approach to interspe-
cific problems such as the identification of cryptic species and life stages.
EcoRI adaptor
Restriction enzymes
+ (Msel + EcoRI) and + AATTG
DNA ligase C
T T A A G A A T T C
T C T T A A G
A A T
G AA T T G
CAT C
G T A A G A A T C G
C A T T C T T A A C
Msel primer 1
C A T T G T A
G T A A C A T C G A G A A T T G
C A T T G T A G C T C T T A A C
C G A G A A T T G
EcoRI primer 1
Fig. 22.1 Basis of AFLP analysis using the restriction enzymes MseI and EcoRI.
M13 M13
GA/TC GA/TC
Hom
11
130
128
Hom
Fig. 22.2 Gel analysis of two microsatellite loci in 20 individuals. M13 GA/TC are
size markers, Hom show examples of homozygotes. Lanes with two band sets are
heterozygotes. Other numbers are reference samples.
in the same genus (Primmer and Merila 2002). It is nevertheless always worth
trying primers from a previously characterized close relative if these are avail-
able (Molecular Ecology Resources and Conservation Genetics publish details for
dozens of new species every year), but often it is necessary to start from scratch.
This means creating a genomic library enriched for microsatellite sequences by
selective hybridization, screening clones for microsatellites, sequencing positive
clones, and then designing primers from the flanking sequence information
(Zane et al. 2002). All of this is expensive and time-consuming. The rewards,
however, can be great. Both RAPD and AFLP analyses generate “dominant”
markers, in which heterozygotes cannot be distinguished from homozygotes.
22 Genetics in field ecology and conservation | 417
to ensure that the loci are in linkage equilibrium. Microsatellites often have tens
of alleles, and mean heterozygosity estimates across loci in large populations are
commonly more than 0.6.
Low levels of genetic diversity are possible indicators of inbreeding depression
or high genetic load. This was true in a study of natterjack toads Bufo calamita
(Rowe and Beebee 2003). However, there are important caveats on this inter-
pretation. Firstly, the estimates must be low relative to other populations, for
exactly the same suite of loci. Second, low genetic diversity at neutral loci can arise
for various reasons apart from inbreeding, such as proximity to a range edge after
postglacial colonization, as with Italian agile frogs Rana latastei (Ficoleta et al.
2007). Such populations were fully viable, whereas isolated populations with
equally low genetic diversity showed fitness reductions (Figure 22.3). It is there-
fore important to test directly for low fitness in suspect populations (for example
by measuring embryonic or larval survival and growth rates) rather than relying
solely on genetic diversity measures. However, another test can be applied to
any population with data from multiple codominant loci. When population size
declines rapidly, allelic richness is lost much faster than average heterozygosity.
BOTTLENECK quantifies this “heterozygote excess,” which is transient but
extends over several generations during and after a decline (Luikart et al. 1998).
This approach accurately identified known recent population crashes in natter-
jack toads, B. calamita (Beebee and Rowe 2001).
0.98
1.00
0.89 0.90
0.80 0.77
0.75
0.54
0.48
Hatch rate
0.38
0.50
0.25
0.00
AL CU MZ T1 T2 T3 A1 A2
Fig. 22.3 Low fitness of isolated Rana latastei populations in Italy. Three isolated
populations ( 3 km from the nearest neighbor) are shown as gray bars, and five
non-isolated populations as white bars. Standard errors are shown, based on more
than 25 egg clutches per population.
22 Genetics in field ecology and conservation | 419
Rana luteiventris exemplifies this approach (Funk et al. 2005). Programs such
as GENELAND explicitly combine both genetic and geographical informa-
tion for landscape studies. Figure 22.4 shows an example of population differ-
entiation in the long-toed salamander Ambystoma macrodactylum, determined
using seven microsatellite loci, mediated primarily by altitude in Montana, USA
(Giordano et al. 2007). Relevance to conservation is exemplified by a study of
eastern red-backed salamander (Plethodon cinereus) populations, which were
much more fragmented in semi-urban than in pristine habitats in Canada (Noel
et al. 2007). This approach can therefore reveal the significance of urban devel-
opment as disrupters of population connectivity.
MEL
Rock
RM3
RM4 Meadow
P = 0.002
CG
RM1 LEL
RM2
SP Bear
BAM 410 M
SL MM 26 BOHS
RCM RCP High vs. low
altitude
PRT P = 0.01255
TBP Bull river
BLH
Wpt 35 P = 0.0004
used to estimate effective population size, Ne, once populations are defined.
Ne is virtually always lower than census size N, on average by a factor of 10 in
wildlife populations. Ne can be estimated most accurately by two samplings of
a population, preferably separated by an interval of several generations (so 5
years for most amphibians), followed by genotyping each sample set across
multiple loci. The larger the change in allele frequencies, the greater has been
the effect of genetic drift, and thus the smaller is the Ne. Programs such as
NeESTIMATOR and TM3 are available for these calculations. An alterna-
tive is to estimate Nb, the numbers of successfully breeding adults in a single
generation, by collecting genetic data from parents and offspring in the same
year. This is more convenient than waiting for several generations, but has
large error limits especially when the parental sample includes more than one
age cohort. Most estimates of Ne in amphibian populations have been low,
often in the tens or hundreds (Beebee and Griffiths 2005).
(a) DD
Q
C
63 G
63
F
20
19 E
Palouse Range
29 B Bitterroot Range
48
6 P Clearwater Mts
61 <5 Seven-Devils Mts
J Northern Rocky Mts
55
D Salmon River Mts
Blue Mts
A
22
20
AA
CC
95 L
100
BB
Inland 100 K
66
100 64 M
East Fk. of the South Fk. o
98 N the Salmon R.
100
O
21
87 Z
97 Northern Cascade Mts
18 H
88 X
88 85 Y Olympic Mts
93 31
40 W
93
R
Coastal 99 98 S Siskiyou Mts
98 T
96 U Oregon Coast Range
100 V
0.005 substitutions/site
(b)
2-3 2-6 1-15 T
R 0 0 0 0 0 0 0 0
1-16
101 steps 1-10
S O N M
3-5
BB
23 steps
X 22 steps 1-8 0
1-18 0
3-4
3-3 2-4
0
2-11 0 0
0 0 0 W 0 0 0 0 0 0 0 0 0 0 0 0 0 0 Z
0
B 1-4 P 1-1 3-2
0 0
1-13
2-51-9 0 J 2-1
2-10 0 2-9 1-17 2-7 0
2-7 L 0 0 0 A E F G
0 26 steps 0 D 0 1-3
U 0 0 V 2-8 0 C 0
1-19 1-11 1-12 2-3 Q 1-2
0
Y AA
0 DD
0 1-6
0 0
1-14
1-7 2-2
Coastal H
K
01-5
CC
4-1
3-1
Fig. 22.5 Relationships among tailed frog populations, based on mtDNA. (a)
Phylogenetic tree. (b) Haplotype network in nested clades.
22 Genetics in field ecology and conservation | 423
offspring to parents from different source populations. The genetic and envir-
onmental contributions to critical attributes such as larval growth rates and
size at metamorphosis can be determined in this way. There are also statistical
approaches for estimating population differentiation based on such quantita-
tive traits, using Q ST as an analogue of FST. Comparisons of these two statistics
can be particularly informative. Where Q ST FST, the implication is that selec-
tion rather than random drift dominates any genetic differentiation (and vice
versa). Work along these lines has been very informative with R. temporaria
(Laugen et al. 2005)
Breeding experiments are time-consuming and, like molecular methods,
require specialized laboratory facilities. It can also be problematic to work on a
large scale, and sacrifice of animals (especially males to provide sperm) is often
necessary. There are, however, also non-destructive molecular approaches to
the study of adaptive variation. The use of highly polymorphic microsatellites
can, after analysis with programs such as KINSHIP, determine kin relations
retrospectively in samples from wild populations and thus reduce the need for
controlled crosses (although this cannot be applied to interpopulation compari-
sons). Fortunately it is now possible to identify loci involved in adaptive vari-
ation and investigate them directly. One such approach is genome scanning. In
this method, AFLPs are used initially to generate multiple polymorphic bands.
Each band (locus) is then analysed separately, in cross population comparisons,
to estimate pairwise FST values. Most loci will be neutral, but those with FST
estimates significantly greater than predicted under neutrality can be identi-
fied. These (or sequences in linkage disequilibrium with them) may be under
directional selection, between (for example) different habitat types. Once iden-
tified, diversity at these loci can be quantified to generate a “population adaptive
index”. These anonymous loci represent a valuable comparison with definitely
neutral loci and are thus a significant step forward. This method has been used
successfully with R. temporaria, tentatively identifying loci associated with
adaptation to different altitudes (Bonin et al. 2007).
Identifying loci definitely under selection is still problematic for most species,
but an exception is the major histocompatibility complex (MHC) of proteins
involved in the cellular immune response. These are the most polymorphic
protein-encoding genes known in vertebrates, and are under strong selection
to maintain resistance to pathogens. Understanding this defence mechanism
is particularly relevant in an age of global amphibian declines at least partly
mediated by chytrid fungi. PCR primers for amplification of the most variable
part of MHC class II loci (exon 2) are now available for some species such as
North American ambystomatid salamanders (Bos and DeWoody 2005), and
22 Genetics in field ecology and conservation | 425
population studies will certainly follow. There are technical difficulties with
MHC, largely because most animals have multiple loci that can be hard to
distinguish, but this gene family has great prospects for estimating adaptive
variation in wild populations.
Genetic studies have already told us a lot about amphibians that we could not
have discovered without them. Much more is undoubtedly in store, and this will
remain an exciting and challenging research area for decades ahead.
22.5 References
Babik, W., Branicki, W., Sandera, M., Lutrinchuk, S., Borkin, L. J., Irwin, J. T., and
Rafinski, J. (2004). Mitochondrial phylogeography of the moor frog, Rana arvalis.
Molecular Ecology, 13, 1469–80.
Bardsley, L., Smith, S., and Beebee, T. J. C. (1998). Identification of Bufo larvae by
molecular methods. Herpetological Journal, 8, 145–8.
Beebee, T. J. C. (2005). Amphibian conservation genetics. Heredity, 95, 423–7.
Beebee, T. J. C. and Rowe, G. (2001). Application of genetic bottleneck testing to the
investigation of amphibian declines: a case study with natterjack toads. Conservation
Biology, 15, 266–70.
Beebee, T. J. C. and Griffiths, R. A. (2005). The amphibian decline crisis: a watershed for
conservation biology? Biological Conservation, 125, 271–85.
Beebee, T. J. C. and Rowe, G. (2008). An Introduction to Molecular Ecology, 2nd edn.
Oxford University Press, Oxford.
Bonin, A., Nicole, F., Pompanon, F., Miaud, C., and Taberlet, P. (2007). Population
adaptive index: a new method to help measure intraspecific genetic diversity and priori-
tise populations for conservation. Conservation Biology, 21, 697–708.
Bos, D. H. and DeWoody, J. A. (2005). Molecular characterisation of major histocom-
patibility complex class II alleles in wild tiger salamanders (Ambystoma tigrinum).
Immunogenetics, 57, 775–81.
Burns, E. L., Eldridge, M. D. B., and Houlden, B. A. (2004). Microsatellite variation and
population structure in a declining Australian Hylid Litoria aurea. Molecular Ecology,
13, 1745–57.
Estoup, A., Wilson, I. J., Sullivan, C., Cornuet, J. M., and Moritz, C. (2001). Inferring
population history from microsatellite and enzyme data in serially introduced cane
toads, Bufo marinus. Genetics, 159, 1671–87.
Evans, B. J., Kelley, D. B., Tinsley, R. C., Melnick, D. J., and Cannatella, D. C. (2004). A
mitochondrial DNA phylogeny of African clawed frogs: phylogeography and implica-
tions for polyploid evolution. Molecular Phylogenetics & Evolution, 33, 197–213.
Ficoleta, G. F., Garner, T. W. J., and DeBernardi, F. (2007). Genetic diversity, but not
hatching success, is jointly affected by postglacial colonisation in the threatened frog,
Rana latastei. Molecular Ecology, 16, 1787–97.
Funk, W. C., Blouin, M. S., Corn, P. S., Maxell, B. A., Pilliod, D. S., Anrish, S., and
Allendorf, J. W. (2005). Population structure of Columbia spotted frogs (Rana luteiv-
entris) is strongly affected by the landscape. Molecular Ecology, 14, 483–96.
426 | Amphibian ecology and conservation
Giordano, A. R., Ridenhour, B. J., and Storfer, A. (2007). The influence of altitude and
topography on genetic structure in the long-toed salamander (Ambystoma macrodacty-
lum). Molecular Ecology, 16, 1625–37.
Jehle, R. and Arntzen, J. W. (2002). Microsatellite markers in amphibian conservation
genetics. Herpetological Journal, 12, 1–9.
Jehle, R., Burke, T., and Arntzen, J. W. (2005a). Delineating fine-scale genetic units in
amphibians: probing the primacy of ponds. Conservation Genetics, 6, 227–34.
Jehle, R., Wilson, G. A., Arntzen, J. W., and Burke, T. (2005b). Contemporary gene flow
and the spatio-temporal genetic structure of subdivided newt populations (Triturus
cristatus, T. marmoratus). Journal of Evolutionary Biology, 18, 619–28.
Laugen, A. T., Kruuk, L. E. B., Laurila, A., Rasanen, K., Store, J., and Merila, J. (2005).
Quantitative genetics of larval life-history traits in Rana temporaria in different envir-
onmental conditions. Genetical Research, 86, 161–70.
Liebgold, E. B., Cabe, P. R., Jaeger, R. G., and Leberg, P. L. (2006). Multiple paternity in
a salamander with socially monogamous behaviour. Molecular Ecology, 15, 4153–60.
Livia, L., Antonella, P., Hovirag, L., Mauro, N., and Panara, F. (2006). A non-destructive,
rapid, reliable and inexpensive method to sample, store and extract high-quality DNA
from fish body mucus and buccal cells. Molecular Ecology Notes, 6, 257–60.
Luikart, G., Sherwin, W. B., Steele, B. M., and Allendorf, F. W. (1998). Usefulness of
molecular markers for detecting population bottlenecks via monitoring genetic change.
Molecular Ecology, 7, 963–74.
Mikulicek, P. and Pialek, J. (2003). Molecular identification of three crested newt species
(Triturus cristatus superspecies) by RAPD markers. Amphibia-Reptilia, 24, 201–7.
Monsen, K. J. and Blouin, M. S. (2004). Extreme isolation by distance in a montane frog
Rana cascadae. Conservation Genetics, 5, 827–35.
Morin, P. A., Luikart, G., and Wayne, R. K. (2004). SNPs in ecology, evolution and con-
servation. Trends in Ecology and Evolution, 19, 208–16.
Newman, R. A. and Squire, T. (2001). Microsatellite variation and fine-scale population
structure in the wood frog (Rana sylvatica). Molecular Ecology, 10, 1087–1100.
Nielson, M., Lohman K., and Sullivan, J. (2001). Phylogeography of the tailed frog
(Ascaphus truei): implications for the biogeography of the Pacific Northwest. Evolution,
55, 147–60.
Noel, S., Ouellet, M., Galois, P., and Lapointe, F. J. (2007). Impact of urban fragmen-
tation on the genetic structure of the eastern red-backed salamander. Conservation
Genetics, 8, 599–606.
Noonan, B. P. and Gaucher, P. (2006). Refugial isolation and secondary contact in the
dyeing poison frog Dendrobates tinctorius. Molecular Ecology, 15, 4425–35.
Pidancier, N., Miquel, C., and Miaud, C. (2003). Buccal swabs as a non-destructive tissue
sampling method for DNA analysis in amphibians. Herpetological Journal, 13, 175–8.
Plötner, J., Kohler, F., Uzzell, T., and Beerli, P. (2007). Molecular systematics of amphib-
ians. In H. Heatwole (ed.), Amphibian Biology, vol. 7, Systematics, pp. 2672–2756.
Surrey Beatty & Sons, Chipping Norton, NSW.
Primmer, C. R. and Merila, J. (2002). A low rate of cross-species microsatellite amplifica-
tion success in Ranid frogs. Conservation Genetics, 3, 445–9.
22 Genetics in field ecology and conservation | 427
Riberon, A., Miaud, C., Guyetant, R., and Taberlet, P. (2004). Genetic variation in an
endemic salamander, Salamandra atra, using amplified fragment length polymorph-
ism. Molecular Phylogenetics & Evolution, 31, 910–14.
Rowe, G. and Beebee, T. J. C. (2003). Population on the verge of a mutational meltdown?
Fitness costs of genetic load for an amphibian in the wild. Evolution, 57, 177–81.
Rowe, G. and Beebee, T. J. C. (2007). Defining population boundaries: use of three
Bayesian approaches with microsatellite data from British natterjack toads (Bufo calam-
ita). Molecular Ecology, 16, 785–96.
Shaffer, G., Fellers, G. M., Magee, A., and Voss, R. (2000). The genetics of amphibian
declines: population substructure and molecular differentiation in the Yosemite Toad,
Bufo canorus (Anura, Bufonidae) based on single-strand conformation polymorph-
ism analysis (SSCP) and mitochondrial DNA sequence data. Molecular Ecology, 9,
245–57.
Smith, M. A. and Green, D. M. (2004). Phylogeography of Bufo fowleri at its northern
range limit. Molecular Ecology, 13, 3723–33.
Vences, M. and Wake, D. B. (2007) Speciation, species boundaries and phylogeography of
amphibians. In H. Heatwole (ed.), Amphibian Biology, vol. 7, Systematics, pp. 2613–71.
Surrey Beatty & Sons, Chipping Norton, NSW.
Vieites, D. R., Nieto-Roman, S., Barluenga, M., Palanca, A., Vences, M., and Meyer, A.
(2004). Post-mating clutch piracy in an amphibian. Nature, 431, 305–8.
Walsh, P. S., Metzger, D. A., and Higuchi, R. (1991). Chelex R100 as a medium for sim-
ple extraction of DNA for PCR-based typing from forensic material. Biotechniques, 10,
506–13.
Whitlock, M. C. and McCauley, D. E. (1999). Indirect measures of gene flow and migra-
tion: F-ST not equal 1/(4Nm+1). Heredity, 82, 117–25.
Zane, L., Bargelloni, L., and Patarnello, T. (2002). Strategies for microsatellite isolation:
a review. Molecular Ecology, 11, 1–16.
Zeisset, I. and Beebee, T. J. C. (2001). Determination of biogeographical range: an
application of molecular phylogeography to the European pool frog Rana lessonae.
Proceedings of the Royal Society of London Series B Biological Sciences, 268, 933–8.
This page intentionally left blank
Part 7
Monitoring, status, and trends
This page intentionally left blank
23
Selection of species and sampling
areas: the importance to inference
Paul Stephen Corn
23.1 Introduction
Inductive inference, the process of drawing general conclusions from specific
observations, is fundamental to the scientific method. Platt (1964) termed
conclusions obtained through rigorous application of the scientific method
as “strong inference” and noted the following basic steps: generating alterna-
tive hypotheses; devising experiments, the results of which will exclude one
or more hypotheses; conducting the experiment to get a “clean result”; and
repeating the process with revision based on the information obtained. Every
student is exposed to these basics in introductory courses, and a consider-
able proportion of a modern graduate education in the sciences is devoted
to acquiring the analytic (statistical) skills necessary to apply the scientific
method. Not even considering the field of mathematical statistics or applied
statistics in disciplines such as social sciences, library shelves groan under the
weight of texts on applied statistics, ranging from introductory (Hayek and
Buzas 1997) to advanced (Williams et al. 2002), for conducting research in
ecology, and new works are published every year. Much effort is currently
devoted to the mechanisms of analysis and the issues involved in choosing
among statistical methods; specifically, traditional hypothesis testing versus
information-theoretic or Bayesian approaches (Hobbs and Hilborn 2006).
This chapter does not address these topics, but instead discusses some of
the issues related to selection of study sites and species necessary to obtain
a “clean result.” Regardless of the method of analysis, successful inference
relies on correctly designed data collection, meaning that the observations
represent the population of interest. As Anderson (2008, p. 7) puts it, valid
432 | Amphibian ecology and conservation
23.2 Sampling
In all but a tiny number of exceptional cases, it is impossible to survey every
possible habitat or catch every individual in a population. Therefore, inference
depends on statistics generated from a subset of individuals or habitats drawn
from the population. Ideally, a sample has three qualities (Williams et al. 2002):
it comprises separate individual units, and these share the same underlying dis-
tribution and are statistically independent. These criteria are more difficult to
satisfy in inventory and monitoring studies than in a carefully designed experi-
ment, but a good study design should try to come as close as possible.
Before a sample can be taken, the sampling units must be defined and be
available for selection. Sample units can be individual animals, artificially
bounded areas (quadrats), or natural habitat features, such as ponds, streams,
or drainage basins. For population studies, where the individual is the sam-
ple unit, the availability for selection is often assumed, but this assumption is
tested when the data are analyzed. For inventory or monitoring studies, site
delineation and selection has become easier with the increasing availability
of tools such as satellite imagery and Geographic Information Systems (GIS)
software. A good example is the study by Kroll et al. (2008), which took advan-
tage of detailed GIS data layers to select stream reaches for sampling. However,
detailed GIS coverage exists mainly where there is an economic use for the
data (e.g. the commercial forests surrounding the streams studied by Kroll
et al. 2008), and even when GIS data are available the amphibian habitats may
be poorly described. Temporary wetlands are notoriously underrepresented in
aquatic data layers, and studies of lentic-breeding amphibians often will have
more success in defining sampling units if areas are used instead of specific
habitats, for example, 1 km2 blocks (Johannson et al. 2006) or drainage basins
(Corn et al. 2005).
The most basic form of probabilistic sampling is a simple random sample.
All the available sample units are put into a metaphorical hat (the mechanics
of choosing a sample now usually involve a computer) and the desired num-
ber of units are selected in random order. The number of sample units should
be sufficient to estimate parameters of interest with sufficient precision, but
samples that are too large should be avoided. Determining the desired sample
size requires knowing the variability of the parameter to be estimated, the mag-
nitude of the effect to detect (e.g. a 10% difference or a 5% annual trend), and
the strength of the inference. The formula for determining sample size depends
on the sampling scheme. Hayak and Buzas (1997) and Williams et al. (2002)
describe the details for determining sample size.
434 | Amphibian ecology and conservation
If, as is usually the case, habitats are not uniform, species are patchily dis-
tributed, or the number of samples is small relative to the area of interest, sim-
ple random sampling may be inadequate to characterize the variability of the
system being studied. There are several slightly more complex designs that can
be used to achieve a more representative sample. In the case where the num-
ber of samples is relatively small, strictly random selection can result in sample
sites clumped together instead of dispersed throughout the study area. If the
study area is not uniform, this is not desirable, and a common modification is
to employ a systematic random sample (Figure 23.1), in which, after a random
start, every nth unit is selected, where the number of available units divided by n
equals the desired sample size. Williams et al. (2002) cautioned that systematic
sampling risks a biased result if the units are arranged in such a way that envir-
onmental gradients are correlated with the order of the sampling units.
0 10 20 km
If a study area is not homogenous and differences within the area can be
described (e.g. altitudinal gradients, different types of wetlands, differences in
ability to access potential sample units) then a stratified design is usually pref-
erable to a simple random or systematic random sample. Once the strata and
sampling units within them are delineated, then simple or systematic random
samples can be drawn from each stratum. The formulas for calculating various
statistics vary among sampling designs. See Williams et al. (2002) for a concise
description of these. For long-term monitoring studies, strata should be based
on features that are not expected to change significantly during the course of the
study. For example, strata based on geological differences would be preferable to
strata based on vegetative land cover.
This section is not a comprehensive treatment of sampling design. Advanced
schemes, such as adaptive sampling (Thompson 1992; Williams et al. 2002),
are beyond the scope of this chapter, but may be necessary or desirable in many
cases. Sources listed above should be consulted when beginning the design of
any study.
Any valid sampling design must incorporate replication of sample units. A study
that compared toad populations in only two ponds could only describe the dif-
ferences in abundance between the two ponds. Tests of hypotheses of the causes
of the differences are not appropriate because there is no replication (Underwood
1998). Most inventory or monitoring studies include numerous sample units, but
care must be taken to ensure that external factors are dealt with in the design stage
(e.g. through stratification). If variation among sample units can be attributed to
something other than random processes, then the study would suffer from pseu-
doreplication (Hurlbert 1984). Large field studies are vulnerable to temporal pseu-
doreplication (sample units vary in some systematic fashion over time), because
it is seldom possible to visit all sample units simultaneously. This is a particular
problem for studies of amphibians. For example, surveys that use breeding activ-
ity are affected by the changing composition of breeding choruses over time, both
among and within species, and external factors, such as weather, that influence
behavior of individuals. Studies that focus on larval stages must deal with growth
and development and changes in abundance that may influence detection. It is
always a good practice to minimize the time span of field surveys where possible.
but common reasons are to choose study sites that are most accessible, or where
the target species are most abundant. The latter case poses particular problems for
analysis of trends in abundance, because there may be a built-in bias for detecting
declining populations (Alford and Richards 1999; but see Green 2003).
The effects of convenience sampling on inference can be subtle. Results of
ecological experiments are often interpreted to demonstrate the generality of
effects or even causality, but experiments by themselves are insufficient to explain
complex ecological phenomena; such an effort requires integrating observation
and theory with experimentation (Werner 1998). Broad inference is limited
when experiments, even those that are internally well designed, are conducted
at locations that are convenient for the researcher. The results may be useful for
investigating possible mechanisms, but cannot be generalized without making
the unsupportable assumption that the study sites are an unbiased represen-
tation of the habitats in question. For example, Blaustein et al. (1994) found
that ambient ultraviolet-b (UV-B) radiation caused higher mortality of amphib-
ian embryos than when embryos were shielded from UV-B at four lakes in the
Cascade Mountains in Oregon, USA. This and subsequent research formed the
basis of a theory of how climate change, UV-B, and pathogens might play a sig-
nificant role in global amphibian declines (Kiesecker et al. 2001; Pounds 2001;
Blaustein and Kiesecker 2002). The issue of amphibian declines and UV-B has
been controversial with a large literature that I will not delve into here. Relevant
to this chapter, the sites used for the UV-B studies were not chosen with respect
to the potential for exposure to UV-B, but were a convenience sample. The pri-
mary study site turned out to have water much more transparent to UV-B trans-
mission than most amphibian breeding habitats in the Pacific Northwest (Palen
et al. 2002), severely restricting the generality of the proposed hypothesis.
Convenience sampling is almost always done without any intention to prod-
uce a biased result. The water chemistry and UV-B transmission at the study
sites in the Oregon UV-B studies were not known beforehand. These sites were
known to the researchers to have suitable amphibian populations and included
locations where long-term studies had been conducted. Similarly, Corn and
Bury’s (1989) study of the effects of logging on stream amphibians in western
Oregon did not include conscious bias in selection of streams to be sampled,
but it was none-the-less based on a convenience sample. In this study, Bruce
Bury and I identified likely sample locations on topographic maps beforehand,
but the decision to sample was made in the field after inspecting each stream.
We attempted to select typical streams and sample reaches based on our know-
ledge of the characteristics of headwater streams in the region. However, we
did not begin with a well-defined population of streams, and we did not apply
23 Selection of species and sampling areas | 437
any probabilistic sampling. As stated by Hayek and Buzas (1997, p. 113), “If
the basis for inclusion in a sample is judgment, regardless of how expert, we
will not have a reproducible measure of our field study’s usefulness.” Corn and
Bury (1989) found strong differences in abundance and diversity of amphibians
between streams in logged and unlogged forest. The paper was an early influ-
ence on what has become a spate of research on stream amphibians and forest
management in the Pacific Northwest. Although the original results are largely
supported by subsequent work (Olson et al. 2007), the convenience sampling
design employed limits the scope of the conclusions and Corn and Bury (1989)
should be viewed as hypothesis-generating work rather than a definitive demon-
stration of differences between streams in managed and unmanaged forests.
Cautions about convenience sampling apply equally to the sample frame, the
pool from which the sample is drawn, as to individual study sites. Study areas,
meaning the regions containing the sample frames, are almost never chosen at
random. Study areas may be defined by a relevant management question (e.g.
what is the status of amphibians in a National Park?), or they might be contain
habitats or species of interest, yet be located conveniently near a researcher’s
institution. Probabilistic methods can be used properly to select study sites, but
if the frame is defined by convenience, then inference is restricted to the study
area and the generality of the results may be limited. Two recent studies, con-
ducted less than 200 km apart in Switzerland, illustrate this point. Pellet et al.
(2004) found that urbanization and roads had a strong negative influence on
presence of European treefrogs (Hyla arborea). Conversely, Van Buskirk (2005)
found only weak support for landscape variables (including urbanization) to
explain occurrence and abundance of treefrogs. Both studies were well designed,
and although different methods of analysis were used (logistic regression versus
information theoretic analysis) the different conclusions likely resulted from
intrinsic differences between the study areas.
The North American Amphibian Monitoring Program (NAAMP; Weir and
Mossman 2005), patterned after the North American Breeding Bird Survey
(BBS; Peterjohn et al. 1995), also suffers from a sampling frame defined by con-
venience. NAAMP conducts manual calling surveys of breeding amphibians
on prescribed road routes. Species are identified by their breeding calls, and
data are collected mainly by volunteers. Survey routes are generated through a
random process, but the goal of the program is to monitor trends in amphib-
ian populations throughout the region where routes are conducted (Weir and
Mossman 2005). Reliance on roadside observations limits the inference to
those areas accessed by roads, or requires investigators to make an additional,
untested assumption that the roadside amphibian population experiences the
438 | Amphibian ecology and conservation
same trends in abundance as those found in populations away from roads. The
biases that potentially undermine this assumption can be related to both habitat
and the observations themselves (Peterjohn et al. 1995). In Chapter 16 in this
volume Dorcas et al. discuss assumptions about auditory observations. Habitat
assumptions require that roadside wetlands reflect the same conditions found
at wetlands away from roads, and that habitat condition changes over time in
the same direction and at a similar pace alongside and away from roads. These
assumptions are more likely to be violated than satisfied. For example, Keller
and Scallan (1999) found that land cover types were similar near and away from
BBS routes in Maryland and Ohio, but that in Maryland, urbanization was
proceeding at a more rapid pace along roads. If urbanization is associated largely
with existing road networks, then roadside habitats may diverge from distant
habitats more rapidly in Maryland than in Ohio.
A more immediate and concrete difference between wetlands near and dis-
tant from roads (and also a difference between amphibians and birds) is that
roads, especially paved roads with higher volumes of traffic, are a signifi-
cant source of mortality of adult amphibians moving in and out of breeding
ponds. Studies in North America and Europe have found negative relation-
ships between traffic intensity and amphibian mortality and breeding activity
(Fahrig et al. 1995), habitat occupancy (Pellet et al. 2004), and species richness
and abundance (Eigenbrod et al. 2008). Similarly in New York, Karraker et al.
(2008) found a twofold increase in density of egg masses of two amphibian spe-
cies away from roads compared to roadside ponds. This may have been more an
indirect road effect than from direct mortality. Demographic models showed
significant negative effects on these species in roadside ponds resulting from the
application of salt for de-icing (Karraker et al. 2008). Road de-icers epitomize
a confounding variable contributing to differences that would be very difficult
to model in analyses of NAAMP data. Application of de-icers varies among
geographic regions, states, road types, and chemistry from year to year, and the
data quantifying application are likely to be extremely difficult to obtain. The
assumption that roadside habitats are equivalent to habitats away from roads is
not supported by research to date.
The road studies illustrate the point that concerns about convenience sam-
pling also apply to choosing variables for explaining patterns or trend. Data may
be readily available in GIS data layers (e.g. road density), but less available data
(e.g. traffic intensity, de-icer application) may be more important for generating
well-supported models. Anderson (2008) emphasizes that considerable effort
should be devoted to generating hypotheses before data are collected; this obvi-
ously applies to selection of the variables that make up the candidate models.
23 Selection of species and sampling areas | 439
plethodontids have found low detection probabilities, often less than 0.05 (Jung
et al. 2000; Welsh and Droege 2001; Dodd and Dorazio 2004). Low capture
probabilities are not a problem if they are relatively uniform among habitats
and observers (Welsh and Droege 2001), but there is considerable reason to
doubt that this is true. Detection probabilities of plethodontids varied from
0.06 to 0.41 among years (Dodd and Dorazio 2004), and.between 0.01 and
0.58 among sampling occasions in one model evaluated by Bailey et al. (2004a).
Low detection magnifies any bias due to variation among observers, habitats, or
years. For example, an increase in detection from 0.50 to 0.55 would result in a
10% increase in numbers of animals observed, but an increase from 0.06 to 0.38
would result in 6.5 times as many observations in the second count. This could
produce a result similar to that illustrated in Schmidt (2004, figure 1), where
counts were similar among years, but capture probability varied, with the result
that counts did not reflect a large increase in actual abundance.
Concerns about use of count data apply more broadly than to just terrestrial
salamanders. Johannson et al. (2006) used uncorrected counts of common frog
(Rana temporaria) egg masses as an index to population size, and concluded
that population size declined with increasing latitude and smaller populations
had less genetic variability. Johannson et al. conceded that egg mass counts
likely underestimated true abundance of breeding females but contended it was
an unbiased index, because sampling was the same at all study sites. However,
counts often fail to detect all egg masses present in a pond for a variety of rea-
sons, such as differences in habitat complexity, weather conditions that might
affect visibility, or variation in ability among observers. Grant et al. (2005)
found detection probabilities of ranid egg masses to vary between 0.78 and 1.0.
Variation in detection introduces uncertainty into conclusions about popula-
tion size; this uncertainty is magnified if detection varies in a systematic manner
across a study area.
Count data are incorporated into many indices of species diversity (Hayek
and Buzas 1997), but calculation of these indices for amphibian assemblages
are not appropriate, unless unbiased estimates of abundance are used instead of
the raw counts. It has long been known that number of captures varies among
sampling methods, so that a diversity index that included counts of species
made using different techniques (for example, pitfall traps for one species and
time-constrained searches for another) would not be valid (Corn 1994). Hyde
and Simons (2001) demonstrated sampling efficiency varied among methods,
but also among habitat types for some species of plethodontids when the same
method was employed. Interpretations of diversity indices suffer from the same
problems as interpretations about abundance.
23 Selection of species and sampling areas | 441
23.5 Conclusions
The path to strong inference leads through good study design that incorporates
probabilistic sampling from a well-defined population. Inventory and, especially,
monitoring studies stray from this path when scientific rigor is sacrificed to logis-
tic constraints and convenience in data collection. Tension often exists between
the field biologist and the consulting statistician regarding the requirements of
good study design and the logistical realities of data collection. Having been on
the field biologist’s side of the argument, I can testify that the attitude summa-
rized by “Yes, we realize valid sample selection is important, and it would be nice,
but we have to collect data from the real world”, is fairly common. Constraints in
site selection can be incorporated into study design, such as by stratifying based
on accessibility, and the resulting analysis can test hypotheses about whether
populations that are easily accessible differ from those that are not.
The perils of convenience sampling also apply to choice of life stage to study
or explanatory variables to incorporate in a model. The easiest life stage to study
23 Selection of species and sampling areas | 443
may not be the same one that is most sensitive to external factors, and variables
should not be included in a model simply because the data are available. There is
no “magic bullet” for sampling amphibians. No single technique encompasses
the variety of life histories of amphibians or the habitats in which they can be
found. Occupancy analysis provides a useful tool for avoiding the pitfalls of
using simple count data or the logistic difficulties of obtaining unbiased esti-
mates of abundance, but it is not a panacea. Ultimately, the design that allows
the strongest inference will be one that avoids convenience sampling and min-
imizes untested assumptions when the data are analyzed.
23.6 Acknowledgments
I thank Larissa Bailey, Blake Hossack, Benedikt Schmidt, and Susan Walls for
suggestions that greatly improved the manuscript.
23.7 References
Alford, R. A. and Richards, S. J. (1999). Global amphibian declines: a problem in applied
ecology. Annual Review of Ecology and Systematics, 30, 130–65.
Anderson, D. R. (2001). The need to get the basics right in wildlife field studies. Wildlife
Society Bulletin, 29, 1294–7.
Anderson, D. R. (2008). Model Based Inference in the Life Sciences: a Primer on Evidence.
Springer, New York.
Bailey, L. L., Simons, T. R., and Pollock, K. H. (2004a). Estimating detection probabil-
ity parameters for plethodon salamanders using the robust capture–recapture design.
Journal of Wildlife Management, 68, 1–13.
Bailey, L. L., Simons, T. R., and Pollock, K. H. (2004b). Estimating site occupancy
and species detection probability parameters for terrestrial salamanders. Ecological
Applications, 14, 692–702.
Bart, J., Droege, S., Geissler, P., Peterjohn, B., and Ralph, C. J. (2004). Density estima-
tion in wildlife surveys. Wildlife Society Bulletin, 32, 1242–7.
Blaustein, A. R. and Kiesecker, J. M. (2002). Complexity in conservation: lessons from
the global decline of amphibian populations. Ecology Letters, 5, 597–608.
Blaustein, A. R., Hoffman, P. D., Hokit, D. G., Kiesecker, J. M., Walls, S. C., and Hays, J. B.
(1994). UV repair and resistance to solar UV-B in amphibian eggs: a link to population
declines? Proceedings of the National Academy of Sciences USA, 91, 1791–5.
Brown, G. W., Scroggie, M. P., Smith, M. J., and Steane, D. (2007). An evaluation of
methods for assessing the population status of the threatened alpine tree frog Litoria
verreauxii alpina in southeastern Australia. Copeia, 2007, 765–70.
Bury, R. B. and Adams, M. J. (1999). Variation in age at metamorphosis across a latitu-
dinal gradient for the tailed frog, Ascaphus truei. Herpetologica, 55, 283–91.
Cochran, W. G. (1977). Sampling Techniques, 3rd edn. John Wiley & Sons, New York.
444 | Amphibian ecology and conservation
Johnson, D. H. (2008). In defense of indices: the case of bird surveys. Journal of Wildlife
Management, 72, 857–68.
Jung, R. E., Droege, S., Sauer, J. R., and Landy, R. B. (2000). Evaluation of terrestrial
and streamside salamander monitoring techniques at Shenandoah National Park.
Environmental Monitoring and Assessment, 63, 65–79.
Karraker, N. E., Gibbs, J. P., and Vonesh, J. R. (2008). Impacts of road deicing salt on the
demography of vernal pool-breeding amphibians. Ecological Applications, 18, 724–34.
Keller, C. M. E. and Scallan, J. T. (1999). Potential roadside biases due to habitat changes
along Breeding Bird Survey routes. The Condor, 101, 50–7.
Kiesecker, J. M., Blaustein, A. R., and Belden, L. K. (2001). Complex causes of amphib-
ian population declines. Nature, 410, 681–4.
Kroll, A. J., Risenhoover, K., McBride T., Beach E., Kernohan B. J., Light, J., and Bach, J.
(2008). Factors influencing stream occupancy and detection probability parameters of
stream-associated amphibians in commercial forests of Oregon and Washington, USA.
Forest Ecology and Management, 255, 3726–35.
Link, W. A. (2003). Nonidentifiability of population size from capture-recapture data
with heterogeneous detection probabilities. Biometrica, 59, 1123–30.
MacKenzie, D. I., Nichols, J. D., Hines, J. E., Knutson, M. G., and Franklin, A. B.
(2003). Estimating site occupancy, colonization, and local extinction when a species is
detected imperfectly. Ecology, 84, 2200–7.
MacKenzie, D. I., Nichols, J. D., Royle, J. A., Pollock, K. H., Bailey, L. L., and Hines, J. E.
(2006). Occupancy Estimation and Modeling: Inferring Patterns and Dynamics of Species
Occurrence. Academic Press, Burlington, MA.
Mazerolle, M. J., Desrochers, A., and Rochefort, L. (2005). Landscape characteris-
tics influence pond occupancy by frogs after accounting for detectability. Ecological
Applications, 15, 824–34.
Muths E., Jung, R. E., Bailey, L. L., Adams, M. J., Corn, P. S., Dodd, Jr, C. K., Fellers, G. M.,
Sadinski, W. J., Schwalbe, C. R., Walls, S. C. et al. (2005). The U.S. Department
of Interior’s Amphibian Research and Monitoring Initiative: a successful start to a
national program. Applied Herpetology, 2, 355–71.
Olson, D. H., Anderson, P. D., Frissell, C. A., Welsh, Jr, H. H., and Bradford, D. F.
(2007). Biodiversity management approaches for stream–riparian areas: perspectives
for Pacific Northwest headwater forests, microclimates, and amphibians. Forest Ecology
and Management, 246, 81–107.
Palen, W. J., Schindler, D. E., Adams, M. J., Pearl, C. A., Bury, R. B., and Diamond, S. A.
(2002). Optical characteristics of natural waters protect amphibians from UV-B in the
U. S. Pacific Northwest. Ecology, 83, 2951–7.
Pellet, J., Guisan, A., and Perrin, N. (2004). A concentric analysis of the impact of urban-
ization on the threatened European tree frog in an agricultural landscape. Conservation
Biology, 18, 1599–1606.
Peterjohn, B. G., Sauer, J. R., and Robbins, C. S. (1995). Population trends from the North
American Breeding Bird Survey. In T. E. Martin and D. M. Finch (eds), Ecology and
Management of Neotropical Migrants, pp. 3–39. Oxford University Press, New York.
Platt, J. R. (1964). Strong inference. Science, 146, 347–53.
Pounds, J. A. (2001). Climate and amphibian declines. Nature, 410, 639–40.
446 | Amphibian ecology and conservation
24.1 Introduction
Understanding the distribution and abundance of organisms is frequently stated
as the objective of ecological investigations (Elton 1927; Krebs 1972). Similarly,
distribution and abundance are primary criteria used to classify the status of
species (e.g. threatened, endangered) for conservation purposes (Gardenfors
et al. 2001). Thus, both scientific and conservation perspectives lead us to select
abundance and occupancy as reasonable state variables for investigation. We
define abundance as the number of organisms present in some area of interest
and occupancy as the proportion of sample units or, more generally, of area, that
is occupied by a species of interest.
Estimation of these quantities is typically based on count statistics of some
sort. For example, abundance estimation may be based on the number of animals
caught in traps, detected by visual encounter surveys, or perhaps by auditory
means. Estimation of occupancy is typically based on detection/non-detection
data (frequently referred to as presence/absence data) representing counts of
sample units at which the focal species is (or is not) detected. Use of count sta-
tistics to draw inferences about such quantities as abundance and occupancy
requires attention to two important sources of variation in counts (Pollock et al.
2002; Williams et al. 2002; Lancia et al. 2005).
The first important source of variation is spatial, with counts, abundance,
and occupancy of organisms varying over space. It is frequently not possible to
count organisms over the entire area about which inference is sought, requiring
selection of a subset of geographic sample units for survey. Selection of these
units must be done in such a way that counts on these units permit inference
about the units not selected, even in the face of geographic variation. Many
448 | Amphibian ecology and conservation
such sampling designs have been developed (e.g. simple random sampling,
stratified random sampling, adaptive cluster sampling), and we refer the reader
to sources that describe these designs in some detail (Cochran 1977; Williams
et al. 2002).
The second important source of variation concerns the recognition that
counts, whether of organisms or occupied sample units, are nearly always incom-
plete, in that some organisms and species go undetected no matter how intensive
the sampling effort. Detection probability associated with the count must be
estimated to provide information needed to translate the count(s) into an infer-
ence about abundance or occupancy. Various means of estimating detection
probabilities have been developed for both counts of organisms (Seber 1982;
Williams et al. 2002; Lancia et al. 2005) and counts of sample units occupied by
a focal species (MacKenzie et al. 2002, 2006; Royle and Dorazio 2008).
Both scientific and conservation interests frequently focus not only on the
values of relevant state variables, but also on the estimation of the vital rates
responsible for the dynamics of those state variables. Rates of survival, reproduc-
tion, and movement in and out of the population are responsible for all changes
in abundance of organisms and are thus important topics of study. Similarly,
rates of local extinction and colonization determine the dynamics of species
occurrence across units of the landscape. Ecological investigations of popula-
tion and occupancy dynamics frequently focus on the environmental and eco-
logical variables that induce changes in the relevant vital rates. Conservation
requires actions that influence vital rates in the desired manner. As with esti-
mation of abundance, inference about vital rates requires models that explicitly
incorporate detection probabilities that reflect the incompleteness of virtually
all wildlife sampling.
In this chapter we describe methods that we believe should be useful for
estimating abundance, occupancy, and their respective vital rates in studies of
amphibian populations. Our intention is to provide the reader with information
about how these methods work, as well as an entrée to the scientific literature
that describes these methods in detail.
24.2.2 Capture–mark–recapture
Capture–mark–recapture studies typically involve one to a few study areas or
populations. The duration of the sampling period, the intensity of sampling
within the period, and the length of time between successive sampling periods
can vary substantially depending on study objectives. These features (objectives
and sampling design) are critical components that determine which capture–
mark–recapture methods are most appropriate for a given study. Capturing,
marking, and recapturing individuals within a single day have been used to esti-
mate amphibian abundance at different life-history stages: egg masses (Grant
et al. 2005) and tadpoles (Jung et al. 2002). Alternatively, many studies of
amphibian populations involve capturing individuals during the species’ breed-
ing season over multiple years, where each sampling period may be 2–12 weeks
(Schmidt and Anholt 1999; Fretey et al. 2004; Church et al. 2007).
Data resulting from capture–mark–recapture studies are summarized into
individual capture histories, which are simply rows of ones and zeros indicating
whether an individual was captured (1) or not (0) during each sampling occasion.
For example, a study of terrestrial or stream-side salamanders might sample an area
for four consecutive days yielding individuals with the following capture history:
0 1 1 0. These individuals were first captured and marked on day 2, recaptured on
the third day, but not encountered on the fourth and final day of sampling. If we
assume that the population exposed to sampling does not change over the 4-day
period, then these data can be modeled using closed-population models (closed
450 | Amphibian ecology and conservation
Since all individuals in the exposed population are assumed to be present at the
initial sampling occasion, (1 p1)p2 denotes the probability a salamander was
missed on the first day and captured for the first time on day 2. The probability
of encountering an individual may change after first capture: thus the remain-
der of the expression indicates the probabilities associated with the events of
recapture on day 3, c3, but not on day 4, (1 c4).
We would compile similar probability expressions for each captured sala-
mander. The product of these probabilities over all captured salamanders would
constitute a model (i.e. the likelihood) for the entire data set, and estimates of
the capture and recapture probabilities could be obtained. In this model, the
likelihood is conditioned on the total number of individuals captured during
the entire study, MK 1, where K indexes the final sample occasion. Abundance
N can be estimated as a derived parameter, specifically as:
M K 1
Nˆ
⎡⎣1(1 pˆ1 )(1 pˆ2 ) (1 pˆ3 ) (1 pˆ4 )⎤⎦ (24.1)
The first two terms, 2p3, represent the probability the amphibian survives
between periods 2 and 3 and is recaptured in year 3 (corresponding to the events 1
1 in the capture history). The remainder of the expression [3 (1 p4) (1 3)]
represents two possibilities: (1) the amphibian survived and remained in the study
area from year 3 to 4, but was not captured in year 4, or (2) the amphibian died
24 Abundance estimation and occupancy models | 453
or permanently emigrated from the study area between year 3 and 4. Probability
statements such as this are compiled for each capture history (i.e. each individual)
and combined into a likelihood function. Estimates can be obtained via max-
imum likelihood using computer software such as programs MARK (White and
Burnham 1999) or M-SURGE (Choquet et al. 2005).
Many extensions are available for this sort of modeling. Parameters can be
constrained to be constant over time, they can be modeled as functions of group
(e.g. males and females) membership, and they can be modeled as functions of
time-specific or individual-specific covariates (Lebreton et al. 1992). So-called
unconditional modeling also permits the estimation of time-specific abun-
dance, recruitment, and even population growth rate (see Williams et al. 2002).
Program MARK (White and Burnham 1999) provides user-friendly software
for such modeling.
These models for open populations are capable of dealing with population
gains and losses between sampling periods, but they do not provide a means of
distinguishing movement from other processes. Indeed, under the CJS model,
the processes of emigration and immigration are confounded with survival
and capture probabilities. For example, if individuals permanently emigrate
from the study area, then the complement of ‘apparent survival’ probability
will include both mortality and permanent emigration (e.g. Burnham 1993). In
the case of random temporary emigration, CJS estimates of capture probability
represent the product of the probability that the individual is not a temporary
emigrant and the probability that the individual is caught, conditional on pres-
ence (Kendall et al. 1997; see also Chapter 25 in this volume). Temporary emi-
gration is particularly important in amphibian ecology, as individuals are often
unavailable for capture during specific life-history phases (i.e. post-metamorphic
juveniles or adults that skip breeding opportunities). Movement processes can
be separately estimated in some cases using multistate models (see below) or
models that incorporate extra sources of information (Burnham 1993; Kendall
et al. 1997; Fujiwara and Caswell 2002; Kendall and Nichols 2002; Williams
et al. 2002; Bailey et al. 2004a; Schaub et al. 2004).
Multistate models are open-population capture–mark–recapture models that
can be used when animals are categorized by some state variable that can be
assessed each time an animal is encountered (Arnason 1972, 1973; Brownie et al.
1993; Schwarz et al. 1993). When location or study site is the state, direct infer-
ences about movement among sampled locations are possible (e.g. Hestbeck et al.
1991; Brownie et al. 1993). In addition to location, state variables can include
characteristics of the captured animals themselves. For example, size class is a
reasonable state variable in some amphibian studies (Wood et al. 1998).
454 | Amphibian ecology and conservation
Multistate models can also be used when one or more of the states may be
unobservable. Such uses focus on situations in which animals move back and
forth between study sites, where capture and resighting efforts are made, and
other locations at which no sampling effort is expended. For example, when
amphibians are sampled at breeding sites, non-breeding animals are tempor-
ary emigrants and thus unobservable. In the simple two-state case, the model
contains parameters that are state-specific, r, where r O denotes the observ-
able or breeding state, and r U denotes the unobservable or non-breeding
o
state. Parameters include: pt , the aforementioned capture probability which is
non-zero for breeders only ( ptU 0, by definition); Str , the probability that an
individual in state r at time t survives until time t 1; and ct , the probability
rs
Pr (0 11 0 1 | released in year 2 )S2O c2OO p3O S3O ⎡⎣(c3OO (1 p4O )S4O c4OO c3OU S4U cUO
4 )⎤
O
⎦ p5
The fact that the individual was not captured in year 4 could result from one of
two processes: either the amphibian was in the study area but not captured, or
the amphibian was not in the study (or breeding) area in year 4, but returned to
the area in year 5. Note that the above parameterization assumes that survival
between sampling periods t and t 1 depends on state at occasion t, e.g. Str .
For situations in which this assumption is not appropriate, it is possible to com-
bine the survival and transition parameters as ft rs Str ctrs and to estimate this
survival-transition parameter ( ft rs ) directly.
Parameters of these multistate models can be estimated using maximum
likelihood methods and modeled as a function of relevant covariates. Useful
recent software includes programs MARK (White and Burnham 1999) and
M-SURGE (Choquet et al. 2005). Practitioners using these models should ver-
ify that all parameters are uniquely identifiable as there are some cases where
multiple parameter values will yield the same maximized likelihood values
(Kendall and Nichols 2002; Gimenez et al. 2004; Schaub et al. 2004; Bailey
et al. 2009; Hunter and Caswell 2009). Simplifying assumptions, such as
equating parameters over time or state, may be necessary (Kendall and Nichols
2002; Schaub et al. 2004; Bailey et al. 2009; Hunter and Caswell 2009), but the
models offer tremendous flexibility and the ability to estimate important demo-
graphic parameters that have long eluded researchers (Biek et al. 2002).
24 Abundance estimation and occupancy models | 455
The original robust design (Pollock 1982; Kendall et al. 1997) combines open
and closed models and involves sampling at two temporal scales: each primary
period (e.g. year) consists of two or more secondary occasions that are closely
spaced in time, such that the population may be considered demographically
and geographically closed. For example, during the breeding season individuals
at a pool may be sampled multiple times over the course of a few days. Each inde-
pendent sample of the breeding population would be considered a secondary
occasion, and this process would be repeated for several years (primary periods),
over which time survival, breeding and/or movement probabilities could be esti-
mated (see Kendall 2004 for review of robust design models). The robust design
typically yields improved precision of vital rate parameter estimates, but can
also be used to estimate period- (year-) specific abundance, while accounting
for modeled or unmodeled heterogeneity in capture probabilities. The robust
design also permits direct estimation of probabilities of temporary emigration
(Kendall et al. 1997) and of the separate contributions of recruitment from
immigration and in situ reproduction (Nichols and Pollock 1990; Nichols et al.
2000; see Schmidt et al. 2005 for an amphibian example).
Pr (0 1 1 0)c(1 p1 ) p2 p3 (1 p4 )
24 Abundance estimation and occupancy models | 457
The site is clearly occupied by the target species, detected during the second and
third visit, but not detected during the first and fourth visit. A detection history
consisting of all zeros, 0 0 0 0, would have two possible explanations: either the
site was occupied, but the species was not detected during any visit, or the site
was unoccupied. Written as a mathematical expression:
Again, the detection data and the corresponding probability model are com-
bined to form a likelihood function, and estimates can be obtained via software
such as program MARK (White and Burnham 1999) or program PRESENCE
(MacKenzie et al. 2006; www.mbr-pwrc.usgs.gov/software/).
An alternative approach to estimation and modeling is to view the above
models hierarchically. One model component concerns the true spatial process
of species occurrence across the sampled sites. Conditional on this true spatial
process, a sampling component models the detection process. These hierarch-
ical components (process and sampling) are combined under a Bayesian frame-
work, and statistical inference is achieved using Markov chain Monte Carlo
(MCMC) methods (Royle and Dorazio 2008). Using either procedure, it is
possible to model occupancy and detection probability as functions of measured
covariates. Models incorporating different combinations of covariates represent
competing hypotheses about factors believed to influence amphibian distribu-
tion or probability of detection. These occupancy models have gained popular-
ity in amphibian studies around the globe (e.g. Bailey et al. 2004c; Royle 2004;
Mazerolle et al. 2005; Muths et al. 2005; Pellet and Schmidt 2005; Royle and
Link 2005) and represent an important improvement over logistic regression
models of amphibian–habitat relationships that ignore imperfect detection (Gu
and Swihart 2004; MacKenzie et al. 2006).
24.5 Disclaimer
Any use of trade, product, or firm names is for descriptive purposes only and
does not imply endorsement by the US Government.
24.6 References
Arnason, A. N. (1972). Parameter estimates from mark-recapture experiments on two pop-
ulations subject to migration and death. Researches on Population Ecology, 13, 97–113.
Arnason, A. N. (1973). The estimation of population size, migration rates, and survival
in a stratified population. Researches on Population Ecology, 15, 1–8.
Bailey, L. L., Kendall, W. L., Church, D. R., and Wilbur, H. M. (2004a). Estimating
survival and breeding probability for pond-breeding amphibians: A modified robust
design. Ecology, 84, 2456–66.
Bailey, L. L., Simons, T. R., and Pollock, K. H. (2004b). Comparing population size
estimators for plethodontid salamanders. Journal of Herpetology, 38, 370–80.
Bailey, L. L., Simons, T. R., and Pollock, K. H. (2004c). Estimating site occupancy
and species detection probability parameters for terrestrial salamanders. Ecological
Applications, 14, 692–702.
460 | Amphibian ecology and conservation
Bailey, L. L., Hines, J. E., Nichols, J. D., and MacKenzie, D. I. (2007). Sampling design
trade-offs in occupancy studies with imperfect detection: examples and software.
Ecological Applications 17, 281–90.
Bailey, L. L., Kendall, W. L., and Church, D. R. (2009). Exploring extensions to multi-
state models with multiple unobservable states. In D. L. Thomson, E. G. Cooch, and
M. C. Conroy (eds), Modeling Demographic Processes in Marked Populations, pp. 693–
710. Springer Science+Business Media, New York.
Biek, R., Funk, W. C., Maxell, B. A., and Mills, L. S. (2002). What is missing in amphib-
ian decline research: Insights from ecological sensitivity analysis. Conservation Biology,
16, 728–34.
Brownie, C., Hines, J. E., Nichols, J D., Pollock, K. H., and Hestbeck, J. B. (1993).
Capture-recapture studies for multiple strata including non-Markovian transitions.
Biometrics, 49, 1173–87.
Bruce, R. C. (1995). The use of temporary removal sampling in a study of population
dynamics of the salamander Desmognathus monticola. Australian Journal of Ecology,
20, 403–12.
Burnham, K. P. (1993). A theory for combined analysis of ring recovery and recapture
data. In J. D. Lebreton and P. M. North (eds), Marked Individuals in the Study of Bird
Populations, pp. 199–213. Birkhauser-Verlag, Basel.
Burnham, K. P. and Anderson, D. R. (2002). Model Selection and Multimodel Inference.
Springer-Verlag, New York.
Chao, A. and Huggins, R. M. (2005). Modern closed-population capture-recapture
models. In S. C. Amstrup, T. L. McDonald, and B. F. J. Manly (eds), Handbook of
Capture-Recapture Analysis, pp. 58–87. Princeton University Press, Princeton, NJ.
Choquet, R., Reboulet, A. M., Pradel R., Gimenez, O., and Lebreton, J. D. 2005.
M-SURGE 1–8 User’s Manual. CEFE, Montepellier. ftp.cefe.cnrs.fr/biom/soft-cr/.
Church, D. R., Bailey, L. L., Wilbur, H. M., Kendall, W. L., and Hines, J. E. (2007).
Iteroparity in the variable environment of the salamander Ambystoma tigrinum. Ecology,
88, 891–903.
Cochran, W. G. (1977). Sampling Techniques. Wiley, New York.
Cormack, R. M. (1964). Estimates of survival from the sighting of marked animals.
Biometrika, 51, 429–38.
Devineau, O., Choquet, R., and Lebreton, J. D. (2006). Planning capture-recapture stud-
ies: straightforward precision, bias, and power calculations. Wildlife Society Bulletin,
34, 1028–35.
Elton, C. (1927). Animal Ecology. Sidgwick & Jackson, London.
Fretey, T., Cam, E., Le Garff, B., and Monnat, J. Y. (2004). Adult survival and temporary
emigration in the common toad. Canadian Journal of Zoology, 82, 859–72.
Fujiwara, M. and Caswell, H. (2002). A general approach to temporary emigration in
mark-recapture analysis. Ecology, 83, 3266–75.
Gardenfors, U., Hilton-Taylor, C., Mace, G. M., and Rodriguez, J. P. (2001). The appli-
cation of IUCN red list criteria at regional levels. Conservation Biology, 15, 1206–12.
Gimenez, O., Viallefont, A., Catchpole, E. A., Choquet, R., and Morgan, B. J. T.
(2004). Methods for investigating parameter redundancy. Animal Biodiversity and
Conservation, 27, 561–72.
24 Abundance estimation and occupancy models | 461
Grant, E. H. C., Jung, R. E., Nichols, J. D., and Hines, J. E. (2005). Double-observer
approach to estimating egg mass abundance of pool-breeding amphibians. Wetlands
Ecology and Management, 13, 305–20.
Gu, W. and Swihart, R. K. (2004). Absent or undetected? Effects of non-detection of spe-
cies occurrence on wildlife-habitat models. Biological Conservation, 116, 195–203.
Hestbeck, J. B., Nichols, J. D., and Malecki, R A. (1991). Estimates of movement and site
fidelity using mark resight data of wintering Canada geese. Ecology, 72, 523–33.
Huggins, R. M. (1989). On the statistical analysis of capture experiments. Biometrika,
76, 133–40.
Huggins, R. M. (1991). Some practical aspects of conditional likelihood approach to cap-
ture experiments. Biometrics, 47, 725–32.
Hunter, C. M. and Caswell, H. (2009). Rank and redundancy of multistate mark-
recapture models for seabird populations with unobservable states. In D. L. Thomson,
M. C. Conroy, and E. G. Cooch (eds), Modeling Demographic Processes in Marked
Populations, pp. 799–828. Springer Science+Business Media, New York.
Jolly, G. M. (1965). Explicit estimates from capture-recapture data with both death and
immigration-Stochastic model. Biometrika, 52, 225–47.
Jung, R. E., Dayton, G. H., Williamson, S. J., Sauer, J. R., and Droege, S. (2002). An
evaluation of population index and estimation techniques for tadpoles in desert pools.
Journal of Herpetology, 36, 465–72.
Jung, R. E., Royle, J. A., Sauer, J. R., Addison, C., Rau, R. D., Shirk, J. L., and Whissel, J. C.
(2005). Estimation of stream salamander (Plethodontidae, Desmognathinae and
Plethodontinae) populations in Shenandoah National Park, Virginia, USA. Alytes, 22,
72–84.
Kendall, W. L. (2004). Coping with unobservable and mis-classification state in capture-
recapture studies. Animal Biodiversity and Conservation, 27, 97–107.
Kendall, W. L. and Nichols, J. D. (2002). Estimating state-transition probabilities for
unobservable states using capture-recapture/resighting data. Ecology, 83, 3276–84.
Kendall, W. L., Nichols, J. D., and Hines, J. E. (1997). Estimating temporary emigration
using capture-recapture data with Pollock’s robust design. Ecology, 78, 563–78.
Krebs, C. J. 1972. Ecology. Harper and Row, New York.
Lancia, R. A., Kendall, W. L., Pollock, K. H., and Nichols, J. D. (2005). Estimating
the number of animals in wildlife populations. In C. E. Braun (ed.), Techniques for
Wildlife Investigations and Management, 6th edn, pp. 106–53. The Wildlife Society,
Bethesda, MD.
Lebreton, J. D., Burnham, K. P., Clobert, J., and Anderson, D. R. (1992). Modeling
survival and testing biological hypotheses using marked animals—a unified approach
with case-studies. Ecological Monographs, 62, 67–118.
MacKenzie, D. I., Nichols, J. D., Lachman, G. B., Droege, S., Royle, J. A., and Langtimm,
C. A. (2002). Estimating site occupancy rates when detection probabilities are less than
one. Ecology, 83, 2248–55.
MacKenzie, D. I., Nichols, J. D., Hines, J. E., Knutson, M. G., and Franklin, A. B. (2003).
Estimating site occupancy, colonization and local extinction probabilities when species
are not detected with certainty. Ecology, 84, 2200–7.
462 | Amphibian ecology and conservation
MacKenzie, D. I., Nichols, J. D., Royle, J. A., Pollock, K. H., Bailey, L. L., and Hines, J. E.
(2006). Occupancy Estimation and Modeling: Inferring Patterns and Dynamics of Species
Occurrence. Academic Press, Boston, MA.
Mazerolle, M. J., Desrochers, A., and Rochefort, L. (2005). Landscape characteris-
tics influence pond occupancy by frogs after accounting for detectability. Ecological
Applications, 15, 824–34.
Muths, E., Jung, R. E., Bailey, L. L., Adams, M. J., Corn, P. S., Dodd, Jr, C. K., Fellers,
G. M., Sadinski, W. J., Schwalbe, C. R., Walls, S. C. et al. (2005). Amphibian Research
and Monitoring Initiative (ARMI): a successful start to a national program in the
United States. Applied Herpetology, 2, 355–71.
Nichols, J. D. and Pollock, K. H. (1990). Estimation of recruitment from immigration
versus in situ reproduction using Pollock’s robust design. Ecology, 71, 21–6.
Nichols, J. D., Hines, J. E., Lebreton, J.-D., and Pradel, R. (2000). Estimation of contri-
butions to population growth: a reverse-time capture-recapture approach. Ecology, 81,
3362–76.
Norris, J. L. and Pollock, K. H. (1996). Nonparametric MLE under two closed capture
recapture models with heterogeneity. Biometrics, 52, 639–49.
Otis, D. L., Burnham, K. P., White, G. C., and Anderson, D. R. (1978). Statistical-
inference from capture data on closed animal populations. Wildlife Monographs, 62,
1–135.
Pellet, J. and Schmidt, B. R. (2005). Monitoring distributions using call surveys: esti-
mating site occupancy, detection probabilities and inferring absence. Biological
Conservation, 123, 27–35.
Pledger, S. (2000). Unified maximum likelihood estimates for closed capture- recapture
models using mixtures. Biometrics, 56, 434–42.
Pollock, K. H. (1982). A capture-recapture design robust to unequal probability of cap-
ture Journal of Wildlife Management, 46, 757–60.
Pollock, K. H., Nichols, J. D., Simons, T. R., Farnsworth, G. L., Bailey, L. L., and
Sauer, J. R. (2002). Large scale wildlife monitoring studies: statistical methods for
design and analysis. Environmetrics, 13, 105–19.
Rexstad, E. and Burnham, K. P. (1991). User’s Guide for Interactive Program CAPTURE.
www.mbr-pwrc.usgs.gov/software/. Colorado Cooperative Fish and Wildlife Unit,
Fort Collins, CO.
Royle, J. A. (2004). Modeling abundance index data from anuran calling surveys.
Conservation Biology, 18, 1378–85.
Royle, J. A. and Link, W. A. (2005). A general class of multinomial mixture models for
anuran calling survey data. Ecology, 86, 2505–12.
Royle, J. A. and Dorazio, R. M. (2008). Hierarchical Modelling and Inference in Ecology.
Academic Press, San Diego, CA.
Schaub, M., Gimenez, O., Schmidt, B. R., and Pradel, R. (2004). Estimating survival
and temporary emigration in the multistate capture-recapture framework. Ecology, 85,
2107–13.
Schmidt, B. R. and Anholt, B R. (1999). Analysis of survival probabilities of female com-
mon toads, Bufo bufo. Amphibia-Reptilia, 20, 97–108.
24 Abundance estimation and occupancy models | 463
Schmidt, B. R., Feldmann, R., and Schaub, M. (2005). Demographic processes under-
lying population growth and decline in Salamandra salamandra. Conservation Biology,
19, 1149–56.
Schwarz, C. J., Schweigert, J. F., and Arnason, A. N. (1993). Estimating migration rates
using tag-recovery data. Biometrics, 49, 177–93.
Seber, G. A. F. (1965). A note on the multiple-recapture census. Biometrika, 52, 249–59.
Seber, G. A. F. (1982). The Estimation of Animal Abundance and Related Parameters.
Macmillian, New York.
Stanley, T. R. and Burnham, K. P. (1998) Estimator selection for closed-population
capture-recapture. Journal of Agricultural Biological and Environmental Statistics, 3,
131–50.
Stanley, T. R. and Burnham, K. P. (1999). A closure test for time-specific capture-recapture
data. Environmental and Ecological Statistics, 6, 197–209.
White, G. C. and Burnham, K. P. (1999). Program MARK: survival estimation from
populations of marked animals. Bird Study, 46, 120–39.
White, G. C., Anderson, D. R., Burnham, K. P., and Otis, D. L. (1982). Capture-recapture
and Removal Methods for Sampling Closed Populations. Los Alamos National Lab, Los
Alamos, NM.
Williams, B. K., Nichols, J. D., and Conroy, M. J. (2002). Analysis and Management of
Animal Populations. Academic Press, San Diego, CA.
Wood, K. V., Nichols, J. D., Percival, H. F., and Hines, J. E. (1998). Size-sex variation in
survival rates and abundance of pig frogs, Rana grylio, in northern Florida wetlands.
Journal of Herpetology, 32, 527–35.
This page intentionally left blank
25
Quantifying abundance: counts,
detection probabilities, and estimates
Benedikt R. Schmidt and Jérôme Pellet
E (C ) Np
where N is the true value of the parameter of interest (i.e. number of individuals,
density, number of populations in an area, or species richness) and p is the detec-
tion probability (Gill 1985; Yoccoz et al. 2001; Pollock et al. 2002; Schmidt
2004). E( ) denotes a statistical expectation. The expectation E(C) is the average
of the count C over repeated realizations of the sampling process. This equation
has three major implications that we discuss below.
Sampling a population should be viewed as a stochastic process because it
involves a probability of detection. This is why there is a statistical expect-
ation E(C) in the above equation. Even under identical conditions we should
not expect to obtain the same result if we sample the same population mul-
tiple times. We should therefore expect variability in the counts. Variability in
counts (C) does not imply variation in abundance (N) or detection probability
(p); it can simply be random variation. Technically, counts are random varia-
bles. This is illustrated in Figure 25.1. The figure shows that, as expected under
25
p = 0.2 p = 0.5 p = 0.8
20
Frequency
15
10
0
0 5 10 15 20 0 5 10 15 20 0 5 10 15 20
Count
Fig. 25.1 Expected variation in counts of individuals under identical conditions for
100 repeated counts, a true population size N 20 and detection probabilities
p 0.2, 0.5, and 0.8. Data were simulated using the R code rbinom(n 100,
size 20, p x), where x had the values 0.2, 0.5, and 0.8, respectively).
25 Quantifying abundance | 467
Table 25.1 Results of linear regressions between counts (C), population size esti-
mates (N̂), and censuses for tadpoles of two anuran species. R2 and F tests were
calculated using PROC GLM in SAS. Asterisks indicate significance at 0.05. The
Lincoln–Peterson estimator was used to estimate abundance. Table adapted from
Schmidt (2004). Data are from Jung et al. (2002).
Species Studied in Intercept Slope R2 F
Table 25.2 Counts and estimates of male European tree frogs (Hyla arborea) in
two breeding aggregations. Maximum and mean chorus size, as well as number of males
captured, correlate only weakly with actual (estimated) male population size, demon-
strating the effects of imperfect detection on population size estimates. Adapted from
Pellet et al. (2007).
Year Maximum Mean Total males Modeled male Proportion of
chorus size chorus captured population calling males
size size SE
pc 1(1 p )n
25 Quantifying abundance | 469
_
where p is the average per-visit detection probability and n is the number of
visits or capture events. Per-visit (i.e. per day) detection probabilities of the frog
Colostethus stepheni in the study of Funk and Mills (2003) ranged from 0 to 1
with a mean of 0.58 (W.C. Funk, personal communication). Because Funk and
Mills (2003) had six capture events during a short time period when the popu-
lation did not change, the cumulative capture probabilities were greater than
0.987. This implies that Funk and Mills (2003) achieved an almost complete
census of the frogs and suggests that multiple capture events are an obvious way
to deal with imperfect detection because cumulative detection probabilities can
be quite high.
E (T )C1 /C 2 N 1 p1 /N 2 p2
capture salamanders (Ambystoma tigrinum) on their way to and from the pond.
Detection probabilities were high in most years (greater than 90%) but inexplicably
lower in one year (76%). Although detection probabilities of egg masses of Rana
sylvatica and Ambystoma maculatum were generally high (usually 80%), Grant
et al. (2005) documented substantial spatial and temporal variation in detection
probabilities that could bias estimates of population trends based on unadjusted
counts.
E (C ) Npe
are placed along the edge of a pond and some individuals spend all their time
near the center of the pond. Non-exposure to sampling may also result from
breeding phenology. Some individuals may breed early and some may breed
later in the season (e.g. Sinsch 1988). If the population is sampled only early in
the season, then many individuals will not be exposed to sampling. This might
be a case where both biology and method cause non-exposure to sampling.
Non-exposure to sampling can have profound consequences for abundance
estimation (Kendall 1999). Imagine that the goal of a study is to estimate abun-
dance in a particular area. If amphibians do not move in and out of the study
area, then an abundance estimator such as the Lincoln–Peterson estimator will
provide an unbiased estimate the number of amphibians in the area. However,
if some individuals move randomly in and out of the area (i.e. if they are not
always exposed to sampling), then the very same estimator will estimate a differ-
ent quantity. It will now estimate the size of the superpopulation where the super-
population is defined as the total number of amphibians exposed to sampling at
least once. This includes all amphibians that are residents in the study area, but
also all individuals that move in and out or that move through the study area.
Bailey et al. (2004b) encountered a related problem in their study of salamanders
in the Appalachian Mountains, USA. They argued that short-term studies where
salamander movement was negligible yielded estimates of the “surface” popu-
lation (i.e. salamanders exposed to sampling only). Long-term studies, where
salamanders had time to move from the surface to deeper ground and vice versa,
gave estimates of the total number of salamanders in the sampled area. Kinkead
and Otis (2007) describe a similar situation with breeding and non-breeding
ambystomatid salamanders that were sampled at the breeding site.
If many individuals are not exposed to sampling (low e), then the mismatch
between the spatial or temporal scale at which a population is sampled and the
desired temporal or spatial scale of inference is likely to be great. In conclusion,
study design (both the spatial and the temporal scales) and species behavior can
jointly determine which biological entity is quantified.
Nˆ C / pˆ
472 | Amphibian ecology and conservation
This equation is the conceptual basis for all kinds of abundance estimators, be
they mark–recapture, distance sampling, point count, removal, or other meth-
ods (Williams et al. 2002; Mazerolle et al. 2007; see also Chapter 24 in this
volume). When non-exposure to sampling is a problem, then N̂ can only be
estimated if one knows the fraction of the population that is exposed to sam-
pling as well as the probability of detecting exposed individuals, populations,
or species:
Nˆ C / pe
ˆˆ
increase the number of repeat visits such that cumulative detection probabilities
are increased. However, standard errors and confidence intervals are not a nuis-
ance. Rather, they are an advantage of estimation methods. Standard errors and
confidence intervals are a measure of uncertainty that allow an assessment of
the estimates’ reliability. Consequently, we view wide confidence intervals as an
honest statement whether a particular estimate is or is not particularly reliable
and useful. Wide confidence intervals are no reason to discard mark–recapture
estimates and to prefer the simple counts (C; e.g. Alford and Richards 1999).
There is also uncertainty associated with counts (because detection probability
p is unknown) but it is not explicit and is, in fact, unknowable.
Heterogeneity in detection probabilities among individuals can be a problem
in mark–recapture studies (Link 2003). Heterogeneity usually leads to nega-
tive bias in abundance estimates and in the worst case it may not be possible to
identify a best model that should be used for inference. We believe that amphib-
ian ecologists and conservationists should attempt to minimize detection prob-
ability heterogeneity among individuals by adopting methods that account
for variation in detectability among individuals (i.e. grouping individuals into
homogeneous sets by sexes, colour morphs, age classes).
temporal trend in detection probability (Bart et al. 2004; tacitly, this is also the
reasoning of Hairston and Wiley 1993). In such a situation, variation in detec-
tion probability likely causes extra variability (sampling variability) in the counts
(in comparison to variation in absolute abundance which is the phenomenon of
biological interest). Detectability-induced extra variation in the counts means
that a monitoring program loses power to detect temporal trends. However, one
should keep in mind that detection probabilities likely show temporal trends.
Reasons include, but are not limited to, habitat succession or changes through
time in the ability of the observer to detect the study species (Link and Sauer
1997).
One commonly held view is that field methods can be standardized to the
extent where detection probability is constant. If this is the case, then counts
or other estimates of relative abundance should serve well as proxies for abso-
lute abundance. Unfortunately, variation in detection probabilities is the rule
rather than the exception (MacKenzie and Kendall 2002). Whenever detec-
tion probabilities of amphibians have been estimated, they were found to be
variable both within and across seasons (Bailey et al. 2004a, Kinkead and Otis
2007; Mazerolle et al. 2007; Pellet et al. 2007; Schmidt et al. 2007). This
was the case even when researchers used standard(ized) methods; even when
drift fences were used—where the assumption is that detection probability is
1—there was variation in detection probability (Bailey et al. 2004a). Pellet
et al. (2007) used the same methods at two sites yet detection probabilities dif-
fered between sites by a factor of approximately two. Hyde and Simons (2001)
showed that counts obtained from applying four standard methods gave results
that were only weakly correlated. That is, the use of standardized methods does
not guarantee that detection probabilities are constant. We believe that stand-
ardization of field methods is important because it can help to keep variation
in detection probabilities within bounds, but it should certainly not be viewed
as a panacea.
Standardization of field methods is one solution to limit variation in detec-
tion probabilities. Another solution is to measure covariates that may affect
detection probabilities and use these covariates at the analysis stage to adjust
counts (Link and Sauer 1997, 1998). This approach may work well as long as the
important covariates are known and has been successfully used in large-scale
bird monitoring programs. However, it may be that the effect of a covariate on
the counts varies from one site to the next. Lauber (2004) counted alpine sala-
manders (Salamandra atra) along fixed transects at four sites and tested whether
weather covariates could be used to predict the salamander counts. An analysis
of covariance found no main effect of air humidity on counts but there was
25 Quantifying abundance | 475
100
Study sites:
Cholerenschlucht
Weissenburgbad
80 Rüschegg-Heubach
Salamander count Grünenbergpass
60
40
20
0
60 70 80 90 100
Air humidity (%)
25.6 Software
There are many computer programs freely available to estimate population
abundance while incorporating detection probability. The most versatile and
widely used of them is program MARK (White and Burnham 1999). This
software, available at www.phidot.org/, allows the analysis of a wide range of
capture–recapture-based data sets. Every new version incorporates the latest
development in capture–recapture and thus allows the user to choose from a
wide range of models the one that will fit its data best. As the name implies,
DISTANCE (www.ruwpa.st and.ac.uk/ distance/) is the software tool that
allows one to design and analyse distance sampling surveys. More recent devel-
opments have been integrated in statistical software such as R and WinBugs.
These tools have the inconvenience of being less user-friendly than the previously
listed programs, which benefit from a graphical user interface. There is soft-
ware that can be used when planning a mark–recapture study (Devineau et al.
2006; Zucchini et al. 2007).
25 Quantifying abundance | 477
25.7 Outlook
Imperfect detection is the feature that the vast majority of all amphibian surveys
have in common. Complete censuses where all amphibians that are present at
the study site are captured are not impossible, but require a lot of work (drift
fences: Bailey et al. 2004a; many capture events: Funk and Mills 2003; Pellet
et al. 2007). We argue that detection probabilities can be low and highly variable
among years and/or sites. While counts that are not adjusted for imperfect detec-
tion can certainly indicate negative population trends (Laurance et al. 1996),
variability in imperfect detection can seriously bias inference from surveys.
Amphibian ecologists and conservationists should therefore estimate detection
probabilities as the best tool to calibrate a survey and use robust methods for
abundance estimation (Williams et al. 2002; Chapter 24). Unfortunately, the
use of such methods is not yet widespread (Alford and Richards 1999).
The number of methods available for estimation of abundance that account
for imperfect detection has increased tremendously in the recent past. Existing
methods are constantly being refined, while new methods are being developed
(e.g. Royle 2004; Royle and Dorazio 2006). Still, all methods need to be used
with care as sampling design, the behavior of the species, and the estimator used
all determine which biological quantity is being estimated. Notwithstanding,
the quality of inference from methods that adjust for imperfect detection will be
stronger than inference from any other kind of method.
In the future, we ought to be able to estimate abundance with a precision and
freedom from bias that was not achievable in the past. We should now be able
to determine which factors influence abundance rather than study patterns of
an inseparable combination of abundance, detectability, and exposure to sam-
pling (such as counts). This will help us gain new insights into fundamental and
applied aspects of amphibian ecology and conservation.
25.8 References
Alford, R. A. and Richards, S. J. (1999). Global amphibian declines: a problem in applied
ecology. Annual Reviews of Ecology and Systematics, 30, 133–65.
Bailey, L. L., Kendall, W. L., Church D. R., and Wilbur, H. M. (2004a). Estimating
survival and breeding probability for pond-breeding amphibians: a modified robust
design. Ecology, 85, 2456–66.
Bailey, L. L., Simons, T. R., and Pollock, K. H. (2004b) Comparing population size esti-
mators for plethodontid salamanders. Journal of Herpetology, 38, 370–80.
Bart, J., Droege, S., Geissler, P., Peterjohn, B., and Ralph, C. J. (2004). Density estima-
tion in wildlife surveys. Wildlife Society Bulletin, 32, 1242–7.
478 | Amphibian ecology and conservation
Condit, R., Le Boeuf, B. J., Morris, P. A., and Sylvan, M. (2007). Estimating population
size in asynchronous aggregations: A Bayesian approach and test with elephant seal
censuses. Marine Mammal Science, 23, 834–55.
Devineau, O., Choquet, R., and Lebreton, J. D. (2006). Planning capture-recapture stud-
ies: straightforward precision, bias, and power calculations. Wildlife Society Bulletin,
34, 1028–35.
Friedl, T. W. P. and Klump, G. M. (2005). Sexual selection in the lek-breeding European
treefrog: body size, chorus attendance, random mating and good genes. Animal
Behaviour, 70, 1141–54.
Funk, W. C. and Mills, L. S. (2003). Potential causes of population declines in forest frag-
ments in an Amazonian frog. Biological Conservation, 111, 205–14.
Funk, W. C., Almeida-Reinoso, D., Nogales-Sornosa, F., and Bustamante, M. R. (2003).
Monitoring population trends of Eleutherodactylus frogs. Journal of Herpetology, 37,
245–56.
Gill, D. E. (1985). Interpreting breeding patterns from census data: a solution to the
Husting dilemma. Ecology, 66, 344–54.
Grant, E. H. C., Jung, R. E., Nichols, J. D., and Hines, J. E. (2005). Double-observer
approach to estimating egg mass abundance of pool-breeding amphibians. Wetlands
Ecology and Management, 13, 305–20.
Hairston, Sr, N. G. and Wiley, R. H. (1993). No decline in salamander (Amphibia: Caudata)
populations: a twenty-year study in the southern Appalachians. Brimleyana, 18, 59–64.
Hyde, E. J. and Simons, T. R. (2001). Sampling plethodontid salamanders: sources of
variability. Journal of Wildlife Management, 65, 624–32.
Jung, R. E., Dayton, G. H., Williamson, S. J., Sauer, J. R., and Droege, S. (2002). An
evaluation of population index and estimation techniques for tadpoles in desert pools.
Journal of Herpetology, 36, 465–72.
Kendall W. L. (1999). Robustness of closed capture-recapture methods to violations of
the closure assumption. Ecology, 80, 2517–25.
Kinkead, K. E. and Otis, D. L. (2007). Estimating superpopulation size and annual prob-
ability of breeding for pond breeding salamanders. Herpetologica, 63, 151–62.
Lauber, A. (2004). Methodenevaluation zum Monitoring der Alpensalamanderpopulation.
Diploma thesis, Eidgenössische Technische Hochschule Zürich (ETHZ), Zürich.
Laurance, W. F., McDonald, K. R., and Speare, R. (1996). Epidemic disease and the cata-
strophic decline of Australian rain forest frogs. Conservation Biology, 10, 406–13.
Link, W. A. (2003). Nonidentifiability of population size from capture-recapture data
with heterogeneous detection probabilities. Biometrics, 59, 1123–30.
Link, W. A. and Sauer, J. R. (1997). Estimation of population trajectories from count
data. Biometrics, 53, 488–97.
Link, W. A. and Sauer J. R. (1998). Estimating population change from count data:
Application to the North American Breeding Bird Survey. Ecological Applications, 8,
258–68.
MacKenzie, D. I. and Kendall, W. L. (2002). How should detection probability be incor-
porated into estimates of relative abundance? Ecology, 83, 2387–93.
Mazerolle, M. J., Desrochers, A., and Rochefort, L. (2005). Landscape characteris-
tics influence pond occupancy by frogs after accounting for detectability. Ecological
Applications, 15, 824–34.
25 Quantifying abundance | 479
Mazerolle, M. J., Bailey, L. L., Kendall, W. L., Royle, J. A., Converse, S. J., and Nichols, J. D.
(2007). Making great leaps forward: accounting for detectability in herpetological field
studies. Journal of Herpetology, 41, 672–89.
Nichols, J. D., Thomas, L., and Conn, P. B. (2008). Inferences about landbird abun-
dance from count data: recent advances and future directions. In D. L. Thomson,
E. G. Cooch, G. Evan, and M. J. Conroy (eds), Modeling Demographic Processes in
Marked Populations, pp. 203–38. Springer ScienceBusiness Media, New York.
Pechmann, J. H. K., Scott, D. E., Semlitsch, R. D., Caldwell, J. P., Vitt, L. J., and
Gibbons, J. W. (1991). Declining amphibian populations: the problem of separating
human impacts from natural fluctuations. Science, 253, 892–5.
Pellet, J., Helfer, V., and Yannic, G. (2007). Estimating population size in the European
tree frog (Hyla arborea) using individual recognition and chorus counts. Amphibia-
Reptilia, 28, 287–94.
Pollock, K. H., Nichols, J. D., Simons, T. R., Farnsworth, G. L., Bailey, L. L., and
Sauer, J. R. (2002). Large scale wildlife monitoring studies: statistical methods for
design and analysis. Environmetrics, 13, 105–19.
Pollock, K. H., Marsh, H., Bailey, L. L., Farnsworth, G. L., Simons, T. R., and
Alldredge, M. W. (2004). Separating components of detection probability in abun-
dance estimation: an overview with diverse examples. In W. L. Thompson (ed.),
Sampling Rare or Elusive Species, pp. 43–58. Island Press, Washington DC.
Royle, J. A. (2004). N-mixture models for estimating population size from spatially rep-
licated counts. Biometrics, 60, 108–15.
Royle, J. A. and Dorazio, R. M. (2006). Hierarchical models of animal abundance
and occurrence. Journal of Agricultural, Biological, and Environmental Statistics, 11,
249–63.
Schmidt, B. R. (2004). Declining amphibian populations: the pitfalls of count data in
the study of diversity, distributions, dynamics and demography. Herpetological Journal,
14, 167–74.
Schmidt, B. R. and Pellet, J. (2005). Relative importance of population processes and
habitat characteristics in determining site occupancy of two anurans. Journal of Wildlife
Management, 69, 884–93.
Schmidt, B. R., Schaub, M., and Steinfartz, S. (2007). Apparent survival of the sala-
mander Salamandra salamandra is low because of high migratory activity. Frontiers in
Zoology, 4, e19.
Sinsch, U. (1988). Temporal spacing of breeding activity in the natterjack toad, Bufo
calamita. Oecologia, 76, 399–407.
White, G. C. and Burnham, K. P. (1999). Program MARK: survival estimation from
populations of marked animals. Bird Study, 46, 120–38.
Williams, B. K., Nichols, J. D., and Conroy, M. J. (2002). Analysis and Management of
Animal Populations. Academic Press, San Diego, CA.
Yoccoz, N. G., Nichols, J. D., and Boulinier, T. (2001). Monitoring of biological diversity
in space and time. Trends in Ecology and Evolution, 16, 446–53.
Zucchini, W., Borchers, D. L., Erdelmeier, M., Rexstad, E., and Bishop, J. (2007). WiSP
1.2.4. Institut für Statistik und Ökonometrie, Georg-August-Universität Göttingen,
Göttingen.
This page intentionally left blank
26
Disease monitoring and biosecurity
D. Earl Green, Matthew J. Gray, and Debra L. Miller
26.1 Introduction
Understanding and detecting diseases of amphibians has become vitally import-
ant in conservation and ecological studies in the twenty-first century. Disease is
defined as the deviance from normal conditions in an organism. The etiologies
(causes) of disease include infectious, toxic, traumatic, metabolic, and neoplas-
tic agents. Thus, monitoring disease in nature can be complex. For amphib-
ians, infectious, parasitic, and toxic etiologies have gained the most notoriety.
Amphibian diseases have been linked to declining amphibian populations, are
a constant threat to endangered species, and are frequently a hazard in captive
breeding programs, translocations, and repatriations. For example, a group of
viruses belonging to the genus Ranavirus and the fungus Batrachochytrium den-
drobatidis are amphibian pathogens that are globally distributed and responsible
for catastrophic population die-offs, with B. dendrobatidis causing known spe-
cies extinctions (Daszak et al. 1999; Lips et al. 2006; Skerratt et al. 2007). Some
infectious diseases of amphibians share similar pathological changes; thus, their
detection, recognition, and correct diagnosis can be a challenge even by trained
veterinary pathologists or experienced herpetologists.
This chapter will introduce readers to the most common amphibian dis-
eases with an emphasis on those that are potentially or frequently lethal, and
the techniques involved in disease monitoring. It will also outline methods of
biosecurity to reduce the transmission of disease agents by humans. We start by
covering infectious, parasitic, and toxic diseases. Next, surveillance methods
are discussed, including methods for sample collection and techniques used in
disease diagnosis. Finally, biosecurity issues for preventing disease transmission
will be covered, and we provide protocols for disinfecting field equipment and
footwear.
482 | Amphibian ecology and conservation
1
Chlamydomonas sp. is a symbiotic blue-green alga in the egg capsule of Ambystoma maculatum in
eastern North America and Ambystoma gracile in western North America, and not considered a
disease agent.
2
Watermold infections (oomycetes of several genera) referred to as saprolegniasis.
(Green and Converse 2005a). However, it is important to note that red-leg dis-
ease is a gross descriptor of a specific lesion (i.e. swollen red legs) and not specific
for a particular etiology. Many pathogens (e.g. Ranavirus, A. hydrophila, alveo-
lates) can cause edema (i.e. swelling) and erythema (reddening) in amphibians
(Figure 26.1a). This emphasizes the importance of diagnostic testing to deter-
mine the correct pathogen causing the disease.
Finally, numerous fungal and fungus-like organisms (Converse and Green
2005a, 2005b; Green and Converse 2005a, 2005b) and newly characterized
pathogens (Davis et al. 2007) are known to cause catastrophic mortality of
amphibian populations. B. dendrobatidis (Figure 26.1b) has resulted in global
population declines and species extinctions (Wake and Vredenburg 2008).
The newly discovered alveolate organism has only been diagnosed in a few
Table 26.2 Significant diseases of larval amphibians.
Disease agent Common host species Mortality rate Organ of choice Test methods
Ranaviruses Bufo spp., Hyla spp., Low in adults; variable Liver, skin ulcers Culture at 20–25°C;
Pseudacris spp., Rana spp., in recently mesonephroi PCR on liver, spleen, skin
Ambystoma spp., metamorphosed ulcers, mesonephroi;
Notophthalmus spp. amphibians electron microscopy
Lucke tumor Rana pipiens Variable in >2 yr old Mesonephroi Histology of tumors
herpesvirus Rana pipiens only
Batrachochytrium Most anuran genera Very high in many anurans Skin of pelvic patch, Histology; PCR; culture;
dendrobatidis especially at high elevations in toe webs electron microscopy
tropical latitudes
Ichthyophonus sp. Rana spp., Very low Skeletal muscle Histology
Ambystoma spp.,
Notophthalmus spp.
Amphibiothecum Bufo spp. 0 Ventral skin nodules Cytology of discharge; histology
penneri1 of nodule
Hepatozoon spp. Rana spp. Low Blood smear; liver Cytology; histology
Ribeiroia ondatrae Bufo spp., Pseudacris spp., 0 Skin around vent, Examination of metacercaria by a
Rana spp., Ambystoma spp. proximal hindlimbs parasitologist; PCR; radiographs of
and at tip of urostyle malformations
Rhabdias spp. Bufo spp. Unknown Lungs Visible at dissection; histology;
(nematode lungworm) Rana spp. examination by a
parasitologist
1
Amphibiothecum (formerly Dermosporidium) penneri, referred to as dermosporidiosis.
486 | Amphibian ecology and conservation
(a)
(b)
(c)
Fig. 26.1 (a) Tadpoles with swollen bodies and swollen red legs (arrow) are often
diagnosed as red-leg disease but the etiology is varied and may include Aeromonas
hydrophila, Ranavirus, and alveolates. (b) The amphibian fungus, Batrachochytrium
dendrobatidis (arrows), infects the keratin-producing cells of amphibians. Tadpole skin
is not keratinized; rather, only their ‘teeth’ contain keratin. Grossly, this is seen by loss
of pigmentation (upper inset) of the tooth rows. Lower inset is of a normal tadpole
for comparison. (c) Trematode cercaria encyst within the skin (arrows) and body
cavities of amphibians serving as a secondary host and may be easily seen grossly.
Histologically, the organisms are found in thin-walled cysts (inset).
isolated geographical areas so far (Davis et al. 2007). Still other organisms,
such as the watermolds Saprolegnia, may be beneficial (e.g. by facilitating
decomposition of dead eggs) but also have the potential to be opportunistic
pathogens of amphibians at any life stage (Converse and Green 2005a, 2005b;
Green and Converse 2005a, 2005b).
26 Disease monitoring and biosecurity | 487
26.2.3 Toxins
Contaminants in the environment may kill larvae or post-metamorphs (Relyea
2005, 2009), and may have non-lethal impacts including reducing growth,
impacting metamorphosis, disrupting gonadal development and secondary sex
characteristics, or causing musculoskeletal, skin, and visceral malformations
(Boone and Bridges 2003; Davidson et al. 2007; Ouellet et al. 1997; Storrs
and Semlitsch 2008). Often these changes are not detected by external exami-
nations until metamorphosis is complete or until the animals attain a size for
reproduction. Amphibians are often considered sentinels or bio-indicators of
environmental quality because they are sensitive to toxins and many species
have the potential to be exposed to stressors in aquatic and terrestrial systems
due to their typical biphasic life cycle (Blaustein and Wake 1995).
been essentially non-existent, unlike for domestic livestock and pets and
some wild mammals (e.g. cervids). The OIE has established guidelines for
surveillance and requirements necessary for countries to declare Ranavirus-
free status (www.oie.int/eng/norms/fcode/en_chapitre_2.4.2.htm#rubrique_
ranavirus) and B. dendrobatidis-free status (www.oie.int/eng/normes/fcode/
en_chapitre_2.4.1.htm#rubrique_batrachochytrium_dendrobatidis). The
OIE-approved methods for conducting surveillance and diagnosis of infection
are in development. In the meantime, guidelines from the 2006 OIE Manual of
Diagnostic Tests for Aquatic Animals (www.oie.int/ eng/normes/fmanual/A_
summry.htm) and from the fisheries industry (USFWS and AFS-FHS 2005)
can be helpful for general monitoring of amphibian population health. In gen-
eral, the criteria of a population health assessment should include (1) determin-
ation of the status and trends of amphibian pathogens, (2) determination of
the risk of disease for threatened or endangered amphibians, (3) investigation
of unexplained population declines, (4) evaluation of populations following a
morbidity or mortality event, (5) detection of pathogens in non-indigenous spe-
cies, (6) evaluation of a site or population prior to translocation, (7) evaluation
of sympatric amphibians prior to release of captive-raised animals, and (8) the
potential for amphibians and their diseases to “piggy back” with fish translocation.
Disease testing should not focus on one pathogen. For surveillance programs,
we recommend that animals are tested for infection by at least the OIE patho-
gens: ranaviruses and B. dendrobatidis. For diagnosis of morbid or dead individ-
uals, we encourage a full diagnostic work-up (i.e. necropsy, histology, bacterial
culture, virus testing, and parasite testing) to attempt to identify all etiologic
agents. It is important to note that simultaneous infection by multiple patho-
gens is possible. Further, histological examination of organs often is required to
determine which of the pathogens identified are causing the changes responsible
for the diseased state (Miller et al. 2008, 2009). Histological examination is also
important in discovering introduced pathogens or pathogens that have not been
described previously (Longcore et al. 1999; Davis et al. 2007).
Population health assessments can include non-lethal or lethal collection of
tissue samples from individual amphibians (Greer and Collins 2007), and col-
lection of environmental samples (e.g. water, soil; Walker et al. 2007). Ideally,
we recommend that tissue samples are collected from all species in a community
and from pre- and post-metamorphic life stages. Amphibian species differ in sus-
ceptibility to pathogens, and some age classes may serve as a reservoir (e.g. larval
for B. dendrobatidis and adults for Ranavirus; Daszak et al. 1999; Brunner et al.
2004; Schock et al. 2008). Further, some infectious diseases become evident
only after the post-metamorph has overwintered (e.g. Lucke tumor herpesvirus,
Amphibiothecum (formerly Dermosporidium) penneri). The lack of gross signs
26 Disease monitoring and biosecurity | 489
of disease also does not imply healthy populations. We and others have found
tadpoles with no signs of illness but that are infected with ranaviruses (Gray et al.
2007a; Harp and Petranka 2006; Miller et al. 2009). Laboratory studies have
demonstrated that amphibian pathogen infection and mortality rates frequently
track each other (e.g. Brunner et al. 2007); thus, high prevalence in a population
could signal that a die-off is imminent.
In some cases, it may not be possible to collect sufficient tissue for dis-
ease testing. For example, small amphibians (e.g. Bufo larvae) may not have
adequate tissue for tests, especially for toxicological analysis. Also, non-lethal
testing may be required because a species is listed as a conservation concern.
We found that testing for Ranavirus from tail clips results in about 20% false-
negatives (D. L. Miller and M. J. Gray, unpublished results). In cases when a
small amount of tissue is collected, multiple individuals within a species could
be pooled to acquire sufficient tissue for testing. If contaminants are suspected
as the cause of a die-off, we also recommend collecting and testing water and
sediment at the amphibian breeding site.
Monitoring for malformations can be challenging, because typically mal-
formed individuals have low survival. Although amphibian malformations
have been documented for many years (Rostand, 1958), an increase in mal-
formation rates occurred in the late twentieth century (Johnson and Lunde
2005). Generally, malformation studies have targeted recently metamorphosed
amphibians (Meteyer et al. 2000), because metamorphs with prominent abnor-
malities are quickly removed from the population by predation or starvation.
Additionally, the bony skeleton of metamorphosed amphibians is more con-
ducive for radiographically visualizing deformities compared to the cartilagin-
ous skeleton of larvae. However, monitoring of larval abnormalities is needed
because it is likely that some abnormalities prevent metamorphosis, thus are not
detected in post-metamorphic cohorts.
Finally, comprehensive disease surveillance should include captive amphib-
ians in zoological and ranaculture facilities, because disease transmission can
occur between captive and free-ranging populations. Maintenance of health
in zoological facilities is especially important for rare species or in captive
breeding populations intended for release. High densities in ranaculture facil-
ities, pet shops, and stores that sell amphibians (e.g. Ambystoma tigrinum) for
fishing bait can be cauldrons for disease transmission and pathogen evolution
(Picco and Collins 2008). Ranaviruses isolated from ranaculture facilities and
bait shops appear to be more virulent than wild strains (Majji et al. 2006;
Storfer et al. 2007). This emphasizes the importance for disease monitoring at
facilities with captive amphibians. In the event of a die-off in a captive facil-
ity, freshly dead animals should be submitted for diagnostic evaluation. Live
490 | Amphibian ecology and conservation
100 45 23
500 55 26
2000 60 27
10 000 60 30
APPL, assumed pathogen prevalence level.
(i.e. desiccated) carcasses with dry, leathery, and stiff digits or limbs usually have
limited diagnostic usefulness.
Dead amphibians should be collected, put individually in plastic bags (e.g.
Nasco Whirl-Pak® bags), and placed on ice for transport. Live amphibians can be
placed in separate plastic containers and humanely euthanized (Baer 2006) via
transdermal exposure for 10 min to tricaine methanesulfonate (100–250 mg/L)
or benzocaine hydrochloride ( 250 mg/L or 20% benzocaine over-the-counter
gel; Oragel, Del Paharmaceuticals, Uniondale, New York, USA) after returning
from the field. It is important that amphibians are bagged separately to prevent
cross-contamination of samples. Biologists that are experienced in blood collec-
tion may collect blood antemortem from the ventral vein in adult anurans or tail
vein in salamanders, or collect blood antemortem or immediately postmortem
from the heart of larvae or adults (Wright and Whitaker 2001). Blood can be
tested for various biochemical parameters and examined for cellular compos-
ition, blood parasites, and viral inclusions (discussed in section 26.3.4).
We recommend that half of the individuals collected are frozen immediately
for cultures and molecular tests. Samples can be frozen in a standard 20°C
freezer if stored for short duration (
1 month); otherwise, samples should be
stored in a 80°C freezer. The other half of samples should be promptly fixed
in 75% ethanol or 10% neutral buffered formalin for histology. For the first
2–4 days of fixation, the volume of fixative should be 10 times the volume of the
animals. After this initial fixation, carcasses can be stored in a smaller volume of
fixative that is sufficient to cover the tissues. The body cavity of amphibians that
are more than 1 g in body mass should be cut along the ventral midline prior to
immersion in fixative to assure rapid fixation of internal organs. Body cavities of
frozen individuals should not be opened.
Special processing is required for amphibians with skin, digital, limb, head,
or vertebral abnormalities. Whenever possible, amphibians with suspected
malformations should be submitted alive for examinations. Dead individuals
should be promptly frozen until time exists to properly fi x individuals. Fixation
can be done with ethanol or formalin but should be done in a pan so that car-
casses can be positioned on a flat surface with limbs and digits extended from
the body during fi xation. Positioning amphibians in the standard museum
configuration is ideal. Digits and limbs may be taped in position prior to fi x-
ation. Amphibians should be covered with fi xative and additional fi xative
added if a significant amount evaporates. Placing a cover over the pan will help
reduce evaporation. After 2–4 days of fi xation, the carcass and limbs will be
hardened in position and may be stored in a smaller volume of fi xative. The
26 Disease monitoring and biosecurity | 493
26.3.4 Diagnostics
Several tools are available for diagnosing amphibian diseases but generally
require some level of specialized expertise to perform. Commonly used diagnos-
tic tools include necropsy, histology, cytology, bacterial culture, virus isolation,
fecal floatation, electron microscopy, molecular modalities, and radiology. Most
of these tests can be performed on samples collected from dead or live amphib-
ians. Fresh or frozen tissues can be used for most tests, and are necessary for virus
isolation. Frozen tissues are not appropriate for histology or cytology; rather,
preserved tissues are used. Although formalin-fixed specimens are preferred for
histological examination, ethanol-fixed specimens may also be used. Blood can
be used for cell counts to assess immune function and to look for inclusion bod-
ies that can be diagnostic for certain pathogens. Blood also may be tested for the
presence of antibody response to various diseases. Examples of laboratories that
currently test for amphibian diseases in Australia, New Zealand, Europe, and
the USA include Australian Animal Health Laboratories (AAHL), Geelong,
Victoria, Australia (www.csiro.au/places/aahl.html), Gribbles Veterinary
Pathology, Australia and New Zealand (www.gribblesvets.com/), Exomed,
Berlin, Germany (www.exomed.de/), Hohenheim University (R. Marschang),
Stuttgart, Germany, Wildlife Epidemiology, Zoological Society of London
(ZSL), London, UK, The University of Georgia Veterinary Diagnostic and
Investigational Laboratory, Tifton, GA, USA (www.vet.uga.edu/dlab/tifton/
index.php), University of Florida (J. Wellehan), College of Veterinary Medicine,
Gainesville, FL, USA, and National Wildlife Health Center, Madison, WI,
USA (www.nwhc.usgs.gov/).
There are advantages and disadvantages to the various tests available
(Table 26.6). Necropsy allows for identification and documentation of exter-
nal and internal gross changes. Histological and cytological examination
allows for identification of changes at the cellular level and is generally neces-
sary to document disease versus infection. Virus isolation is the process of
culturing a virus which is necessary to determine the presence of live virus
and to perform some molecular tests used in identifying viral species (e.g.
sodium dodecyl sulphate/polyacrylamide-gel electrophoresis (SDS/PAGE)
and restriction fragment length polymorphism (RFLP)). One caveat is that
some viruses are difficult to culture, thus infection cannot be ruled out based
solely on negative isolation results. Electron microscopy is used for identify-
ing key features of parasites or other infectious agents (e.g. B. dendrobatidis,
Ranavirus, herpesvirus), documenting intracellular changes or changes to the
cellular surface, and confirmation of cultured virus. Electron microscopy can
be performed on fresh, fi xed, or paraffin-embedded tissues. Radiology allows
26 Disease monitoring and biosecurity | 495
Table 26.6 Advantage and disadvantages of diagnostics tests for amphibian pathogens
given the type of sample.
Sample type Tests Pathogen Advantages Disadvantages
and disinfect boots, waders, nets, and field equipment and change clothes before
entering a vehicle and leaving the site (discussed in section 26.4.2).
Few infectious diseases of amphibians are contagious to humans. Potential
zoonotic diseases that may be carried by amphibians include certain Salmonella
spp., Yersinia spp., Chlamydophila spp. (formerly Chlamydia)), and some tox-
in-producing mycobacteria (e.g. Mycobacterium liflandii) that can cause skin
ulceration. In addition, Gray et al. (2007c) demonstrated that Rana catesbe-
iana metamorphs were suitable hosts for the human pathogen Escherichia coli
O157:H7. We also demonstrated recently that tadpoles could maintain this
pathogen in aquatic mesocosms (M. J. Gray and D. L. Miller, unpublished
results). Thus, disposable gloves should be worn whenever handling amphib-
ians, and hands washed thoroughly with soap and warm water after removing
gloves. In the field, hands can be soaked in a 2% clorhexidine solution for 1 min
or disposable antibacterial wipes used. Avoid exposure of surface water to soaps
and disinfectants, as they may negatively affect local flora and fauna. Clothing
that becomes stained with feces or skin secretions should be removed as soon as
possible and washed in color-safe bleach.
The skin secretions of many amphibians contain potent irritants and toxins.
For example, newts (Salamandridae), toads (Bufonidae), and poison-dart frogs
(Dendrobatidae) exude toxic skin secretions. Skin secretions of certain newts (e.g.
Taricha) may cause temporary blindness lasting several hours if the secretions get
into the eyes. The parotoid gland secretions of giant toads (Bufo marinus), if
ingested, can rapidly cause heart malfunction in humans and animals. When
handling toads, it is best to avoid touching the parotoid glands. After handling
amphibians, avoid touching your eyes or mouth prior to washing hands.
and effective against most bacteria and many viruses. The US Fish and Wildlife
Service and American Fisheries Society – Fish Health Section (USFWS and
AFS-FHS) (2005) recommend 10 min of exposure of a 0.05% bleach solution
(i.e. 28.4 g of 6.15% sodium hypochlorite in 3.8 L of clean water) for disinfection
of field equipment and surfaces for B. dendrobatidis, and, although not conclu-
sive, a 0.5% solution (i.e. 312 g of 6.15% sodium hypochlorite per 3.8 L of water)
is recommended to destroy myxosporeans. However, bleach is not very effective
at inactivating Ranavirus, and requires at least a 3% concentration (Bryan et al.
2009). It should be noted that this concentration can be toxic to amphibians. In
contrast, chlorhexidine used at a dosage that is safe for amphibians (0.75% for
a 1 min exposure) has been shown to inactivate Ranavirus (Bryan et al. 2009).
Further, it is important to keep in mind that the shelf-life of bleach solutions is
influenced by exposure to light, air, and organic material, and solutions should
be discarded after 5–7 days. After disinfection, equipment may be allowed to air
dry or rinsed with fresh, clean water. Alternatively, if carrying large quantities of
water is not possible because multiple fields sites are to be visited, surface water
from the subsequent site (i.e. where the equipment will be used next) can serve as
the rinse water. If mountain systems with stream watersheds are sampled, we rec-
ommend that researchers begin sampling at higher elevations and work towards
lower sites. If a disease agent is present at higher elevations, it is likely to be at
lower elevations due to downstream transmission. Hence, if accidental transmis-
sion occurs during travel on fomites, it is less likely to be a novel introduction.
discarded at the same or other sites because this may facilitate the spread or
persistence of infectious diseases. Dead amphibians that are not used for testing
should be placed in double-layered plastic trash bags and disposed by burial or
incineration. Removal of carcasses is a good strategy to help thwart the spread
of infectious diseases.
While some serious infectious diseases of amphibians (e.g. B. dendrobatidis,
nematode lungworms (Rhabdias spp.)) are readily treated and eliminated
from captive populations, some important infectious diseases have no known
treatments (e.g. ranaviruses, alveolates) or no practical treatment in the wild.
Treatment of any disease varies by the pathogen involved as well as the host.
Some pathogens are resistant to many treatments (e.g. antibiotic-resistant bac-
teria) and some hosts may be sensitive to a particular treatment (e.g. Methylene
Blue may be toxic to tadpoles at concentrations over 2 mg/ml). Generally, it is
best to contact a veterinarian with experience in amphibians for proper treatment
of disease. However, some treatments (i.e. elevated temperature for B. dendroba-
tidis or dermosporidium, sea salt or Methylene Blue for Saprolegnia, chlorhexi-
dine for bacteria and Ranavirus) may be attempted by the non-veterinarian and
treatment guidelines can be found in Wright and Whitaker (2001) and Poole
(2008). As a general rule, treatment for disease is only applicable to captive
environments; however, it can be a valuable conservation tool for amphibians
slated for release.
In the event that animals destined for release test positive for a treatable dis-
ease, the animal and any others that may have been exposed should be treated.
Following treatment, a minimum of two negative test results with 1 month
between tests should be obtained. If the animal does not test negative, the treat-
ment should be repeated. Only animals that test negative should be released into
the wild. In addition, if one animal in a group of 10 housed together tests posi-
tive for a pathogen, all of the animals should be treated, regardless of individual
test results. Current guidelines for treatment and release have been established
by the Association of Zoos and Aquariums (Poole 2008). Testing at the appro-
priate life stage for the host and disease agent is important.
26.5 Conclusions
Amphibians are declining globally and emerging infectious diseases are one of
the causes. Natural resource agencies and conservation organizations should
consider establishing amphibian disease surveillance programs that moni-
tor populations for at least the two pathogens linked to catastrophic die-offs:
Ranavirus and B. dendrobatidis. Further, the OIE has listed these pathogens as
26 Disease monitoring and biosecurity | 501
26.6 References
Baer, C. K. (2006). Guidelines for Euthanasia of Nondomestic Animals. American
Association of Zoo Veterinarians, Yulee, FL.
Blaustein, A. R. and Wake, D. B. (1995). The puzzle of declining amphibian populations.
Scientific American, 272, 52–7.
Boone, M. D. and Bridges, C. M. (2003). Effects of carbaryl on green frog (Rana clami-
tans) tadpoles: timing of exposure versus multiple exposures. Environmental Toxicology
and Chemistry, 22, 2695–2702.
Brem, F., Mendelson, III, J. R., and Lips, K. R. (2007). Field-Sampling Protocol for
Batrachochytrium dendrobatidis from Living Amphibians, using Alcohol Preserved
Swabs, version 1.0 (18 July 2007). www.amphibianark.org/pdf/Field%20sampling%
20protocol%20for%20amphibian%20chytrid%20fungi%201.0.pdf. Conservation
International, Arlington, VA.
Brunner, J. L., Schock, D. M., Davidson, E. W., and Collins, J. P. (2004). Intraspecific reser-
voirs: complex life history and the persistence of a lethal ranavirus. Ecology, 85, 560–6.
Brunner, J. L., Richards, K., and Collins, J. P. (2005). Dose and host characteristics influ-
ence virulence of ranavirus infections. Oecologia, 144, 399–406.
Brunner, J. L., Schock, D. M., and Collins, J. P. (2007). Transmission dynamics of the amphib-
ian ranavirus Ambystoma tigrinum virus. Diseases of Aquatic Organisms, 77, 87–95.
Bryan, L. K., Baldwin, C. A., Gray, M. J., and Miller, D. L. (2009). Efficacy of select
disinfectants at inactivating Ranavirus. Diseases of Aquatic Organisms, 84, 89–94.
502 | Amphibian ecology and conservation
Carey, C., Bradford, D. F., Brunner, J. L., Collins, J. P., Davidson, E. W., Longcore, J. E.,
Ouellet, M., Pessier, A. P., and Schock, D. M. (2003). Biotic factors in amphibian
declines. In G. Linder, S. K. Krest, and D. W. Sparling (eds), Amphibian Declines: an
Integrated Analysis of Multiple StressorEffects, pp. 153–208. Society of Environmental
Toxicology and Chemistry, Pensacola, FL.
Cashins, S. D., Alford, R. A., and Skerratt, L. F. (2008). Lethal effect of latex, nitrile, and
vinyl gloves on tadpoles. Herpetological Review, 39, 298–301.
Converse, K. A. and Green, D. E. (2005a). Diseases of tadpoles. In S. K. Majumdar,
J. Huffman, F. J. Brenner, and A. I. Panah (eds), Wildlife Diseases: Landscape
Epidemiology, Spatial Distribution, and Utilization of Remote Sensing Technology,
pp. 72–88. Pennsylvania Academy of Science, Easton, PA.
Converse, K. A. and Green, D. E. (2005b). Diseases of salamanders. In S. K. Majumdar,
J. Huffman, F. J. Brenner, and A. I. Panah (eds), Wildlife Diseases: Landscape
Epidemiology, Spatial Distribution, and Utilization of Remote Sensing Technology,
pp. 118–30. Pennsylvania Academy of Science, Easton, PA.
Cunningham, A. A., Hyatt, A. D., Russell, P., and Bennett, P. M. (2007). Experimental
transmission of a ranavirus disease of common toads (Bufo bufo) to common frogs
(Rana temporaria). Epidemiology and Infection, 135, 1213–16.
Daszak, P., Berger, L., Cunningham, A. A., Hyatt, A. D., Green, D. E., and Speare, R.
(1999). Emerging infectious diseases and amphibian population declined. Emerging
Infectious Diseases, 5, 735–48.
Davidson, C., Benard, M. F., Shaffer, H. B., Parker, J. M., O’Leary, C., Conlon, J. M.,
and Rollins-Smith, L. A. (2007). Effects of chytrid and carbaryl exposure on survival,
growth and skin peptide defenses in foothill yellow-legged frogs. Environmental Science
and Technology, 41, 1771–6.
Davis, A. K., Yabsley, M. J., Keel, M. K., and Maerz, J. C. (2007). Discovery of a novel
alveolate pathogen affecting southern leopard frogs in Georgia: description of the dis-
ease and host effects. EcoHealth, 4, 310–17.
Forson, D. D. and Storfer, A. (2006). Atrazine increases ranavirus susceptibility in the
tiger salamander, Ambystoma tigrinum. Ecological Applications, 16, 2325–32.
Gray, M. J., Miller, D. L., Schmutzer, A. C., and Baldwin, C. A. (2007a). Frog virus 3
prevalence in tadpole populations inhabiting cattle-access and non-access wetlands in
Tennessee, U.S.A. Diseases of Aquatic Organisms, 77, 97–103.
Gray, M. J., Smith, L. M., Miller, D. L., and Bursey, C. R. (2007b). Influence of agricul-
tural land use on trematode occurrence in southern Great Plains amphibians, U.S.A.
Herpetological Conservation and Biology, 2, 23–8.
Gray, M. J., Rajeev, S., Miller, D. L., Schmutzer, A. C., Burton, E. C., Rogers, E. D.,
and Hickling, G. J. (2007c). Preliminary evidence that American bullfrogs (Rana cat-
esbeiana) are suitable hosts for Escherichia coli O157:H7. Applied and Environmental
Microbiology, 73, 4066–8.
Green, D. E. and Converse, K. A. (2005a). Diseases of amphibian eggs. In S. K. Majumdar,
J. Huffman, F. J. Brenner, and A. I. Panah (eds), Wildlife Diseases: Landscape
Epidemiology, Spatial Distribution, and Utilization of Remote Sensing Technology,
pp. 62–71. Pennsylvania Academy of Science, Easton, PA.
26 Disease monitoring and biosecurity | 503
Mauel, M. J., Miller, D. L., Frazier, K., and Hines, II, M. E. (2002). Bacterial pathogens
isolated from cultured bullfrogs (Rana catesbeiana). Journal of Veterinary Diagnostic
Investigation, 14, 431–3.
Meteyer, C. U., Loeffler, I. K., Fallon, J. F., Converse, K. A., Green, E., Helgen, J. C.,
Kersten, S., Levey, R., Eaton-Poole, L., and Burkhart, J. G. (2000). Hind limb malfor-
mations in free-living northern leopard frogs (Rana pipiens) from Maine, Minnesota,
and Vermont suggest multiple etiologies. Teratology, 62, 151–71.
Miller, D. L., Bursey, C. R., Gray, M. J., and Smith, L. M. (2004). Metacercariae of
Clinostomum attenuatum in Ambystoma tigrinum mavortium, Bufo cognatus and Spea
multiplicata from west Texas. Journal of Helmithology, 78, 373–6.
Miller, D. L., Rajeev, D., Brookins, M., Cook, J., Whittington, L., and Baldwin, C. A.
(2008). Concurrent infection with Ranavirus, Batrachochytrium dendrobatidis, and
Aeromonas in a captive anuran colony. Journal of Zoo and Wildlife Medicine, 39, 445–9.
Miller, D. L., Gray, M. J., Rajeev, S., Schmutzer, A. C., Burton, E. C., Merrill, A. and
Baldwin, C. (2009). Pathological findings in larval and juvenile anurans inhabiting
farm ponds in Tennessee, U. S.A. Journal of Wildlife Diseases, 45, 314–24.
Muths, E., Gallant, A. L., Campbell, E. H. C., Battaglin, W. A., Green, D. E., Staiger, J. S.,
Walls, S. C., Gunzburger, M. S., and Kearney, R. F. (2006). The Amphibian Research
and Monitoring Initiative (ARMI): 5-year report. U.S. Geological Survey Scientific
Investigations Report 5224. U.S. Geological Survey, Reston, VA.
Ouellet, M., Bonin, J., Rodrigue, J., DesGranges, J. L., and Lair, S. (1997). Hindlimb
deformities (ectromelia, ectrodactyly) in free-living anurans from agricultural habitats.
Journal of Wildlife Diseases, 33, 95–104.
Pallister, J., Gould, A., Harrison, D., Hyatt, A., Jancovich, J., and Heine, H. (2007).
Development of real-time PCR assays for the detection and differentiation of Australian
and European ranaviruses. Journal of Fish Diseases, 30, 427–38.
Picco, A. M. and Collins, J. P. (2008). Amphibian commerce as a likely source of patho-
gen pollution. Conservation Biology, 22, 1582–9.
Poole, V. A. (2008). Amphibian Husbandry Resource Guide, edn 1.1. Association of Zoos
and Aquarium’s Amphibian Taxon Advisory Group, Yulee, FL.
Relyea, R. A. (2005). The lethal impact of Roundup on aquatic and terrestrial amphib-
ians. Ecological Applications, 15, 1118–24.
Relyea, R. A. (2009). A cocktail of contaminants: how mixtures of pesticides at low con-
centrations affect aquatic communities. Oecologia, 159, 363–76.
Rostand. J. (1958). Les Anomalies des Amphibiens Anoures. Sedes, Paris.
Schock, D. M., Bollinger, T. K., Chinchar, V. G., Jancovich, J. K., and Collins, J. P.
(2008). Experimental evidence that amphibian ranaviruses are multi-host pathogens.
Copeia, 2008, 133–43.
Schotthoefer, A. M., Koehler, A. V., Meteyer, C. U., and Cole, R. A. (2003). Influence of
Ribeiroia ondatrae (Trematoda: Digenea) infection on limb development and survival
of Northern leopard frogs (Rana pipiens): Effects of host-stage and parasite exposure
level. Canadian Journal of Zoology, 81, 1144–53.
Skerratt, L. F., Berger, L., Speare, R., Cashins, S., McDonald, K. R., Phillott, A. D.,
Hines, H. B., and Kenyon, N. (2007). Spread of chytridiomycosis has caused the rapid
global decline and extinction of frogs. EcoHealth, 4, 125–34.
26 Disease monitoring and biosecurity | 505
Skerratt, L. F., Berger, L., Hines, H. B., McDonald, K. R., Mendez, D., and Speare, R.
(2008). Survey protocol for detecting chytridiomycosis in all Australian frog popula-
tions. Diseases of Aquatic Organisms 80, 85–94.
Storfer, A., Alfaro, M. E., Ridenhour, B. J., Jancovich, J. K., Mech, S. G., Parris, M. J., and
Collins, J. P. (2007). Phylogenetic concordance analysis shows an emerging pathogen
is novel and endemic. Ecology Letters, 10, 1075–83.
Storrs, S. I. and Semlitsch, R. D. (2008). Variation in somatic and ovarian develop-
ment: predicting susceptibility of amphibians to estrogenic contaminants. General and
Comparative Endocrinology, 156, 524–30.
USFWS and AFS-FHS (US Fish and Wildlife Service and American Fisheries Society – Fish
Health Section) (2005). AFS-FHS Blue Book: suggested procedures for the detection and
identification of certain finfish and shellfish pathogens, 2005 edition. American Fisheries
Society – Fish Health Section, Bethesda, MD.
Wake, D. B. and Vredenburg, V. T. (2008). Are we in the midst of the sixth mass extinc-
tion? A view from the world of amphibians. Proceedings of the National Academy of
Sciences USA, 105, 11466–73.
Walker, S. F., Salas, M. D., Jenkins, D., Garner, T. W.J., Cunningham, A. A., Hyatt, A. D.,
Bosch, J., and Fisher, M. C. (2007). Environmental detection of Batrachochytrium den-
drobatidis in a temperate climate. Diseases of Aquatic Organisms, 77, 105–12.
Whiles, M. R., Lips, K. R., Pringle, C. M., Kilham, S. S., Bixby, R. J., Brenes, R.,
Connelly, S., Colon-Gaud, J. C., Hunte-Brown, M., Huryn, A. D., Montgomery, C.,
and Peterson, S. (2006). The effects of amphibian population declines on the struc-
ture and function of Neotropical stream ecosystems. Frontiers in Ecology and the
Environment, 4, 27–34.
Williams, T., Barbosa-Solomieu, V., and Chinchar, V. G. (2005). A decade of advances
in iridovirus research. In K. Maramorosch and A. Shatkin (eds), Advances in Virus
Research, vol. 65, pp. 173–248. Academic Press, New York.
Wright, K. M. and Whitaker, B. R. (2001). Amphibian Medicine and Captive Husbandry.
Krieger Publishing, Malabar, FL.
Yip, M. J., Porter, J. L., Fyfe, J. A. M., Lavender, C. J., Portaels, F., Rhodes, M., Kator, H.,
Colorni, A., Jenkin, G. A., and Stinear, T. (2007). Evolution of Mycobacterium ulcerans
and other mycolactone-producing mycobacteria from a common Mycobacterium mari-
num progenitor. Journal of Bacteriology, 189, 2021–9.
Yuan, J. S., Reed, A., Chen, F., and Stewart, Jr, C. N. (2006). Statistical analysis of real-
time PCR data. BMC Bioinformatics, 7, 85–97.
This page intentionally left blank
27
Conservation and management
C. Kenneth Dodd, Jr
27.1 Introduction
Considering the large number of amphibian species and the diversity of life
histories, potential conservation strategies are likewise diverse and may be com-
plex. Beyond a few guiding principles, such as the need for planning and set-
ting objectives, understanding life-history constraints, and protecting existing
habitats and ecosystem function, there are no standardized approaches that
will be applicable to all situations. In the brief discussion that follows, I outline
some important considerations and various management approaches that have
proved successful in particular instances, and I provide references that con-
tain more extended discussions. Specific information on management options
for amphibian conservation are available in Beebee (1996), Gent and Gibson
(1998), Scoccianti (2001), Kingsbury and Gibson (2002), Semlitsch (2003),
Bailey et al. (2006), and Mitchell et al. (2006).
Amphibian conservation requires an integrated landscape approach to man-
agement, rather than solely a species-oriented focus. The reason for this is simple:
amphibians do not live in a biotic or physical vacuum in nature. Lindenmayer
et al. (2007) present a number of conceptual ideas for landscape conservation
that are apropos when focusing on amphibians; these apply equally to all man-
agement options (Table 27.1). Amphibian conservation options range from the
rather simple and inexpensive to the very complex and expensive. When plan-
ning, the overriding consideration should be “first, do no harm” to the species,
its community, or its habitat. High-technology-based approaches may work no
better than simple and inexpensive approaches, and care should be exercised to
maximize the benefits from the human resources and funds available. The pri-
mary objectives of management should always focus on the species or commu-
nity of concern, and not on peripheral or extended objectives, such as positive
publicity.
508 | Amphibian ecology and conservation
Table 27.1 Conceptual ideas for landscape conservation that are important when
planning for amphibian management. Adapted from Lindenmayer et al. (2007) and
Gardner et al. (2007).
Conceptual ideas important for amphibian management
Plan for habitat connectivity, particularly between aquatic and terrestrial habitats
Develop a long-term approach, both in management and in the evaluation of objectives
Manage all aspects of the habitat, not individual components, regarding landscape features,
species, or stressors
Manage for change, not a static condition
Manage species and ecosystems using explicit experimental approaches and at multiple
scales, rather than resorting to “cookbook” approaches
lands. In all cases, however, management plans must ensure that long-term
research and monitoring is carried out, and that land use occurs in accord-
ance with initial agreements. All land acquisition and alternative conservation
programs require early, careful, fiscal planning in order to ensure long-term
continuity of protection, management and assessments of how well objectives
are being achieved.
People residing adjacent to protected lands can be incorporated into long-
term management programs to provide local or regional economic incentives to
protection. Such an approach would be especially desirable where key amphib-
ian habitats are located in proximity to economically disadvantaged human
settlements. Providing educational programs, outlets for local arts, and even
minimal health care could go a long way to ensure that crucial habitats are
protected from vandalism or incursion. This approach works well in both devel-
oped and underdeveloped regions, for example, where local residents have been
hired as caretakers. Residents thus come to have a stake in the protected area and
in the species being protected.
New South Wales, Australia 10 frogs Constructed farm ponds Hazell et al. (2004)
Prince Edward Island, Canada 6 frogs Dredged wetlands Stevens et al. (2002)
3500 ponds, Denmark 7 frogs, 2 salamanders Dredging existing ponds, dug ponds Fog (1997)
Samsø, Denmark Bufo viridis Dredged ponds, vegetation removal Amtkjaer (1995)
Lolland, Denmark Hyla arborea Dredging existing ponds, dug ponds Hels and Fog (1995)
France Pelobates fuscus Use of bentonite and plastic-liners Eggert and Guyetant (2002)
Muensterland, Germany 5 frogs, 3 salamanders Dug and dredged ponds Schwartze (2002)
Salzburg, Germany 3 frogs, 4 salamanders Foil-lined permanent ponds Kyek et al. (2007)
Arribes del Duero, Spain 4 frogs, 4 salamanders Permanent dug ponds Alarcos et al. (2003)
Gipuzkoa, Spain Hyla meridionalis Permanent dug ponds Rubio and Etxezarreta (2003)
UK Rana temporaria Dug ponds, mostly seasonal Williams (2005)
Arkansas, USA Rana sylvatica Permanent dug wildlife ponds Cartwright et al. (1998)
Colorado, USA Bufo boreas Dug ponds, 1.5–6 m deep, little or no vegetation Pearl and Bowerman (2006)
restoration
Illinois, USA 11 frogs, 6 salamanders Earthen dams in shallow valley, ponds permanent Palis (2007)
Minnesota, USA 7 frogs, Ambystoma tigrinum Filled ditches, reflooded basins Lehtinen and Galatowitsch (2001)
North Carolina, USA Rana sylvatica, Seasonal and permanent ponds, modified golf Petranka et al. (2003)
Ambystoma maculatum course ponds, blocked stream channels
Ohio, USA 7 frogs, 2 salamanders Complex ponds with shallow and deep sections, Weyrauch and Amon (2002)
revegetation, zooplankton and some amphibians stocked
South Carolina, USA 13 frogs, 4 salamanders Plastic-lined permanent dug ponds Pechmann et al. (2001)
514 | Amphibian ecology and conservation
are permanent and require vigilance to ensure that fish predators are not inad-
vertently (through sheet flow) or purposely introduced. Weyrauch and Amon
(2002) provided an innovative U-shaped design for created ponds whereby one
arm of the U is dug shallow and the other deep. Natural evaporation allows the
water to fluctuate in hydroperiod between the arms, and provides amphibians
with a choice of water depth.
Once ponds are dug, they are filled naturally by precipitation. Thus, ponds
are best created prior to the season when most precipitation occurs, allowing
pond chemistry and limnology to stabilize before the breeding season. Water
may be added to ensure that at least some amount is available between con-
struction and the onset of rainy periods. Most invertebrates and amphibians
will naturally colonize newly created ponds, often very rapidly. However, some
researchers have combined pond creation with stocking or relocation in order
to speed-up the colonization process or to target certain declining species. As
Petranka et al. (2003) noted, perturbations (disease, drought) sometimes hin-
der efforts to establish amphibians at created ponds, and monitoring needs to
continue for approximately 5 years or more to evaluate success. The longer the
pond is maintained, the more likely that additional species will colonize it, and
that amphibian populations will become established.
Existing breeding ponds may be restored or enhanced, usually by mechanic-
ally deepening the pond to remove infill or muck and by removing encroaching
vegetation. Trees, such as willows, are cut within the pond basin and along
the pond’s perimeter. Herbicides must be used with extreme care at amphibian
breeding sites, since these or their surfactants may have detrimental effects on
amphibians. In fire-dependent habitats, controlled burns through pond basins
may help to periodically re-establish breeding ponds by burning off muck and
vegetation. Typically, pond restoration is carried out during the non-breeding
time of the year when individuals are dispersed. Biebighauser (2002) has pre-
pared a free guidebook to creating and maintaining vernal ponds.
Apart from the critical conservation area, zones of protection may extend
outward depending on the extent to which amphibians travel away from the
site. The critical area requires a buffer to ensure its integrity, and the “core”
habitat to be protected becomes the critical area plus its associated buffer zones.
For wetland-breeding amphibians, a core area for conservation would include
the wetland site, plus both aquatic and terrestrial buffer zones that would
ensure a functioning community (see Figure 27.1). For a terrestrial salamander
habitat specialist, a core conservation area might include the immediate steep
slopes or rock face on which it resides, plus terrestrial buffer zones to allow for
complete canopy cover over the slope or rock face, allowing retention of shade
and high humidity. Thus, the concept of a “buffer” zone should be expanded
to include as much habitat as necessary to maintain functionality, rather than
F2 F1
AB CH
CW
Wil
dlife
Cor
rido
r
TB
F3 F4
be synonymous with a minimum area necessary to protect the critical area from
destruction or alteration.
Certainly core areas and associated buffer zones need to be managed essen-
tially free from adverse effects on resident amphibians. However, some human
uses may be allowed within a series of outermost concentric buffer zones such
that layers of allowed uses surround the core conservation area. For example, a
core protection area might extend 100 m from a wetland breeding site (a core area
plus aquatic and terrestrial buffer zones) where no forestry would be allowed.
However, very selective cutting could be allowed at distances of 100–150 m,
with even less restrictions on the types of activity at further than 150 m (see
Figure 27.1). Given that amphibians are not evenly distributed around a core site
or present at all times of the year, developing an effective core plus buffer zone
conservation and management area requires good data on spatial distribution.
A common question is, how much habitat must be included as conservation
buffer zones for amphibians? As Semlitsch and Jensen (2001) noted, there is no
one-size-fits-all value, and the size of buffer zones will vary depending on spe-
cies and community. In temperate habitats, amphibians often routinely travel
more than 500 m from breeding sites; thus, buffer zones may need to extend
much farther than some previous authorities have suggested. Based on extensive
life-history and movement data, for example, Semlitsch and Jensen (2001) sug-
gested that a core plus buffer zone of 164 m would provide substantial protec-
tion for 95% of the salamander community that they studied.
27.4.2 Silviculture
One of the most contentious issues in amphibian conservation management is the
effect of silvicultural practices on amphibians (deMaynadier and Hunter 1995).
However, there are still important principles which should be kept in mind.
Certainly, practices which completely destroy wetland breeding sites should be
518 | Amphibian ecology and conservation
prevented, and buffer zones need to be implemented around streams and wet-
lands to avoid desiccation and changes in thermal regimes and water chemistry.
The most extreme form of forestry is clear-cutting all trees and vegetation.
Many species of terrestrial amphibians, particularly salamanders and ground-
dwelling tropical frogs, decline or disappear after clear-cutting, and surviving
populations may take many decades to return to pre-cut abundance. If left to
succession in the absence of mechanical roller chopping, however, even such
severely degraded sites can recover if amphibian source populations are nearby
and wetlands with buffers remain intact. However, these sites cannot recover
if wetlands are destroyed, large areas are clear-cut, roller-chopping eliminates
all subsurface retreats, such as root tunnels, and all surface debris is eliminated
manually or by burning. Herbicides and fertilizers also may be detrimental to
many species.
One solution to these problems is to manage for a mosaic of habitats. Some
areas (including breeding sites and associated buffer zones) would be off limits
to cutting, whereas other areas would be cut on a rotational basis, thus pro-
viding different aged stands at various locations in the landscape (see Figure
27.1). Patches would have to be sufficiently large to be economical yet retain the
diversity of amphibian species and allow for inter-patch migration. If plantation
forestry is involved, patches of native vegetation could be retained within the
landscape mosaic. A second solution to forest management would be to employ
selective cutting that allows for canopy cover, yet does not significantly affect
surface and subsurface conditions.
(a) (b)
(c) (d)
successfully cross a highway, but others may require culverts of more than 1 m
diameter. The length of the culvert may have important effects of the propen-
sity of an amphibian to enter and cross successfully. For long culverts, as across
more than four lanes, the underpass might have to be constructed in sections,
and access to light and moisture become critical. Entrances should allow easy
access in both directions, and brush or other woody debris may provide retreat
sites and refuges from predators. Excellent discussions showing diagrams and
photographs of various types of amphibian passages are provided in Langton
(1989), ALASV (1994), Percsy (1995), Glandt et al. (2003), Brodziewska
(2005), and in the Proceedings of the International Conferences on Ecology
and Transportation (www.icoet.net/) (also see Figure 27.2a–d). Overpasses
generally do not work as well for amphibians as they do for mammals.
been unsuccessful, or the ultimate fate of the RRT is still unknown (Dodd
and Seigel 1991; Dodd 2005). There have been some successful RRT programs
(Román 2003; Rubio and Etxezarreta 2003; Bell et al. 2004; Kinne 2004), and
such projects are still often advocated.
There are a number of criteria to consider, in addition to funding and staff-
ing, prior to undertaking RRT projects (Dodd and Seigel 1991): (1) the causes
of the original decline should be known, and steps must be taken to rectify
them, (2) the biological constraints on conservation relating to life history, habi-
tat, and demography should be understood, (3) population genetic and social
structure should be evaluated, (4) individuals should be screened for disease,
and (5) long-term habitat protection and monitoring should be a part of the
RRT project. These criteria need to be considered for both the donor and recipi-
ent individuals and populations, where appropriate, especially when a resident
population remains at or in the vicinity of a recipient site. To avoid problems of
half-way technology, RRT should be considered only after other less manipula-
tive options have been considered.
Various life stages have been used in RRT programs and to augment exist-
ing populations. Captive-reared animals may been used as alternative sources
for all life stages, particularly juveniles and adults (Bloxam and Tonge 1995)
for RRT, rather than removing individuals from existing natural populations.
Most projects use larvae or eggs for RRT, as these are the easiest life-history
stages to obtain and most abundant for mass movement. Presumably, larvae
metamorphosing from a breeding site are more likely to remain in the vicinity
than adults or subadults found or reared in one location but released in another.
Adult mortality may be high in unfamiliar surroundings, and adults generally
are scarcer or less available than eggs or larvae for RRT.
Most successful RRT programs employ a stage-based release protocol through
time, rather than planning for a single release. This helps ensure that individuals
become established, especially when presented with stochastic environmental
conditions. Animals may have to be moved for years prior to successful estab-
lishment. Pre-movement activities could include population censuses, assess-
ment and preparation of suitable habitat at recipient sites, disease and genetic
screening, predator removal, and education.
27.7 Conclusion
Preventing the decline of amphibians is challenging in the global economy of
the twenty-first century, one that requires biologists to use a complex array of
innovative and integrated approaches. The greatest threat to amphibians remains
the loss or alteration of habitats, so primary emphasis should be placed on main-
taining large tracts of undisturbed areas and to link already fragmented habitats.
Conservationists must use tested and scientifically based options, and research
must address questions in the field, laboratory, and through statistical model-
ling. Approaches need to focus on the parsimonious (what is most likely to work)
rather than on the latest technological advancement, unless they are synonym-
ous. Amphibian conservation is in triage mode; researchers are competing with
many others for scarce resources. It is our responsibility to use these resources
wisely for amphibians as we navigate through the Earth’s sixth great extinction.
27.8 References
Alarcos, G., Ortiz, M. E., Lizana, M., Aragón, A., and Fernández Benéitez, M. J. (2003).
La colonización de medios acuáticos por anfibios como herramienta para su conser-
vación: el ejemplo de Arribes del Duero. Munibe, 16, 114–27.
524 | Amphibian ecology and conservation
ALASV(ArbeitsgruppeunterLeitungderAbteilungStraßenbaudesVerkehrsministeriums)
(1994). Amphibienschutz. Leitfaden für Schutzmaßnahmen an Straßen. Schrifttenreihe
der Straßenbauverwaltung Baden Württemberg, Heft, 4.
Amtkjaer, J. (1995). Increasing populations of the green toad (Bufo viridis) due to a pond
project on the island of Samsø. Memoranda Societatis pro Fauna et Flora Fennica, 71,
77–81.
Anonymous (undated). Garden Ponds as Amphibian Sanctuaries. British Herpetological
Society, Montrose.
Aplin, K., Paino, A., and Sleep, L. (2000). Building Frog Friendly Gardens. Western
Australian Museum, Perth.
Bailey, M. A., Holmes, J. N., Buhlmann, K. A., and Mitchell, J. C. (2006). Habitat man-
agement guidelines for amphibians and reptiles of the southeastern United States. Technical
Publication HMG-2. Partners for Amphibian and Reptile Conservation, Montgomery,
AL.
Beattie, R. C., Aston, R. J., and Milner, A. G. P. (1993). Embryonic and larval survival
of the common frog (Rana temporaria) with particular reference to acidic and limed
ponds. Herpetological Journal, 3, 43–8.
Beebee, T. J. C. (1996). Ecology and Conservation of Amphibians. Chapman & Hall,
London.
Bell, B. D., Pledger, S., and Dewhurst, P. L. (2004). The fate of a population of the
endemic frog Leiopelma pakeka (Anura: Leiopelmatidae) translocated to restored habi-
tat on Maud Island, New Zealand. New Zealand Journal of Zoology, 31, 123–31.
Bellemakers, M. J. S. and van Dam, H. (1992). Improvement of breeding success of the
moor frog (Rana arvalis) by liming of acid moorland pools and the consequences of lim-
ing for water chemistry and diatoms. Environmental Pollution, 78, 165–71.
Biebighauser, T. R. (2002). A Guide to Creating Vernal Ponds. U.S. Department of
Agriculture, Forest Service, Morehead, KY.
Bloxam, Q. M. C. and Tonge, S. (1995). Amphibians: suitable candidates for breeding-
release programmes. Biodiversity & Conservation, 4, 636–44.
Brodziewska, J. (2005). Wildlife tunnels and fauna bridges in Poland: past, present,
and future, 1997–2013. International Conference on Ecology and Transportation 2005
Proceedings, pp. 448–60.
Brown, C. J., Blossey, B., Maerz, J. C., and Joule, S. J. (2006). Invasive plant and experi-
mental venue affect tadpole performance. Biological Invasions, 8, 327–38.
Cartwright, M. E., Trauth, S. E., and Wilhide, J. D. (1998). Wood frog (Rana sylvat-
ica) use of wildlife ponds in northcentral Arkansas. Journal of the Arkansas Academy of
Science, 52, 32–4.
Chan-McLeod, A. C. A. and Moy, A. (2007). Evaluating residual tree patches as stepping
stones and short-term refugia for red-legged frogs. Journal of Wildlife Management, 71,
1836–44.
Davis, A. K., Yabsley, M. J., Keel, M. K., and Maerz, J. C. (2007). Discovery of a novel
alveolate pathogen affecting southern leopard frogs in Georgia: description of the dis-
ease and host effects. EcoHealth, 4, 310–17.
deMaynadier, P. G. and Hunter, Jr, M. L. (1995). The relationship between forest manage-
ment and amphibian ecology: a review of the North American literature. Environmental
Review, 3, 230–61.
27 Conservation management | 525
Semlitsch, R. D. (2000b). Size does matter: the value of small isolated wetlands. National
Wetlands Newsletter, 22, 5–6, 13.
Semlitsch, R. D. (ed.) (2003). Amphibian Conservation. Smithsonian Books, Washington
DC.
Semlitsch, R. D. and Jensen, J. B. (2001). Core habitat, not buffer zone. National Wetlands
Newsletter, 23, 5–6, 11.
Smith, K. G. (2005). Effects of nonindigenous tadpoles on native tadpoles in Florida:
evidence of competition. Biological Conservation, 123, 433–41.
Stevens, C. E., Diamond, A. W., and Gabor, T. S. (2002). Anuran call surveys on small
wetlands in Prince Edward Island, Canada restored by dredging of sediments. Wetlands,
22, 90–9.
Trauth, J. B., Trauth, S. E., and Johnson, R. L. (2006). Best management practices and
drought combine to silence the Illinois Chorus Frog in Arkansas. Wildlife Society
Bulletin, 34, 514–18.
Weyrauch, S. L. and Amon, J. P. (2002). Relocation of amphibians to created seasonal
ponds in southwestern Ohio. Ecological Restoration, 20, 31–6.
Williams, L. R. (2005). Restoration of ponds in a landscape and changes in common frog
(Rana temporaria) populations, 1983–2005. Herpetological Bulletin, 94, 22–9.
This page intentionally left blank
Index
abiotic factors 96, 152, 216, 218, 268, 285, agar models 370f20.1, 378, 390, 391, 392–3,
286, 300, 302–3, 304, 307, 309 394f21.2, 397f21.4, 399, 400,
abundance 24, 28, 55, 76, 79, 233f13.3, 287–8, 400f21.6, 401, 401f21.7
300, 302, 309, 321, 325, 330, see also models; physical models
334, 436, 438, 439, 440, 441, agile frogs 418
449, 458, 518 Agranat, I. D. 292
amphibians 143, 152, 233–4, 235–6, 240–1, agriculture 16, 185, 341, 511
248–50, 254, 255, 256, 259, 271, Aguirre, A. A. 267
307, 364, 437, 448–9, 452, 465 Aichinger, M. 271
estimation (counts) 287, 294, 329, 333, 432, Aide, T. M. 334
447, 448, 449, 450–2, 455, 456, Alford, R. A. 25, 143, 148, 157, 198, 207,
458, 465–6, 467, 468, 469, 470, 215, 367, 388, 390, 395, 400,
471–4, 476, 477; 436, 473, 477,
indices 235, 241, 288, 322 algae 16, 73, 76, 77–8, 79, 80, 82, 95, 107
relative 247, 248, 251, 259, 271, 288, 326, algal blooms/food 73, 307
452, 474, 475 alkalinity 110, 111, 308
sampling 451–2, 455 alleles 407, 414–15, 417, 418, 421
species 25, 60, 259, 281–2, 326, 327, Allen, C. R. 287
329, 346 Allmon, W. D. 250
transformation, effect of 322–3, 323t18.2 alpine newts 352
true 271, 467 alpine salamanders 414, 474–5
variation 441, 452, 467, 469 alternative hypotheses 22, 24
see also frogs, abundance; salamanders, Altig, R. 39, 40, 42, 49, 50, 51, 71, 72, 74,
abundance 82, 105
Acevedo, M. M. A. 189 Altwegg, R. 204, 207, 219, 222
ACI (amphibian calling index) 284, 288, 291 aluminium 111, 113, 309
acid rain 23, 110 alveolates 483, 486f26.1, 496, 497, 500, 522
acidity 517 Alytes 12, 45
Acris 11, 218 Alytes muletensis 512, 521
Acris crepitans 179, 284, 285f16.1 Ambystoma 43f3.2, 44, 151, 166, 216
blanchardi 179 Ambystoma laterale 353
Adams, M. H. 143, 145 Ambystoma macrodactylum 131, 420
Adams, M. J. 64, 65, 231, 442 Ambystoma maculatum 8, 43f3.2, 44, 100, 133,
adaptive variation 423–5 135–6, 145, 150f9.2, 151, 154–5,
Adenomera 14 197, 198, 348f19.5, 348f19.5, 352,
Adis, J. 175 470, 475
Adolf, E. F. 372 Ambystoma mexicanum 8, 44
Aeromonas 522 Ambystoma opacum 8, 135
Aeromonas hydrophila 284, 483, 486f26.1 Ambystoma talpoideum 44, 59f4.1, 217f12.4
AFDM (ash-free dry mass) 75, 76, 77 Ambystoma tigrinum 8, 106, 189, 196, 196f11.3,
AFLP (amplified fragment length 232, 353, 470, 472, 482, 489, 501
polymorphism) 411, 413, 415f22.1, Ambystomatidae 39, 42, 44, 45, 62, 77, 106,
416–17, 424 153, 220, 236, 314, 350f19.6, 424,
Africa 5, 7, 11, 13, 14, 15, 168, 250, 421 472, 473
African clawed frogs 10, 112 American bullfrogs 46f3.3, 106, 108, 132,
Afrixalus 13 511, 523
Agalychnis 12 American toads 215f12.3
530 | Index
environment 151, 363 Farrand, L. 189, 192, 193, 195, 197, 199
adaptation to 364 fatty acids 74, 75, 78, 81–2, 177–8
and ARS 288–9 fauna 272, 275
changing 364, 380 Fauth, J. E. 29
conditions 105, 273, 293, 295, 389 fecundity 221–2, 521
stressors 203 Feder, M. E. 105, 106, 363
variables 452 feeding patterns 168, 71–2, 167, 170, 179, 375
Ernst, R. 143, 270, 271 tadpoles 72–3
Escherichia coli 498 Fellers, G. M. 237, 267
Estoup, A. 423 Felton, A. 156
EstimateS 331, 333 Ferner, J. W. 32, 123, 130, 132t8.2, 136t8.7
estivation 5, 8, 108, 374 fertilizers 100, 114, 518
Eterovick, P. C. 174 Fiala, A. C. S. 311, 312
ethanol 74, 137, 170, 171–2, 173, 174, 178, Ficetola, G. F. 144, 273, 344, 349, 418
189, 191, 273, 410, 492 field cages 94, 97, 209
ethics 123 field enclosures 203–4, 207, 215f12.3, 216, 223
and toe clipping 129–30 construction 208–16
use of formalin 174 escape from 216–18
Etxezarreta, J. 522 response metrics 221–3
Euclidean metrics 330, 333, 350, 353 study species 218–19
Eupsophus 11 field experiments 204, 351, 381–2
Eurasia 5, 8 field studies 21–2, 24–6, 32, 91, 101, 123, 124,
Europe 8, 9, 11–12, 15, 130, 144, 145, 153, 180, 230, 274, 365
419, 421, 423, 437, 467, 460t25.2, objectives 30–1
468t25.2, 469, 472 SMART approach 27–9
Eurycea bislineata 133, 135 Finlay, J. C. 79, 81
Eurycea longicauda 133 fish 93, 177, 365, 482, 490, 514, 523
Eurycea lucifuga 133 fish 511
euthanasia of study animals 74, 169, 512 lobe-finned 3
eutrophication 16 Fitch, H. S. 366
Evans, B. J. 421 Fitzgerald, L. E. 314
Evans, C. J. 193 flagship species 26
Evans-White, M. A. 76, 77 Flectonotus 13
evenness 329; see also species, diversity foam nests 11, 12f1.2, 13, 14, 145
evolution 17, 156 Fogarty, J. H. 253, 273
current theory 21 Foltz, K. D. 292
mesocosms 91–2 food 71
evolutionary studies 99 collection of 73
EWL (evaporative water loss) 387, 389, 390–1, ingestion of 74
396, 401 organic carbons 113
physical models 399 food webs 29, 79, 80–1, 82, 93, 167, 178, 321,
see also cutaneous evaporative water loss 380, 487
experiments 29, 29f2.2, 30, 204, 207, 209, 218, aquatic 71, 96, 307
220–2, 223 freshwater 78
area-based surveys 247 natural 90
explosive-breeding species 150–1, 152, 153, 290 stream 79–80
extinction 15–16, 71, 448, 458, 481, 483, foraging activity/patterns 71, 72, 74, 75, 169,
521, 523 171, 172–3, 179, 519
extrapolation 326, 475 Ford, E. D. 22
forest 15, 45, 116, 203, 206, 247, 251, 349
F statistics 419, 424 counts 257, 257f14.1, 258f14.2, 259
Faber, J. 49 cover 347
Faccio, S. D. 189, 192, 199 habitats 341
Fahrig, L. 346, 349, 438 management 91, 100–1
Farnsworth, E. J. 31 tropical 175, 249, 250
538 | Index
Johnson, S. A. 64, 232, 234 188, 193, 198, 204, 235, 239, 308,
Johnson, W. S. 176 309, 373, 374, 381–2, 388, 489, 494
Johnston, G. F. 39, 40, 50 equipment 408–9
Jolly, G. M. 452 Lajtha, K. 380
Joly, P. 169, 346, 349, 351, 352 lakes 8, 10, 55, 62, 82, 90, 93, 113, 145
Jorgensen, C. B. 375 LaLiberte, G. D. 76, 77
Juncá, F. A. 174 Lamberti, G. A. 73, 80
Jung, R. E. 63, 65, 66, 440, 441, 449, 451, 455 Lamoureux, V. S. 187, 188, 195, 197, 199
Just, J. J. 382 Lampo, M. 267
Lancia, R. A. 447, 448
Karraker, N. E. 438 land acquisition 275
Kasier, K. 283 landscape 354–5, 519
Kam, Y.-C. 374 configuration 349, 356
Kaplan, H. M. 130, 131 habitats 514
Karasov, W. H. 307 permeability 351–2
Karraker, N. E. 112 perspective 340
Kayanne, H. 81 landscape ecology 339, 340, 340f19.1, 356
Keller, C. M. E. 438 amphibian conservation 341–2, 507,
Kendall, W. L. 453, 454, 455, 469, 470, 508t27.1
471, 474 population 346
Kennedy, A. D. 387, 388, 403 landscape genetics 352–3, 419
Kenya 353 landscape-as-population 340
keratin 44, 47, 49–50, 179, 364, 372, landscape-composition thresholds 349, 355–6
486f26.1, 496 Langton, T. E. S. 521
Kerr, J. T. 275 Lannoo, M. J. 42, 46, 47, 50, 51
Kéry, M. 441 Laposata, M. M. 222
keystone species 26, 98 larvae 5, 6, 9, 12, 39, 40, 50, 56, 59f4.1, 61, 522
Kie, J. G. 193 acquatic 10, 15, 56
Kiesecker, J. M. 87 diets 73
Kilham, S. S. 79 toxins 487
Kimball, K. D. 87, 100 see also caecilians, larvae; frogs, larvae;
Kimmel, C. A. 496 salamanders, larvae
Kingsbury, B. 507 larval development 5, 57, 96, 304, 382, 510, 512
Kinkead, K. E. 471, 472, 474 larval stages 6, 29, 58, 59f4.1, 65, 101, 207, 252,
Kinne, O. 522 306, 288, 435
Kirlin, M. 284, 286, 293 aquatic 5, 8, 203, 374
Klump, G. M. 472 field enclosures 219
Knapp, R. A. 354 growth rates 423–4
Knutson, M. G. 282, 350 larviform 39
Köhl, M. 30 lateral line system, see neuromasts
Kouamé, N. G. 265, 268, 269, 271, 273, 274 Lauber, A. 474
Koussoroplis, A. M. 82 Lauer, A. 380
Kraaijeveld-Smit, F. J. L. 525 Laugen, A. T. 424
Krebs, C. J. 21, 22, 26, 30, 78, 321, 330, Laurance, W. F. 477
333, 447 Lavilla-Pitogo, C. R. 490
Kriger, K. M. 493, 496 Layman, C. A. 81
Kroll, A. J. 433, 442 leaf litter 92f6.3, 95, 100, 174, 175, 217, 218,
Kupfer, A. 169 220, 249, 250, 272, 517
Kupferberg, S. 72 plots 249–50, 253
Kutrup, B. 180 quadrats 250
Kyriakopoulou-Sklavounou, P. 144 species 223, 250
surveys 250–1, 252
labial folds 41, 42 leaf litterbags 59f4.1, 62–3, 237
laboratory experiments 29, 88, 101, 116, 124, Lee, A. K. 364
Index | 543
hierarchic 457, 458 movements 185, 195, 197, 199, 209, 215, 222–3,
likelihood 450, 451, 453, 454, 457, 458 303, 341, 356, 448, 453
multi-season 458 barriers to migration 346, 407–8, 419, 519
see also agar models; physical models between breeding sites 205, 516, 519
Mohr, J. R. 288 detection probabilities 452
moisture 209, 234, 237, 238, 239, 252, 309, effects of transmitters 198
310, 313, 314, 387, 402, 520, 521 patterns 342f19.2
environments 387, 388 see also frogs, movements; salamanders,
microenvironments 399 movements
soil 250, 251 MS-222 381
molecular ecology 407, 408, 423 mtDNA (mitochondrial DNA) 412–13, 417,
mollusks 167 419, 421, 422f22.5
monitoring programs/schemes 5, 55, 143, mud nests 12f1.2
145, 153, 158, 197, 239–40, 247, Mugglestone, M. A. 354
264, 275, 334, 388, 432, 433, Muller-Navarra, D. C. 82
434f23.1, 435, 473–4, 145, multistate models 453–4
152, 153, 197, 234, 247, 435, Murphey, T. G. 187, 197, 199
441, 481 muscles 106, 190, 366, 379, 496
design 239 mutation 412, 414, 417, 423
devices 368–9t20.1 Muths, E. 151, 292, 441, 456, 457, 459, 482
malformations 489 mycobacteria 482
see also frogs, monitoring programs Myers, N. 15
Monsen, K. J. 419 Mycobacterium lifl andii 482, 498
Moore, G. E. 339 Mycobacterium ulcerans 482
Moore, J. 363 myobatrachids 11, 151
Moore, J. D. 237, 238
morbidity 482, 488, 499–500 NAAMP protocol 283, 284, 287, 437, 438
Moriau, L. 263 Nagai, T. 373
Morin, O. J. 90, 93, 98, 204 Nagendra, H. 269
Morin, P. A. 423 Nagy, K. A. 375, 387
Morisita 330–1 Nanjappa, P. 135
Morisita, M. 311 Napolitano, G. E. 82
Morisita-Horn 333 Narendran, T. C. 27
morphology 168, 365 Narins, P. M. 287
larval 39, 44–5, 51 natterjack toads 353, 415, 415f22.1, 418, 420
tadpoles 40, 45–50 Natrix maura 512
variation 135 natural-cover surveys 250–1
see also caecilians, larval natural selection 156, 408
morphometry 306 Navas, C. A. 310, 370, 372, 373, 378, 388, 391, 400
Morrison, M. L. 25, 300 Naveh, Z. 339
mortality 215–16, 221, 409, 438, 452, 482, Nectophrynoides 14
483, 488, 489, 490–2, 497–8, Nectophrynoides occidentalis 14
499, 501, 522 Nelson, G. L. 288
caused by metals 115 nematode lungworms 500
and embryos 111 neophytes 60
high temperature 209 nested clade analysis 421, 422f22.5
study animals 169 nested designs 249, 257–8
and toe clipping 128 nests 149, 155
Morton, M. L. 145 artificial sites 149
Moseley, K. R. 220, 223 attendance 147–8, 155–6
Mossman, M. J. 282, 283, 437 construction 147, 148
Moulton, C. A. 237, 238 counts 143, 148–9, 153, 154, 157, 158
Mourão, G. 267 embryos 156
mouthparts, see oral apparatus desertion 155–6
546 | Index
sampling methods 197–8, 236, 238f13.4, satellite imagery 269, 339, 355433
239, 254, 450 Sauer, J. R. 474
seepage 9 Sayre, R. 264, 265, 266, 268, 269
skin 8, 131, 187, 189 Scallan, J. T. 438
stream 249 Scaphiopodidae 46, 47, 50
surgery 190 Scaphiopus 13, 98, 99, 237, 286, 290, 293
tails 5, 7, 131 Scaphiophus couchii 108, 152
terrestrial species 9, 169, 237, 249, 250–1, Scaphiopus holbrookii 98
439, 440, 441, 442, 449–50, Schabetsberger, R. 188, 189, 199
515, 518 Schaub, M. 255, 453, 454
toe clipping 129–30, 131, 133 Schindler, D. W. 87, 101
see also under individual species Schismaderma 49
Salamandra atra 9, 414 Schmidt, B. 272
air humidity 475, 475f25.2 Schmidt, B. R. 283, 286, 350, 439, 440, 441,
detection probabilities 474–5 449, 455, 457, 465, 466, 468,
Salamandra salamandra 9 469, 474
detection probabilities 449 Schmuck, R. 389
Salamandridae 498 Schneider, D. C. 24
Salek, L. 312 Schoener, T. W. 174
salinity 105, 112 Schotthoefer, A. M. 496
Salthe, S. N. 6 Schoutedenella xenodactyloides 353
Salmonella 498 Schwarz, C. J. 453
Salvidio, S. 268, 270, 274, 334 Schwarzkopf, L. 388, 390, 400
Sammouri, R. 41 scientific investigation 21, 24
sampling methods 26–7, 30, 31, 56–8, 205, cycle 23f2.1, 24
256, 323, 325, 374, 381–2, 388, goals 22–3
433, 456, 458, 466 applied 23; non-applied 23
amphibian larvae 55–6, 56t4.1, 60–1 objectives 23, 27–8
availability 470 effects and causes 23; relevance 27–8;
design 432, 435, 447–8, 458–9, 472, 477 time 28
food sources 73–8 Scinax boulengeri 153
larvae 252 Scolemorphidae 7
non-exposure 470–1, 472, 476 Scott, C. T. 30
plans 268–74 Scott, D. E. 94, 137, 205, 232
prey 167, 168, 169–70, 171, 172–4, 179–80 Scott, J. M. 321
sites 269, 271, 275, 283, 347, 353, 420, 436 Scott, N. J. 270, 271, 272, 274
size 433 Scott, W. A. 283
surgery 189–90, 190f11.1, 191 Seale, D. 65, 152
techniques 58, 59f4.1, 60–8, 80–1, 178–9, seasons 400, 401
229, 255, 260, 265, 267, 269, Seber, G. A. F. 253, 448, 452
270–2, 324, 327, 331, 407, 440 Seebacher, F. 388, 400
time (seasonal) 270, 324 Seigel, R. A. 511, 522
see also adaptive cluster sampling; area-based seine 61–2, 64
surveys; ARS; coverboards; selection 424
distance sampling; frogs, sampling Semlitsch, R. D. 93, 100, 101, 109, 129, 187,
methods; MCS; photography; 188, 189, 191, 193, 197, 198, 199,
RA; radiotelemetry; radiotracking; 205, 206, 207, 216, 217, 221, 222,
salamanders, sampling methods; 223, 230, 235, 340, 341, 349, 351,
traps 356, 487, 507, 510, 514, 516
Sanzone, D. M. 80 Senar, J. C. 22
Saprolegnia 131, 486, 522 Serenidae 7, 42–3
Sarasin, P. 41, 41f3.1 larvae 43, 44
Sarasin, S. 41, 41f3.1 Service, P. M. 131
Sargent, L. G. 287 Seutin, G. 273
552 | Index
thermometers 109, 303, 314, 366–7, 370, 371, transformation see abundance, transformation
373–4 transients 470
thermoregulation 366, 376, 377, 388 transmitters, see radiotransmitters
Thibaudeau, D. G. 39 traps 64–5, 174–5, 217, 229, 232, 235, 236,
Thomas, C. 389, 402 238, 288, 325–6, 447, 467
Thomas, C. D. 388, 403 aquatic 234
Thomas, D. L. 300, 301 active 229, 236, 237, 238, 239, 241
Thomas, E. 45 capture 235
Thomas, L. 253, 273 funnel 220, 229, 230, 230f13.1, 231, 232,
Thompson, G. G. 326 233f13.2, 234, 470–1
Thompson, S. K. 253, 432, 435 minnow 64, 234, 270, 470–1
Thoms, C. 60 passive 229–33, 233f13.3, 234, 235–6, 239
Thorpe, J. H. 73 pitfall 220, 229, 230, 231, 231f13.2, 232,
Thorson, T. B. 364, 372 233f13.3, 236, 270, 352
tiger salamanders 8, 106, 108, 198, 353 Trauth, J. B. 511
time-constrained surveys 248 Travis, J. 65, 88, 93
Timm, R. M. 263 tree (overstory) measurements 310–12
Tischendorf, L. 349 treefrogs 100–1, 237, 238, 270, 419, 437, 467
toads 5, 341, 498, 498 Cuban 511
acquatic larvae 12 detection probabilities 469, 472
burrows 376 gray 237
egg masses 151 pine barren’s 111
eggs 12, 13–14 trees 421, 514, 518
fossorial species 13–14, 237 trematodes 105, 487, 522
larvae 40 Trenham, P. C. 283, 340, 341, 351
physical models 378f20.2, 400 Triplehorn, C. A. 172
radiotransmitters 197 Triturus alpestris 352
reproduction 11–12 Triturus cristatus 135, 236, 419
sampling methods 187, 233, 435 Triturus marmoratus 419
tadpoles 13–14, 50 Triturus vulgaris 135
see also under individual species trophic niche/status:
Tobler, W. R. 353 breadth 176–7
Todd, B. D. 204, 206, 223, 231, 233, 235, 236 larval amphibians 71–2, 74, 77, 78
Todd, M. J. 284, 286, 288, 292 overlap 177
toe clipping 32, 123, 124, 125, 220, 273 prey availability 174
alphanumeric code for 126–7t8.2 relationships 167
and digit–regeneration 128, 128t8.3, 129 tadpoles 72–3
and ethics 129–30 stable-isotope data 81
see also markings; frogs, toe clipping; structure 176–7
salamanders, toe clipping tropical amphibians 145, 223, 421
Toft, C. A. 167, 168, 174, 1798, 180, 247 tropical frogs 252, 518
Toledo, R. C. 373, 375, 389 Trueb, L. 6, 40, 44, 49, 55, 73, 168, 170, 174,
Tonge, S. 522 285, 365
total dissolved solids, see TDS TSS (total suspended solids) 113
total suspended solids, see TSS Tucker, G. 31
tooth rows 47, 49, 50 tunnels 520–1
Townsend, D. S. 149, 151, 252 turbidity 114
toxicology 100 Turner, M. G. 339
toxins 188, 487, 497, 498 turtles 106
tracking 388, 395; see also radiotracking Typhlomolge 8
Tracy, C. R. 365, 366, 367, 372, 376, 377, 379, Typhlonectidae 6–7
387, 388, 389, 390, 400, 401, 402
trailing devices 133–4, 134t8.6 Uganda 250, 270
Trammell, C. 496 Ultsch, G. R. 106, 108, 112, 374
transect surveys, see area-based surveys, transect umbrella species 26
Index | 555
Weir, L. A. 26, 143, 152, 282, 283, 286, 287, 437 Woolley, H. P. 131, 133
Wellborn, G. A. 306 Woosley, L. B. 154, 155
Wells, K. D. 5, 6, 144, 145, 151, 152, 153 World Animal Health Organization, see OIE
Welsh, H. H. 300, 304, 307, 439, 440 World Charter for Nature 17
Wengert, G. M. 134 worms 179
Werner, E. E. 58, 60, 61, 143, 174, 340, 436 Wright, A. A. 282
Werneria 11 Wright, A. H. 282
Western Alps 26 Wright, K. M. 482, 487, 492, 500
wetlands 55, 58, 60, 65, 88, 93, 96, 105, 110, Wu, Z. J. 169
113, 143, 144–5, 152, 153, 185, Wuycheck, J. C. 76
232, 233–4, 236, 237, 239, 341, Wygoda, M. L. 387, 389, 390, 402
433, 438, 510–15, 515f27.1, 516, Wyman, R. L. 314
517–18
Weyrauth, S. L. 514 xenobiotics 364
Whiles, M. R. 72, 79, 82, 263, 501 Xenohyla truncata 168
Whitaker, B. R. 482, 487, 492, 500 Xenopus 10, 49, 168, 421
White, G. C. 193, 453, 454, 455, 457, 476 Xenopus laevis 112, 375
Whiteman, H. H. 189, 190, 191 xerophilic frogs 373
Whitfield, S. M. 334
Whitford, W. G. 364 Yamamuro, M. 81
Whitlock, M. C. 419 Yamasaki, M. 237
Whunam, J. 31 Yamashita, K. 44
Wieczorek, A. M. 352, 353 yellow-legged frog 354
Wiens, J. A. 300, 339, 341 Yersinia 498
Wilbur, H. M. 29, 88, 90, 93, 98, 216, 222, 237 Yilmaz, Z. C. 180
Wiley, R. H. 465, 469, 474 Yip, M. J. 482
Williams, A. K. 216 Yoccoz, N. G. 30, 466, 473, 476
Williams, B. K. 95, 143, 154, 155, 269, 270, Yosemite toads 145, 419
272, 431, 432, 433, 434, 435, Young, B. 274, 275
447, 448, 450, 451, 453, 470, Young, C. A. 237
472, 477 Young, J. E. 153, 387, 389
Williams, T. 497 Yuan, J. S. 495
Wilson, E. O. 15 Yurewicz, K. L. 143, 216, 222
wind 303–4
Windmiller, B. S. 132, 154 Zahradnik, D. 312
Winkler, C. 134 Zamboni, D. S. 173
With, K. A. 348 Zamudio, K. R. 352, 353
Withers, P. C. 326, 375, 387, 399, 402 Zane, L. 416
Wolff, J. O. 21, 22, 26 Zanini, F. 346
Wood, K. V. 453 Zar, J. H. 179
wood frogs 66, 108, 111, 150f9.2, 198, 419 Zeisset, I. 423
wooden boards 237 Zimmermann, N. E. 346, 347
Woodhams, D. C. 387, 388 zooplankton 93, 96, 512
Woolbright, L.L. 254, 282 Zucchini, W. 476