You are on page 1of 202

EFFECT OF ANTHROPOGENIC DISTURBANCE AND

LANDSCAPE STRUCTURE ON BODY SIZE, DEMOGRAPHICS, AND

CHAOTIC DYNAMICS OF SOUTHERN HIGH PLAINS AMPHIBIANS

by

MATTHEW JAMES GRAY, B.S., M.S.

A DISSERTATION

IN

WILDLIFE SCIENCE

Submitted to the Graduate Faculty


of Texas Tech University in
Partial Fulfillment of
the Requirements for
the Degree of

DOCTOR OF PHILOSOPHY

Approved

May, 2002
©2002, Matthew J. Gray
ACKNOWLEDGEMENTS

Foremost I thank my major advisor, Dr. Loren M. Smith, for providing the

opportunity to continue my education and professional development at Texas Tech

University (TTU). Dr. Smith not only provided me with this great academic opportunity:

his guidance honed me as a scientist. I now am capable of perceiving the edge of science

and conducting research independently. Dedicating my dissertation to him is not enough

(see page v). Without question, my future research endeavors will be a product of his

mentorship. I hope my science will make him proud. Thank you so much, Loren!

Next, I thank my minor professor. Dr. Linda J. S. Allen of the Department of

Mathematics and Statistics at TTU. Dr. Allen sculptured my quantitative abilities and

burgeoned my interest in biomathematics. I now understand and appreciate the

usefulness of mathematics in explaining ecological phenomena. Indeed, her tutelage

changed my perception of science. Thank you. Dr. Allen, for making me quantitatively

literate. I also want to thank my other instructors in the math department: Drs. Kamal

Chanda, Wijesuriya Dayawansa, William Gustafson, Ira Wayne Lewis, and James Surles.

Each of these individuals helped me develop important quantitative skills.

In addition to Drs. Smith and Allen, I thank my committee members: Drs. Lou

Densmore (Department of Biological Sciences, TTU), David Haukos (U.S. Fish and

Wildlife Service), and David Wester (Department of Range, Wildlife and Fisheries,

TTU). These individuals provided incredible personal and professional guidance during

my four years at TTU. My dissertation is a testament of their devotion and support.

Thank you for making my doctoral pursuit a stimulating and intellectual experience!

n
I also want to thank Dr. James Surles of the Department of Mathematics and

Statistics at TTU for quantitative guidance on se\'eral anal> ses in my dissertation and

numerous stimulating discussions on experimental design and statistics. I am grateful to

Drs. Kevin Pope, Mark Wallace, and Gene Wilde of the Department of Range. Wildlife,

and Fisheries at TTU and Dr. Ray Semlitsch of the University of Missouri (Di\ision of

Biological Sciences) for editorial comments on my dissertation. I also thank the

following TTU graduate students for reviewing chapters: J. Avey, J. Brunjes. K. Brunjes,

and S. Sebring. I thank Stephanie Dupree of the Geospatial Technology Lab at TTU for

helping construct Figure 2.1. I am grateful to Dr. Ernest Fish, Cherise Ginsburg. and

Raquel Leyva for instructing me in ERDAS®, ESRI®, and FRAGSTATS® software.

I am deeply indebted to the landowners of my study playas: A. Arajo, L. Crum, K.

Gabel, B. Harrell, R. Hunter, B. Kimbrough, M. Marley, S. McPhearson, J. Merritt, G.

Olson, S. Olson, D. Schulte, E. Schulte, C. Sellers, E. Verett, and M. Wilmeth. Their

gratefulness allowed me to take my research thoughts to the field. I also thank M. Cerda.

T. Grand, L. Lloyd, and A. Wallace of the U.S. Department of Agriculture (USDA) Farm

Service Agency (FSA) for assistance in locating aerial slides and FSA Farm Folders for

my study playas.

This study would not have been possible without field and lab assistance from

many individuals. I gratefully thank: A. Anderson, J. Anderson, J. Avey, R. Brenes, W.

Dayawansa, E. Edwards, J. Emrich, D. Garza, D. Ghioca, W. Huskisson, A. Kauffman,

R. Kostecke, P. Lemons, A. Marshall, R. McCaffrey, D. Miller, K. Offill, J. Pautz, C.

Perchellet, R. Phillips, A. Quarels, D. Ray, W. Redell, S. Sebring, C. Shavlik, and C.

ill
Smith. I especially am indebted to Roberto Brenes (now at the University of Texas-

Tyler)—my right-hand man and savior. Roberto assisted me in the field for two years.

Honestly, I do not understand how he tolerated me—and, I am still quite amazed that we

survived. Roberto was more than a technician to me—he was a colleague and great

friend. I am confident my research would have suffered without him. I thank him for all

his dedication, patience, and herpetological insight. I doubt if I will ever be able to repay

him equally. Muchisimo gracias y pura vida, Sementall

This study was financially supported primarily by the Caesar Kleberg Foundation

for Wildlife Conservation and the Department of Range, Wildlife, and Fisheries

Management at TTU. Student research grants also were provided by the Society of

Wetland Scientists and Ducks Unlimited, Inc., and Home Depot donated buckets for

pitfall traps. I thank Dr. Gene Wilde of TTU for use of various sampling equipment and

Program CANOCO for multivariate ordination analyses. I also am very grateful to Dr.

Rob Mitchell of TTU for allowing me use his ATV on days when the west Texas roads

were impassable. Lastly, I thank Dr. Nick Parker of TTU and Dick Lubke of the Texas

Parks and Wildlife Department for use of aquariums and a PIT tag scanner, respectively.

Finally, I am grateful for the incredible support of my family: James S. Gray

(father), Scott H. Gray (brother), Sharon L. Malloy (mother), and Carly L. Pratt (sister).

No one gets anywhere by himself or herself My family has been an amazing nucleus of

love and support. Without their encouragement, my 12-year academic journey would

have been impossible. I also thank my very dear friend, Jessica L. Emrich. Her love and

patience gave me the strength to succeed.

IV
I dedicate my dissertation to my four academic mentors. In order of academic

interaction: Drs. Harold H. Prince (Michigan State Universit>). Richard M. Kaminski

(Mississippi State Universit>'). Francisco J. Vilella (Mississippi Cooperati\ e Fish and

Wildlife Research Unit), and Loren M. Smith (Texas Tech University). Dr. Prince

trained me in the scientific methods. Dr. Kaminski fueled m\ passion for wetlands and

research and de\ eloped my wxiting skills. Dr. \ ilella sparked my interest in conservation

biology, and Dr. Smith directed me in ecological theor>' and landscape ecology. These

individuals made me the person and scientist that I am toda}'. Indeed, I am indebted.

Thank \ ou so much gentlemen!

\'
TABLE OF CONTENTS

ACKNOWLEDGEMENTS ii

ABSTRACT ix

LIST OF TABLES xi

LIST OF FIGURES xiii

CHAPTER

I. INTRODUCTION 1

II. EFFECT OF LANDUSE AND RAINFALL ON POSTMETAMORPHIC


BODY SIZE OF PLAYA AMPHIBIANS 5

Abstract 5
Introduction 6
Study Area and Amphibians 9
Methods 13
Experimental Design and Units 13
Sampling Techniques 14
Terrestrial capture 14
Biological processing 15
Statistical Analyses 15
Results 16
Landuse Type and Year Interaction Effects 16
Adults 16
Subadults 17
Metamorphs 17
Landuse Type Main Effect 19
Adults 19
Subadults 19
Metamorphs 19
Year Main Effect 21
Adults 21
Subadults 21
Metamorphs 23
Discussion 23
Landuse Effect 23
Year Effect 27
Competing Body Size Hypotheses 29

VI
Conclusions ^^

III. EFFECT OF LANDUSE AND YEAR ON POPULATION


DEMOGRAPHICS OF SOUTHERN HIGH PLAINS
AMPHIBIANS 32

Abstract 32
Introduction 33
Methods 36
Results 41
Abundance and Diversity 41
Source-Sink Dynamics 43
Discussion 45
Landuse Effect 47
Year Effect 52
Source-Sink Dynamics 53
Conservation Implications 54

IV. INFLUENCE OF LANDSCAPE STRUCTURE ON COMMUNITY


COMPOSITION AND RELATIVE ABUNDANCE OF
AMPHIBIANS 56

Abstract 56
Introduction 57
Methods 59
Results 63
Discussion 76
Conservation Implications 78

V. TEMPORAL NICHE PARTITIONLNG IN SOUTHERN HIGH PLAINS


AMPHIBIAN POPULATIONS 80

Abstract 80
Introduction 81
Methods 83
Results 85
Discussion 91

VI. CHAOS AND REGULATING MECHANISMS IN A SOUTHERN HIGH


PLAINS AMPHIBIAN ASSEMBLAGE 106

Abstract 106
Introduction 107
Methods 110

Vll
Results 117
Discussion 128
Ecological and Mathematical Implications 133

VII. CONCLUSION 135

LITERATURE CITED 140

APPENDIX

A. PERCENT COVERAGE OF COVER TYPES IN LANDSCAPES


SURROUNDLNG PLAYA WETLANDS 167

B. ANOVA TABLES FOR CHAPTER III 169

C. AGE CLASS ANOVA TABLES FOR CHAPTER III 172

D. CROPLAND AND GRASSLAND LANDSCAPES 175

Vlll
ABSTRACT

Amphibian populations are declining globally. Anthropogenic disturbance of

landscapes surrounding wetlands may affect fitness, demographics, and dynamics of

amphibian populations. Spatial positioning and relati\'e connectedness of wetlands also

may influence population demographics. Thus, I examined the effect of anthropogenic

landscape use (cultivation vs. grassland) and structure on postmetamorphic body size (a

fitness correlate), demographics, and dynamics of amphibians at 16 playa wetlands on the

Southern High Plains (SHP) of Texas during 1999 and 2000. Amphibian populations

were monitored using drift fence and pitfall traps, landscape structure was quantified

using spatial analysis software, and dynamics were assessed using difference equations.

Postmetamorphic body size of all amphibian species and age classes generally

was greater at grassland than cropland playas, and in 1999 (i.e., a wetter year) than 2000.

Abimdance of New Mexico and plains spadefoots (Spea multiplicata and S. bombifrons)

generally was greater at cropland than grassland playas, and greater for barred tiger

salamanders {Ambystoma tigrinum mavortium) in 1999 than 2000. Mean daily

abundance of amphibians also was positively related to landscape structure indices

representing geometric complexity and spatial positioning of wetlands. In general, as

landscapes became more complex (e.g., numbers of edges increased) and inter-playa

distance decreased, mean daily abundance of amphibians increased. Additional

demographic anah ses indicated that temporal niche partitioning existed in SHP

amphibian populations; however, no differences existed between landuses. Lastly,

biological chaos in the amphibian assemblage existed at 1 of 8 cropland and 7 of 8

ix
grassland playas. A stochastic density-dependent Ricker function predicted chaotic

dynamics most accurately.

Anthropogenic disturbance surrounding wetlands affects body size,

demographics, and dynamics of SHP amphibians. Spatial positioning of wetlands and

landscape complexity may be as or more important than general landuse in affecting

amphibian demographics. Armual differences in body size and abimdance suggest

rainfall may be important in influencing amphibian populations. Although spadefoot

abundance was positively influenced by anthropogenic disturbance, I recommend

retention and restoration of grasslands surrounding playa wetlands because landscape

cultivation decreased body size and altered amphibian demographics and dynamics from

an undisturbed state. These results have important implications in conservation biology,

landscape ecology, and basic ecological and mathematical theory.


LIST OF TABLES

2.1 Adult body size of New Mexico spadefoot {Spea multiplicata, NSF), plains
spadefoot {S. bombifrons, PSF), Great Plains toad {Bufo cognatus, GPT),
and barred tiger salamander {Ambystoma tigrinum mavortium, BTS) between
landuse types and years (fixed factorial effects) at 16 playa wetlands (random
nested effect) on the Southern High Plains of Texas, April-September 1999
and 2000 18

2.2 Subadult body size of New Mexico spadefoot {Spea multiplicata, NSF), plains
spadefoot {S. bombifrons, PSF), and Great Plains toad {Bufo cognatus, GPT)
between landuse types and years (fixed factorial effects) at 16 playa wetlands
(random nested effect) on the Southern High Plains of Texas, April-
September 1999 and 2000 20

2.3 Metamorph body size of New Mexico spadefoot {Spea multiplicata, NSF), plains
spadefoot {S. bombifrons, PSF), Great Plains toad {Bufo cognatus, GPT), and
barred tiger salamander {Ambystoma tigrinum mavortium, BTS) between
landuse types and years (fixed factorial effects) at 16 playa wetlands (random
nested effect) on the Southern High Plains of Texas, April-September 1999
and 2000 22

3.1 Relative daily abundance and diversity of amphibians between landuse types
and years and among species at 16 playa wetlands on the Southern High
Plains, Texas, May-October 1999 and April-August 2000 42

3.2 Relative daily abundance of age classes between landuse types and years for
amphibian species that differed significantly in daily abundance (see Table
C.2) at 16 playa wetlands on the Southern High Plains, Texas, May-October
1999 and April-August 2000 44

4.1. Pearson coefficients of correlations and p-values (located beneath coefficients)


from parametric testing {t distribution) of linear associations between
landscape metrics^ at 16 playa wetlands on the Southern High Plains,
Texas, 1999 and 2000 64

4.2 Intra-set correlations of landscape* metrics associated with the first two axes
generated by a canonical correspondence analysis of mean daily abundance of
New Mexico spadefoot {Spea multiplicata), plains spadefoot {S. bombifrons),
barred tiger salamander {Ambystoma tigrinum mavortium), and Great Plains toad
{Bufo cognatus) at 16 playa wetlands on the Southern High Plains, Texas, 1999
and 2000 66

XI
4.3 Univariate Pearson coefficients of correlations between mean daily abundance
of amphibians and landscape* metrics at 16 playa wetlands on the Southern
High Plains, Texas, 1999 and 2000 70
4.4 Multiple linear regression between mean daily abundance of amphibians and
landscape* metrics at 16 playa wetlands on the Southern High Plains, Texas,
1999 and 2000 74

6.1 Stability of amphibian assemblage trajectories at 16 playa wetlands during 16


May-17 October 1999 and 19 April-18 August 2000 on the Southern High
Plains, Texas 122

6.2 Occurrence of chaos between landuses and models in amphibian assemblages


at 16 playa wetlands during 16 May-17 October 1999 and 19 April-18
August 2000 on the Southern High Plains, Texas 126

6.3 Least-squares estimates of parameters for the discrete stochastic Ricker


equation* from amphibian assemblages at 16 playa wetlands during 16
May-17 October 1999 and 19 April-18 August 2000 on the Southern High
Plains, Texas 127

6.4 Iterative and least-squares estimates of parameters for deterministic difference


equations used to model amphibian assemblages at 16 playa wetlands during 16
May-17 October 1999 and 19 April-18 August 2000 on the Southern High
Plains, Texas 129

6.5 Estimates of parameters between landuses for difference equations used to


model amphibian assemblages at 16 playa wetlands during 16 May-17
October 1999 and 19 April-18 August 2000 on the Southern High Plains,
Texas 130

A. 1 Percent coverage of cover types in landscapes* surrounding 16 playa wetlands


between landuses on the Southern High Plains, Texas, 1999 and 2000 168

B. 1 Statistics for ANOVAs testing differences in relative daily abundance and


diversity of amphibians between landuse types (cropland vs. grassland) and
years and among species at 16 playa wetlands on the Southern High Plains,
Texas, May-October 1999 and April-August 2000 170

C.l Statistics for ANOVAs for species (by age class) that differed significantly in
relative daily abundance between landuse types (cropland vs. grassland) or years
(see Table B.l) at 16 playa wetlands on the Southern High Plains, Texas,
May-October 1999 and April-August 2000 173

xii
LIST OF FIGURES

2.1 Location of study playa wetlands on the Southern High Plains of Texas,
U.S.A 12

3.1 Effect of landuse (top graph) and year (bottom graph) on frequency of directional
movement of amphibian species at 16 playa wetlands on the Southern High
Plains, Texas, May-October 1999 and April-August 2000 46

4.1 Canonical correspondence analysis of mean daily capture (natural-log transformed)


of amphibians and uncorrelated metrics in landscapes (i.e., 2,830-ha circular plot)
associated with 16 study playas on the Southern High Plains, Texas, 1999 and
2000 67

4.2 Simple linear regression and 95% confidence bands of mean daily abundance
(natural-log transformed) of New Mexico spadefoot {Spea multiplicata) and
percent aerial coverage of playa wetlands (PP), interspersion/juxtaposition index
(IJI, McGarigal and Marks 1995:103), edge density (ED, McGarigal and Marks
1995:106), playa edge density (PED [i.e., mean number of edges to cross from
study playa to surrounding playas), landscape shape index (LSI, McGarigal and
Marks 1995:109), and landuse richness (LR, McGarigal and Marks 1995:119) in
landscapes (i.e., 2,830-ha circular plot) associated with 16 study playas on the
Southern High Plains, Texas, 1999 and 2000 71

4.3 Simple linear regression and 95% confidence bands of mean daily abundance
(natural-log transformed) of plains spadefoot {Spea bombifrons) and percent
aerial coverage of playa wetlands (PP), interspersion/juxtaposition index (IJI,
McGarigal and Marks 1995:103), edge density (ED, McGarigal and Marks
1995:106), and landscape shape index (LSI, McGarigal and Marks 1995:109) in
landscapes (i.e., 2,830-ha circular plot) associated with 16 study playas on the
Southern High Plains, Texas, 1999 and 2000 73

5.1 Time series plots for pairs of amphibian species residing at 16 playa wetlands
in 2 landuses and years on the Southern High Plains of Texas 87

5.2 Percent occurrence pairs of standardized species-specific linear trajectories (see


Figure G.l) exceeded each other during 3 sequential and equal time intervals
{n=22 and 17 days for 1999 and 2000) within 2 landuses and years on the
Southern High Plains, Texas 92

6.1 Time series plots of amphibian assemblage dynamics at 8 playa wetlands during
16 May-17 October 1999 on the Southern High Plains, Texas 118

Xlll
6.2 Time series plots of amphibian assemblage dynamics at 8 playa wetlands during
19 April-18 August 2000 on the Southern High Plains, Texas 120

D.l Anthropogenic and natural cover types in landscapes (i.e., 2,830-ha [3-km radius]
circular plots) surrounding 16 playa wetlands on the Southern High Plains, 1999
and 2000 176

xiv
CHAPTER I

INTRODUCTION

Amphibians are declining globally and anthropogenic disturbance of landscapes

surrounding wetlands has been implicated as a primary cause (Blaustein et al. 1994.

Houlahan et al. 2000). Deforestation and agricultural cultivation in watersheds of

wetlands have been considered ultimate reasons for decline (Alford and Richards 1999).

Interestingly, few data exist in the United States that suggest amphibian populations are

negatively affected by cultivation. Two of the three studies (Knutson et al. 1999,

Kolozsvary and Swihart 1999) indicate that amphibian populations can be elevated in

wetlands surrounded by cropland, and one (Anderson et al. \999a) documented no

differences between disturbed and undisturbed wetlands. In Canada and Europe, most

studies suggest that landscape cultivation significantly decreases amphibian abundance

and species richness, and negatively affects fitness correlates (Laan and Verboom 1990,

Oldham and Swan 1991, Hecnar and M'Closkey 1996, Bonin et al. I997a,b). However,

most of these studies were conducted using call surveys during the breeding season,

which have less demographic resolution than other methods (e.g., pitfall sampling, Heyer

et al. 1994). Drift fence and pitfall traps facilitate the most comprehensive examination

of dynamics in non-arboreal amphibian populations (Heyer et al. 1994). Thus, previous

studies may not have sufficiently documented population responses to anthropogenic

disturbance, particularly considering that few sampling efforts were performed outside

the breeding season.


It has been hypothesized that anthropogenic disturbance affects amphibian

populations by altering abiotic and biotic conditions in and around wetlands (Semlitsch

2000). Anthropogenic disturbance can negatively affect water qualit>, which might

affect trophic structuring in the aquatic environment, and ultimately growth and survival

of larvae (Boone and Semlitsch 2001). Agricultural chemicals also can reduce

immunocompetence, mobility, and survival of larvae (Carey et al. 1999, Bridges and

Semlitsch 2000). Sedimentation in wetlands can decrease hydroperiods (Luo et al. 1997).

thus presumably breeding opportunities for adults and de\ elopmental time of larvae.

Finally, landscape disturbance may cause direct mortality, reduce terrestrial habitat and

subterranean retreats, and affect movement of postmetamorphic amphibians among

spatially structured populations (Semlitsch 2000).

Amphibians are important ecological entities of aquatic and terrestrial ecosystems

(Wilbur 1984). Larval amphibians can significantly affect trophic structuring and

nutrient cycling in wetlands via consumption and excrement of fine and course

particulate organic matter and micro- and macro-invertebrates (Scale 1980, Hoff et al.

1999, Petranka and Kennedy 1999). Also, amphibian larvae are prey for various

invertebrate and vertebrate species (Skelly 1997). Postmetamorphic amphibians consume

numerous invertebrates, some of which are agricultural pests, and they are food items for

various terrestrial predators (Duellman and Trueb 1994, Anderson et al. 19996). Larval

and postmetamorphic amphibians also readily absorb chemicals via osmoregulation

(Boutilier et al. 1992, Shoemaker et al. 1992), which can affect their physiology and

bioaccumulate (Semlitsch 2000). Consequently, amphibians are considered good


indicators of ecosystem health (Beebee 1996). Thus, it is important to discern the effect

of human disturbance on dynamics, demographics, and fitness correlates of amphibians

(Beebee 1996).

Amphibians exist primarily in spatially structured playa wetlands on the Southern

High Plains (Bolen et al. 1989). Cultivation and grassland surround most playas; thus,

they are ideal systems to comparatively and simultaneously investigate the effect of

human disturbance and landscape structure on amphibian populations. Therefore, I

proposed the following research. My goal was to determine the effect of anthropogenic

disturbance and landscape structure on body size and population demographics of

Southern High Plains amphibians. I accomplished this goal through 5 simultaneous

studies at 16 playa wetlands in 2 anthropogenic landuse types (cultivation and grassland)

over 2 years (1999 and 2000, 4 playas per landuse per year). My specific research

objectives were to determine the effect of (1) landuse and yearly rainfall on

postmetamorphic body size (a fitness correlate) of amphibians (Chapter II), (2) landuse

and year on demographics of amphibians (Chapter II), (3) landscape structure on

community composition and relative abundance of amphibians (Chapter IV), (4) landuse

on temporal niche partitioning and chaotic dynamics in amphibian communities

(Chapters V and VI), and to (5) identify potential regulating mechanisms in an amphibian

assemblage (Chapter VI). I used pitfall sampling, the geographic information system and

associated software, and difference equations to investigate these objectives. Results

presented herein provide insight into the effect of anthropogenic disturbance and

landscape structure on amphibian populations and mechanisms structuring their


communities. These results have important implications in amphibian and conservation

biology, landscape ecology, applied and theoretical mathematics, and basic ecological

theo^>^
CHAPTER II

EFFECT OF LANDUSE AND RAINFALL ON POSTMETAMORPHIC

BODY SIZE OF PLAYA AMPHIBIANS

Abstract

Postmetamorphic body size of amphibians is positively related to evolutionary

fitness. Landuse surrounding wetlands and rainfall may affect body size by influencing

proximate factors (e.g., hydroperiod, food resources) that affect larval and

postmetamorphic growth rates. I hypothesized that cultivation of terrestrial habitats

surrounding wetlands would affect body size of amphibian, but the magnitude of its

effect would depend on yearly differences in rainfall. I compared postmetamorphic body

size between two landuse types (cultivation versus grassland) surrounding playa wetlands

and two years (1999 versus 2000) with different rainfall (1999 > 2000) for 4 species

{Spea multiplicata, S. bombifrons, Bufo cognatus, and Ambystoma tigrinum mavortium)

and 3 age classes (metamorph, subadult, and adult) of amphibians on the Southern High

Plains of Texas. Sixteen playas (4 per landuse type per year) were partially enclosed with

drift fence and pitfall traps, and mass and snout-vent length (SVL) measured from a

subsample of captured individuals {n = 5,188). Mass and SVL generally were greater at

grassland than cropland playas for all species and age classes. Mass and SVL also were

greater in 1999 (i.e., the wetter year) than in 2000 for most species and age classes.

Moreover, the effect of landuse type and year on postmetamorphic body size generally

was non-additive; cultivation had a greater negative effect on body size in 2000 (i.e., the

drier year) than in 1999 for some of the age classes and species. Differences in body size
between landuse types and > ears may have been related to densit>' independent and

dependent factors, such as reduced wetland hydroperiod, food availabilit\\ and predators

or increased chemicals and density of conspecifics in cropland pla>as. Anthropogenic

modifications of landscapes surrounding wetlands may negati\eh influence

postmetamorphic body size of amphibians, especially during drier \ears.

Introduction

Postmetamorphic body size of amphibians can influence e\olutionary fitness b>

affecting survi\'al and reproduction (Wilbur 1984, Werner 1986, Semlitsch et al. 1988).

Probability of survival and reproduction in amphibians is positiveh related to

postmetamorphic body size (Smith 1987, Berven 1990. Goater 1994, Morey and Reznick

2001). Large individuals may have greater survival than small indi\ iduals, because they

can use a greater moisture range of microhabitats (Newman and Dunham 1994). can

consume a greater range of prey sizes (Flowers and Graves 1995. Newman 1999). are

more efficient foragers (Newman 1999), and have lower specific metabolic rates

(Hutchison et al. 1968, Ultsch 1973, 1974). Large individuals also take more time to

reach critical levels of dehydration than small individuals (Newman and Dunham 1994).

In addition, large individuals have greater sprint speeds, jumping ability, and endurance

than small individuals (Taigen and Pough 1981, John-Alder and Morin 1990, Goater et

al. 1993, Beck and Congdon 2000). Mating success of male and female amphibians also

may be positively related to body size (Berven 1981, Verrell 1982, 1985. Howard and

Young 1998, Marco et al. 1998). For territorial amphibians, territory size and quality

(i.e., density of food resources) can be positively correlated with body size (Howard
1978, Mathis 1990, Gabor 1995). Large females also may exhibit greater fecundity than

small females and spawn multiple times (Clarke 1974, Howard 1978, Berven 1982,

Semlitsch 1985a, Krupa 1986, 1994, Fontenot 1999). Finally, large individuals may

reproduce at an earlier age than small individuals (Berven and Gill 1983, Smith 1987.

Bervenl990, Scott 1994).

Postmetamorphic body size of amphibians can be affected by duration of larval

development and pre- and post-metamorphic conditions that differentially affect growth

and mortality rates of individuals in aquatic and terrestrial habitats (Wilbur and Collins

1973, Werner 1986). Generally, postmetamorphic body size is correlated with body size

at metamorphosis (Smith-Gill and Berven 1979. Semlitsch et al. 1988, Scott 1994).

Wetland hydroperiod, water quality, and density of food resources, conspecifics, and

predators can individually or synergistically affect body size at metamorphosis (Travis

1984, Wilbur 1984, Semlitsch 1987fl, Alford and Harris 1988, Pfennig et al. 1991,

Bradford et al. 1992, Werner and Anholt 1993). Postmetamorphic body size also is

affected by density of food resources and conspecifics in the terrestrial landscape (Goater

1994, Morey and Reznick 2001). Indeed, small body size at metamorphosis may be

compensated for by high density of terrestrial food resources (as theoretically suggested

by Werner 1986), allowing small individuals to attain a body size at first reproduction

similar to individuals that were larger at metamorphosis (e.g.. Goater 1994, Morey and

Reznick 2001).

Landuse type surrounding wetlands can affect larval and postmetamorphic

amphibians (Hecnar and M'Closkey 1996, Bonin et al. 1997a,b, Dodd 1997, Alford and
Richards 1999, Semlitsch 2000), presumably by influencing some of the aforementioned

mechanisms that affect growth and mortality rates of individuals in aquatic and terrestrial

environments (Werner 1986). Landscape cultivation (i.e., arable cropland) may localize

amphibians in wetlands, resulting in unexpected positive species associations or high

densities (Knutson et al. 1999, Kolozsvary and Swihart 1999, Chapter III).

Consequently, landscape cultivation may affect postmetamorphic body size of

amphibians by affecting density of conspecifics (Oldham 1985). Cultivation of the

watershed also can increase sedimentation in wetlands, which decreases hydroperiods

(Martin and Hartman 1987, Corn and Bury 1989, Luo et al. 1997), and may reduce

duration of larval development (Brady and Griffiths 2000). Insecticides and herbicides

associated with agricultural practices may affect postmetamorphic body size by reducing

available food resources. Similarly, agricultural chemicals (e.g., nitrates, ammonia,

organophosphates) can bioaccumulate and reduce foraging activity and growth of larval

amphibians (Hall and Kolbe 1980, Baker and Waights 1993, 1994, Hecnar 1995,

Semlitsch et al. 1995), and perhaps postmetamorphic body size.

Rainfall can affect growth rates of larval and postmetamorphic amphibians

(Tinsley and Tocque 1995, Camp et al. 2000) and potentially postmetamorphic body size

and condition (Bruce and Hairston 1990, Reading and Clarke 1995). Rainfall can

increase pond hydroperiod hence growth rates and duration of larval development

(Semlitsch et al. 1996, Reading and Clark 1999, Camp et al. 2000). Rainfall also can

increase prey availability in the terrestrial landscape (Dimmitt and Ruibal 1980a, Jaeger

8
1980). Moreover, rainfall reduces probability of desiccation and increases foraging

activity for amphibians (Jaeger 1980, Grover 1998, Newman 1999).

No studies have been conducted in North America examining the effect of

landuse type on postmetamorphic body size of amphibians. I hypothesized that

cultivation of terrestrial landscapes surrounding wetlands would negatively affect body

size of amphibians, but the magnitude of its effect would depend on rainfall. Thus, my

objective was to measure the effect of two landuse types (i.e., cultivation and grassland

[control]) over two years (i.e., 1999 and 2000) on postmetamorphic body size of

amphibians. I quantified the effect of landuse and year for 3 age classes (i.e., metamorph,

subadult, and adult) and 4 species (i.e., Spea multiplicata, S. bombifrons, Bufo cognatus,

and Ambystoma tigrinum mavortium) of amphibians using playa wetlands on the

Southern High Plains. This information is particularly important for species currently in

decline, such as amphibians (Houlahan et al. 2000), because body size can potentially

influence local population persistence.

Study Area and Amphibians

My study was conducted on the Southern High Plains (SHP) of Texas. The SHP

(i.e., the Llano Estacado) is the largest continuous plateau in the United States (Sabin and

Holliday 1995), encompassing ca. 8.9 million ha, of which 5.9 million ha are in

northwestern Texas (Haukos and Smith 1994). Primary land uses on the SHP are

agricultural cultivation (2.7 million ha, 46%)) and grassland (2.9 million ha, 49%) [native

and replanted lands combined], Haukos and Smith 1994). The most significant

topographic formations on the SHP are playa wetlands (Bolen et al. 1989).
Playa wetlands (or simply playas) occur at the highest density (ca. 19,340) in

Texas, comprising ca. 2% of the Texas SHP landscape (Haukos and Smith 1994). Playas

are circular wetlands (xsize=6.3 ha) that form dynamically and continually via an

interaction of eolian and alluvial processes (Osterkamp and Wood 1987, Gustavson et al.

1994). Playas are the lowest points on the relatively level SHP landscape, and their

watersheds are not interconnected (Guthery and Bryant 1982). Playas also are not

directly connected to the underlying Ogallala Aquifer (Bolen et al. 1989). Consequently,

playa hydrology is dependent on precipitation and watershed runoff (Haukos and Smith

1994). Because the climate on the SHP is semi-arid (i.e., Xrainfaii=33-45 cm/year,

><evaporation^200-250 cm/ycar) and playa basins are shallow, playa hydroperiods are

variable (Bolen et al. 1989). Playas usually have intermittent to seasonal or annual

hydroperiods, and they are oases of biodiversity in a predominately agricultural landscape

(Bolen et al. 1989, Haukos and Smith 1994).

Amphibians can be the most abundant vertebrate in playas during summer (Bolen

et al. 1989). Thirteen species exist on the SHP; I selected the 4 most abundant for this

study: New Mexico spadefoot {Spea multiplicata Cope), plains spadefoot {S. bombifrons

Cope), Great Plains toad {Bufo cognatus Say), and barred tiger salamander {Ambystoma

tigrinum mavortium Green) (Rose and Armentrout 1974, Anderson 1997:74, Chapter III).

The above anurans are explosive breeders (i.e., complete breeding in 1-3 d, Sullivan and

Fernandez 1999), although Great Plains toads may breed for several weeks and spawn

multiple times if ambient conditions are favorable (Krupa 1986, 1994). Barred tiger

salamanders can be more prolonged breeders, taking ca. 2-4 weeks to breed (Petranka

10
1998:111). Breeding for these species is concentrated during and after seasonal rains

March-August; however, barred tiger salamander may breed during any month on the

SHP if sufficient water is present (Webb and Roueche 1971. Rose and Armentrout 1976).

Duration of larval development for these amphibians is rapid (i.e.. 2-6 weeks) and likely

depends on phenotypic responses to competition, predation, hydroperiod, and water

quality (Rose and Armentrout 1976, Newman 1992, 1994a. Werner and McPeek 1994,

Skelly 1997, Denver et al. 1998, Morey and Reznick 2000).

Sixteen playas (4 per landuse type per year) were selected via aerial and ground

reconnaissance in March 1999 and 2000 according to surrounding landuse (cultivated or

grassland), presence of water (i.e., all playas were inundated initially), landowner

cooperation, and study playa juxtaposition (i.e., clusters of playas within the same

landuse type were sought to increase sampling efficiency). Playas were considered to be

in grassland if >75%) of the surrounding landscape (i.e., <3 km from the playa center) was

undisturbed and vegetated, and cultivation if >75% of the surrounding landscape was

arable cropland (Anderson et al. 1999(^:760, Chapter III). Grassland could have included

native, replanted (via the Conservation Reserve Program, CRP), or grazed lands with

intact grass. I chose the 3-km threshold for landuse designation, because it probably was

the near maximum dispersal distance for my study species (cf Gehlbach 1967, Gehlbach

et al. 1969, Sinsch 1990, 1997, Miaud et al. 2000).

Grassland playas were located in Castro County northeast of Dimmitt, Texas USA

(Figure 2.1). Playas surrounded by cropland were located in Hale and Floyd counties

11
34M5'0CrN
"'"100"52'3Cr W
Si- / •

.%^'
amoiitt*©
A ^
- • .•• ; ••• •/
• <...
Ca stre <\ ••••' lllv»rWfiV_,
A
1

0 0 0 J
BHn»M ^J-.
-A
"V.
o %
Hale F10 y d ^'••^.
- A
^
Lltti«l«id
A. V

- - I

J
0^ -
8 0 y
<.'^-
/ • '

L»v«llind
Lubboet ._ Crasb^s
• /
":.\.
33*22-3 (TN 1 •;:•
102*37'30W

•*• Major Qties Rateau Escarpment


o Cropland Playas Study Area Counties
© Grassland Playas

20 10 0 20 40 60 80 100
Kilometers

Figure 2.1. Location of study playa wetlands on the Southern High Plains of Texas,
U.S.A.

12
near Plainview, Texas in 1999 and north of Ralls, Texas in Crosby County in 2000.

Although grassland playas were located northwest (x=97 km, SD=11) of cropland playas,

soils and historical vegetation were similar between locations (Allen et al. 1972, D. A.

Haukos and L. M. Smith, U. S. Fish and Wildlife Service and Texas Tech University,

unpublished data). Moreover, rainfall (and I assume other ambient variables because

elevation and latitudinal change is minimal) was similar between sites (x=5.5 cm, SE=1.8

[cultivation]; x=6.9 cm, SE=2.7 [grassland]) during field sampling. Thus, I assumed a

possible landuse effect would not be confounded by geographic differences in sampling

location. Average rainfall at the study playas during sampling was greater in 1999

(x=15.1 cm, SE=1.6) than in 2000 (x=7.3 cm, SE=3.5).

Methods

Experimental Design and Units

Independent factors were arranged as a 3-way factorial-nested design

(Montgomery 2001:569-573). Landuse types and years were crossed factors and

considered fixed effects. Playas were nested within landuses and years and treated as a

random effect. I used individuals captured at nested playas as experimental units of

landuse and year. Individuals might be considered pseudoreplicates of the crossed effects

(Hurlbert 1984); thus, I restricted my inferences to the population of amphibians residing

in these 2 landuse types and years on Southern High Plains. I also used an unrestricted

mixed model, which allows covariance (i.e., dependency) between observations within

levels of the nested random effect (Hocking 1973). Consequently, results of ratio

significance tests using this model remained unbiased and accurate, despite potential

13
interdependencies between observations (Hocking 1973, Montgomer> 2001:526-527).

The aforementioned further supports my use of individuals as experimental units, because

potential interdependencies between individuals within nested playas were modeled in

the response.

Recent research (e.g., Morey and Reznick 2001) also suggests that terrestrial food

resources might differentially affect postmetamorphic body size of genders. Thus. 1 also

tested the effect of gender and landuse type on body size of adults in preliminary

multivariate analyses. These analyses were not conducted on the other age classes (i.e.,

metamorph and subadult), because captured individuals were not dissected for gender

determination. Body size of females was greater than males for adult New Mexico

spadefoot, plains spadefoot and Great Plains toads. However, because the gender effect

did not interact with landuse type and mean sex ratio across species was similar between

landuses, I combined genders for final statistical analyses (discussed later).

Sampling Techniques

Terrestrial capture. Playas were partially enclosed (i.e., 25% of circumference)

with 60-cm high drift fence (xiength= 330 m per playa) and 19-L pitfall traps (Dodd and

Scott 1994). Pitfall traps were placed on alternate sides of the fence at 10-m intervals

with openings flush to the ground (Dodd and Scott 1994). Pitfall traps in each landuse

type were checked alternate days for captures from 16 May-17 October 1999 and 19

April-18 August 2000. Pitfall traps within one landuse type were opened near dusk (e.g.,

1800-2100 hrs) before the night of intended capture and processed during the next

14
afternoon (e.g., 1400-1800 hrs). Traps were subsequently closed and reopened in 24 hrs.

Simultaneous with closing of pitfall traps in one landuse type, pitfall traps in the other

landuse type were opened.

Biological processing. Captured individuals were enumerated by species and age

(metamorph [juveniles <1 yr], subadult [>1 yr but not displaying secondary sexual

characters], adult [>1 yr and possessing reproductive morpholog) such as vocal sacs,

nuptial excrescences, or eggs for anurans, and enlarged cloacae for the salamander];

Duellman and Trueb 1994:33-38, 52-60). I did not designate a subadult category for

barred tiger salamander because distinction is based on color (Webb and Roueche 1971).

which can be hydroperiod-dependent in playas (Rose and Armentrout 1976). Snout-vent

length (i.e., tip of snout to posterior of vent; SVL) and mass were measured using calipers

and a spring scale for the first 5 individuals captured per playa per species per age class

per day (« = 2816 and 2372 in cultivation and grassland [years combined]). Individuals

were toe-clipped uniquely using sterilized scissors (soaked in 0.01% chlorhexidine

gluconate); recaptured individuals were not measured. Sampling techniques in this and

subsequent chapters followed approved Texas Tech University Animal Use and Care

Committee protocol (permit #99843).

Statistical Analyses

Multivariate analysis of variance (MANOVA) was used to test for differences (a

= 0.05) in body size within species and age classes between landuse types and years and

among levels of the random nested effect (i.e., playas). I assumed the matrix composed

of SVL and body mass response vectors represented a multivariate description of body

15
size. I used an unrestricted mixed MANOVA model with an unstructured covariance

matrix for factorial-nested designs to account for between- and within-subject

heterogeneity (Littell et al. 1991:120-126, Littell et al. 1996:269, 293, Montgomery

2001:526-527). Consequently, it was not necessary to assume sphericity or other

covariance structures (Littell et al. 1996:99). Because sample size generally was large

(i.e., «>30 individuals per level per effect), I assumed average body size per effect

followed approximately a bivariate normal density (Milton and Arnold 1995:241,

Johnson and Wichem 1998:187). Despite these precautions, I also used Pillai-Bartlett

trace test statistic, because it is more powerful and robust than other statistics (e.g.,

Wilk's likelihood ratio) when multivariate assumptions are violated (Olson 1974, 1976,

1979). I tested additivity of fixed main effects by including an interaction term in the

mixed model (Montgomery 2001:526); analyses were sorted by simple main factors (i.e.,

landuse type by year and year by landuse) with playas nested when this assumption was

violated. Westfall-Young (1993:113-121) multivariate multiple comparison method

(hereafter STEPBOOT) was used to test for simple effect differences in body size

between levels of fixed factors when MANOVA was significant (Westfall et al.

1999:227-239). I did not present results of the random effect (i.e., differences in body

size among playas nested in landuse and year) herein.

Results

Landuse Type and Year Interaction Effects

Adults. Landuse and year effects were not additive for adult Great Plains toad

(MANOVA F=12.6; 2, 593 df; P<0.001), New Mexico spadefoot (MANOVA F=49.8; 2.

16
1541 df; P<0.001), and barred tiger salamander (MANOVA F=7.5; 2, 361 df; i'<0.001);

therefore, I separated analyses by main factors (see next 2 sections). Landuse and year

effects were additive for aduft plains spadefoots (MANOVA F=2.5; 2, 240 df; ^^=0.09);

thus, analyses were not separated by main factors for this species. Body size of adult

plains spadefoots was greater at grassland than at cropland playas (MANOVA F=15.0; 2,

240 df; /'<0.001) and greater in 1999 than in 2000 (MANOVA F=4.1; 2, 240 df;

P=0.02): both SVL and mass were significant (STEPBOOT P<0.001, Table 2.1).

Subadults. Landuse and year effects were not additive for subadult Great Plains

toad (MANOVA F=\9A; 2, 325 df; P<0.001) and New Mexico spadefoot (MANOVA

F=12.3; 2, 383 df; P<0.001); therefore, I separated analyses by main factors. Additivity

of main effects was not tested for subadult plains spadefoot because no individuals were

captured in cropland in 1999; thus, these analyses were separated immediately by main

factors. Analyses were not performed for barred tiger salamander, because there was no

subadult category for this species.

Metamorphs. Landuse and year effects were not additive for metamorph New

Mexico spadefoot (MANOVA F=12.2; 2, 577 df; P<0.00\) and barred tiger salamander

(MANOVA F=7.8; 2, 462 df; P<0.001); therefore, I separated analyses by main factors.

Additivity of main effects was not tested for metamorph Great Plains toad and plains

spadefoots, because no individuals were captured in grassland in 2000. Consequently, I

separated analyses immediately by main factors for these species.

17
Table 2.1. Aduk body size of New Mexico spadefoot {Spea multiplicata. NSF). plains
spadefoot {S. bombifrons. PSF), Great Plains toad {Bufo cognatus. GPT). and barred tiger
salamander {Ambystoma tigrinum mavortium, BTS) between landuse types and \ears
(fixed factorial effects) at 16 playa wetlands (random nested effect) on the Southern High
Plains of Texas, April-September 1999 and 2000.
Size Landuse
Variable Species Year' Cultivation Grassland
n X SE P<0.05' n X SE P<0.05
Snout- NSF 1999 536 42.1 0.1 Aa 409 44.4 0.2 Ba
vent 2000 401 38.7 0.2 Ab 206 43.7 0.2 Bb
length PSF 1999 45 46.1 0.6 Aa 57 49.2 0.7 Ba
(mm) 2000 77 43.9 0.4 Ab 72 46.9 0.4 Bb
GPT 1999 256 72.4 0.5 Aa 164 80.2 0.6 Ba
2000 76 64.9 0.9 Ab 108 75.3 0.8 Bb
BTS 1999 63 92.6 1.8 Aa 107 101.4 0.9 Ba
2000 66 81.3 0.9 Ab 100 86.5 1.2 Bb
Mass NSF 1999 536 8.5 0.1 Aa 409 9.8 0.1 Ba
(g) 2000 401 7.2 0.1 Ab 206 9.4 0.2 Bb
1999 45 11.4 0.5 Aa 57 12.2 0.4 Ba
PSF 2000 77 9.7 0.3 Ab 72 11.5 0.3 Bb
1999 256 51.5 1.1 Aa 164 70.1 1.7 Ba
GPT 2000 76 40.5 1.7 Ab 108 53.8 1.8 Bb
1999 63 39.7 2.8 Aa 107 48.1 1.6 Ba
BTS 2000 66 22.3 0.9 Ab 100 27.1 1.4 Bb
^Analyzed by year because landuse and year main effects were not additi\e (P<0.001) for
all species except PSF (P=0.09); PSF statistics were presented by year for tabular
parsimony.
Within-species means in the same row with unlike uppercase letters are different (i.e..
simple effect test between landuses within years); means in the same column within size
variables and species with unlike lowercase letters are different (i.e., simple effect test
between years within landuses); all analyses performed using Westfall-Young
(1993:113-121) multivariate multiple comparison test after the simple main effect
MANOVA for factorial-nested designs (Montgomery 2001:569-573) was significant
(P<0.05).

18
Landuse Type Main Effect

Adults. Body size was greater at grassland than at cropland playas for adult Great

Plains toad (MANOVA F=47.2; 2, 411 df; P<0.00\), New Mexico spadefoot (MANOVA

F=9A; 2, 936 df; P<0.001), and barred tiger salamander (MANOVA F=12.5; 2, 161 df;

P<0.001) in 1999; SVL and mass were significant for all species (STEPBOOT P<0.005,

Table 2.1). Body size also was greater at grassland than at cropland playas for adult

Great Plains toad (MANOVA F=18.3; 2, 175 df; P<0.001), New Mexico spadefoot

(MANOVA F=90.6; 2, 598 df; P<Om\), and barred tiger salamander (MANOVA

F=5.4; 2, 157 df; /*<0.005) in 2000; SVL and mass were significant for all species

(STEPBOOT P<0.03, Table 2.1).

Subadults. Body size was greater at grassland than at cropland playas for

subaduh Great Plains toad (MANOVA F=\3.7; 2, 256 df; P<0.001), New Mexico

spadefoot (MANOVA F=\ 12.2; 2, 298 df; P<0.001), and plains spadefoot (MANOVA

F=24.8; 2, 64 df; P<0.00\) in 2000; SVL and mass were significant for all species

(STEPBOOT P<0.001, Table 2.2). Body size did not differ between landuses for

subadult Great Plains toad (MANOVA F=0.7; 2, 62 df; P=0.49) and New Mexico

spadefoot (MANOVA F=2.9; 2, 80 df; P=0.07) in 1999.

Metamorphs. Body size was greater at grassland than at cropland playas for

metamorph Great Plains toad (MANOVA F=l 14.1; 2, 335 df; P<0.001), New Mexico

spadefoot (MANOVA F=19.2; 2, 494 df; P<0.00\), and plains spadefoot (MANOVA

F=21.9; 2, 156 df; P<0.001) in 1999; SVL and mass were significant for all species

19
Table 2.2. Subadult body size of New Mexico spadefoot {Spea multiplicata, NSF), plains
spadefoot {S. bombifrons, PSF), and Great Plains toad {Bufo cognatus, GPT) between
landuse types and years (fixed factorial effects) at 16 playa wetlands (random nested
effect) on the Southern High Plains of Texas, April-September 1999 and 2000.
Size Landuse
Variable Species Year' CLiltivation Grassland
n X SE P<0.05'' n X SE P<0.05
Snout- NSF 1999 26 34.1 0.9 Aa 61 36.2 0.5 Aa
vent 2000 221 32.5 0.2 Ab 86 39.1 0.4 Bb
length PSF 1999 0 0 0 NT 15 38.1 0.8 a
(mm) 2000 55 34.2 0.5 A 18 43.2 0.7 Bb
GPT 1999 41 58.4 0.9 Aa 30 56.5 1.3 Aa
2000 64 47.1 0.8 Ab 201 59.0 0.6 Ba
Mass NSF 1999 26 4.8 0.3 Aa 61 5.3 0.3 Aa
(g) 2000 221 4.6 0.2 Aa 86 6.6 0.2 Bb
PSF 1999 0 0 0 NT 15 6.5 0.4 a
2000 55 4.7 0.2 A 18 7.8 0.4 Bb
GPT 1999 41 23.3 0.9 Aa 30 24.4 1.8 Aa
2000 64 14.7 0.7 Ab 201 26.9 0.7 Ba
Analyzed by year because landuse and year main effects were not additive (P<0.001) for
NSF and GPT; PSF analyzed for 2000 only.
Within-species means in the same row with unlike uppercase letters are different (i.e.,
simple effect test between landuses within years); means in the same column within size
variables and species with unlike lowercase letters are different (i.e., simple effect test
between years within landuses); all analyses performed using Westfall-Young
(1993:113-121) multivariate multiple comparison test after the simple main effect
MANOVA for factorial-nested designs (Montgomery 2001:569-573) was significant
(/*<0.05); NT indicates no test performed because SE=0 for >1 level of the effect.

20
(STEPBOOT P<0.001, Table 2.3). Body size did not differ between landuses for

metamorph barred tiger salamander (MANOVA F-0.5; 2, 267 df; P=0.62) in 1999.

Body size was greater at grassland than at cropland playas for metamorph barred tiger

salamander (MANOVA F=18.4; 2, 188 df; P<0.001) in 2000; SVL and mass were

significant (STEPBOOT P<0.005, Table 2.3). Body size was greater at cropland than at

grassland playas for metamorph New Mexico spadefoot (MANOVA F=7.9; 2, 78 df;

P<0.001) in 2000; SVL and mass were significant (STEPBOOT P<0.006, Table 2.3).

Year Main Effect

Adults. Body size was greater in 1999 than in 2000 for aduk Great Plains toad

(MANOVA F-15.6; 2, 323 df; P<0.001), New Mexico spadefoot (MANOVA F=l 10.6;

2, 928 df; P<0.001), and barred tiger salamander (MANOVA F=22.5; 2, 120 df;

P<0.001) at cropland playas; SVL and mass were significant for all species (STEPBOOT

P<Om\, Table 2.1). Body size also was greater in 1999 than in 2000 for aduU Great

Plains toad (MANOVA F=63; 2, 263 df; 7^=0.002), New Mexico spadefoot (MANOVA

F=3.1; 2, 606 df; P=0.05), and barred tiger salamander (MANOVA F-65.1; 2, 198 df;

P<Q.OO\) at grassland playas; SVL and mass were significant for all species

(STEPBOOT P<0.05, Table 2.1).

Subadults. Body size was greater in 1999 than in 2000 for subaduh Great Plains

toad (MANOVA F=12.4; 2, 96 df; /'<0.001) and New Mexico spadefoot (MANOVA

F=3.7; 2, 239 df; P=0.03) at cropland playas; SVL and mass were significant for Great

Plains toad (STEPBOOT P<0.001), and SVL was significant for New Mexico spadefoot

(STEPBOOT P=0.04, Table 2.2). Body size was greater in 2000 than in 1999 for

21
Table 2.3. Metamorph body size of New Mexico spadefoot {Spea multiplicata. NSF).
plains spadefoot {S. bombifrons, PSF), Great Plains toad {Bufo cognatus, GPT). and
barred tiger salamander {Ambystoma tigrinum mavortium, BTS) between landuse types
and years (fixed factorial effects) at 16 playa wetlands (random nested effect) on the
Southern High Plains of Texas, April-September 1999 and 2000.
Size Landuse
Variable Species Year" Cultivation Grassland
n X SE P<0.05'' n X SE P<0.05
Snout- NSF 1999 384 28.1 0.2 Aa 119 31.9 0.4 Ba
vent 2000 51 24.5 0.4 Ab 34 22.9 0.4 Bb
length PSF 1999 76 31.8 0.4 Aa 89 36.5 0.4 B
(mm) 2000 27 25.9 0.4 b 0 0 0 NT
GPT 1999 142 25.1 0.7 Aa 202 38.2 0.7 B
2000 39 25.5 1.1 a 0 0 0 NT
BTS 1999 103 77.4 1.0 Aa 173 78.5 0.6 Aa
2000 76 72.4 0.7 Ab 121 79.4 0.7 Ba
Mass NSF 1999 384 2.8 0.1 Aa 119 3.7 0.1 Ba
(g) 2000 51 2.3 0.1 Ab 34 1.2 0.1 Bb
1999 76 3.6 0.2 Aa 89 5.6 0.2 B
PSF 2000 27 2.7 0.1 b 0 0 0 NT
1999 142 2.5 0.2 Aa 202 8.6 0.5 B
GPT 2000 39 2.8 0.3 a 0 0 0 NT
1999 103 21.4 0.9 Aa 173 22.0 0.6 Aa
BTS 2000 76 15.4 0.9 Ab 121 18.8 0.8 Bb
^Analyzed by year because landuse and year main effects were not additive (P^O.OOl) for
NSF and BTS; PSF and GPT analyzed for 1999 only.
Within-species means in the same row with unlike uppercase letters are different (i.e.,
simple effect test between landuses within years); means in the same column within size
variables and species with unlike lowercase letters are different (i.e., simple effect test
between years within landuses); all analyses performed using Westfall-Young
(1993:113-121) multivariate multiple comparison test after the simple main effect
MANOVA for factorial-nested designs (Montgomery 2001:569-573) was significant
(/*<0.001); NT indicates no test performed because SE=0 for >1 level of the effect.

22
subaduh New Mexico spadefoot (MANOVA F=3.2; 2. 139 df: P-0.04) and plains

spadefoot (MANOVA F=3.5; 2, 24 df; P=0.05) at grassland playas; SVL and mass were

significant (STEPBOOT P<0.01, Table 2.2). Body size did not differ between years for

subadult Great Plains toad (MANOVA F=0.2; 2, 222 df; P=0.78) at grassland playas.

Metamorphs. Body size was greater in 1999 than in 2000 for metamorph New

Mexico spadefoot (MANOVA F=17.9; 2, 427 df; P<0.00\), plains spadefoot (MANO\'A

F=37.5; 2. 96 df; P<0.001), and barred tiger salamander (MANOVA F=10.8: 2. 170 df

P<0.00\) at cropland playas: SVL and mass were significant for all species (STEPBOOT

P<0.006, Table 2.3). Body size did not differ between years for metamorph Great Plains

toad at cropland playas (MANOVA F=0.3; 2, 174 df; ^=0.73). Body size was greater in

1999 than in 2000 for metamorph New Mexico spadefoot (MANOVA F=16.4; 2. 145 df:

P<0.001) and barred tiger salamander (MANOVA F=20.7; 2, 285 df; P<0.001) at

grassland playas; SVL and mass were significant for New Mexico spadefoot

(STEPBOOT P<0.00\), and mass was significant for barred tiger salamander

(STEPBOOT P=0.002, Table 2.3).

Discussion

Landuse Effect

Body size of amphibians generally was greater at grassland playas than at playas

surrounded by cultivation for all age classes (metamorph, subadult, and adult) on the

Southern High Plains. To my knowledge, this is the first documentation of the potential

effect of anthropogenic landuse on body size of amphibians in North America. Beebee

(1983) and Oldham (1985) documented in the United Kingdom that body size of Bufo

23
bufo and two species of newts {Triturus vulgaris, T. helveticus) at wetlands in cultivation

was greater than at wetlands with an uncultivated watershed (i.e.. opposite of my results).

Oldham (1985) related his resuhs to postmetamorphic density, where spring influx of

adults was 14X greater at the uncultivated wetland than at the wetland surrounded by

cropland. Similarly, with respect to relative abundance, mean daily capture of

individuals in pitfall traps (all species combined) at cropland playas (x=648. SE=226

[1999]: x=157. SE=29 [2000]) was greater than (P<0.03, Wilcoxon 2-sample

nonparametric test, Conover 1980:216) at grassland playas (x=135, SE=20 [1999]; x=90.

SE=23 [2000]) both years (also see Chapter III). Density of conspecifics, and possibly

congeners, can affect postmetamorphic growth rate and body size of amphibians (Goater

1994, Pechmann 1994. Morey and Reznick 2001) by increasing competition for food

resources.

Beebee (1983) and Oldham (1985) also speculated that body size might be a

consequence of differential food abundance. Although I did not quantify invertebrate

abundance, capture of beetles, springtails, and spiders appeared greater in pitfall traps at

grassland playas than at cropland playas (M. J. Gray and R. Brenes. Texas Tech

University, personal observation). Moreover, mean total mass (g) of invertebrates in

stomachs {n=70) of metamorph Great Plains toads was greater (P=0.005, Wilcoxon test)

for individuals I collected at grassland playas (x=0.047, SE=0.008) than at cropland

playas (x=0.009, SE=0.002). Runoff or aerial drift of insecticides and scarification of

vegetation and soil may negatively affect terrestrial food resources available for

24
amphibians at wetlands surrounded by cultivation (Freemark and Boutin 1995); however,

this h\ pothesis needs to be tested for playa wetlands.

Postmetamorphic body size also may have been affected by conditions in the

larval em ironment (Wilbur and Collins 1973, Werner 1986, Brady and Griffiths 2000,

More> and Reznick 2001). Hydroperiod can affect postmetamorphic body size b}

reducing duration of larval development (e.g., Newman 1988, 1989. Denver et al. 1998).

Mean h\ droperiod (i.e., measured from the date amphibian sampling started until playas

dried) of grassland playas (x=139 d, SE=14) was greater than (P=0.05, Wilcoxon test)

cropland playas (x=89 d. SE=22) in 1999. Although 1 did not detect differences (F=0.18.

Wilcoxon test) between landuse types, mean hydroperiod in grassland playas (x=107 d.

SE=13) was greater than in cropland playas (x=69 d, SE=18) during 2000. Because

rainfall at playas was similar {P>0.32, Wilcoxon test) between landuses (x=6.4 cm,

SE=1.3 [cultivation, 1999]; x=9.2 cm, SE=1.5 [grassland. 1999]; x=4.4 cm, SE=2.2

[cultivation, 2000]; x=4.5 cm, SE=3.8 [grassland, 2000]) both years, hydroperiod

probably was shorter in cropland playas because of differential sedimentation (Luo et al.

1997).

Although I did not quantify potential herbaceous (e.g., nektonic algae. Savage

1952) and animal (e.g., anostracan shrimp. Pfennig 1992) foods between landuse types,

amphibian densities may have been lower in cropland playas because of increased

mortality of these organisms associated with agricultural runoff or drift (Freemark and

Boutin 1995). Reduced aquatic food resources can negatively influence larval growth

(e.g.. Wilbur 1977, Steinwascher 1979, Newman 19946, Walls 1998), and consequently

25
postmetamorphic body size. Sediments and agricultural chemicals also ma> have altered

the percent composition of aquatic invertebrate taxa in cropland playas (Hall 1997:68-69.

85-87), which could have influenced potential predators (Skelly 1997). If predator)

insect density was reduced in cropland playas, competition for food resources may have

been greater than in grassland playas (e.g., Werner and McPeek 1994). perhaps

contributing to reduced postmetamorphic body size. Lastly, agricultural chemicals can

directh reduce foraging acti\'ity and growth of larval amphibians, which can decrease

body size at metamorphosis (Semlitsch et al. 1995, Fioramonti et al. 1997. Boone and

Semlitsch 2001).

Amphibian predator density and size also may have been related positively to

hydroperiods (Skelly 1996). Mean daily capture of barred tiger salamander larvae and

neonates (a potential predator and cannibal. Rose and Armentrout 1976, Collins and

Holomuzki 1984) in seine net samples (0.002 ha sample area) was greater (P<0.01.

Wilcoxon test) in grassland playas (x=108, SE=16 [1999]; x=l 1. SE=4 [2000]) than in

cropland playas (x=26, SE=6 [1999]; x=0.2, SE=0.1 [2000]). Predators can positiveh

influence postmetamorphic body size of anurans by reducing density of conspecific

larvae, thus competition for food resources (Wilbur 1984). Alternatively, predators may

negati\'ely affect postmetamorphic body size of individuals by decreasing foraging

activity and size at metamorphosis (Skelly and Werner 1990), yet this effect may be

species-dependent (Werner 1991, Werner and McPeek 1994) and secondary. I

hypothesize, if predators are influencing postmetamorphic body size in playas, their

presence is primarily reducing competition (e.g., Morin 1983), thus resulting in large

26
postmetamorphic body size associated with predator-rich wetlands (i.e., grassland

playas).

One different resuU existed in my landuse results; body size of metamorph New

Mexico spadefoots was greater at playas surrounded by cultivation than at grassland

playas in 2000. I believe this was a consequence of timing at metamorphosis. New

Mexico spadefoot tadpoles metamorphosed ca. one month earlier from playas in

cultivation (06/28/00) than from grassland playas (07/25/00), which was ca. 25 and 52 d

following breeding, respectively. Thus, metamorph New Mexico spadefoots had more

time to grow postmetamorphically in cropland than in grassland, which likely positiveh

affected body size. Developmental plasticity in response to wetland hydroperiod is well

documented for spadefoot tadpoles (see references in Newman 1992 and Denver 1997).

Presumably, New Mexico spadefoot tadpoles assayed the ratio between their size-specific

mortality and growth rates (Wilbur 1984, Werner 1986) and delayed metamorphosis in

grassland playas, perhaps because water levels were more stable than in cropland playas

(as suggested by my hydroperiod data). Indeed, body size (e.g., SVL) of newly

metamorphosed individuals (i.e., captured the first day) at grassland playas (x=21.9 mm,

SE=0.4) was slightly greater than at cropland playas (x=20.7 mm, SE=0.4), but it

apparently did not compensate for the longer duration of postmetamorphic growth of

individuals in cultivation.

Year Effect

Postmetamorphic body size of amphibians was greater in 1999 (i.e., the wetter

year) than in 2000 generally for all species and age classes. Moreover, the effect of

27
landuse type and year on body size generally was non-additive. Differences in body size

between landuse treatments increased 33-545% between 1999 and 2000 for adult and

subadult New Mexico spadefoot and Great Plains toads, and metamorph barred tiger

salamanders. Rainfall can affect postmetamorphic growth rates (Tinsley and Tocque

1995) and presumably body size (Bruce and Hairston 1990, Reading 1990) by reducing

probability of desiccation (Newman and Dunham 1994) and increasing foraging (Jaeger

1980) and movement activity (e.g., Hurlbert 1969, Semlitsch \9S5b, Sinsch 1988). Mean

number of individuals immigrating and emigrating daily (all species and movement

directions combined) at my playas was greater {P<0.00\, Wilcoxon test) in 1999 (x=391,

SE=115) than in 2000 (x=129, SE=20) for both landuses. Increased opportunity for

movement may increase foraging success of amphibians (Jaeger 1980). Prey abundance

also may increase with rainfall (Dimmitt and Ruibal \9S0a). Diet of metamorphs and

adults of my study species was primarily composed of beetles on the Southern High

Plains (Anderson et al. 19996, L. M. Smith and M. J. Gray, unpublished manuscript).

Coleoptera reproduction in semi-arid environments often is correlated with rain (Crowson

1981:380-381,391-396).

Rainfall also can affect postmetamorphic body size by influencing duration of

larval development (Reading and Clarke 1999, Camp et al. 2000). Mean larval

emergence at my playas was 43.5 and 38.5 d after breeding commenced (determined as

per ancillary call surveys) in 1999 and 2000. Interestingly, this difference in larval

duration between years is attributed to the reduction in developmental time in cropland

(x=36 d, SE=4 [1999], 25 d, SE=2 [2000]) not in grassland playas (x=51 d, SE=2 [1999],

28
52 d, SE=2 [2000]). Thus, postmetamorphic body size may have been different between

years, and landuse and year main effects non-additive, because of greater available food

resources and duration of larval development associated with 1999 and grassland playas.

Body size of subadult New Mexico and plains spadefoots was greater in 2000

than in 1999 at grassland playas, which was contradictory to the above results. Because

body size of subadult anurans reflects growth during the previous year (see Turner 1960),

body size of subadults may have been elevated in 2000 because of additional growth

experienced during more favorable ambient conditions in 1999. However, if this was

true, then subadults captured at cropland playas should have been larger in 2000 than in

1999 (which was not observed). Possibly, the negative effect of cultivation on

postmetamorphic body size of subadults during both years counteracted the positive

benefits of higher rainfall in 1999.

Competing Body Size Hypotheses

Wilbur and Collins (1973) suggested that body size at metamorphosis would be

maintained through postmetamorphic development thus confer adult fitness. Several

empirical studies (e.g., Semlitsch et al. 1988, Berven 1990, Scott 1994) have supported

their prediction. Werner (1986) suggested that body size at metamorphosis (and

ultimately adult body size and fitness) would be a consequence of size-specific mortality

(u) and growth rates {g) in aquatic and terrestrial habitats. Thus, an individual that

metamorphosed at a small body size due to a high larval \i/g ratio might exhibit ''catch-

up" growth because of a low postmetamorphic \i/g ratio. Morey and Reznick (2001) and

Goater (1994) supported this prediction (at least in part). Although I did not measure the

29
same indi\'iduals through time, because individuals were random samples from the

population residing within main effect treatments and body size of amphibians can var>'

seasonally in response to environmental conditions (Reading and Clarke 1995 and

references therein). I can assume that postmetamorphic bod\ size for each age class

represented mean population response of age-specific growth rate to main effects. Thus,

my results seemingly provide support for U'ilbur and Collins' (1973) h\ pothesis. because

differences in body size between levels of main effects generally were maintained

throughout all age classes (i.e.. metamorphs, subadults, and adults were consistenth

larger at grassland playas [thus no catch-up growth exhibited in cultivation]).

Conclusions

Landuse type surrounding wetlands and rainfall can affect postmetamorphic body

size of amphibians on the Southern High Plains of Texas. Individuals captured at

grassland playas generally were larger than those caught at playas surrounded by

agricultural cultivation, and individuals in 1999 (i.e., a wetter year) usually were larger

than those captured in 2000. Moreover, I found the effect of landuse type and year

usually was non-additi\'e; cultivation negati\ eh affected body size greater during the

drier year. Thus, my results suggest that anthropogenic modification of terrestrial

landscapes surrounding wetlands may negatively influence postmetamorphic bod> size of

amphibians, especially during drier years. Density independent and dependent factors

both may be related to landuse and year effects, and proximally responsible for

influencing body size of amphibians. This information is important because

postmetamorphic bod> size can be correlated with individual survival and fitness, thus

30
population persistence of amphibians. Indeed, retention of grasslands surrounding

wetlands may positively benefit amphibians.

31
CHAPTER III

EFFECT OF LANDUSE AND YEAR ON POPULATION

DEMOGRAPHICS OF SOUTHERN HIGH PLAINS AMPHIBIANS

Abstract

Antliropogenic disturbance of landscapes surrounding wetlands is considered a

factor in local and global amphibian declines. Few data exist on the effects of landscape

cultivation on amphibians, and results from previous studies are contradictory. My

objective was to test the effect of landuse (cultivation vs. grassland [control]) on

population demographics of 7 species and 3 age classes of amphibians during 2 years

(1999 vs. 2000) on the Southern High Plains of Texas. I partially enclosed 16 playa

wetlands (4 per landuse per year) with drift fence and pitfall traps, and monitored relative

abundance and diversity daily from 16 May-17 October 1999 and 19 April-18 August

2000. I also monitored immigration and emigration rates to investigate the effect of

landuse and year on possible source-sink dynamics. In general, abundance (i.e., daily

capture rate) of New Mexico and plains spadefoots {Spea multiplicata, S. bombifrons)

was greater at cropland than grassland playas; abundance of other species and diversity of

the amphibian assemblage was not affected by landuse. Also, abundance generally was

greater in 1999 than 2000 for metamorph spadefoots and barred tiger salamanders

{Ambystoma tigrinum mavortium). Frequency of days emigration exceeded immigration

(i.e., source dynamics) was greater at grassland than cropland playas and in 1999 than

2000. Differences in spadefoot abundance between landuse types may have been related

to low species-specific vagility, resulting in increased nestedness within disturbed

32
landscapes and absence of a keystone intraguild predator (i.e.. neotenic or large larval

tiger salamanders) in cropland playas. Yearly variation in metamorph abundance and

frequency of direcfional movements likely was related to armual rainfall (i.e., 1999 >

2000). Landscape disturbance surrounding wetlands influences demographics and

source-sink dynamics of amphibian populations.

Introduction

Amphibian populations are declining globally (Houlahan et al. 2000), and

identifying factors associated with declines is a necessary prelude to conservation efforts.

Anthropogenic modifications of landscapes surrounding wetlands can affect resident

organisms (Findlay and Houlahan 1997). Most research suggests species richness and

abundance of amphibians are negatively associated with anthropogenic disturbance (e.g.,

timber harvesting, agricultural cultivation) of uplands adjacent to wetlands (e.g., Petranka

et al. 1993, Hecnar and M'Closkey 1998, Scribner et al. 2001). However, some recent

studies have documented unexpected positive or neutral associations of amphibians in

disturbed landscapes (Anderson et al. 1999a, Knutson et al. 1999, Kolozsvary and

Swihart 1999). Presumably, landscape disturbance affects amphibians by influencing

probability of survival, reproduction, or interdemic movement.

Landscape disturbance probably affects amphibian population parameters by

physically altering the aquatic (i.e., wetland) and terrestrial (i.e., upland) environments

(Wilbur 1980, Semlitsch 2000). For example, agricultural cultivation can affect water

quality and increase sedimentation in wetlands (Hanson et al. 1994, Luo et al. 1997).

Pesticides in drift and runoff can directly increase mortality of amphibian larvae or

33
indirectly affect survivorship by reducing swimming mobility, thus probability of

predator avoidance and food item acquisition (Bridges 1997, Bridges and Semlitsch

2000, Relyea and Mills 2001). Pesticides also may reduce invertebrate densities in

wetlands (Boone and Semlitsch 2001), which can function as predators, competitors, and

food for amphibian larvae (Morin et al. 1988, Skelly 1997). Herbicides in runoff can

reduce densities of phytoplankton and periphyton (Freemark and Boutin 1995), which

could reduce foraging carrying capacity of amphibian larvae (Wilbur 1997). In contrast,

nitrogen fertilizers can increase net primary productivity in wetlands (Hanson et al.

1994), which might increase foraging opportunities for larval amphibians (Leibold and

Wilbur 1992). Finally, sedimentation decreases wetland volume and hydroperiod (Luo et

al. 1997), which may decrease time available for larvae to develop (Chapter II).

Landscape disturbance also may affect metamorphosed amphibians by

mechanically or chemically altering the terrestrial environment. For example, soil

cultivation can cause direct mortality of fossorial amphibians or it can unbury and expose

them to inhospitable ambient conditions (M. J. Gray, personal observation). Mechanical

disturbance also reduces live and detrital vegetation, which can be important in retaining

soil moisture and ftjnction as foraging, retreat, and borrowing sites for amphibians (Dodd

1996, deMaynadier and Hunter 1998, Herbeck and Larsen 1999, Naughton et al. 2000).

Additionally, mechanical manipulations of the landscape may increase soil compaction

(Welsh 1990), thus decrease burrowing efficiency of amphibians (Jansen et al. 2001),

which can be particularly important for fossorial species inhabiting xeric environments

(McClanahan et al. 1994).

34
Landscape disturbance also can affect connectivity of spatially structured

populafions by influencing permeability to movement (Turner et al. 1989, Wiens 1997,

Gibbs 1998a). For example, agricultural landscapes may be more complex (e.g., contain

greater number of edges to cross. Chapter IV), thus be less permeable to interdemic

movement (Stamps et al. 1987). Agricultural fields also may contain vegetation with less

vertical structure than grasslands or forests thus be less viscous and more permeable to

movement (Wiens et al. 1993, Vos and Chardon 1997). Additionally, pesticides applied

to crops or silvicultural plots may reduce invertebrate abundance for postmetamorphic

amphibians. Finally, the aforementioned factors in the aquatic and terrestrial

environments may individually or synergistically reduce postmetamorphic body size of

amphibians in disturbed landscapes (Chapter II). Populations composed of smaller

individuals can have a greater probability of extinction, because smaller individuals are

less evolutionarily fit than larger individuals (Semlitsch et al. 1988, Berven 1990).

Several studies have been conducted in North America examining the effects of

forest disturbance (e.g., clearcutting, timber thinning) on amphibian populations (e.g.,

deMaynadier and Hunter 1995, Gibbs \99Sb, Herbeck and Larsen 1999, Naughton et al.

2000). Interestingly, few data exist (cf Knutson et al. 1999, Kolozsvary and Swihart

1999) on the effects of agricultural cultivation on amphibians (Dodd 1997:178). It

generally is assumed that agriculture negatively influences populations (Semlitsch 2000).

However, recent evidence suggests that some amphibian populations may become nested

in wetlands within disturbed landscapes (Hecnar and M'Closkey 1997, Wright et al.

1998), thus be positively associated with agricultural cultivation on a local scale

35
(Knutson et al. 1999. Kolozsvary and Swihart 1999). Addifionally. some species (e.g.,

American toad, Bufo americanus Holbrook) may actually benefit from disturbance

(Kolozsvary and Swihart 1999).

Two primary landuse types exist on the Southern High Plains (SHP) of Texas:

cultivation and intact grassland (Haukos and Smith 1994). Amphibians on the SHP

inhabit spatially structured playa wetlands (Bolen et al. 1989). Because playas are

numerous, similar structurally, and exist in their own watershed (Smith and Haukos

2002), they are ideal systems to test the effects of anthropogenic disturbance on

amphibians. My objective was to determine the effect of landscape cultivation around

playa wetlands on amphibian demographics during 1999 and 2000. In agreement with

the status quo (e.g., Berger 1989, Laan and Verboom 1990, Oldham and Swan 1991:147.

Jung 1993, Bonin et al. 1997a, Hecnar and M'Closkey 1998), I hypothesized that playa

amphibian populations would be negatively affected by agricultural cultivation. I also

hypothesized that populations in playas surrounded by cultivation would exhibit sink

dynamics (i.e., immigration > emigration) more frequently than source dynamics (i.e..

emigration > immigration, PuUiam 1988). Thus, I examined whether relative daily

abundance, diversity, immigration, and emigration of 7 species and 3 age classes of

amphibians residing in playa wetlands on the SHP varied by landuse (cultivation vs.

grassland) and between years (1999 vs. 2000) that had different rainfall (Chapter II).

Methods

This study was conducted on the Southern High Plains of Texas (Sabin and

Holliday 1995). I used playa wetlands as experimental units of landuse and year effects.

36
On average, playas located in grassland and cropland were surrounded by >75% natural

or replanted grass and agricultural vegetation, respecfively (Appendix A). I located 16

playas (4 per landuse per year) via aerial and ground reconnaissance in March 1999 and

2000. Because wildlife population dynamics can be highly variable among playas (Smith

and Haukos 2002), I restricted experimental inferences to these landuses and years and

my study playas. A more detailed description of my study area, playas, and years was

presented in Chapter II.

Playas were partially enclosed (i.e., 25% of circumference) with 60-cm high drift

fence and 19-L pitfall traps (Dodd and Scott 1994); a cardinal quadrant was randomly

selected for placement. Fence and pitfalls were placed ca. 10 m upslope from and

parallel to the playa edge (i.e., clay-silt loam line, Luo et al. 1997). One 10-m

perpendicular lead centered at the main fence existed at each end. Vegetation underneath

the main fence and leads was removed, and the bases covered with soil to reduce

probability of trespass (Dodd and Scott 1994). Pitfalls were placed on alternate sides of

the fence at 10-m intervals with openings flush to the ground (Dodd and Scott 1994).

Water and sponges were placed in pitfalls to reduce desiccation of captured individuals

(Daoust 1991). Pitfall traps were checked alternate days in each landuse for captures

from 16 May-17 October 1999 and 19 April-18 August 2000. Pitfalls were open for

23,520 and 12,460 trap nights in 1999 and 2000, respectively.

Captured individuals were enumerated by species and age (i.e., metamorph,

subadult, and adult). Morphological characters used to distinguish among age classes

have been described (Chapter II). Also, I did not designate a subadult category for

37
captured barred tiger salamanders (Chapter II). Individuals were toe clipped uniquely

with respect to study playa and direction of movement (i.e., immigration, emigration)

using sterilized scissors (i.e., soaked in 0.01% chlorhexidine gluconate). Individuals

were assumed to be immigrating and emigrating if captured on the upslope and wetland

side of the fence, respectively (Dodd and Scott 1994). Individuals were released on the

opposite side of the fence after marking (Dodd and Scott 1994).

I used mean daily capture per playa as an estimate of relative daily abundance,

and captures on the upslope and wetland side of the fence as an estimate of relative daily

immigration and emigration of amphibians (Dodd and Scott 1994 and references therein.

Heyer et al. 1994:77). I assumed if any bias existed by using mean daily capture to

estimate relative daily abundance, immigration, and emigration, it w as constant between

landuses and years. It follows that results of ratio tests of significance between levels of

effects would not be affected as per the rules for expectation (Milton and Arnold

1995:53). Mean daily diversity was calculated using the Shannon-Weaver algorithm

(Hair 1980:273). Because individuals were marked uniquely with respect to direction of

movement at capture, trespass rate per direction could be estimated from subsequent

captures. For example, if an individual emigrated during its first and second captures, it

trespassed in the immigrating direction between captures. This enabled me to estimate

percent trespass per direction (M. J. Gray, unpublished data) and correct immigration and

emigration rates prior to analysis. Frequency of days that emigration exceeded

immigration (for all species combined), immigration exceeded emigration, and

emigration equaled immigration was recorded per playa to examine source-sink dynamics

38
of the amphibian assemblage (Pulliam 1988, 1996). I assumed any bias associated with

combining species to examine source-sink dynamics was constant between landuses and

years, thus did not affect resuhs of ratio tests of significance as discussed previously.

The experiment was arranged as a 3-factor factorial design (Montgomery

2001:194-196), where landuse, year, and species captured were crossed-fixed effects. 1

analyzed population demographics for 7 species {Spea multiplicata Cope [New Mexico

spadefoot, NSF], S. bombifrons Cope [plains spadefoot, PSF], Bufo cognatus Say [Great

Plains toad, GPT], B. woodhousii Girard [Woodhouse's toad, WHT], Pseudacris clarkii

[spotted chorus frog, SCF], Rana blairi Mecham et al. 1973 [plains leopard frog, PLF],

Ambystoma tigrinum mavortium Green [barred tiger salamander, BTS]). Two additional

species were captured {B. debilis Girard [green toad], Gastrophryne olivacea Hallowell

[Great Plains narrowmouth toad]); however, there was insufficient capture data for

analysis so they were not used.

I used a 3-factor analysis of variance (ANOVA) model with days as subsamples

to test for differences (a = 0.05) in relative abundance of individuals (i.e., daily capture)

between landuse types and years and among species (Montgomery 2001:194-196). I

used a 2-factor ANOVA model (landuse and year main effects) with days as subsamples

to analyze diversity, because species were combined to calculate the Shannon-Weaver

index thus it was not available as an effect. Nonadditivity was tested by including all 2-

way and the 3-way interaction in the models (Montgomery 2001:194). When

nonadditivity was violated (i.e., >1 interaction significant), analyses were separated by

species to test for differences between landuses and years (i.e., 2-way ANOVAs) and by

39
landuse and year (i.e., 1-way ANOVAs) to test the species effect. For species that

differed in response between levels of landuse or year, I performed a post-hoc set of 2-

way ANOVAs using landuse and year as crossed factors, but by age class, to discern

which demographic category was responsible for the observed abundance differences for

that species. Age was not included as a factor for the initial analyses, because there was

no subadult category for BTS (i.e., design was not balanced) and I assume levels of age

were not independent. Also, this analytical approach conserved experimentwise error

rate at the lowest probability for the main effect ANOVAs, because only significant

species were analyzed in the post-hoc tests for each age class. Tukey's HSD multiple

comparison test was used to test for differences among species when the main effect

ANOVA was significant (Milliken and Johnson 1992:36-38, Westfall et al. 1999:179). 1

natural-log transformed all data to meet linear model assumptions of normality and

homoscedasticity {P > 0.08 as per Shapiro-Wilk and Levene's tests, Milliken and Johnson

1992); means and standard errors were back-transformed for presentation.

I used logistic regression with a generalized logit model to test for differences (a

= 0.05) in frequency distributions of « days that emigration (EM) exceeded immigration

(IM), IM exceeded EM, and EM equaled IM between landuses and years; nonaddivity

was tested by including an interaction term in the model (Agresti 1990:307, Stokes et al.

2000:257- 259). I assumed the categorical response of « days followed a multinomial

distribution per effect, and frequency distributions were independent (Agresti 1990:306).

Normalized chi-square tests of homogeneity (i.e., Z-tests), Bonferroni corrected at a =

0.015, were used to test pairwise differences in frequency distributions when the main

40
effect logit model was significant (Zar 1984:395-396, Milton and Arnold

1995:667-670).

Results

Abundance and Diversity

Mean daily abundance of amphibians was different between landuses and years

and among species (Appendix B). Cropland playas (x=71.5, SE=1.4) had greater

abundance of amphibians than grassland playas (x=22.4, SE=1.4), and abundance was

greater in 1999 (x=67.8, SE=1.4) than 2000 (x-23.5, SE=1.4). However, the species

effect interacted with landuse and marginally interacted with year (Appendix B);

therefore, analyses were separated by species.

Mean daily abundance of NSF and PSF was greater at cropland than grassland

playas, and abundance was greater in 1999 than 2000 for BTS (Appendix B, Table 3.1).

Mean daily abundance also differed among species across and within levels of landuse

and year main effects. New Mexico spadefoot were more abundant than all species at

cropland playas. Mean daily abundance of BTS, GPT, and PSF was greater than SCF,

PLF, and WHT at cropland playas. New Mexico spadefoot, BTS, GPT, and PSF were

more abundant than SCF, PLF, and WHT at grassland playas. Mean daily abundance of

NSF was greater than all other species except BTS in 1999 and 2000. Barred tiger

salamanders were more abtmdant than all other species except NSF and GPT in 1999,

and they were more abundant than other species except NSF, GPT, and PSF in 2000.

Mean daily abundance of GPT was greater than SCF, PLF, and WHT in 1999 and 2000.

41
en
^ p c« W -o -a
o < en < < < < < <
V II

'o
o »y^mmr^ooO'— 'C'r-
a, 04
00
on

C •n- 0 0 — — — —
<N
o ^»y-ir^r<^ — — — —
a
G
t-c
o
o
o
cj _o
< ^ < ^ < <
o -a -O -O "O
<
s^l
o V
o
T3 (N
C *^
c3 oi
D c/5
ON<Nm(N — — — —
C/)

O —
^ CI. • ^ '^^ f S VO - ^ (N (N
IX ^ C^ ^ ^ _ _ _
C <
iH T3
c
cd
rt ca rt C3 X5 X3 X5
2 ] < < CQ < < <
^ ON

on <U
J2
a o 00 T t m - ^ f S — 0 0 —

:a o
C cd
cd >^ — n T;!- p rsj — —, —
IX
O vd S "^ ("^ — —' ^ —
.ti X c

>
p CO x> x> X) u u u
o < < < < < < < £-c
T3 C V
cd ex. 3 II
-a T3
C
c
cd 00 a. u^ r i m m — — o
(U
o Jd i i
o
c U
cd
C
;3
^ — — — mcNcsm
cd IX ro r^ w^ ^- ^ _• ^ ^
>% o
cd
T3
t i - C/D H tin tl.1 UH C
X J«
> O 00 H Cu c/3 U J ^
o a. ; z CQ O cu C/D Cu ^
C/O

cd
o
c
Cd
« cd
.2 T3 (/I
C
3
cd ^ > X>
< •
Plains spadefoot were more abundant than PLF and WHT in 2000. No differences were

detected in amphibian diversity between landuses and years (Appendix B, Table 3.1).

Mean daily abundance of adult NSF was greater at cropland than grassland

playas; landuse and year effects interacted for metamorph and subadult NSF thus

analyses were separated by main effects (Appendix C, Table 3.2). Metamorph NSF were

more abundant at cropland than grassland playas in 1999, and they were more abundant

in 1999 than 2000 at cropland and grassland playas. Mean daily abundance of subadult

NSF was greater at cropland than grassland playas in 2000, and abundance was greater in

2000 than 1999 at cropland playas. Abundance of metamorph PSF was greater in 1999

than 2000, and adult PSF were more abundant in 2000 than 1999; landuse and year

effects interacted for subadult PSF thus analyses were separated by main effects. Mean

daily abundance of subadult PSF was greater at cropland than grassland playas in 2000,

and abundance was greater in 2000 than 1999 at cropland playas. Metamorph BTS were

more abundant in 1999 than 2000. No additional differences were detected between

landuses and years in mean daily abundance for age classes of significant species

(Appendix C, Table 3.2).

Source-Sink Dynamics

Frequency distributions of n days for the amphibian assemblage that emigration

(EM) exceeded immigration (IM), IM exceeded EM, and EM equaled IM were different

between landuses (xV26.1, P<0.001) and years (xV27.1, P<0.001); main effects did

not interact (% 2^2.8, P=0.25). Percent days EM exceeded IM (i.e., source dynamics)

43
Table 3.2. Relafive daily abundance of age classes between landuse types and years for
amphibian species that differed significantly in daily abundance (see Appendix B) at 16
playa wetlands on the Southern High Plains, Texas, May-October 1999 and
April-August 2000.
Landuse
Species^ Age Year^ Cultivation Grassland
X SE' P<0.05^ X SE P<0.05
NSF Metamorph 1999 100.5 1.7 Aa 3.1 1.8 Ba
2000 1.6 1.5 Ab 2.2 2.1 Aa
Subadult 1999 1.2 1.0 Aa 1.3 1.2 Aa
2000 11.8 1.4 Ab 2.1 1.1 Bb
Aduh 1999 8.9 1.2 Aa 3.5 1.5 Ba
2000 12.2 1.2 Aa 3.2 1.5 Ba
PSF Metamorph 1999 2.6 1.2 Aa 1.9 1.3 Aa
2000 1.3 1.2 Ab 1.0 1.0 Ab
Subadult 1999 1.0 1.0 Aa 1.1 1.0 Aa
2000 3.3 1.4 Ab 1.2 1.0 Ba
Adult 1999 1.3 1.2 Aa 1.3 1.1 Aa
2000 2.9 1.3 Ab 1.6 1.1 Ab
BTS Metamorph 1999 8.9 1.7 Aa 14.5 1.1 Aa
2000 2.6 1.2 Ab 4.7 1.5 Ab
Adult 1999 2.2 1.2 Aa 2.4 1.3 Aa
2000 2.7 1.3 Aa 3.2 1.4 Aa
^ S F = New Mexico spadefoot {Spea multiplicata), PSF = plains spadefoot {S.
bombifrons), and BTS = barred tiger salamander {Ambystoma tigrinum mavortium).
''Analyzed by year for metamorph NSF and subadult NSF and PSF because landuse and
year main effects interacted (Appendix C); statistics for remaining ages and species
presented by year for tabular parsimony.
'^n=A playa wetlands per landuse per year; xs and SEs were back-transformed from
natural-logs for presentation; transformation was necessary to meet linear model
assumptions of ANOVA.
''within-species means in the same row (i.e., within ages and years) with unlike
uppercase letters are different (i.e., landuse effect test); means in the same column within
species and ages with unlike lowercase letters are different (i.e., year effect test).

44
was greater at grassland than cropland playas (Zi=2.7, /'=0.007, Figure 3.1), and greater

in 1999 than 2000 (Zi=2.7, P<0.001). Percent days IM exceeded EM (i.e., sink

dynamics) was greater at cropland than grassland playas (Zi=4.9, P<0.00\), and greater

in 2000 than 1999 (Zi=2.4, P=0.015). Percent days EM equaled IM (i.e., neutral

dynamics) was greater at grassland than cropland playas (Zi=3.1, P=0.002), and greater

in 2000 than 1999 (Zi=2.9, P=0.004). At cropland playas, percent days IM exceeded EM

was greater than percent days EM exceeded IM and EM equaled IM (Zi=8.4-13.5,

P<0.001). Also, percent days EM exceeded IM was greater than percent days EM

equaled IM at cropland playas (Zi=5.6, /*<0.001). At grassland playas and in 1999,

percent days EM exceeded IM and IM exceeded EM were greater than percent days EM

equaled IM (Zi=5.4-10.6, P<0.001), but the former did not differ (Zi=0.9-1.5,

P=0.147-0.322) between each other. In 2000, percent days IM exceeded EM was greater

than percent days EM exceeded IM and EM equaled IM (Zi=8.3-9.3, P<0.00\), but latter

did not differ (Zi=l.l, P=0.292) between each other (Figure 3.1).

Discussion

Mean daily abundance of amphibians was greater at playa wetlands surrounded

by cultivation than at grassland playas, and abundance was greater in 1999 than 2000.

However, when I combined data for all species, frequency of days emigration exceeded

immigration for the amphibian assemblage was greater at grassland than cropland playas;

no yearly differences existed. In contrast, frequency of days immigration exceeded

emigration was greater at cropland than at grassland playas and greater in 2000 than

1999. In general, New Mexico spadefoots were more abundant than all other species,

45
0.6
0.5 :
!
ays 0.4 Ba
Q
Aa Cropland
0.3
Percei

Grassland
#»f#
0.2

0.1
4 •, * ]

0 -—^ * *

EM IM EQ
Directional Movement

11999
12000

EM IM EQ
Directional Movement

Figure 3.1. Effect of landuse (top graph) and year (bottom graph) on frequency of
directional movement of amphibians at 16 playa wetlands on the Southern High Plains,
Texas, May-October 1999 and April-August 2000; EM = percent days emigration
exceeded immigration (i.e., source dynamics), IM = percent days immigration exceeded
emigration (i.e., sink dynamics), EQ = percent days emigration equaled immigration (i.e.,
neutral dynamics); unlike uppercase letters within directional movements and unlike
lowercase letters within landuses are different (P<0.007) by a main effect logistic
regression (generalized logit model) and pairwise chi-square tests of homogeneity
Bonferroni corrected at a = 0.015 (Milton and Arnold 1995:667-670, Stokes et al.
2000:257- 259).

46
followed numerically by barred tiger salamander and Great Plains toad. No difference in

mean daily diversity of amphibians was detected between landuses and years.

Landuse Effect

Abundance of spadefoots may have been elevated at cropland playas, because

landscape disturbance may have confined individuals to the remnant available habitat

(i.e., the playa, Kolozsvary and Swihart 1999). Spadefoots may have become more

nested in cropland wetlands than other species, because of their relative vagility and

perception to patch viscosity and edge permeability in adjacent agricultural fields (Wiens

1997). Indeed, geometric complexity (i.e., mean fractal dimension, FD) and edge density

(i.e., m edge/ha, ED) were greater {P<0.0\, Wilcoxon 2-sample nonparametric test,

Conover 1980:216) in cropland (XFD=1.32, S E = 0 . 0 1 ; XED=64.3, S E = 2 . 0 ) than grassland

(XFD=1-28, S E = 0 . 0 1 ; XED=40.6, S E = 3 . 5 ) landscapes (also see Chapter IV). Road density

also was greater in cropland than grassland landscapes (Appendix A), which can decrease

permeability (Yanes et al. 1995). Thus, dispersing spadefoots may have been unable to

penetrate cropland landscapes, resulting in increased nestedness and abundance at natal

wetlands (Kolozsvary and Swihart 1999).

Spadefoot vagility may have been less than other species on the Southern High

Plains, because of their relatively small postmetamorphic body size (With and Crist

1995). Species-specific perception to landscape context and migration distance is

positively correlated with body size for various organisms (e.g., Peters 1983:89-91, Crist

et al. 1992, With 1994). Excepting the spotted chorus frog, spadefoots were smaller

physically than other species that I monitored (Degenhardt et al. 1996:61, 81, Chapter II).

47
I would expect the spotted chorus frog to be affected similarly by landscape disturbance

if population demographics in playa wetlands were exclusively dictated by the relative

connectivity between them. Differences may not have been detected in demographics

between landuses for spotted chorus frog, because they have relatively low abundance on

the Southern High Plains (Anderson et al. 1999fl, this study). Alternatively, other

proximate factors may be individually or synergistically responsible for demographic

responses to landuse.

Elevated spadefoot abundance in cropland playas also may have been a

consequence of changes in the trophic structure of the aquatic environment. Barred tiger

salamander may fiinction as a keystone intraguild predator {sensu Paine 1969, Polls and

Holt 1992) in playas similar to different species in other wetland systems (e.g.,

Ambystoma opacum. Walls and Williams 2001), because they can establish larval and

neotenic populations prior to anurans if water is present during winter or early spring

(Rose and Armentrout 1974, 1976). Indeed, density of larval and neotenic barred tiger

salamanders was greater in playas surrounded by grassland than in cropland playas,

presumably due to longer hydroperiods associated with grassland playas (Chapter II).

Grassland playas generally have greater volume and longer hydroperiods than cropland

playas because of differential sedimentation (Luo et al. 1997). Consequently, barred tiger

salamander populations became established in grassland playas before anurans bred both

years (M. J. Gray, unpublished data), which may have resulted in top-down control on

spadefoot populations via predation on their eggs and larvae (Henrickson 1990, Walls

and Williams 2001). Insect predators (e.g., Anax, Dytiscus) also may have been abundant

48
in grassland playas due to their longer hydroperiods (Schneider and Frost 1996). In

contrast, cropland playas had shorter hydroperiods and were dry until intense rains early

summer both years, resulting in simultaneous breeding and larval development of

salamanders, anurans, and insects. Thus, salamander and insect larvae likely interacted

with spadefoot larvae as competitors not predators in cropland playas (Morin 1983.

Morin et al. 1988). Other predators of amphibian larvae (e.g., wading birds, fish) occur

infrequently in playa wetlands during summer (Bolen et al. 1989); thus, cropland and

grassland playas may have represented relatively predator-free and -rich environments,

respectively (Heyer et al. 1975). The aforementioned may have facilitated greater

recruitment of metamorphs into spadefoot populations located in cropland than in

grassland.

Spadefoot larvae may have been more susceptible to predation in grassland pla\ as

than other species on the Southern High Plains, because of their relative palatabilit\ and

escape probability (Alford 1999:260-264). Spadefoots typically inhabit xeric

environments worldwide and exploit ephemeral water sources with few vertebrate or

invertebrate predators (Bragg 1965, Dimmitt 1975). Consequently, spadefoots have

evolved physiological mechanisms for rapid growth (Richmand 1947, Newman 1992) but

lack the ability (e.g., toxicity, cryptic coloration, altered microhabitat use) to efficienth

escape predators (e.g., Spea bombifrons, Kruse and Francis 1977). For example,

spadefoot tadpoles are more active than tadpoles of certain Bufo, Rana, and Pseudacris

species; thus, they are more conspicuous and vulnerable to predators (Morin 1983.

Woodward 1983, Dayton and Fitzgerald 2001). Spadefoot tadpoles also lack dermal

49
toxins and cryptic coloration unlike many species of Bufo, Rana, and Pseudacris (e.g.,

Walters 1975, Formanowicz and Brodie 1982, Peterson and Blaustein 1991, Smith and

VanBuskirk 1995).

Spadefoots also may be superior competitors to other species on the Southern

High Plains (e.g., Morin 1983, Wilbur 1987, Dayton and Fitzgerald 2001). Thus, in the

absence of predation, spadefoots might outcompete other species for food and other

resources resulting in increased recruitment (Woodward 1982, Wilbur 1987, Dayton and

Fitzgerald 2001). Indeed, several elegant lab experiments have documented that

spadefoot tadpoles are competitively dominant in anuran communities when aquatic

predators are absent (e.g., Morin 1981, 1983, Wilbur 1987). Unfortunately, no studies

have been performed on the relative palatability, escape probability, or competitive

ability among all my species simultaneously; thus, I only can speculate that

aforementioned hypotheses are responsible for differential demographics in spadefoot

populations between landuses on the Southern High Plains.

Additional modifications of the aquatic and terrestrial envirormients also may

have positively influenced spadefoot populations at cropland playas. Smith and Haukos

(2002) noted that plant diversity and structure could be greater in cropland playas, which

could increase food resources and escape cover for larvae. Nitrogen influx from

fertilizers may have increased food resources for larval amphibians in cropland playas

(Leibold and Wilbur 1992). Additionally, pesticide drift and runoff into cropland playas

may have reduced aquatic insect densities (Boone and Semlitsch 2001), which can

compete with amphibian larvae for food resources (Morin et al. 1988). Similarly,

50
cultivafion and chemical application may have reduced invertebrate densities and cover

in the surrounding landscape near cropland playas (Freemark and Boutin 1995), inducing

greater nestedness in remnant available habitat within disturbed landscapes as mentioned

previously (Kolozsvary and Swihart 1999). However, if any of the aforementioned

suggestions were strictly true, I should have observed similar demographic responses by

species other than spadefoots. Thus, these hypotheses probably are not ultimate factors

responsible for observed demographics in our disturbed and undisturbed landscapes.

Lastly, subadults were the only age class of plains spadefoots that was greater in

cropland than grassland playas. I am uncertain why the other age classes did not respond

similarly considering the aforementioned logic concerning spadefoot response; however,

it may have been a consequence of delayed development or competition. Subadults were

individuals >1 year that did not exhibit sexual characteristics thus probably did not breed

(Chapter II). Amphibians may forego breeding if resource conditions are not adequate

(Semlitsch et al. 1996). Perhaps, cropland landscapes contained less food and other

resources than grassland landscapes (Freemark and Boutin 1995), resulting in greater

number of individuals at cropland playas that did not breed. However, if this was true, I

should have observed similar elevated subadult levels with other species. Alternatively,

increased New Mexico spadefoot abundance at cropland playas may have induced

delayed breeding of plains spadefoots, because these species are known to compete for

resources (Pfermig and Murphy 2000). Thus, abundance of subadult plains spadefoots

may have been a consequence of interspecific competition with New Mexico spadefoot

(see Chapter V).

51
Year Effect

Yearly differences in abundance occurred because of a greater number of

metamorph New Mexico and plains spadefoots, and barred tiger salamanders in 1999

than 2000. Demographic differences between years likely were related to rainfall

(Berven 1995, Semlitsch et al. 1996). Average monthly rainfall at my playas was greater

in 1999 than 2000 (Chapter II). Rainfall stimulates amphibian emergence (Dimmitt and

Ruibal \9S0b) and positively affects intra- and inter-demic movement (Hurlbert 1969,

Sinsch 1988). Adult survival also can be positively related to rainfall (e.g., Berven

1990), presumably because density of food resources increases (Dimmitt and Ruibal

1980a) and probability of desiccation decreases (Jaeger 1980, Newman and Dunham

1994). Probability of breeding, larval survival, and juvenile recruitment also increases

with rainfall, because wetland hydroperiods are positively correlated with rainfall

(Pechmann et al. 1989, Sinsch and Seidel 1995, Semlitsch et al. 1996). Indeed, mean

duration of hydroperiods at my study playas was greater in 1999 than 2000 (Chapter II).

Lastly, the difference in yearly abundance of metamorph spadefoots occurred only at

cropland playas, suggesting that disturbance may affect recruitment greater during drier

years (Lannoo 1998).

Three inconsistencies existed in my year results; abundance of subadult New

Mexico and plains spadefoots was greater in 2000 than 1999 at cropland playas, and adult

plains spadefoots were more abundant in 2000 than 1999. Subadult spadefoot abundance

probably was elevated in 2000 because of increased recruitment experienced in 1999 at

cropland playas (Semlitsch et al. 1996, Meyer et al. 1998). Also, food densities may have

52
been lower during the drier year (i.e., 2000) at cropland playas (Dimmitt and Ruibal

1980a), resulting in more individuals not breeding and consequently classified as

subadults. I am uncertain why adult plains spadefoots were more abundant during 2000

in both landuses, but perhaps drier conditions during this year facilitated competitive

dominance over some species.

Source-Sink Dynamics

Frequency of days emigration exceeded immigration (i.e., source dynamics) for

the assemblage was greater at grassland than cropland playas and in 1999 (i.e., the wetter

year) than 2000. Grassland playas may have served as sources more often than cropland

playas, because of their longer hydroperiods (Chapter II). Consequently, larvae may

have metamorphosed more continuously from grassland than cropland playas, facilitating

the former to function as a more constant source of emigrating propagules. Also, abiotic

conditions (e.g., relative humidity) may have been more ideal in landscapes surrounding

grassland than cropland playas for interdemic movement, particularly the exodus of

individuals (Bardsley 1998).

Source dynamics likely were greater in 1999 than 2000, because favorable

ambient conditions resulted in greater production of individuals (Semlitsch et al. 1996).

Lannoo (1998) suggested that source dynamics would occur more frequently during

relatively wet years in seasonal and semipermanent wetlands. Indeed, catastrophic

mortality or reduced recruitment of individuals may occur in ephemeral wetlands if they

dry before larvae can attain a critical body-size threshold for metamorphosis (Semlitsch

19876, Newman 1992, Denver 1997).

53
Conservation Implications

Anthropogenic disturbance around wetlands can affect demographics and

community structure of amphibian populations. Less vagile species, such as spadefoots

and chorus frogs, may be affected more by disturbance, resulting in greater nestedness of

their populations near natal wetlands. Thus, I caution exclusive interpretation of my

abundance results, because elevated values could be a consequence of area-restricted

movement or reduced home ranges. Moreover, more individuals does not necessarily

imply better conditions for a population (Yeargers et al. 1996). Populations with elevated

abundance may be less stable, particularly if they exceed their carrying capacity or have a

high intrinsic rate of increase (Edelstein-Keshet 1988). Pathogenic incidence also can be

positively correlated with population density (Hastings 1996). Additionally, landscape

disturbance may negatively affect genetic structure of populations (Reh and Seitz 1990)

and fitness correlates, such as postmetamorphic body size of amphibians (Chapter II).

My source-sink results for combined species also suggest that playas in cultivated

landscapes may serve as sinks more often than sources, perhaps because of wetland

hydroperiod and connectivity is less than in undisturbed landscapes. Considering the

aforementioned, amphibian assemblages in disturbed landscapes may have a greater

probability of local or metapopulation extinction than those positioned in natural

landscapes (see Chapter VI also).

Additional research is needed to discern proximate and ultimate mechanisms

driving population responses to anthropogenic disturbance. For example, the effect of

chemical and mechanical disturbance in and around wetlands on the trophic structure of

54
invertebrate and amphibian communities needs to be investigated. Pathogenic incidence

in amphibians (i.e., viruses, fungi, bacteria, and parasites) also should be quanfified in

disturbed and undisturbed wetland systems. Estimates of landuse type and boundary

permeability to amphibian movement also are needed. This latter suggestion would

enable landscape ecologists to assign permeability and hardness indices to cover types

and boundaries in spatial analysis programs (e.g., FRAGSTATS, McGarigal and Marks

1995), and subsequently estimate relative connectivity of habitat patches and extinction

probabilities for local populations. Finally, my year results suggest that annual rainfall

may be an important variable to include in extincfion probability models for amphibians.

55
CHAPTER IV

INFLUENCE OF LANDSCAPE STRUCTURE ON COMMUNITY

COMPOSITION AND RELATIVE ABUNDANCE OF AMPHIBIANS

Abstract

Landscape structure can influence local dynamics of spatially structured

populations, particularly with less \agile organisms such as amphibians. I examined the

effect of landscape structure on communit}' composition and relative abundance of

amphibians inhabiting the Southern High Plains of Texas. Amphibian populations were

monitored using pitfall traps and drift fence located at 16 playa wetlands (8 playas/year)

in 1999 and 2000. I quantified landscape structure surrounding each playa via estimating

13 spatial metrics that represented playa isolation and landscape complexity using remote

sensing from aerial photographs and geocorrection and spatial analysis software.

Multivariate ordination and univariate correlations and regressions indicated that

landscape structure influenced {P<0.05) community composition and mean daily

abundance of amphibians. Spadefoots {Spea multiplicata, S. bombifrons) generally were

positively associated with metrics representing playa juxtaposition and landscape

complexity. Great Plains toads {Bufo cognatus) and barred tiger salamanders

{Ambystoma tigrinum mavortium) usually were negatively associated with spadefoots but

not affected by landscape structure. Differences in community composition and relati\ e

abundance probably were related to species-specific vagility and perception to patch

viscosity and edge permeability. Spatial separation of these species in the ordination also

may have been a consequence of differential competitive ability for food or other

56
resources in aquatic and terrestrial environments. Results in Chapter III suggested

general landuse (i.e., cultivation or grassland) could influence demographics of

amphibians. These additional findings indicate that landscape structure may be as or

more important than general anthropogenic landuse in influencing demographics of

amphibian populations. I also present multiple regression models for predicting relative

species-specific daily abundance using field estimates of spatial metrics. This

information can be used by landscape ecologists to locate potential sites for amphibian

conservation initiatives.

Introduction

Landscape structure can influence demographics of spatially structured

populations via influencing probability of interdemic movement (Fahrig and Paloheimo

1988). Components of landscape structure include spatial positioning of habitat patches

and the geometric matrix in which they are embedded (Fahrig and Merriam 1994).

Relative abundance of wildlife generally declines as habitat patches become isolated or

are not optimally juxtaposed (Lefkovitch and Fahrig 1985, Burel 1989). Moreover, as

landscapes become more geometrically complex, probability of successful dispersal may

decrease because boundary density increases (Stamps et al. 1987, Wiens 1997).

Consequently, isolated populations embedded in complex landscapes may have a greater

probability of extinction than those more optimally positioned within simpler landscapes

(Fahrig and Merriam 1985, Taylor et al. 1993).

Amphibian populations are declining worldwide (Houlahan et al. 2000).

Anthropogenic modifications of habitat patches and the landscape matrix have been

57
implicated in many declines (Blaustein et al. 1994). Various studies have demonstrated

amphibian abundance is negatively related to wetland or terrestrial habitat isolation (e.g.,

Marsh and Pearman 1997. Pope et al. 2000, Scribner et al. 2001). Anthropogenic

modifications of the landscape (e.g., cuhivation) may induce isolation by disrupting

connectivity among patches (Marsh and Trenham 2001). Connectivity in disturbed

landscapes might be negatively affected because of differential viscosity of

anthropogenic cover types or permeability of edges (Durelli et al. 1990. With 1994,

Wiens 1997). Consequently, amphibian abundance and probability of persistence in

wetlands surrounded by disturbance may be lower than in undisturbed wetlands (Sjogren

1991, Sjogren-Gulve 1994). Alternatively, cultivated landscapes may possess greater

edge density and less terrestrial habitat, resulting in increased nestedness and abundance

of amphibians in remaining remnant wetlands (Knutson et al. 1999, Kolozsvary and

Swihart 1999, Chapter III).

On the Southern High Plains (SHP), amphibians exist primarily in playa wetlands

(Bolen et al. 1989). General landscape use (i.e., cultivation vs. grassland) can affect

postmetamorphic body size and demographics of amphibian populations in playas

(Chapters II and III); however, the effect of landscape structure on community

composition and relative abundance of amphibians in playas is unknown. Because

amphibians can interact among playas via dispersal (M. J. Gray, unpublished data),

landscape structure may be an important factor influencing local population dynamics.

Thus, I examined the influence of landscape structure on amphibian populations via

estimating various spatial metrics and relating them to relative daily abtmdance of 4 SHP

species {Spea multiplicata Cope [New Mexico spadefoot], S. bombifrons Cope [plains

58
spadefoot], Bufo cognatus Say [Great Plains toad], Ambystoma tigrinum mavortium

Green [barred figer salamander) at 16 playas (8/year) during 1999 and 2000. I

hypothesized that relative abundance of amphibians would be negatively influenced b>'

playa isolation (Loman 1988, Laan and Verboom 1990, Vos and Stumpel 1995, Lehfinen

et al. 1999). I also hypothesized that landscape complexity would be positively related to

abundance (Knutson et al. 1999, Kolozsvary and Swihart 1999). Understanding the

effects of landscape structure on population dynamics is a ftjndamental construct of

landscape ecology, and a critical prelude to conservation initiatives for amphibians

(Marsh and Trenham 2001).

Methods

This study was conducted at 16 playa wetlands on the Southern High Plains of

Texas. Playas were partially enclosed (i.e., 25% of circumference) with 60-cm high drift

fence and 19-L pitfall traps (Dodd and Scott 1994). Pitfall traps were checked alternate

days for captures from 16 May-17 October 1999 and 19 April-18 August 2000.

Captured amphibians were enumerated by species, marked uniquely, and released. A

more detailed account of my study area, amphibians, and sampling protocol was

presented in Chapters II and III.

I quantified landscape structure surrounding each study playa using remote

sensing from aerial photos and geographical information system and spatial analysis

software. Landscapes («=16) were 2,830-ha circular plots (i.e., 3-km radius) with their

origins positioned at the center of each study playa (Appendix D). I chose a 3-km radius

to delineate the landscape plot, because this distance probably was the near maximum

59
dispersal distance for my species (cf Gehlbach 1967, Gehlbach et al. 1969, Sinsch 1990,

1997, Miaud et al. 2000). Aerial images of my landscapes in summers 1999 and 2000

were obtained from the Farm Service Agency (FSA) of the U.S. Department of

Agriculture in Crosby, Floyd, Hale, and Castro counties, Texas. Ground control points

were identified on images and U.S. Geological Survey 7.5-minute quadrangle maps then

geo-corrected using ERDAS® software. Geo-referenced images were mosaicked in

ERDAS® and cover types (i.e., crop type, replanted Conservation Reserve Program

grass, and natural grass) were identified in each landscape plot using 1999 and 2000 FSA

farm folders. Boundaries of cover types were digitized in ERDAS® then exported as an

ESRI® ARC/INFO coverage. Coverages were cleaned, built, and polygons classified as

per cover types in ESRI® ARC/INFO. Landscape structure was quantified per plot using

FRAGSTATS* ARC® {www.innovativegis.com, McGarigal and Marks 1995); I used the

following 13 spatial metrics. Shape index of the study playa (PSI), study playa size (PS),

mean nearest-neighbor distance from the study playa to surrounding playas (PNN), mean

nearest-neighbor distance from all playas to each other (MNN), number of playas (NP),

percent aerial coverage of playas (PP), and an interspersion/juxtaposition index of playas

(IJI) were used to quantify relative spatial positioning and isolation of playa wetlands

(McGarigal and Marks 1995). Mean number of edges to cross from the study playa to

surrounding playas (PED), edge density (m edge/ha, ED), landscape shape index (i.e.,

measure of geometric complexity, LSI), landuse (i.e., cover type) richness (LR), Shannon

evenness index of landuses (SEI), and Shannon diversity index of landuses (SDI) were

used as measures relative landscape complexity (McGarigal and Marks 1995). I direct

readers to McGarigal and Marks (1995) for spatial metric formulae and algorithms.

60
Unity was assigned to cover type and edge permeability in FRAGSTATS* ARC®

(McGarigal and Marks 1995), because relafive viscosity for my species is unknown.

I used mean daily capture per species per playa as an index of relative daily

abundance (Dodd and Scott 1994). Canonical correspondence analysis (CCA) was used

to examine the effect of landscape structure metrics on relative abundance of species in

the amphibian assemblage (ter Braak 1986, 1994). Correlated metrics (P<0.05) were

removed prior to CCA to reduce probability of an arch effect (ter Braak 1995:139).

Because CCA is sensitive to outliers and bimodally distributed data, mean daily

abundance per species was natural-log transformed prior to analysis (ter Braak 1995); all

resulting distributions were unimodal with <2 outliers. A global Monte Carlo

permutation test was performed to test for existence of a relationship between landscape
V

metrics and species composition (ter Braak and Smilauer 1998:47-49, 124-125). A

dimensionless species-landscape metric biplot was constructed using CANOCO® to

graphically examine the pattern of variation in relative species abundance with landscape

metrics (ter Braak 1995, ter Braak and Smilauer 1998). Biplot axes were interpreted by

examining the sign and magnitude of their respective intra-set correlations and using

subject-matter knowledge (ter Braak 1995:140). Length of metric-specific eigenvectors

(i.e., the arrows) in the biplot was interpreted as the strength of correlation between the

metric and species abundance (ter Braak 1995:141-142). Therefore, long eigenvectors

were most important in affecting the species assemblage (ter Braak 1995:141). Relative

correlative ranking of species with respect to the metrics was graphically represented by

extending each eigenvector through the origin of the biplot and intersecting it with

orthogonal lines drawn from the species (ter Braak 1995:143). Species positioned near

61
the arrow- and blunt-end of the eigenvector were most positively and negafively

correlated with the metric, respecfively (ter Braak 1995:141-143).

Canonical correspondence analysis in CANOCO® does not provide univariate

measures of association between individual metrics and species-specific abundance (ter

Braak and Smilauer 1998). Thus, I also calculated Pearson coefficients of correlafion and

tested the relationship of each metric with species-specific abundance (Milton and Arnold

1995:425-430). Metrics that were correlated (P<0.05) with species-specific abundance

also were regressed linearly using least-squares estimation, and univariate prediction

models developed (Milton and Arnold 1995:386-391). Finally, I used multiple linear

regression with stepwise selection (entry and stay significance level at a=0.15) to develop

models for predicting relative daily abundance per species using all landscape metrics

and 2 indicator variables (i.e., general landuse [O=cropland, l=grassland] and year [0=wet

year, l=dry year]) as potential explanatory variables (Milton and Arnold 1995:497-501,

504-506). These indicator variables were included in prediction model development,

because results in Chapter III suggested that general landuse and annual rainfall could

affect population demographics of amphibians. Linear dependency (i.e., coUinearity

[intercept adjusted]) among predictor variables in the final models was examined via

estimating variance inflation factors (VIF) and condition numbers (CN); VIF > 10 and

CN > 30 were suggestive of coUinearity between >1 variable (Freund and Littell

2000:98-101). Standardized and unstandardized parameter estimates were presented for

linear interpretation and predicting relative daily abundance of amphibians using field

estimates of metrics or indicator variables, respecfively (Myers 1990:384-385).

Coefficients of determination adjusted for number of variables in the model (i.e., R adj)

62
were presented as a measure of model performance (Freund and Littell 2000:23). Mean

daily abundance per species were natural-log transformed for all correlation and

regression analyses to satisfy (P>0.05 as per Shapiro-Wilk tests) linear model

assumpfions (Milton and Arnold 1995:391-392).

Results

Six of the 13 landscape metrics (PNN, NP, PED, ED, LSI, and SDI) were not

used in the canonical correspondence analysis, because they were correlated with other

metrics (Table 4.1). The global Monte Carlo permutation test based on 199 permutations

revealed that species composition was associated (F=2.201, P=0.05) with landscape

metrics. The first two canonical axes were most important in explaining variation in the

species assemblage (i.e., as per respective eigenvalues, ?ii=0.061, ?t2=0.022, >.3=0.001,

X4=0.013); they collectively explained 67.9% of variation in mean daily abundance of

these SHP amphibian species. Intra-set correlations suggested that axes 1 and 2

explained variation in species abundance with respect to wetland positioning and

landscape complexity, respectively, because of the large absolute magnitude of PP and IJI

for axis 1 and SEI for axis 2 (Table 4.2). The dimensionless species-environmental biplot

of axes 1 (ordinate) and 2 (abscissa) suggested that SEI was most positively correlated (as

per eigenvector length) with the pattern of species abundance (Figure 4.1a). Playa size,

PP, IJI, and LR were moderately correlated and PSI and MNN least correlated with

species composition patterns. Orthogonal inferred ranking of species against

eigenvectors suggested that both spadefoot species were most positively associated with

PS, PP, IJI, and LR; barred tiger salamander and Great Plains toad generally were

63
O
to O
OS d
V
oo Os C7\ m
m oo oo o
T^ o so o
t) H d d d d
o m CN -^ •<:l-
o (N r<-) r^ o
o
d d ON CN r<^
V o o o

O C3 ON »o -^ ON
OS O Q -^ »r^ OS vo
OS . O oo csl (N
o ^ O "^ d d d d

"^o*^c:50<50\J2r--S5<N
(-5 »'i v^w -^ 0\ 2 ON
• ^ O ^. "^ "^ ON
^ ' ^ ^ ^ d d o o o o

r - m r - — fNtNONONSIJ-^moN
O N O C N l O C N O V O i / O " — ' o - ^ v o
voot~-or~-om-^^r^ooo
o o o o o o o o o o o

oo -^ t ^ ON o o r -
r-. vo r- — oo m oo
O <N CN ON ^ ON
CO ^ CU 'sf o lO o u-^ voocncs<^m*^r-
o o o o o o o o o o ' ^ o ' ^ o

voONmf<^ONT}-ONONOO^^u-ioo^m^^(N
G- o - ^ v o f N ' ^ o o m O N c n o o ON W-i CO —
c n t N i r j O r n — c o - ^ m c N ^ - vo ON O 0 \
d d d d d d d d d d d d oI d d d
^»r^zroo;3^r~-r"r-^ ^ ^ o
- vo -^ oo
^. O
_ ' p. ON ^ . O
o ^~. u-) <N r«-, CN (^ —;. v_o —- VO O oo
< = ' d < = = ' d < = > d ' = ' d o d < = > d- • ^' " ' S- • -<z>
• d <=i ^

m m '^ c<i <^lCJ<^^o2<N^r^lCiv£»r~-ON>y-^Tt


co£!Do\iQ'^Ho^S^o?Sw->r-ocM'^
-^ ON <in^ CO
"^ (N. p ^ c o < ^ - ^ < ^ ( N P t ^ — lO — v o
d dO d od o d o o o o o
I

-^•^r-^co-^ONcnoooovovovouooor^iriOomr^tN^^
r^ONO-^ — »/^r-)Oor~-mcnoosvocncnr-i^j^^'-'ON
o r ~ - ( N ' ^ o a \ c n c N O O N O o o — ^O'— i r > c s c n < ^ ' ^ ' ^ . vo
d o d d o d d o o o o o o o o o d d p d
*/^ r~- vo ^,
en f!:^voOOcMOsco-<^cnONOO«y^Oscnr^^ — r J t N
t-- !_ o ^ •£'^^cnvOc<^'^NOcnOcnONCNcnS^^O
-: u-^ ^ o
^ i r j P o o - - » n c s c n - ^ ^ o o o o o s g ' ^ <^. e n < ^ <N
'^<d>'^<6ci<6c><zi<d><zi<d><6d><:6 ,- d ^O <zi '^ <z:>
<=> d 9 d

Q
CO 00
a. OH
OH
Q C/3
00
Q
C/3

64
T3
• * - '

O 2 eu
<V

o
<u
•4—'

c
o

<U
O
C/3 a-^
O

Q3 .^

OX)

o
..—t

T3
>-i
c ^<

en N 00

on
OH
00
>^-s
3
o
C/3 CO C/5

»-l II H <L>
*o
I
<U
o
CO
oo
(N
<L)

-a VO

c
o
w
u. (U
ft (/3 -TH
^-H
cd
'St o
<U T3

3
Cd
§ a ^ g | >>&
h-1
H ca Quo c/j Cd DH O

65
Table 4.2. Intra-set correlations of landscape^ metrics associated with the first two axes
generated by a canonical correspondence analysis of mean daily abundance of New
Mexico spadefoot {Spea multiplicata), plains spadefoot {S. bombifrons). barred tiger
salamander {Ambystoma tigrinum mavortium). and Great Plains toad {Bufo cognatus) at
16 playa wetlands on the Southern High Plains, Texas, 1999 and 2000.
Metric^ Correlations^
Axis 1 Axis 2
PSI 0.0202 -0.1854
PP -0.6970 0.1859
IJI -0.7625 -0.0788
MNN 0.1489 -0.0134
LR -0.5559 -0.2485
SEI 0.0092 -0.7729
PS -0.3993 0.4379
Landscapes {n=\6) were 2830-ha circular plots (i.e., 3-km radius) with their origins
positioned at the center of each study pla\a.
PSI = shape index of study playa. PP = percent aerial coverage of playas. IJI =
interspersion/juxtaposition index of playas, MNN = mean nearest-neighbor distance from
all playas to each other. LR = landuse richness, SEI = Shannon evenness index of
landuses, and PS = study playa size (McGarigal and Marks 1995).
'^Intra-set correlations are standardized and dimensionless thus the\ can be interpreted as
the strength and direction of metric-specific correlation and species abundance in the
presence of all other metrics (ter Braak 1995:140): I infer that axes 1 and 2 explained
variation in species abundance with respect to wetland positioning and landscape
complexity because of the large absolute magnitude of PP and IJI for axis 1 and SEI for
axis 2.

66
67
Figure 4.1. Canonical correspondence analysis of mean daily capture (natural-log
transformed) of amphibians and uncorrelated metrics in landscapes (i.e., 2,830-ha circular
plot) associated with 16 study playas on the Southern High Plains, Texas, 1999 and 2000.
a) Species-environmental (i.e., -landscape metric) biplot (ter Braak 1995:142); axes 1
(ordinate) and 2 (abscissa) are dimensionless and represent wetland positioning and
landscape complexity, respectively (see corresponding intra-set correlations in Table 2);
length of eigenvectors (i.e., their respective eigenvalue) indicates the strength of
correlation between the variable and pattem of variation in species composition (ter
Braak 1995:141); species more close to ends of eigenvectors are more positively
correlated with it; PS = study playa size, SEI = Shannon evermess index of landuses, PP
= percent aerial coverage of playas, PSI = shape index of study playa, MNN = mean
nearest-neighbor distance from all playas to each other, IJI = interspersion and
juxtaposition index of playas, and LR = landuse richness (McGarigal and Marks 1995).
b) Inferred ranking of species along variables based on biplot interpretation of Part a of
figure (ter Braak 1995:143); ranking constructed by extending eigenvectors through the
origin and intersecting with orthogonal lines from the species to the vector; the vertical
segment represents the origin (i.e., centroid) of the biplot and is the grand mean of each
variable; species more close to the arrow or blunt end are positively or negatively
correlated with the variable, respectively. NSF = New Mexico spadefoot {Spea
multiplicata), PSF = plains spadefoot {S. bombifrons), GPT = Great Plains toad {Bufo
cognatus), and BTS = barred tiger salamander {Ambystoma tigrinum mavortium).
a)

PS

PP
' ..^^^pspy i GPT


IJI ^,..-'^^^?.Y
I
PSI
LR
-

' jSEI

I r — - - - " - T — - t f T • T ---—,-- f 1 f T T

-1.0 + 1.0
p N G B P G B N
s S P T S P T S
PS F T S F T S F SEI
b)

P N G B p G B N
S S P T s P T S
PP F F T S F T S F PSI
^ - •
N P G B P N G B
S S P T S S P T
IJI F F T S F F T S MNN
^4- —•
N P B G
S S T P
LR F F S T

68
negatively associated with these spatial metrics (Figure 4.\b). In contrast, barred tiger

salamander and Great Plains toad were most positively and the spadefoots most

negatively related with MNN. Plains spadefoot also appeared to be negatively associated

with SEI and PSI (Figure 4.1^?).

Mean daily abundance was correlated linearly with PP, IJI, ED, LSI, PED. and

LR for New Mexico spadefoot and the first 4 aforementioned metrics for plains

spadefoot; significant correlations were not detected for Great Plains toad or barred tiger

salamander (Table 4.3). Approximately 26%, 27%, 45%, 47%, 48%, and 53% of the

variation in mean daily abundance of New Mexico spadefoots was explained by PP

(Fi,14=4.82, P=0.046), PED (Fi,,4=5.14, P=0.04), LR (Fi.,4=11.41, 7^=0.005), LSI

(Fi.14=12.49, P=0.003), ED (Fi,i4=13.04, P=0.003), and IJI (Fi,i4=15.6, F=0.001),

respectively (Figure 4.2). Approximately 30%, 33%, 35%, and 35% of the variation in

mean daily abundance of plains spadefoots was explained by ED (Fi,14=6.03, F=0.028),

LSI (Fi,14=6.78, F=0.021), IJI (Fi,i4=7.46, F=0.016), and PP (Fi,i4=7.55, F=0.016),

respectively (Figure 4.3).

Multiple linear regression models explained 38-78% of the variation in mean

daily abundance for these amphibian species (Table 4.4). Approximately 78% of the

variation in mean daily abundance of plains spadefoot was explained by PP, SEI, PSI, IJI,

and PNN (F5,io=l 1 -62, F<0.001). Approximately 73% of the variation in mean daily

abundance of New Mexico spadefoot was explained by LU, SDI, MNN, and PSI

(F4,i 1=11.02, F<0.001). Forty-one percent of the variation in mean daily abundance of

Great Plains toad was explained by MNN, PSI, NP, and PNN (F4,ii=3.59, F=0.042).

69
Table 4.3. Univariate Pearson coefficients of correlation between mean daily abundance
of amphibians and landscape^ metrics at 16 playa wetlands on the Southern High Plains.
Texas, 1999 and 2000.
Species'^
Variable'' NSF PSF BTS GPT
r P r P r P r P
PSI -0.138 0.612 -0.175 0.515 -0.102 0.708 -0.284 0.286
PS -0.011 0.970 0.274 0.304 -0.477 0.062 -0.138 0.611
PNN -0.199 0.459 -0.284 0.287 0.102 0.708 -0.166 0.538
MNN -0.405 0.119 -0.299 0.259 -0.322 0.224 -0.414 0.111
NP 0.432 0.095 0.388 0.138 -0.023 0.933 0.082 0.763
PP 0.506 0.046 0.592 0.016 -0.352 0.182 -0.073 0.787
IJI 0.726 0.002 0.589 0.016 -0.036 0.896 0.063 0.816
PED 0.518 0.039 0.457 0.075 -0.353 0.180 -0.186 0.492
ED 0.694 0.003 0.548 0.028 -0.234 0.383 0.048 0.858
LSI 0.687 0.003 0.571 0.021 -0.256 0.338 0.046 0.863
LR 0.670 0.005 0.249 0.351 -0.085 0.754 0.247 0.356
SEI 0.219 0.414 -0.473 0.065 0.104 0.701 -0.077 0.777
SDI 0.421 0.105 -0.297 0.263 0.055 0.838 0.032 0.905
landscapes (« = 16) were 2830-ha circular plots (i.e., 3-km radius) with their origins
positioned at the center of each study playa.
PSI = shape index of study playa, PS = study playa size, PNN = mean nearest-neighbor
distance from study playa to surrounding playas, MNN = mean nearest-neighbor distance
from all playas to each other, NP = number of playas, PP = percent aerial coverage of
playas, IJI = interspersion/juxtaposition index of playas, PED = mean number of edges to
cross from study playa to surrounding playas, ED = edge density (m edge/ha), LSI =
landscape shape index (i.e., measure of geometric complexity), LR = landuse richness,
SEI = Shannon evenness index of landuses, and SDI = Shannon diversity index of
landuses (McGarigal and Marks 1995).
"^NSF = New Mexico spadefoot {Spea multiplicata), PSF = plains spadefoot {S.
bombifrons), BTS = barred tiger salamander {Ambystoma tigrinum mavortium), and GPT
= Great Plains toad {Bufo cognatus).

70
71
Figure 4.2. Simple linear regression and 95% confidence bands of mean daily abundance
(natural-log transformed) of New Mexico spadefoot {Spea multiplicata) and percent
aerial coverage of playa wetlands (PP), interspersion/juxtaposition index (IJI, McGarigal
and Marks 1995:103), edge density (ED, McGarigal and Marks 1995:106), playa edge
density (PED [i.e., mean number of edges to cross from study playa to surrounding
playas), landscape shape index (LSI, McGarigal and Marks 1995:109), and landuse
richness (LR, McGarigal and Marks 1995:119) in landscapes (i.e., 2,830-ha circular plot)
associated with 16 study playas on the Southern High Plains, Texas, 1999 and 2000.
6 —r

0}
o a>
c o
ca c
^—^ OJ
• • . ^ — - ' • ' ' ' ^
•D
c ^^^-^"''''^ ' . C
13
jQ

ailyA
< 3 —
ro
2 —
c Q
to
c
nj
Y = 0 6905 + 0.7074X
R-square = 0.256 Y = 0 4447 + 0.0534X
R-square = 0.527
0 —
~~\ I
10 20 30 40 50 60
PP
IJI

6 —r

(V 5 —
o <u
c o
ro c
c T3
Q
< c
3 —
TO
<
2 —
TO 2 —
C
CD Q
i? 1
Y = -0.92 + 0 0726X c 1 — Y = 0.2842 + 0 3078X
TO R-square = 0.269
R-square = 0 482

"T" —r 0 —
T 1 r I
30 40 so 60 70 6 7 8 13
9 10 11 12 14

ED PED

6 —r

5 —
0) 5 —
O
c
CO
4 — •D
c c
< 3 —
<
CO
:o Q
c
(0 1 —
1 —
Y =-1.3868+ 0 3254X
Y = -0.84 + 0.4482X 0 — R-square = 0.449
R-square = 0.472

I I I r
9 10 11 13 1S

LSI LR

72
25 — /! S —

O •
C 2.0 —
O 20 — CO
C TD
(0 C
•D 3
C 15 —
13 1.5 —
n <
^^,-^^^
. . * • • • '

<
'co 1.0 —
= 10 Q
(0 •• .
c
Q CO
c CU 05 — • ,.-"
CO 05
0) •
Y = -0.033 + 0 3498X Y = 0.2139+ 0.0184X
00 — R-Square = 0 348
R-Square = 0 35
0.0 — 1 1 1 1
10 20 30 40 50 60 70

PP IJI

0) 2 — (U 2 —
O
o c
c
m CO
T3
"D C
c 3
3 XI
.Q <
< CO
Q Q
c c
CO CO
0)
Y =-0 22 + 0.0243X o
Y = -0.26 + 0 1577X
0 — R-Square = 0.301
R-Square = 0.326

"T" "T
30 40 50 60 70 10 11
ED LSI

Figure 4.3. Simple linear regression and 95% confidence bands of mean daily abundance
(natural-log transformed) of plains spadefoot {Spea bombifrons) and percent aerial
coverage of playa wetlands (PP), interspersion/juxtaposition index (IJI, McGarigal and
Marks 1995:103), edge density (ED, McGarigal and Marks 1995:106), and landscape
shape index (LSI, McGarigal and Marks 1995:109) in landscapes (i.e., 2,830-ha circular
plot) associated with 16 study playas on the Southern High Plains, Texas, 1999 and 2000.

73
Table 4.4. Multiple linear regression between mean daily abundance of amphibians and
landscape^ metrics at 16 playa wetlands on the Southern High Plains, Texas, 1999 and
2000.
Estimates*^ Collinlearity'
Species^ Variable"'^ Parameter Standardized / P VIF CN
NSF Intercept 9.869 0 2.35 0.039 0 0
LU -2.077 -0.703 -4.91 <0.001 1.13 1.0
SDI 1.203 0.319 2.24 0.046 1.12 1.14
MNN -0.002 -0.278 -1.98 0.073 1.08 1.39
PSI -5.634 -0.244 -1.67 0.124 1.19 1.58
PSF Intercept 9.199 0 4.53 0.001 0 0
PP 0.190 0.322 2.23 0.049 1.418 1.0
SEI -2.299 -0.458 -3.56 0.005 1.125 1.453
PSI -6.469 -0.665 -4.23 0.002 1.642 1.546
IJI 0.012 0.379 2.48 0.033 1.598 2.175
PNN -0.0005 -0.372 2.26 0.047 1.843 2.587
BTS Intercept 2.619 0 11.51 <0.001 0 0
YR -0.939 -0.615 -2.92 0.011 1.0 1.0
GPT Intercept 17.36 0 3.72 0.003 0 0
MNN -0.003 -0.759 -2.74 0.019 1.957 1.0
PSI 9.220 -0.756 -2.87 0.015 1.754 1.231
NP -0.157 -0.599 -2.14 0.056 1.990 2.285
PNN -0.001 -0.516 -1.92 0.082 1.838 2.731
^Landscapes {n= 16) were 2830-ha circular plots (i.e., 3-km radius) with their origins
positioned at the center of each study playa.
hSlSF = New Mexico spadefoot {Spea multiplicata), PSF = plains spadefoot {S.
bombifrons), BTS = barred tiger salamander {Ambystoma tigrinum mavortium), and GPT
= Great Plains toad {Bufo cognatus).
"^Landscape metrics retained as per stepwise selection using default entry and stay
significance level at a=0.15; all overall F-tests on final models were significant
{P<Q.042) and coefficients of determination adjusted for number of variables in the
model (i.e., /?%) = 0.728, 0.779, 0.378,and 0.409 for NSF, PSF, BTS, and GPT,
respectively.
PSI = shape index of study playa, PNN = mean nearest-neighbor distance from study
playa to surrounding playas, MNN = mean nearest-neighbor distance from all playas to
each other, NP = number of playas, PP = percent aerial coverage of playas, IJI =
interspersion and juxtaposition index of playas, SEI = Shannon evenness index of
landuses, and SDI = Sharmon diversity index of landuses (McGarigal and Marks 1995);
LU (landuse) and YR (year) were indicator variables using 0 = cropland watershed and 1
= grassland watershed, and 0 = 1999 (wet year) and 1 = 2000 (dry year), respectively.

74
Table 4.4. Continued.

^Relative natural-log (LN) abundance per species can be estimated via multiplying field
estimates of landscape metrics or indicator variables by parameter estimates, and
summing them and the intercept within species; predictions can be back-transformed to
original units by exponentiation; magnitude and sign of species-specific standardized
estimates can be interpreted as the strength and relation of the variable with LN
abundance in the presence of the other retained explanatory variables (Myers
1990:384-385).
CoUinearity diagnostics (intercept adjusted) were variance inflation factors (VIF) and
condition numbers (CN); VIF > 10 and CN > 30 are suggestive of a linear dependency
between >1 variable (Freund and Littell 2000:98-101).

75
Finally, 38% of the variation in mean daily abundance of barred tiger salamander was

explained by YR (Fi,14=8.52, P=0.011). No linear dependencies existed among variables

(Table 4.4).

Discussion

Canonical correspondence analysis (CCA) revealed that landscape structure

influenced the composition of the amphibian species assemblage at playa wetlands.

Subsequent univariate analyses indicated New Mexico and plains spadefoots generally

were positively associated with spatial metrics representing optimal spatial positioning of

playas (i.e., interspersion/juxtaposition index, percent aerial coverage of playas) and

landscape complexity (i.e., landscape shape index, edge density). Great Plains toad and

barred tiger salamander usually were negatively associated with spadefoots (as per CCA),

but corresponding univariate analyses indicated they were not associated with spatial

metrics. Thus, these results suggest Great Plains toad and barred tiger salamander may

be less affected by spatial positioning of playas and complexity of the landscape matrix

but perhaps are poorer competitors than spadefoots because of their negative association

with these species.

Spadefoots probably are associated with landscape structure more than the other

species because of their relatively small body size (Chapter II). Body size has been

shown to connote vagility (Peters 1983:89-91, With and Crist 1995), seemingly due to

differential perception to patch viscosity and edge permeability (Crist et al. 1992, With

1994, Wiens et al. 1997, Mclntyre 2000). Although relative vagility has not been

compared among these amphibians, dispersing spadefoots may have been unable to

76
penetrate geometrically complex landscapes, resulting in increased abundance in natal

wetlands. Indeed, several recent studies have documented positive associations of

amphibians at wetlands embedded in complex agricultural landscapes (Knutson et al.

1999, Kolozsvary and Swihart 1999, Chapter III). Similarly, abundance of spadefoots

may have been positively related with less isolated and more optimally juxtaposed playas

because of their lower relative vagility. Probability of successful interdemic dispersal

can increase with decreasing inter-patch distance and increasing patch juxtaposition

(Bascompte and Sole 1996, Hess 1996). Consequently, abundance and persistence of

less vagile species can be positively related to habitat patch positioning thus the exchange

of individuals among spatially structured local populations (Ritchie 1997, Stacey et al.

1997). Therefore, these results support my hypotheses and previous applied and

theoretical research on the relationship of abundance, spatial positioning of populations,

and landscape complexity.

Great Plains toad and barred tiger salamander may have been negatively

associated with spadefoots because of differential competitive ability of their larvae

(Wilbur 1984). Spadefoot larvae have been shown to be competitively dominant over

various genera (e.g., Bufo, Gastrophryne, Hyla, Rana, Morin 1983, Wilbur 1987, Dayton

and Fitzgerald 2001). Increased recruitment of spadefoots also could increase

postmetamorphic interactions of individuals among our species (Wilbur 1980), which

may compete for food resources as suggested by diet composition studies (Anderson et

al. 1999^, L. M. Smith and M. J. Gray, unpublished data). Thus, Great Plains toad and

barred tiger salamander may have been displaced or experienced reduced survival in

77
terrestrial landscapes with greater spadefoot abundance, because of potential diet overlap

and relative competitive ability. Also, if microhabitat characteristics (e.g., vegetation

composition and structure) were correlated with landscape structure, differential

microhabitat use might explain species separation. Spadefoots seem to prefer playas with

less vegetation for breeding unlike Great Plains toad and barred tiger salamander

(Anderson et al. 1999).

Conservation Implications

Although regulating mechanisms are unknown, these results suggest that

landscape structure influences species composition and abundance of amphibians.

Landscape structure may favor certain species as per species-specific vagility and their

relative perception to patch viscosity and edge permeability (Wiens 1997). Competitive

ability also may interact individually or synergistically with landscape structure and

species-specific vagility to affect community composition and abundance of amphibians

in wetlands. Therefore, landscape ecologists should consider species-specific mobility

differences when developing conservation initiatives for spatially structured amphibian

populations (DriscoU 1997, Szacki 1999). Existing software packages (e.g.,

FRAGSTATS* ARC®, RAMAS® GIS) have options for inclusion of relative patch

viscosity, boundary permeability, and species vagility data; however, generally unity or

educated estimates are used because the aforementioned data do not exist. Future

research directives need to focus on estimation of species-specific vagility in

anthropogenic and natural cover types and the relative permeability of their boundaries

for realistic estimation of extrinsic connectivity among habitat patches (e.g., With and

78
Crist 1995, Wiens et al. 1997). Indeed, various theoretical studies (e.g., O'Neill et al.

1988, With and Crist 1995) that used percolation models have shown this information is

critical in estimating viability and persistence spatially structured local populations.

Finally, my prediction models can be used to estimate relative species-specific abundance

of amphibians on the SHP using field estimates of landscape metrics (e.g., Knutson et al.

1999, Scribner et al. 2001). Landscape ecologists may use this and ground-truthing

information to locate potential locations for conservation endeavors.

79
CHAPTER V

TEMPORAL NICHE PARTITIONING IN SOUTHERN

HIGH PLAINS AMPHIBIAN POPULATIONS

Abstract

Temporal niche partitioning reduces interactions among species by decreasing

probability of interference and exploitative competition. Although temporal segregation

of resource use has been recorded among amphibian species, it has not been examined for

amphibians in semi-arid and prairie environments. Moreover, anthropogenic landuse

may affect temporal resource use of amphibians. My objective was to test for temporal

niche partitioning in Southern High Plains amphibians, and examine if landuse

(cultivation vs. grassland) affected temporal resource use among 4 common species {Spea

multiplicata, S. bombifrons, Bufo cognatus, and Ambystoma tigrinum mavortium).

Standardized species-specific daily abundance from pitfall traps at 16 playa wetlands

during 2 years (1999 and 2000) was compared among 3 within-year time periods and

between 2 landuses. Inspection of time series plots and hypothesis testing indicated that

temporal niche partitioning existed in Southern High Plains amphibian populations.

Differences in temporal activity may be a consequence of interspecific competition

associated with dietary or microhabitat niche overlap. No differences were detected in

temporal niche partitioning between landuses, suggesting that phenotypic plasticity in

temporal resource use is unlikely for these amphibian species. Thus, amphibians that use

cropland wetlands later in a year may have greater probability of extinction in grassland

wetlands, because hydroperiods are shorter in wetlands surrounded by disturbed

80
landscapes. It may be beneficial to preserve grassland playas and restore hydroperiods in

cropland landscapes to ensure habitat for later acti\'e amphibian species.

Introduction

Resource partitioning in ecological communities facilitates species coexistence

and allows biological diversity (Schoener 1974). The way in which species partition

resources defines their ecological niche (Pianka 1976). Hutchinson (1957) suggested that

an ecological niche could be conceptualized as an ^-dimensional hypervolume, where

each axis represented abiotic and biotic resources necessary for survival and

reproduction. Niche dimensions also may include axes for time and space (Cody 1968.

Schoener 1974. Rosenzweig 1995), resulting in a pulsating hypervolume (Pianka 1994).

The competitive exclusion principle {sensu Gause 1934, Hardin 1960) states that species

coexistence depends on the dissimilarity of niche hypervolumes (Pianka 1994:249, 272-

273).

Niche hypervolumes must be dissimilar along at least one axis to facilitate species

coexistence (Schoener 1974). The most basic hyper\olume includes axes for a resource,

space, and time (Pianka 1974). Temporal segregation is necessary for coexistence if

species use a common limiting resource, such as food, in the same location (Schoener

1974). Use of shared resources at different times is called temporal niche partitioning

(Huey and Pianka 1983), and a fundamental construct to niche complementarit}' and

biodiversity (Loreau 2000. Loreau and Hector 2001). Presumably, temporal niche

partitioning facilitates species coexistence by reducing interference and exploitative

competition (MacArthur and Levins 1967, Case and Gilpin 1974). Indeed, it has been

81
suggested that temporal niche partitiomng in^lies competition along space or resource

(e.g., food) axes (Huey and Pianka 1983, Schoener 1983).

Temporal niche partitioning is most common in poikilotherms among \'ertebrates

(Schoener 1974). Several studies have documented temporal differences in activit\'

patterns among amphibian species (Blair 1%1, Murphy 1963, Creusure and UTiitford

1976, Eraser 1976flA Bowker and Bov^^er 1979, Toft and Duelhnan 1979. W iest 1982).

Following Pianka (1976) and others (e.g., Schoener 1974), it has been hypothesized that

temporal niche partitioning exists in amphibian communities to reduce competition for

breeding sites or food resources (Wilbur 1980, Toft 1985). For amphibians that

e>q)losively breed in wetlands, temporal niche partitiomng probably reduces competition

most for food resources, because few species establish breeding territories (Duellman and

Trueb 1994). Indeed, dietary overly has been documented in adults and juveniles of

several coexisting explosive breeding species (Whitaker et al. 1977, Dimmitt and Ruibal

1980a, Anderson et al. 1999ft, Newman 1999).

Southern High Plains amphibians exist primarily in spatially structured playa

wetlands and tbey are explosive breeders (Bolen et al. 1989, CHiapters II and HI).

Because diets of Southern High Plains amphibians overlap (Anderson et al. 1999ft, L. M.

Smith and M. J. Oray, itnpublished data), species may exhibit temporal niche partitioning

to reduce exploitative competition. Several Southem High Plains species also use similar

habitat (e.g., vegetation structure, percent water coverage) for breeding (Anderson et al.

1999a); thus, species may temporally segregate to reduce interference competition (albeit

explosive breeders). Anthropogenic landuse also affects demographics and d>Tiamics of

82
Southem High Plains amphibian populations (Chapters III and VI). Therefore, my

objective was to test for the existence of temporal niche partitioning in Southem High

Plains amphibian populations, and examine the effect of landuse (cultivation vs.

grassland) on species-specific temporal activity pattems. I hypothesized that Southern

High Plains amphibians would temporally partition resources, and landuse would affect

temporal activity pattems.

Methods

This study was conducted at 16 playa wetlands (8/year) on the Southern High

Plains of Texas during 1999 and 2000 (Bolen et al. 1989). Playas were considered to be

in grassland if on average >75% of the surrounding landscape (2830-ha circular plot) was

undisturbed and vegetated, and cropland if >75% of the landscape was cultivated

(Chapter III). Playas were partially enclosed (i.e., 25% of the circumference) with 60-cm

high drift fence and 19-L pitfall traps (Dodd and Scott 1994), and pitfalls checked

altemate days from 16 May-17 October 1999 and 19 April-18 August 2000. Captured

amphibians were enumerated by species, marked uniquely, and released. After releasing

individuals, pitfalls were closed for 24 hr to reduce probability of immediate recapture

and possible biasing of species-specific temporal abundance (Turchin 1998). Temporal

niche partitioning was examined for the 4 most common species of amphibians on the

Southem High Plains (New Mexico spadefoot [NSF, Spea multiplicata], plains spadefoot

[PSF, S. bombifrons]. Great Plains toad [GPT, Bufo cognatus], and barred tiger

salamander [BTS, Ambystoma tigrinum mavortium]. Chapter III). A more detailed

83
account of the study area, amphibians, and sampling protocol was described in Chapters

II and III.

I used total daily capture per species (i.e., summed across playas within landuses

and years) as an estimate of daily species-specific abundance (Dodd and Scott 1994).

Summing across playas (i.e., local populations) was a reasonable estimate of relative

population size within landuses each year, assuming playas were mutually independent

(Hanski 1997). Species-specific abundance was standardized (hereafter percent

abundance) by dividing species-specific daily capture {N\t) by species-specific total

capture (i.e., A^i, sum of captures over t days) within landuses each year (Hurlbert 1978).

Thus, species-specific percent abundance represented the proportion of the population of

the /• species present on the t day. Standardization was necessary to remove abundance

effects and compare niche overlap of species equally among time periods and between

landuses (Hurlbert 1978, Paton and Crouch 2002). Time periods (i.e., 1, 2, and 3)

represented 3 equal sequential time series in each year (i.e., 22 and 17 days per period in

1999 and 2000). I chose to divide each yearly time series into 3 periods, because

previous studies in other regions of North America suggested amphibians temporally

partition resources into ca. 2-3 within-year time periods (e.g.. Murphy 1963, Wemer and

McCune 1979, Wilbur 1980). A different temporal division (e.g., 2 or 4) of the within-

year time series may have revealed different results.

Temporal niche partitioning was determined by inspecting time series plots and

comparing frequency distributions of percent abundance for pairs of species among the 3

time periods within years. Frequency distributions also were compared between cropland

84
and grassland landscapes. Frequency distributions were constmcted by tallying number

of days within time periods, landuses and years that percent abundance of pairs of species

exceeded each other. Thus, each frequency distribution represented the pairwise species

comparison of standardized temporal abundance. I considered differences between

species as a reasonable measure of temporal niche partitioning, because standardized

abundance of species pairs was compared to the null distribution of equal relative

frequency among time periods and between years (see below). This approach to

quantifying niche partitioning is similar to Hurlbert (1978).

I tested for differences in frequency distributions (a = 0.05) among time periods

and between landuses by each year using logistic regression with a generalized logit

model (Agresti 1990:307, Stokes et al. 2000:257- 259). Nonaddivity of main effects

(i.e., time and landuse) was tested by including an interaction term in the model (Stokes

et al. 2000:259). Normalized chi-square tests of homogeneity (i.e., Z-tests) Bonferroni

corrected were used to test pairwise differences in frequency distributions when the main

effect logit model was significant (Zar 1984:395-396, Milton and Amold

1995:667-670). If main effects were not additive, analyses were separated by periods

and landuses, and pairwise Z-tests performed. I asstimed the categorical response of t

days followed a multinomial distribution per effect and frequency distributions were

independent (Agresti 1990:306).

Results

Time series plots of pairwise species-specific percent abundance suggested that

amphibian species temporally segregated resources on the Southem High Plains (Figure

85
5.1). The main effect logistic regression indicated that frequency distributions were

different among time periods both years between GPT and NSF (x 2=14.9, P<0.001

[1999]; 5cV9.6, P=0.008 [2000]), PSF and NSF (xV29.9, P<0.001 [1999]; xVl8.4,

P<0.001 [2000]), BTS and NSF (xV40.4,P<0.001 [1999]; x V 13.3,/'<0.001 [2000]),

BTS and PSF (xV31.2, P<0.001 [1999]; xV8.6, P=0.014 [2000]), and GPT and BTS

(xV33.5, P<0.001 [1999]; x V 9 . 1 , P=0.0\ 1 [2000]), and GPT and PSF in 1999

( x V l 6 . 3 , P<0.001 [1999]; x^2=18.4, 7^=0.099 [2000]). Temporal niche partitioning did

not differ between landuses for all species pairs both years (x^i=0.001-1.75,

P=0.19-0.98).

Nonadditivity was violated for GPT and NSF in 1999 (xVl4.1, P<0.001), BTS

and NSF in 2000 ( x V l 1-9, P<0.001), GPT and PSF in 1999 (xV9.4, P=0.009), BTS

and PSF in 2000 (xVl4.4, P<0.001), and GPT and BTS in 2000 ( x V l 1.2, P=0.004);

therefore, these analyses were separated by landuse. Frequency distributions were

different among time periods between GPT and NSF in grassland in 1999 (x^2=15.9,

P<0.001), between GPT and PSF in cropland in 1999 (xVl6.8, P<0.001), and between

BTS and NSF (x^2=15.4, P<0.00\), BTS and PSF (xVl6.8, P<0.001), and GPT and BTS

(X^2=12.9, P=0.002) in cropland in 2000. Differences were not detected in frequency

distributions among time periods for GPT and NSF in cropland in 1999 (x 2=4.3,

P=0.114), GPT and PSF in grassland in 1999 (xV4.6, P=0.099), and for BTS and NSF

(X^2=1.5, P=0.482), BTS and PSF (xVo.65, P=0.723), and GPT and BTS (xVo.76,

P=0.684) in grassland in 2000.

86
87
Figure 5.1. Time series plots for pairs of amphibian species residing at 16 playa wetlands
in 2 landuses and years on the Southem High Plains of Texas. Linear trajectories
represent total daily capture per species summed across playas (i.e., 4 per landuse per
year). Daily capture per species was standardized using species-specific capture for the
entire time series (Hurlbert 1978), thus each node on the trajectory represents the
proportion of the population present during that day (i.e., percent abundance). The solid
line is the first species listed for the pair in the upper right comer of the graph. NSF =
New Mexico spadefoot {Spea multiplicata), PSF = plains spadefoot {S. bombifrons), BTS
= barred tiger salamander {Ambystoma tigrinum mavortium), and GPT = Great Plains
toad {Bufo cognatus).
Cropland, 1999 GPT vs. NSF Grassland, 199£
GPT vs. NSF

c
T3
T3
c J3

€ 2 - <c 2 H
o
(1) CL
a.

10 20 30 40 50 60 30 40
Days Days

Cropland, 2000 GPT vs. NSF Grassland, 2000


GPT vs. NSF
4 -
c
(D a
•o o
C c
3 - ro
<
T3 3 -
c
8 2 <
0)
0.
\mf V /I c
a>
Q-
2

1 \U \ V \W

10 20 30 Days
Days

Grassland, 1999 PSF vs. NSF

0)
o
c
ra
T3 3 -
c
3
<c 2 -

a.

0 -

30n
Days 40
Days

Cropland, 2000 PSF vs. NSF Grassland, 2000 PSF vs. NSF

4 -
8 8
c c
n
•o (D
c 3 - •D
C
3
< <
8
« 0)
Q. a.

••»<o«»

Days Days

88
Cropland, 1999 Grassland, 1999p^ B T S v s . NSF
B T S v s NSF

8 '-^ 3 -
c c
•a
i 3 ^/l ra
T3
c
€ 3
§ 2 <
0)

CL 1 ^

10 20 30 40 50 60 30 r, 40
Days Days

/ - I
Cropland, 2000 Grassland, 2000
BTS vs. NSF BTS vs. NSF

(U
o
c
ra
•D
c
3
<
Q. 2 -

1
\ \! 1/\A ^
i i y 0
0 -
" 1 ~ ~r~ I

10 40 50 Days
20 Days 3 °

Cropland, 1999 Grassland, 1999


GPT vs. PSF GPT vs PSF

c ra
3 T3
3 -

8 c
2 -
0. 8
<u
Q.
1 -

0 -

30 40 30_ 40
Days Days

j, Grassland, 2000
G P T vs. PSF

ra
•o ro
c
3 = 4 -
<

QL

0 -

20 Days 30 20 Days ^0

Figure 5.1. Continued.

89
5 - 5 -'
Cropland, 1999 i
BTSvs PSF

o c
c ra
ra T3
•o c
c
3
3 - ! \ 3
.a 3 -
<
i> •
o
o 2 -
i_
<D

a. I ' r, • . . .
Q.

1 - I
:;' 4 ' J : ' •
0 - n ;i 0 -I

10 20 30 40 50 60 30^ 40
Days Days

Grassland, 2000
BTSvs PSF

u
c
ra ra
•D
•o c II f
c 3
3 1.1 1
<
AV
0)
Q.
a.

V m4:
•J I' \ii J ? vyv
I i 1 i u
10 20 30 40 50 10 20 Days 30 40 50
Days

Cropland, 1999 4 —'


GPTvs BTS

8
c
3 -
ra
•D
C
3
A ra
•o
c
3
.a
< < 2 -

01
^U/1 0)
0.
1 -
a. 1 -

0 _ ( u '— 0 -

10 20 30 40 50 60 10 20 30 40 50 60
Days Days

s —r C r o p l a n d , 2000 Grassland, 2000


G P T vs. BTS G P T vs. BTS

8c
c ro
ro T3
•D 3 - c
c
3 3
n <
<
c 2 H
Q.
a. 1 -

Days Days

Figure 5.1. Continued.

90
In general, pairwise Z-tests indicated that a larger proportion of the GPT

population was present during time 2 than the NSF populafion (Figure 5.2). Also, a

larger proportion of the PSF population was present during time 2 than NSF and GPT

populations. Barred tiger salamander was most active during time 3 compared to NSF,

PSF, and GPT in 1999. However, a larger proportion of the BTS population was present

in time 1 than NSF, PSF, and GPT populations in cropland in 2000 (Figure 5.2).

Discussion

This represents the first documentation of temporal niche partitioning in semi-arid

Great Plains amphibian communities. Even though amphibians on the Southem High

Plains generally emerge and breed simultaneously and clutch times are similar (see

Chapter II life history review. Rose and Armentrout 1974, 1976, Degenhardt 1996:38, 40.

51, Sullivan and Femandez 1999), plains spadefoot was more active during the middle of

both years compared to New Mexico spadefoot and Great Plains toad. Creusere and

Whitford (1976) also observed this temporal relationship with plains spadefoot and Great

Plains toad and a different spadefoot {Scaphiopus couchii) in New Mexico. Plains

spadefoot may be more active after New Mexico spadefoot and Great Plains toad because

of exploitative competition for food resources (Schoener 1974, 1983). Anderson et al.

(1999^) documented some diet overlap among adults of these 3 species. Moreover,

plains and New Mexico spadefoots use similar habitats for breeding and can breed

simultaneously (Bragg 1965, Anderson et al. 1999a, M. J. Gray, unpublished call survey

data); thus, temporal niche partitioning between spadefoots may be a consequence of

interference competition for breeding sites (Wilbur 1980, Toft 1985). Similarly, Great

91
92
-O 1=J

u
"T^ •*—'

C3'

J3
; ^ 0^ Q PH

W3

(/3

O)

C/3
Cd

(/3

<N

<U

UH TS PC -.cj
CO CO
© m o © ©

19%
LL E o u. E

DGP T DNSF
(0 ©^
o -;- E ©^

DNS
•7 00 VO
V
•1 - vo
o>
o
o
o
I—
Q. Q. CM
O
%3

•a T3

81%
D
c
j2 c o c
J2 D JS
Q. (A Q.
ro
OS
^r\
-^
O
O (A
(Tl
1

n O
(N O o w
II II ^ ^
N Q.,
^ ^ o
JC
i ci i i i c C3 A i i i i c
CO CD

aouajjnooo iuaojad aouajjnooo luaojad aouajjnooo tuaojad

CM CM
© Tt o CM
©
IL

E E o ©

24%
DGP T DNSF
©*>
VO
o E
D f^ at at V
at at o
' i - j ^ at , • . >

o
1— at
Q. o
•a CM
o •o cta

76%
©^ c •a
D c
oa
(/)
vo (0 J2
oo vo ra Q.
oo O o o
— o
II II
N C

i A i i A c
C3
CI3 oo t-O -"^ C--.J

aouajjnooo fuaojad aouajjnooo luaojad


aouajjnooo fuaojad

©
"'^4M&'^ ©
E E
©
E
u. >^^ffiss|^
'^'.^^^^ H H
at o
999,

at o
"•'^W' at o
CM
-o
0. '" <u
•a - ::s
O •D IL
c
Ti
a
D ©^ C C
in re w SS re •*->

z IT} (/) a c
iJrop

-^ r^ .—i
n Tj-
o
(A
re
o o
H o O CJ
o o CL vl
O O
II V CN
N (^ U N cx. ir^
(D
S—S—S—3—S" S—S S—S 5" ;-i

C3 oo to •^r Osi ;3
ou
eoudJjnooo l u a o j a d aouajjnooo fuaojad eouajjnooo fuaojad UH

93
CO CO CO

© © ©

82%
E u. E

DPSF DNSF
E
NSF

©^ '©^

DNS
>*^
D *n o VO O)
o at CT>
o at IL at
<N
Q. CO
O •a 0. T3

18%
c C C
D ©^
SS
D j2
SS in
in Q.
Tf in o o fO in
r- -^ re vo rt ro
m r-~ o rt — k-

o o o — O O
II II V II II
N Q. CI, N a.
JC
i c> o A i c t A X A i c;

aouajjnooo fuaojad aouajjnooQ fuaojad aouQjtjnooQ ^uaojad

DPSF DNSF
CM

/'<0.001
CM
Z=5.47

Z=8.94
CM
© o ©
18%
E o ©
DGPT DNSF

DPSF DNSF

10%
©^
OS
o v.. E
V
<SI o 0. at
o at at
o at at
CM at
TJ
82%
71%

T3
•a

90%
C C
J5 SS c
(A a. SS
oo (A o in
in
so o ro ro
o k.
^ o O

A ci <±. A A c c-i C3 O O O C3
CD OO CO ^d- c%j
A
C3
A
oo
A
CO
A A c^

aouajjnooo ^uaojad aouajjnooo tuaojad


aouajjnooo ^uaojad

©
^SjB ©
© E
b
DNSF

E i
*
1- t-
at
'.\o -
99,

o at
o at
o IL * " M *^'fcf
o> X!
CM CO »- . <D
•a Q. •o •D 3
Gra sslan
ntin

c
c
SS
D ©^
VO
re
in fo Q.
in oo VD o o
ra
^
oo o O U
-^ o
II II
N Q. oi
u^
A AOO A A A c3 <D
C3 (-0 • « - c-%1 Ui
:3
tJj
eoudJjnooo ^uaojed eouajjnooo tuaojad eouajjnooo )uaojad |J^

94
LL
CO
CO r^ .
CO
©
^ TT O CO

© D ^' o ©
E E E

DNSF
• ^ « r - < •:•--•
©^ o

BTS
1^
Jl ^
V
t^ o N c ITi
o o D a>
o o at
o LL CM
at
CM CO
Q. •D
TJ
D ^ C ^ •a
C SS
SS ON in IT/ c
a. SS
o ^ o n in
re
ON
a.
o
^ o ^
O

JL JL JL JL JL „ JC JL JL 1 1 1
3C CO DC ;0

aouajjnooo ^uaojad aouajjnooo luaojad aouajjnooo luaojad

LL
LL
CO , , CO
z CM oc CM
z

P<0. 001
m o
n •r © CM
Z=2.6

©
o E D
29%
E OJ
DPSF DNSF

LL ©^

, Tim
6=Z
' • . ' ;

BTS
CO
Q.
<s II
o
D t o
o
o
CM
o
o
CM
D
ON
at
at
at
•D '"
71%

•D*

nd,
^ C C
00 SS SS ^
in
00 a. in re
o re ON

>op
k.
^
o

j r — -T. X •X •Tr—
= oc CO c:; oc CO A ^
c; OC

aouajjnooQ fuaojad aouajjnooo ^uaojad


aouajjnooQ luaojad

r^
00 o ©
o ©
©
T E ^PP^^^^P-
\XX\i

LL II
E ©^
DPSF BNSF

DNS

N c. :t^ 'r ^B&i Vi


fO o 0\ at
o o
o o lEBT-'-^ .. , . , ' . „ „ _ . - „ • • , at ,
o II CM at Ti
CM CO O
0. •6 u. •a 3
67%

t
53%

Tj" CO
C D c c C
SS
SS z ra • ^^
in yr,
a.
o
in
re
• r~ — a
o
r
o
k.
CO rr o
ok- U
O O (D m — o
il i' D il y (N
N 0.- in
c; CD c; c:: c; A S" -TT •nr o
A A A A A c: OC cC •«:- c-^ L<
c: oc 3
CO
eouajjnooo luaojad eouajjnooo ^uaojad eouajjnooo ^uaojad U,

95
"
u. LL
CO r- ^— CO
z ^ o CO CO z CO
D • ^ o P © © D ©
CO
«—1

dV E ©^ E CO ©^
E
1- ^ ex., ©^ h- l>
CQ V) 1 ON CQ o
i o V) o
U at o
at o u o
CM
CM
•o LL
•6
c c ^ c
©^ CO ©^ JS j5
J5
Vi
OS (A zD ci ^—
a (A
(A
(A ON o o
o o o ro
ra CO
H
d II II k.
k. ^—^ oV
m II N ex. O
O D N ex.

A A A A c c5 S" C3 "S—3" A A A o c^ c=
CO
o oo

aouajjnooo luaojad aouajjnooo tuaojad aouajjnooo ;ua3jad

CM IL

TS DNS
CM
© CM
©
E E ©
E
©^
at o Vi
at o CO f*^ o
at o o
CM D o
CM
•o T3"
c
_re C
•6
(A J5 ^ c
in a. V5
re o OS r- VO iS
(A
k.
O
r- p (A
ro
II II
N 0.,

A A A A A c-

aouajjnooo fuaojad eouajjnooo )ua3jad


aouajjnooo )ua3Jdd

DBTS DNSF

/'<0.001
Z= 10.93

©
© ©
LUJi

©^
Vi
E ^ 1 E
OS vo f o o
at o
o o TJ
at
CM
o 3
CM
•D
•6 c
94%

TJ
u.
CO ©^
c
j2
c
j2 _rre . ^^
z V) (A a in
o
D r- ^- (A o re
(A
O
• ^
o re O
CO o
H ^t
CQ <—> <N
II V U^
U N a. <u
LH

•s—s—s—5" A
C3
A
oo
A
CO
A
-^
A
(Tsl
c= ? A—S—S—3—S—5 3
OX)
PH
eouajjnooQ luaojad eouajjnooo luaojad eouajjnooo ^uaojad

96
DGPT DPSF

DGPT DPSF
CO
CO
©

36%l
CO
©

29%
E
DGPT DPSF

E ©

33%
E
at
at
at at o
at o
o
•a

71%
•a c
©^ c •6

67%
SS
SS in c
so -^ vo a. in SS
o •^ o re r3- to Q.
OS o
O
II
o <^ d o
II iT II II
N C
-c i 1 • JC ! 1 1 1

aouajjnoDQ luaojad aouajjnooo luaojad


a o u a j j n o o o ^uaojad

CM LL
CM

PT DPS
© © CM

©^ E LL ©^
E ©

"^ Q. Tf E
1-H at
at
at D . .vo at
at 1-
at O r- o
o
Q.
D o
CM
u. •a •a
CO ©^ c D c •6
Q. — SS SS ^ c
c vo
00 a. oo so in ON SS
D O o o oo o r^i in oo
re fS a.
t^ d o
h-
Q. V
o T II o
tii. N Ci, k.
o
D ^J o
JU JL JL JC,
oc CO oc iO -^ j _ J- JL X
cr; DC ;-C •^^ rx

83uajjn33o juaojad aouajjnooo )uaojad


aouajjnooo tuaojad

u.
CO r~ o
Q. • ^
o ..-..
ime '

D u-i
o ©
II
-^? ©
V
^
E CO ©^ • - E
N H
Q.
00 1- 0. ^ •.
nd. 000,

O ^ 4

at
at D '^
D
^ •

at at
' >.- at t-
at
^ •a
Q.
O
CM o
nd.

^
c
SS D ^
_re
c
<N ra in vo
00 Q. in
re — oo
V) Q.c
O o
o O d d o
Jl
N
II
ci.
o u
Csj
A" - x — X - TTT iri
S S S S S ~ C3 oc
=5

aouajjnooo luaojad aouajjnooo )uaojad aouajjnooo tuaojad

97
il. LL
CO (O (^ ^^

001
CO Q. ro Q. • ^
o
D O ^ o CO

u.
©
E o
CO
© • ^— dV ©
CO CO II ^ E CO II ^k
E
Q. 1- N V H N ex. is
D o m ex. -^ m Vi
o D ^* at D at
1- o at at
Q. CM at at
o •6
D c •o X3
SS ^ c i< c
in SS SS
• ^ i/~, in vo a. V, in
OS ra 00 in
o o ON ro
— o ^
u
N c..

jrr
JC I- JL- JC 1 1 1 1
=— oc ^
re ^^

aouajjnooo iuaojad
aouajjnooo luaojad aouajjnooo tuaojad

CM
CM
© CM
LL ©
E LL ^ E ©
CO ©^ ^ LL E
Q. O O CO ©^
Q.
D Vi o Vi Q. 00
o D at
at D SO at
o
GPT

at
CM CO CO at
H H
•o"
D c
m •D
m T3

%z
C D
SS
in
D o SS C
Vi a. SS
in V/ in
c5 O ra o o o OS — f^- in
1.0

1.0

c ^ o o lO o ra
c o
1 II II II II II
^s! C , N ex, N ex,
-c— 1 1 JC X 1- -C X "— X — -J- JL.
JC nr x X X 1-
re -o
c_; oc CO

aouajjnooo )uaojad aouajjnooo ;uaojad


aouajjnooo )uaojad

© "•••-NP'-;' © ©

E . © ^ . - ••
E ©^
E
H
o os: at OS o>
o at at
o at at -TD
T-
CM <U
•a •o T3 P
DBTS DPSF

C
6%

C
c c u.
CO ^ ro • ^ M
SS ra Q. «A
a c
Z= 10.93

in
(A
in o D ra o
S2 ok. ^ § O
r

CO
o 1-
ffl csi
D N ^
»n
A A A A A
jr- -X- -X- -X- <u
LH
oc
D
00

aouajjnooo tuaojad aouajjnooo tuaojad aouajjnooo )uaojad b

98
miiimimKiiiiit>mtRi.iifJiim!x:f, CO
fO
© ©

DPSF
©
E E E
i- ©^ ©^
o"
Vi
o '^ o
o 0\ at
o CO o at
CM 1- • • " •
at
CM
ffl
DBTS DPSF

•D T3
C D CO T3
©^ ©^ C
SS SS »- C
OS Q. 00 in m r- •—' SS
V)
£ i *—I
o
o
ro
00

"^
Vi (A
ro
D
H
• ^

^
^—
.
o
o
dV
a.
o
o
"^ d dII IId
Q.
O N
II
Q,
N ex. D
JL JL X JL
-J— -X-
A A A A A c= A- • 3 — ^

aouajjnooo juaojad
aouajjnooo )uaojad aouajjnooo )uaojad

CM CM
f ,^ - "i^ y A + A ^ '^•*t ''*• '* ' © CM
LL
©
CO E E ©
©^ V
Q. E
r«< o
D 00 * o
o o at
CO ., o o at
1- CM CM at
m
D 13 •o"
C c T3
00 SS J2 C
a. (A SS
, ^H «A
o a.
ro o
°) o %..
CD
o

A A A A A cr

aouajjnooo fuaojad aouajjnooo fuaojad


aouajjnooo )uaojad

CO
IL i— t^ —
0.00

CO m
14.4

Q. (N -
ON « D
D ©
©
LUJi

CO
^ E 1-
i' y
II E 0. ^
1- t^ij
m O N ex. V)
o D at
D o o at
o
o
o at -a
CM
T- <L»
CM .
TJ" •o n
•D" ^ c
c
c c ra
SS SS Vi ••—>

a c
a. in
in
OS
o o
o ra
k.
k.
O L)
o O
(N
ir^

s—s—s—S—S" <u
;-!
CTJ oo CO •^r c\ 3
DD
aouajjnooo )uaojad aouajjnooo )uaojad tin
aouajjnooo )uaojad

99
oo O
CO so CO
© o CO
ro d ©
E ©

DBTS
©^ V ©^ E E
VO ex. ^-
00 <n
at rs| o o
at o o
n o o
CM
CM
O
(0 c Tj"
^ D •o
C
CQ TJ SS ^ c SS
in OS SS
001

in
D in
ra 1^ a. in
so k.
o re
k. ^
Q. sd o O o
O til^ V
D
^
c;
CD
oc
A A A c: i> 1
C5 CD A ci c.

a o u a j j n o o o )uaojad aouajjnooo )uaojad aouajjnooo tuaojad

CM T CM CM
© SI
R o ©
24%

r -> d ©
E E
DBTS

E
, V
at o
at o o
at o o
y—
a. '^M CM
o
CM
•o
c o ©^ c •a
c
2. SS SS
in
in
D vo
1^ a. in
o in
ra ra
o k.
O

A A A A A c
C^ oo CD

a o u a j j n o o o )uaojad aouajjnooo )uaojad


a o u a j j n o o o )uaojad

(0
H r- ,—,
CD • ^
o
Ii rf o ©
©
©
H •—1
d E E
Q. Jl V ©^ E H
O N ti. V)
o o
D o
en o
o o T 5.
CM
CM <U
•o
•o"
TJ P
c C
SS
CO
H
c c
ro ..—<
.t->
in SS in
tn CQ VO a. in a
2 [1 OS o o re
k.
o
O H d <=? k. U
O
Q.
— o o cs
o i y ir*
D (U
A^ A S S S 5
cr3 "s—s—s—5"
oo CO
c5"
C3 oo CO s—^ CT) oo CO -"3- C^
>-l
3
QU
aouajjnooo )U80jad aouajjnooo )U80Jad aouajjnooo )uaojad

100
Plains toad was more acti\e during the second time period than New Mexico spadefoot.

which may ha\e been a consequence of diet o\erlap, because these species use different

microhabitats for breeding (Anderson et al \999a,b).

Barred tiger salamander was most acti\e later in 1999 than all other species. I

believe activity pattems of barred tiger salamander probably were related to

developmental time or physiological constraints associated with probability of

desiccation not competition (Toft 1985). Although barred tiger salamander can develop

as rapidly as anurans on the Southem High Plains. the> may prolong development or

become neotenic during >ears of favorable rainfall (Petranka 1998). Rainfall and

hydroperiod in playas was greater in 1999 than 2000 (Chapter II): thus, barred tiger

salamander may have been more active later in 1999 than other species because

favorable conditions in the aquatic en\ ironment prolong development hence metamorph

emergence. Rainfall also was greater during the last time period (17.5 cm) than the 2

previous periods (14.5 cm and 10.5 cm for 1 and 2, respectively) in 1999. possibly

facilitating dispersal of all age classes (Petranka 1998). If temporal pattems were a

consequence of evolutionary competition (i.e.. ghost of competition past, Connell 1980),

barred tiger salamander should have been more active later in both years; this was not

observed.

Barred tiger salamanders also were most active earlier than all other species in

2000 at cropland playas. Similarly, I suspect this ma>' be a consequence of differential

hydroperiods between landuses (Chapter II). Two of 4 cropland playas were dry prior to

time period 2, whereas all grassland playas retained water through time period 3 in 2000.

101
Perhaps, early playa drying in cropland landscapes caused barred tiger salamanders to

emigrate in search of altemate water sources, resulting in increased levels of activity

(Semlitsch 19876). I speculate that anurans remained near playa bottoms and aestivated

in subterranean burrows following playa drying (McClanahan et al. 1994).

Temporal niche partitioning implies niche overlap and competitive exclusion in at

least one other niche dimension (Schoener 1974. Huey and Pianka 1983, Pianka 1994).

thus seemingly a competitive dominant exists among Southem High Plains anuran

species. Data from Chapter III suggested that New Mexico spadefoots can become

competitively dominant in the absence of larval and neotenic barred tiger salamanders.

Laboratory studies (e.g., Morin 1983, Wilbur 1987, Dayton and Fitzgerald 2001) indicate

that certain spadefoot tadpoles are superior competitors for food resources compared to

other anurans (e.g., Bufo, Rana). Therefore, New Mexico spadefoots may have displaced

the other 2 anurans from earlier activity over evolutionary time, resulting in plains

spadefoot and Great Plains toad being more active during the middle portion of the

breeding season. The time series plots (Figure 5.1) support this inference, because

species-specific percent abundance was > 0 for most days. However, no experiments

exist comparing the relative competitive ability among these species for food or other

resources in terrestrial landscapes. To test the above hypothesis, a shared limiting

resotirce (e.g., food, cover) would need to be identified among these species (Pianka

1976). Then, an experiment could be conducted where individuals were allowed to

compete for the limiting resource in allopatric and sympatric scenarios (Pianka 1976).

102
Numerical or fitness response could be compared subsequently among species (Pianka

1976).

It also has been hypothesized that interspecific competition in harsh environments

is low or nonexistent compared to more stable habitats (Wiens 1977, Huston 1979,

Strong 1983). This idea was stimulated from the observation that biodiversity is greater

in stable climates (e.g., tropics) than in variable ones (e.g., temperate regions, Pianka

1966). Theoretical studies since then have shown that competition and competitive

exclusion operate independently of environmental stability (May and MacArthur 1972,

Holt 1985, Chesson and Huntly 1997); however, few empirical examples exist. Because

climate on the Southem High Plains is semi-arid and hydroperiod in playa wetlands is

variable (Haukos and Smith 1994), these results seem to support the prediction that

interspecific competition and niche partitioning exist in variable environments.

Moreover, it has been suggested that temporal niche partitioning exists most often with

species, such as insects and larval amphibians, that exploit transient resources (Schoener

1974, Toft 1985). Considering the environmental variability in playas and the explosive

breeding strategy of Southem High Plains amphibians, these results also seem to support

this hypothesis.

Although anthropogenic landuse affected dynamics and demographics of

Southem High Plains amphibians (Chapters III and VI), it did not affect temporal niche

partitioning. Perhaps, small sample size and variability in multinomial distribution

pattems prevented detection of biologically meaningful differences. Altematively,

temporal resource use may not be as phenotypically plastic as polymorphic characters or

103
developmental time in amphibians. For example, spadefoots may develop carnivorous

feeding apparatuses in wetlands with short hydroperiods to capitalize abundant transient

invertebrate food resources (Pfennig 1992). Also, many amphibian species may

accelerate growth and undergo metamorphosis at a smaller body size in wetlands with

shorter hydroperiods (Newman 1992). In Chapter II, I suggested the capability of

phenotypic plasticity in development might explain differences in postmetamorphic body

size among playas with differing hydroperiods.

Because amphibians behaviorally hydroregulate (Duellman and Tmeb 1994), it is

reasonable to hypothesize that activity pattems would follow water availability, hence

temporal resource use be shortened in wetlands with rapid hydroperiods (i.e., cropland

playas, Luo et al. 1997). Thus, temporal niche partitioning should have been affected by

landuse, but this was not observed. Accordingly, I hypothesize that change in temporal

resources use may be beyond the genotypic norm of reaction for Southem High Plains

amphibian species (Via and Lande 1985, Gomulkiewicz and Kirkpatrick 1992).

Conservation implications of possible genotypic rigidity in temporal niche

partitioning are important. Paton and Crouch (2002) and Snodgrass et al. (2000) suggest

that duration and timing of wetland hydroperiods are critical to survival and reproduction

of pond-breeding amphibians because of specific-specific phenology in temporal use; my

results support their suggestions. Species, such as plains spadefoot and Great Plains toad,

that use resources later than other species (e.g., New Mexico spadefoot) may have

reduced fitness in anthropogenically modified wetlands, because shortened hydroperiods

may decrease probability of survival and reproduction. Consequently, these species may

104
have a greater probability of local extinction. I recommend retention and restoration of

grasslands around playa wetlands, because hydroperiods generally are shorter in cropland

landscapes from sedimentation (Luo et al. 1997).

105
CHAPTER VI

CHAOS AND REGULATING MECHANISMS

IN A SOUTHERN HIGH PLAINS AMPHIBIAN ASSEMBLAGE

Abstract

Biological chaos is a dynamic condition of nonlinear ecological time series that

exhibits sensitive dependence on initial conditions. Chaos has been demonstrated in

various discrete biological models; however, it rarely has been documented in natural

ecosystems. Simulations suggest that chaos may be a fundamental characteristic of

spatially structured individuals, and its presence can reduce probability of metapopulation

extinction. It also has been hypothesized that anthropogenic disturbance may influence

occurrence of chaos. Amphibians exist in spatially structured playa wetlands embedded

in natural and anthropogenically disturbed landscapes on the Southern High Plains

(SHP). Thus, I tested the difference in occurrence of biological chaos in a SHP

amphibian assemblage between landuses (i.e., cultivation vs. grassland) at 16 playa

wetlands during 1999 and 2000 (4 playas per landuse per year). Also, deterministic and

stochastic difference equations can be used to model assemblages and determine

regulating mechanisms. Thus, I compared occurrence of chaos in a SHP amphibian

assemblage predicted by 5 common difference equations with actual dynamics. For 3

deterministic equations, I also tested for differences in parameter estimates between

iteration and least squares techniques. The amphibian assemblage time series was

constructed from total daily captures in pitfall traps summed across species at each playa.

Existence of chaos was determined via estimating Liapunov exponents per playa and

106
using published stability estimates. Occurrence of chaos was greater at grassland (7/8)

than at cropland (1/8) playas, indicating that biological chaos was a fundamental

characteristic of the SHP amphibian assemblage and its occurrence was altered by

anthropogenic disturbance. The Ricker logistic function with a stochastic parameter

(M+i=Mexp[r(l- N[IK) + ex\) predicted occurrence of chaos similar to actual dynamics,

suggesting that density dependence and environmental stochasticity may be important

regulating mechanisms in this amphibian assemblage. The stochastic Ricker function

predicted that intrinsic rate of increase at grassland playas was greater than at cropland

playas. Intrinsic rates of increase also were greater using an iterative method rather than

the least squares method for estimation in all difference equations. Anthropogenic

disturbance of landscapes surrounding wetlands altered the dynamics of the spatially

stmctured amphibian assemblage on the Southem High Plains, and perhaps, increased

probability of metapopulation extinction via reducing natural occurrence of biological

chaos. Parameter estimates for the stochastic Ricker function are presented for future

simulation of amphibian assemblage dynamics in disturbed and undisturbed wetland

systems.

Introduction

Examining the dynamical behavior of ecological time series is a critical prelude to

understanding regulating mechanisms of assemblages and populations (Edelstein-Keshet

1988) and initiating conservation endeavors. Most biological theories assume abundance

trajectories approach equilibrium and the population or assemblage is inherently stable

(DeAngelis and Waterhouse 1987). However, complex dynamics, such as chaos, have

107
been demonstrated in simple discrete deterministic models (May 1974. 1975fl. 1976a).

Interestingly though, chaos has been detected in few natural assemblages (Turchin and

Taylor 1992). Assemblages that exhibit chaos may have a greater probability of

extinction than those near positive stable equilibria (Hastings et al. 1993, Shulenburger et

al. 1999). Altematively, chaos may be a critical component of spatially stmctured

assemblages, because it can reduce the probability of metapopulation extinction by

increasing local abundance noise (Allen et al. 1993). Anthropogenic modifications of the

landscape also may induce transition from stable equilibria to chaos or vice versa

(Berryman and Millstein 1989, Berryman 1991). This may be especially critical for

assemblages in decline or near extinction thresholds, such as amphibians (Houlahan et al.

2000).

Populations, and possibly assemblages, are regulated via density dependent or

independent mechanisms (Lotka 1925, Andrewartha and Birch 1954). Densit}

dependence implies that dynamics are a consequence of positive or negative feedback

among individuals (Roughgarden 1998). Competition and predation have been

considered important negative feedback processes in most natural populations and

assemblages (May 1976b). Competition and predation can be modeled over discrete time

intervals using difference equations (Edelstein-Keshet 1988). The discrete logistic

growth and Ricker functions are commonly used to model competition (Hastings 1996).

A negative quadratic term can be added to these equations to incorporate a predation

effect (Edelstein-Keshet 1988). Additionally, a stochastic parameter {st) can be included

to model density independent factors (i.e., environmental stochasticity, Turchin and

108
EUner 2000). Parameters of these models can be estimated using observed data and least-

squares estimation or by solving sequential iterations of the difference equations

(Edelstein-Keshet 1988). Parameter estimates subsequently can be used to predict

population or assemblage dynamics. If predictions are similar to actual dynamics, the

regulating mechanism mathematically associated with the function (e.g., densit}

dependence) may be an important process in the population or assemblage (Turchin and

Ellner 2000). Thus, regulating mechanisms in populations and assemblages can be

discemed using observed data and difference equations (Elkinton 2000).

Amphibians exist in spatially stmctured wetlands called playas on the Southem

High Plains (SHP, Bolen et al. 1989). Anthropogenic landuse (i.e., grassland and

cultivation) affects postmetamorphic body size (a fitness correlate), demographics, and

community stmcture of amphibians (Chapters II and III). Because amphibians are

spatially structured on the SHP, the existence of chaotic dynamics may be critical.

Anthropogenic disturbance of the landscape surrounding playas may dismpt natural

dynamics. Thus, my first objective was to determine the presence of chaos in a SHP

amphibian assemblage at 16 playa wetlands from an existing time series and test for

differences in occurrence of chaos between landuses. I examined assemblage instead of

population dynamics because species-specific analyses may have lower probability of

chaos detection (Turchin 1991, 1993, Turchin and Taylor 1992). My second objective

was to determine regulating mechanisms in the SHP amphibian assemblage via

estimating parameters for deterministic and stochastic versions of the discrete logistic and

Ricker functions and comparing their predictions to actual dynamics. I assumed

109
regulating mechanisms (e.g., density dependence) in populations also operated in

assemblages. Turchin (1991) also made this assumption in his chaotic analysis of multi-

species vole data. Lastly. I quantified differences in parameters between iteration and

least-squares estimation techniques. This study provides insight into determining if

biological chaos was a fundamental characteristic of a spatially structured amphibian

assemblage, and whether density dependence, predation, or environmental stochasticity

were important regulating mechanisms.

Methods

This study was conducted at 16 playa wetlands (8 grassland, 8 cropland) on the

SHP of Texas during 1999 and 2000. Playas were considered to be in grassland if >75%

of the surrounding landscape (i.e., <3 km from the playa center) was undisturbed and

vegetated, and cropland if >75% of the surrounding landscape was cultivated (Chapter

III). Playas were partially enclosed (i.e., 25% of circumference) with 60-cm high drift

fence and 19-L pitfall traps (Dodd and Scott 1994). Pitfall traps were checked altemate

days for captures from 16 May-17 October 1999 and 19 April-18 August 2000. Thus,

the amphibian time series represented altemate equally spaced days nested within 2

years. Captured amphibians were enumerated, marked uniquely, and released. A more

detailed account of the study area, amphibians, and sampling protocol has been described

previously (Chapters II and III).

I used total daily capture per playa (natural-log transformed) as an estimate of

relative assemblage size (i.e., Nt, Dodd and Scott 1994). As in Chapter III, I assumed if

any bias existed by using mean daily capture to estimate relative daily abimdance of the

110
assemblage, it was constant between landuses and years. Daily capture was summed

across species for Nt to incorporate positive and negative feedbacks into dynamics

(Schaffer and Kot 1986). As mentioned, Turchin (1991, 1993) summed across vole

species using data of Henttonen et al. (1984) for chaos analyses. Turchin (1991) also

suggested and later demonstrated (Turchin and Taylor 1992) that modeling a nonlinear

time series at lower dimensionality (e.g., number of species, time lags) than the tme

overall dimension of the system can decrease detection of real chaos. I did not sum

captures across playas (i.e., local assemblages), because simulations have demonstrated

that it can reduce probability of detecting actual chaos (Rohde and Rohde 2001). I also

used da>s as time units instead of years or generations. Thus, I assumed regulating

mechanisms, such as density dependence, operated within years for my amphibian

assemblage (Wilbur 1980).

Time series plots were constmcted to graphically illustrate actual dynamics at

each playa. Then, I assumed the amphibian assemblage was govemed by an underlying

difference equation of the form Nt, Nt+i=F(N(). The presence of chaos was determined

via computing local Liapunov exponents per playa as,

Am{t) = -HAt + rn-\}-J{t + l)J{t)\\ , (6.1)


m

where J(t) was the first derivative of F(N) along the assemblage trajectory and m was the

playa-specific length of the time series (Ellner 2000). Liapunov exponents measure the

average deviation in slope of F(N) over time (Abarbanel 1996). They also measure the

rate of trajectory divergence of F(N), commonly called sensitive dependence on initial

conditions, which is tantamount to chaos (Ellner 2000). Positive ^^{t) indicates the

111
presence of chaos; negative exponents suggest convergence to some stable or periodic

equilibrium (Ellner 2000). Inasmuch as equation (6.1) is the natural-log geometric mean

of DF(N), the presence of chaos merely implied the average slope of the trajectory of

F(N) > 1. For detecting chaos in the SHP amphibian assemblage, I initially assumed no

specific form of F(N) (i.e., Nt+i=F(Nt)), because regulating mechanisms (i.e., the

underlying ftanction) were unknown. Thus, Liapunov exponents were estimated directly

by computing the slope of the linear trajectory between time periods (i.e., equally spaced

days) per playa (Olsen and Degn 1985),

N{t + m-\) N{t + \) N{t) (6.2)


m dN dN dN

I assumed the slope of each linear trajectory was a reasonable estimate of the derivative

of A^,. Differences in occurrence of chaos (i.e., positive 'km{t)s) between landuses was

tested (a=0.05) using a Fisher's exact test (Conover 1980:167).

Subsequently, I assumed that density dependence was the primary regulating

mechanism in the SHP amphibian assemblage and followed discrete logistic growth.

A^/+1 = A^^ 1 + r 1- (6.3)

where, r = intrinsic rate of increase and A'= carrying capacity (May 19756). Equation

(6.3) simplifies to

(6.4)
Nt K

Finally, making the following change of variables {a = \ + r, b = r/K), equation (6.4)

simplifies to
^ ^ = a-bNr, (6.5)
Nt

112
which can be solved using 2 successive iterates (Edelstein-Keshet 1988). For example, if

A^i, N2, and N3 = 4.9, 3.6, and 3.7, it follows that NiINx ~ 0.735 and N^^INi = 1.028,

resulting in 2 equations of the form (6.5) with 2 unknowns (i.e., 0.735 = a-4.9b , 1.028 =

a-3.6b). These equations can be solved for a (0.622) and b (-0.113) and subsequently r

(-0.378) and K (3.36). Using daily capture per playa and equation (6.5), I estimated r

and K for each successive pair of iterates in the amphibian assemblage time series, and

computed the arithmetic mean for each parameter per playa. It should be noted that

because my time series was within years and I used difference instead of differential

equations, r and K may not represent the true intrinsic rate of increase and carrying

capacity of the assemblage; they were merely control parameters of these functions.

Because Mean become negative using equation (6.3) if M > K{\+r)lr, which is

biologically unrealistic, I then assumed the amphibian assemblage followed Ricker

logistic growth,

Nt+\ = Nt^^V
'I'-i (6.6)

which is positive for all positive initial conditions (May 19756). Equation (6.6)

simplifies to
In Nt+\ = '-^Nt • (6.7)
\ Nt J

Finally, making the following change of variables {a = r,b = r/K), equation (6.7)

simplifies to
In ^Nt.i^ = a-bNt , (6.8)
V Nt )

which can be solved using successive iterates as described previously

113
Previous research (e.g., Morin 1981, Chapters II and III) also suggests that

predators may be an important regulating mechanism in amphibian assemblages.

Because predation generally has a negative effect on M , I assumed Ricker logistic

growth then added a negative quadratic term for predation to equation (6.6),

A^/+l = A^/exp ,fl_A^l_,^2 (6.9)


K t

Equation (6.9) simplifies to

In {EI±I\ = f- — Nt-cNf (6.10)


\ Nt J

Finally, making the following change of variables {a = r, b = r/K, and c=c), equation

(6.10) simplifies to

In Nt+\ = a-bNt-cNJ (6.11)


\ Nt )

which can be solved using 3 successive iterates similar to previously described.

Parameters in equations (6.5), (6.8) and (6.11) also can be approximated using

least-squares estimation (e.g., Turchin 1990). Letting, yt = M+i/M for equation (6.5), yt =

ln(M+i/M) for equation (6.8), and x, = M for both, equations (6.5) and (6.8) reduce to

y- = a + bxi . (6.12)

which is a linear function. Parameters a and b can be solved by fitting a line to the entire

data set and using the least-squares algorithm (Milton and Amold 1995:386). Similarly,

letting >^, = \n{Nt+i/Nt) and x, = M, equation (6.11) reduces to

y. = a + bxi-cxj , (6.13)

114
which is a quadratic function. Parameters a, b, and c can be solved similarly by fitting a

quadratic curve to the data set and using least-squares estimation. Thus, I estimated

parameters in equations (6.3), (6.6), and (6.9) per playa using an iterative method and

least-squares estimation, and tested for differences between estimation techniques by

computing the difference {D) between estimates per playa then performing a one-sample

Mest on D (HQ: D=0). The aforementioned was necessary because estimation samples

per playa were not independent (Milton and Arnold 1995:353).

Estimating assemblage dynamics in natural systems using deterministic functions

may not be completely realistic because of inherent environmental stochasticity (Ellner

and Turchin 1995). This might be especially tme for amphibian assemblages inhabiting

semi-arid environments, such as on the SHP, because their aboveground dynamics can be

strongly influenced by ambient conditions. Accordingly, Turchin and Ellner (2000)

suggested a stochastic parameter (ct) be added to deterministic functions to increase

realism and account for variability in dynamics associated with the environment. Thus, I

added St to equations (6.6) and (6.9) to create stochastic versions of the Ricker logistic

and Ricker logistic with predator term, respectively:

Nt+\ = Nt^^P + £t (6.14)


V KJ

and.

Nt+\ = Nt^^P ,i-f)-c;v?.. (6.15)

Parameters in equations (6.14) and (6.15) were approximated by fitting the functions to

the observed data and using least-squares estimation (Turchin and Ellner 2000), which

115
was done using PROC NLIN in SAS® (Freund and Littell 2000). A reasonable estimate

of ft for the function is mean-squared error of the nonlinear regression (Turchin and

Ellner 2000). Residuals can be used as time-specific CtS, which were needed to classify

dynamics (discussed later).

Turchin and Ellner (2000) also suggested density-dependent difference equations

(e.g., equations 6.3, 6.6, 6.9, 6.14, and 6.15) should be constmcted with lags to increase

prediction realism. Lags should correspond to duration of delayed density dependence,

which is the embedding dimension of the time series. Because no information exists on

the presence of delayed density dependence in the amphibian assemblage on the SHP, I

assumed embedding dimension = 1 for all aforementioned difference equations. This is

reasonable for many vertebrate assemblages because density dependence and predation

usually most strongly affects individuals in the adjacent time period (Turchin and Ellner

2000). Moreover, graphs of autocorrelation functions from my time series (M. J. Gray.

unpublished data) suggested lack of correlation after 1 time lag (Turchin 1990, 1996).

Thus, I did not incorporate lags into presented functions.

Predicted dynamics from difference equations were classified as chaotic or not

chaotic using published stability estimates from May (19756) for equations (6.3) and

(6.6), and estimating Liapunov exponents for equations (6.9), (6.14), and (6.15). If r >

2.59 and 2.69, dynamics were classified as chaotic for equations (6.3) and (6.6),

respectively (May 19756). Estimation of Liapunov exponents for the other difference

equations was straightforward using equation (6.1), because F{N^ (i.e., the right hand

side of the difference equations) was a constant (c) after solving using parameter

116
estimates. £t and M (Ellner 2000). It follows. DF(M) was mereh' e' for equations (6.9).

(6.14), and (6.15), because the time-specific derivative {J(t)) of M+i= e'N^ was e' (Ellner

2000). Posifive Liapunov exponents indicated chaotic dynamics (Olsen and Degn 1985).

Predicted occurrence of chaos for each difference equation was tested between

landuses and actual dynamics (i.e., chaos occurrence estimated assuming no specific

F(N)) using a main effect logistic regression with a generalized logit model (Stokes et al.

2000:257). Parameter estimates per playa were presented for functions that predicted

chaotic frequency similar to the actual dynamics. Lastly, estimates of parameters for all

difference equations were tested between landuses using a 2-sample /-test (Milton and

Amold 1995:347).

Results

Time series plots of relative abundance at each playa clearly suggested that

within-year dynamics of the SHP amphibian assemblage were not linear and may not

approach a single positive stable equilibrium (Figures 6.1-6.2). Measures of stability for

actual and predicted dynamics from difference equations indicated that chaos was a

fundamental characteristic of the SHP amphibian assemblage (Table 6.1). Actual

occurrence of chaos was greater at grassland (7/8) than cropland (1/8) playas (Table 6.2).

The stochastic Ricker function predicted occurrence of chaos similar to actual dynamics

(i.e., significant and non-significant landuse and model effect, respectively, Table 6.2);

least-squares estimates of its parameters per playa are presented (Table 6.3). The

deterministic and stochastic Ricker fiinctions with a predator term also predicted

dynamics similar to actual dynamics (i.e., no model effect); however, no landuse effect

117
118
Figure 6.1. Time series plots of amphibian assemblage dynamics at 8 playa wetlands
during 16 May-17 October 1999 on the Southem High Plains, Texas. Relative
abundance was total daily capture (natural-log transformed) of all species in pitfall traps;
days were altemate equally spaced days within the time series. Time series for playas 1
and 4 were delayed because rain caused flooding over pitfalls. Playas 1-4 and 5-8 were
located in cropland and grassland, respectively.
6.5

5.5
8c
I 4.5 nj
•o
c c
< 35 XI
> <
TO >
"S 2.5
cr

1.5

0.5

7 -f

0) c
o
c 5 - as
nj T3
-a c
c
XI
< <(1)
<U >
>
"35
<u
ca
1 -

6 -

(U
5 - 0)
o
o c
T3
nj
C
c XI
D
3 -
< 3 -
< >
I
a:
1 -

0 -
10 20 30 40 50 60
Days

b —(

5 -

c c
to n
•o
J3 c
C
X3 X3
< <
>
s
a:

0 -
10 20 30 40 50 60
Days

119
120
Figure 6.2. Time series plots of amphibian assemblage dynamics at 8 playa wetlands
during 19 April-18 August 2000 on the Southem High Plains, Texas. Relative
abundance was total daily capture (natural-log transformed) of all species in pitfall traps;
days were altemate equally spaced days within the time series. Playas 9-12 and 13-16
were located in cropland and grassland, respectively.
6 -f
6 -

5 -
5 -
0)
o o
c c
01 ro
"O •o
c c:
X3 XI
< 3 - 3 -
> <
ro >
"83 j5 2
OJ

1 -

20 ^ 30
Days

6 -f
1 1J Playa 11
5 - (U
o
c
ro ro
•o •o
c c
3 3
X3 XI
< < 3 -
41
> >
— 9

1 - \ 1 -

1 1 1 1 1
10 20 r^ 30 40 50
Days

lU 5 -
8 o
c
c ro
ro 2 -
•D TJ
c c
3
X3 3
< XI
<
> >
-5
0)
1I ro
2 -
•33
Q:

0 -

0)
o
c c
ro ro
TJ TJ
c 4 - c
3 3
XI X3
< 3 - <
!« >
2 -
(U

cr

20 30
Days

121
Table 6.1. Stability of amphibian assemblage trajectories at 16 playa wetlands during 16
May-17 October 1999 and 19 April-18 August 2000 on the Southem High Plains, Texas.
Stability
F(Ny Year Landuse Playa Parameter Classification'^
NSF 1999 Cropland 1 -0.004 Not Chaotic
2 -0.015 Not Chaotic
3 -0.009 Not Chaotic
4 0.001 Chaotic
Grassland 5 0.003 Chaotic
6 0.001 Chaotic
7 0.001 Chaotic
8 0.008 Chaotic
2000 Cropland 9 -0.007 Not Chaotic
10 -0.014 Not Chaotic
11 -0.019 Not Chaotic
12 -0.029 Not Chaotic
Grassland 13 0.023 Chaotic
14 0.015 Chaotic
15 0.015 Chaotic
16 -0.011 Not Chaotic
DDL 1999 Cropland 1 4.465 Chaotic
2 3.691 Chaotic
3 2.693 Chaotic
4 3.931 Chaotic
Grassland 5 2.808 Chaotic
6 3.242 Chaotic
7 2.791 Chaotic
8 2.841 Chaotic
2000 Cropland 9 2.516 Not Chaotic
10 2.842 Chaotic
11 3.087 Chaotic
12 2.271 Not Chaotic
Grassland 13 3.031 Chaotic
14 2.101 Not Chaotic
15 2.219 Not Chaotic
16 2.861 Chaotic

122
Table 6.1. Continued.
Stability
F(Nf Year Landuse Playa Parameter'' Classification'^
DDR 1999 Cropland 1 3.881 Chaotic
2 3.105 Chaotic
3 3.029 Chaotic
4 3.706 Chaotic
Grassland 5 2.615 Not Chaotic
6 2.922 Chaotic
7 2.665 Not Chaotic
8 3.021 Chaotic
2000 Cropland 9 2.305 Not Chaotic
10 2.706 Chaotic
11 2.817 Chaotic
12 1.848 Not Chaotic
Grassland 13 3.141 Chaotic
14 1.995 Not Chaotic
15 2.064 Not Chaotic
16 2.021 Not Chaotic
DSR 1999 Cropland 1 0.006 Chaotic
2 0.004 Chaotic
3 -0.001 Not Chaotic
4 -0.003 Not Chaotic
Grassland 5 0.009 Chaotic
6 0.009 Chaotic
7 -0.002 Not Chaotic
8 0.008 Chaotic
2000 Cropland 9 -0.001 Not Chaotic
10 -0.007 Not Chaotic
11 -0.012 Not Chaotic
12 -0.027 Not Chaotic
Grassland 13 0.001 Chaotic
14 0.024 Chaotic
15 0.015 Chaotic
16 0.011 Chaotic

123
Table 6.1. Continued.
Stability
F(nf Year Landuse Playa Parameter' Classification'
DRP 1999 Cropland 1 -0.036 Not Chaotic
2 0.023 Chaotic
3 -0.021 Not Chaotic
4 -0.419 Not Chaotic
Grassland 5 -0.093 Not Chaotic
6 -0.037 Not Chaotic
7 0.002 Chaotic
8 0.087 Chaotic
2000 Cropland 9 -0.014 Not Chaotic
10 -0.004 Not Chaotic
11 -0.006 Not Chaotic
12 -0.108 Not Chaotic
Grassland 13 -0.099 Not Chaotic
14 -0.033 Not Chaotic
15 0.101 Chaotic
16 0.252 Chaotic
SRP 1999 Cropland 1 -0.029 Not Chaotic
2 0.021 Chaotic
3 -0.019 Not Chaotic
4 -0.423 Not Chaofic
Grassland 5 -0.082 Not Chaotic
6 -0.029 Not Chaotic
7 0.006 Chaotic
8 0.098 Chaotic
2000 Cropland 9 -0.009 Not Chaotic
10 -0.015 Not Chaofic
11 -0.012 Not Chaotic
12 -0.139 Not Chaotic
Grassland 13 -0.108 Not Chaotic
14 -0.014 Not Chaotic
15 0.099 Chaotic
16 0.248 Chaotic

124
Table 6.1. Continued.

^ S F - no specific form assumed (i.e., Nt+i=F(Ni), tme population dynamics), DDL =


discrete deterministic logistic (M+i=M + rNt[\- Nt/K]), DDR = discrete determinisfic
Ricker (M+i=Mexp[r(l- N^/K)]), DSR = discrete stochastic Ricker (M+i=Mexp[r(l- M / ^
+ etl), DRP = discrete determinisfic Ricker with predator term (M+i=Mexp[r(l- Nt/K) -
cNt ]), and SRP = discrete stochastic Ricker with predator term (M+i=Mexp[r(l- Nt/K) -
cNt^ + et]) {May 1975b).
''Stability parameter was X (Liapunov exponent) for NSF, DSR, DRP, and SRP and r
(intrinsic rate of increase) for DDL and DDR (May 19756, Olsen and Degn 1985).
"Dynamics classified as chaotic when A > 0, and r > 2.59 and r > 2.69 for DDL and DDR,
respecfively (May 19756, Olsen and Degn 1985).

125
Table 6.2. Occurrence of chaos between landuses and models in amphibian assemblages
at 16 playa wefiands during 16 May-17 October 1999 and 19 April-18 August 2000 on
the Southem High Plains, Texas.
Percent Chaos Landuse Effect Model Effect"
F(N/ Cropland Grassland X'^ P X^ P
NSF 12.5 87.5 9.0 0.003 NT NT
DDL 75.0 75.0 0.0 1.0 4.26 0.039
DDR 75.0 37.5 2.29 0.131 6.35 0.012
DSR 25.0 87.5 6.35 0.012 0.18 0.675
DRP 12.5 50.0 2.62 0.106 0.96 0.326
DSP 12.5 50.0 2.62 0.106 0.96 0.326
^ S F = no specific form assumed (i.e., Nt+\=F(Nt). true population dynamics), DDL
discrete deterministic logistic (M+i=M + rNt[\- Nt/K]), DDR = discrete deterministic
Ricker {Nt+i=NtQxp[r{l- Nt/K)]), DSR = discrete stochastic Ricker {Nt+i=Ntexp[r{\- Nt/K)
+ Et]), DRP = discrete deterministic Ricker with predator term {Nt+\=Ntexp[r{\- Nt/K) -
cNt ]), and SRP = discrete stochastic Ricker with predator term {Nt+\^Nttxp\r{\-Nt/K) -
cM^ + et]) (May 19756).
''Model effect was pairwise Fisher's exact tests in percent chaos between NSF (i.e., tme
dynamics) and other functions; no test (NT) was performed on NSF and itself

126
Table 6.3. Least-squares esfimates of parameters for the discrete stochastic Ricker
equation^ from amphibian assemblages at 16 playa wetlands during 16 May-17 October
1999 and 19 April-18 August 2000 on the Southem High Plains, Texas.
Parameter Estimate
Year Landuse Playa r K Et MSreg" F" P
1999 Cropland 1 0.273 4.493 0.057 0.168 5.65 0.022
2 0.292 3.551 0.174 0.798 9.14 0.004
3 0.308 3.795 0.113 0.392 6.60 0.013
4 0.118 3.407 0.116 0.189 2.46 0.123
Grassland 5 0.593 4.064 0.247 1.578 12.71 <0.001
6 0.673 3.187 0.211 2.561 24.27 <0.001
7 0.813 3.924 0.175 2.394 27.35 <0.001
8 0.486 3.251 0.209 1.568 15.02 <0.001
2000 Cropland 9 0.491 3.041 0.228 1.277 11.03 0.002
10 0.346 3.807 0.131 0.421 6.07 0.018
11 0.457 3.996 0.111 0.643 11.04 0.002
12 0.677 2.992 0.317 2.955 18.55 <0.001
Grassland 13 0.525 2.113 0.189 1.186 12.54 0.001
14 0.467 2.731 0.174 1.059 12.19 0.001
15 0.654 3.679 0.189 1.571 16.55 <0.001
16 0.511 3.509 0.159 1.025 12.83 <0.001
^ h e discrete stochastic Ricker equation is A^t+i=Mexp(r[l-A^t/-^ + ^t); parameter
estimates can be used for predicting playa- and time-specific assemblage dynamics.
''Mean-square regression (MSreg) is variation in assemblage dynamics explained by the
discrete stochastic Ricker equation (Freund and Littell 2000:185-189).
'^Numerator, denominator degrees of freedom of F were 2,65 and 2,50 for 1999 and 2000,
respectively.

127
was detected, which was inconsistent with observed dynamics (Table 6.2). The

stochastic Ricker function predicted mean intrinsic rate of increase was greater at

grassland than cropland playas; carrying capacities were similar (Table 6.4). No

differences in parameters between landuses were detected for other functions (Table 6.4).

Lastly, estimates of parameters for the deterministic logistic, Ricker, and Ricker with

predator term functions were different the between iterative method and least-squares

estimation (Table 6.5). Estimates of intrinsic rate of increase were greater using the

iterative method than by least squares (Table 6.5). Parametric assumptions of residual

normality (P>0.79) and homoscedasticity {P>0.23) were satisfied for all ^tests.

Discussion

Occurrence of chaos in the SHP amphibian assemblage was greater at grassland

than cropland playas. These results represent the first empirical evidence of chaos in

spatially stmctured amphibian assemblages, and more importantly, that its frequency can

be affected by anthropogenic disturbance. Occurrence of chaos may have been lower at

cropland playas, because anthropogenic disturbance might have affected normal

assemblage processes, such as trophic stmcturing or interdemic exchange of individuals

among spatially stmctured populations. Agricultural landscape use and stmcture can

affect local population demographics and body size of amphibians (Chapters II-IV). In

particular, mean daily abundance of some species was greater at cropland than grassland

playas, perhaps because of the anthropogenic reduction of an intraguild predator (i.e.,

barred tiger salamander, Ambystoma tigrinum mavortium) in cropland playas. Also,

differential viscosity of anthropogenic crop types may have influenced movements of

128
Table 6.4. Iterafive and least-squares estimates of parameters for deterministic difference
equafions used to model amphibian assemblages at 16 playa wetlands during 16 May-17
October 1999 and 19 April-18 August 2000 on the Southem High Plains, Texas.
Estimation Technique Paired
F(N/ Parameter Iteration Least-Squares Difference'^
X SE X SE X SE /15 P
DDL r 3.17 0.15 0.78 0.09 2.39 0.21 11.42 <0.001
K 3.94 0.16 4.12 0.16 0.17 0.07 2.49 0.025
DDR r 2.74 0.15 0.48 0.05 2.26 0.18 12.57 <0.001
K 3.89 0.15 3.47 0.15 0.43 0.08 5.27 <0.001
DRP r -1.52 4.10 0.89 0.13 11.33 3.02 3.74 0.002
K 2.40 0.26 2.89 0.60 1.48 0.61 2.42 0.029
c 0.10 0.22 -0.03 0.008 0.61 0.17 3.66 0.002
"^DL = discrete determinisfic logisfic (M+i=M + ^ ^ [ 1 - Nt/K]), DDR = discrete
deterministic Ricker {Nt+i=NtQxp[r{\- Nt/K)]), and DRP = discrete deterministic Ricker
with predator term (M+i=Mexp[r(l- Nt/K) -cM^]) (May 19756).
Iterative estimates were computed by making a change of variable {a=\+r and b=r/K for
DDL, a=r and b=r/K for DDR, and a=r, b=r/k, and c=c for DRP), solving 2-3 iterates
using daily capture per playa for M+i and Nt, and computing the arithmetic mean within
then across playas; least-squares estimates were ordinary least-squares computed by
fitting the deterministic function to the entire data set and estimating parameters that
allow the best available fit (Milton and Amold 1995:386).
'^Mean paired difference {D) was the absolute difference between iterative and least-
square estimates averaged across playas; a one-sample /-test (HQ: D = 0 ) was performed on
arithmetic mean of D to test for difference between parameter estimates because samples
were not independent (Milton and Amold 1995:353).

129
Table 6.5. Esfimates of parameters between landuses for difference equations used to
model amphibian assemblages at 16 playa wefiands during 16 May-17 October 1999 and
19 April-18 August 2000 on the Southem High Plains, Texas.
Cropland Grassland
F(nr Parameter X SE X SE tu P
DDL r 3.19 0.27 2.74 0.14 1.48 0.161
k 4.19 0.18 3.75 0.27 1.36 0.196
DDR r 2.92 0.24 2.56 0.17 1.27 0.225
k 4.09 0.19 3.69 0.22 1.35 0.199
DSR r 0.37 0.06 0.59 0.04 3.02 0.009
k 3.64 0.18 2.77 0.23 1.14 0.275
DRP r 4.07 6.78 -7.02 4.21 1.39 0.186
k 2.52 0.46 2.28 0.26 0.46 0.651
c -0.17 0.39 0.37 0.21 1.23 0.239
DSP r 0.62 0.18 1.02 0.16 1.64 0.123
k 1.85 0.32 2.35 0.23 1.25 0.231
c -0.02 0.01 -0.05 0.02 2.06 0.361
"^DL = discrete deterministic logisfic (M+i=M + rNt[\- NJK]), DDR = discrete
deterministic Ricker {Nt+\=NtQxp[r{\- Nt/K)\), DSR = discrete stochastic Ricker
(M+i'^Mexp[r(l- Nt/K) + et]), DRP = discrete deterministic Ricker with predator term
{Nt+\=Nttxp[r{\- Nt/K) -cN^]), and SRP = discrete stochastic Ricker with predator term
(M+i=Mexp[r(l- Nt/K) -cN^ + Et]) (May 19756).

130
individuals among local assemblages. Existence of predafion and interdemic movement

of individuals among grassland playas may act as natural destabilizing mechanisms,

owing to the increased occurrence of chaos in undisturbed grassland playas. Allen et al.

(1993) demonstrated that chaos was an important characteristic of simulated spatially

structured populations, because it can reduce probability of metapopulation extinction by

amplifying local population noise. Thus, a spatially structured amphibian assemblage

surrounded by cultivation may have a greater probability of extinction than one

surrounded by an undisturbed landscape.

Differences in occurrence of chaos in the SHP amphibian assemblage also may

have been related to density dependence and environmental stochasticity. The stochastic

Ricker function predicted occurrence of chaos most similar to actual dynamics,

suggesting that density dependence is an important regulating mechanism in the

assemblage. Density dependence has been hypothesized and shown to regulate some

amphibians (Wilbur 1980, 1984). Density dependence probably is a regulating

mechanism in the SHP amphibian assemblage, because most species are explosive

breeders that emerge simultaneously and produce numerous offspring (see species review

in Chapter II). Consequently, relative abundance of individuals can increase rapidly. For

example, total daily capture increased from 30 to 14,623 individuals in 1 week at 1 playa

in 1999. I suspect competition for food, cover, and mating resources may become

limiting under these and less extreme dynamics, inducing density dependent regulation.

Density independence also seems to be important in regulating the SHP amphibian

assemblage, because the deterministic Ricker model did not predict occurrence of chaos

131
similar to actual dynamics. Thus, these results support Ellner and Turchin (1995)

hypothesis that including a stochastic parameter in completely deterministic models

makes them more realistic.

The stochastic Ricker function also predicted that intrinsic rate of increase was

greater in grassland than cropland playas. Elementary stability analyses demonstrate that

as the control parameter of an underlying function (such as r) increases, probability of

progression to a chaotic state also increases (e.g.. May 1975). The aforementioned may

be the functional response to suspected differential predation, interdemic movement, and

density dependence between landuses, and the mathematical reason for differences in

chaotic occurrence between landuses.

Although no differences were detected, the stochastic Ricker function also

predicted mean carrying capacity was greater at cropland (x=38.1 captures/day, SE=1.19)

than grassland (x=15.9 captures/day, SE=1.26) playas. As I hypothesized in Chapters II

and III, carrying capacity may be lower in grassland playas because larval and neotenic

barred tiger salamanders may be exhibiting top-down control. Indeed, deterministic and

stochastic Ricker functions with predator terms predicted occurrence identical and similar
A

to actual dynamics in cropland and grassland, respectively. Nutrient loading also may be

greater in cropland than grassland playas due to agriculture (Freemark and Boutin 1995),

facilitating greater numbers of individuals in cropland playas via increasing food

abundance (Chapter III). Although abundance and carrying capacity may be higher in

cropland playas, the chaos results herein indicate that dynamics of an amphibian

assemblage in disturbed wetlands are altered from their natural state in grasslands.

132
Estimates of intrinsic rate of increase were greater using the iterati\ e method than

least squares for all difference equations. This probably was a consequence of

methodological differences in their algorithms. As described previously, sets of

difference equations were solved and parameters estimated for 2-3 sequential time

iterates for the iterative technique. Thus, parameters per playa were approximated from

m/2 or m/3 samples (where m is the duration of the time series) for non-predator and

predator difference functions, respectively. The arithmetic average then was calculated

per playa for each parameter estimate. In contrast, the least-squares technique fit the

difference function to the entire data such that residual sum-of-squares v\ as minimized,

and obtained one sample estimate per parameter per playa. Consequently, iteration may

capture the true dynamics of a chaotic time series better than least-squares estimation

because it is based on more samples. However, limited simulations (J. Surles, Texas

Tech University, unpublished data) suggest that least-squares estimates are less \ ariable

and more robust than those from the iterative method.

Ecological and Mathematical Implications

Chaos was a ftjndamental characteristic of a spatially stmctured amphibian

assemblage on the SHP, and anthropogenic disturbance around playa wetlands can

dismpt its occurrence. Chaos may exist naturally in this assemblage because of densit>

dependent and independent regulation. Dynamics also may be a consequence of

predation and interdemic movement (Chapters III and IV). Additional research is needed

to test these hypotheses. Regardless, dynamics of amphibians in cropland wetlands are

altered from their natural state in grasslands. These results and those of Allen et al.

133
(1993) on simulated metapopulations suggest that spatially stmctured amphibians

positioned in cropland may have a greater probability of extinction, because of reduced

natural occurrence of chaos. Future research needs to discem differences in the

mechanisms regulating amphibian assemblages and populations between these

anthropogenic landuse types. Subsequently, we need to evaluate conservation techniques

(e.g., wetland buffer restoration, corridor construction) that restore dynamics and

demographics similar to an undisturbed state. Amphibian dynamics at undisturbed and

disturbed wetlands can be predicted initially by using the stochastic Ricker model and

parameter estimates presented herein (Table 6.3). The true underlying function

describing dynamics and the relationship of regulating mechanisms in the SHP

amphibian assemblage likely is more complex than the stochastic Ricker function. Thus,

future modeling endeavors should incorporate additional parameters, such as immigration

and emigration, and investigate other relationships of density dependence (e.g.,

N,+]^aNt/[l+{bNty], Maynard Smith 1974) and predation. Future models also should

include a stochastic parameter (ct) to improve predictive ability. Additional density

independent state variables (e.g., rainfall, water chemistry) could be added to refine

models. Cross validation techniques and information statistics (e.g., Akaike information

criterion) can be used to compare predictability among difference equations (Turchin and

Ellner 2000). Lastly, species-specific analyses would be useful to discem if chaos is a

ftjndamental characteristic of SHP amphibian populations. Reduced detection of chaos at

the population-level may give insight into the embedding dimension of the amphibian

assemblage, and provide support for combined assemblage analyses as presented here.

134
CHAPTER VII

CONCLUSION

This study provides evidence that anthropogenic disturbance surrounding

wetlands affects body size, demographics, and chaotic dynamics of amphibians on the

Southern High Plains. In particular, postmetamorphic body size of amphibians was

greater at grassland playa wetlands than at playas surrounded by agricultural cultivation

(Chapter II). The presence of biological chaos also was altered by anthropogenic

disturbance (Chapter VI). Amphibian assemblages surrounded by grassland exhibited

chaotic dynamics more often than assemblages in cropland landscapes. Abundance of

New Mexico and plains spadefoots {Spea multiplicata, S. bombifrons) was positively

affected by cultivation; abundance of other species and diversity of the amphibian

assemblage was not affected (Chapter III). Grassland playas exhibited source dynamics

most often, while playas in cultivated landscapes generally were sinks (Chapter III).

Landuse did not affect temporal niche partitioning among amphibian species (Chapter V).

Geometric indices representing spatial positioning of playas and landscape complexity

were positively related with amphibian abundance and affected community composition

(Chapter IV). A stochastic Ricker equation predicted that density dependence and

environmental stochasticity were possible regulating mechanisms in the Southem High

Plains amphibian assemblage (Chapter VI).

Various abiotic and biotic factors probably infiuenced amphibian responses to

anthropogenic disturbance treatments (cultivation vs. grassland). Differences in body

size between cultivated and grassland landscapes may have been related to wetland

135
hydroperiod, agricultural chemicals, availability of food resources, or density of

conspecifics (Chapter II). Occurrence of biological chaos probably was greater in

grassland landscapes, because intrinsic rate of increase was greater at grassland than

cropland playas (Chapter VI). Differences in chaotic dynamics between anthropogenic

landuses may have been related to competition, predation, interdemic movement, or

environmental stochasticity (Chapter VI). Species-specific vagility and anthropogenic

reduction of barred tiger salamanders {Ambystoma tigrinum mavortium) in the aquatic

environment may have positively influenced spadefoot abundance at cropland playas

(Chapter III). Rigidity in the genie norm of reaction for temporal resource use may have

resulted in no differences between landuses in temporal niche partitioning (Chapter V).

Amphibian abundance was positively related to playa juxtaposition, probably because

probability of interdemic movement increases with decreasing dispersal distance

(Chapter IV). Abundance might be positively related to landscape complexity, because

amphibians may perceive complex agricultural landscapes as impervious, resulting in

increased abundance near natal wetlands (Chapter IV). Most of these inferences are

based on results in lab and field experiments on amphibians elsewhere or from

predictions by difference equations; thus, they are speculative. More research is needed

(discussed later). Regardless, results herein strongly suggest that agricultural cultivation

around wetlands affects amphibian populations, thus the relative importance of various

regulating mechanisms (e.g., competition, predation, interdemic movement).

Although spadefoot abundance was positively influenced by cultivation, I

recommend retention and restoration of native grasslands surrotmding playa wetlands.

136
Cropland playas exhibited sink dynamics whereas playas in grassland landscapes were

most often sources of individuals (Chapter III). Inasmuch as body size is positively

related to survival and reproduction (thus evolutionary fitness) in amphibians,

populations surrounded by cropland also may have a greater probability of extinction

(Chapter II). Presence of biological chaos may increase persistence of spatially

structured populations by increasing local population noise; thus, an amphibian

metapopulation in a cropland landscape may have greater probability of exfinction than

one in grassland (Chapter VI). Incidence of disease, parasite infestation and

malformations also can be positively related to amphibian abundance thus increase in

cropland landscapes (Carey et al. 1999). Lastly, I suspect that interdemic movement is

reduced in cropland landscapes, which can have negative demographic and genetic

consequences on amphibian populations (Brown and Kodric-Brown 1977, Hastings and

Harrison 1994).

Additional research is needed on the effect of anthropogenic disturbance on

Southem High Plains (SHP) amphibians. I recommend the following studies: (1)

determine the effect of landscape use and structure on interdemic genetic variability of

SHP amphibians, (2) estimate species-specific vagility of SHP amphibians in natural and

anthropogenic cover types, (3) estimate home ranges, dispersal probabilities, and micro-

habitat use of SHP amphibians, (4) simultaneously examine the effect of anthropogenic

landuse on sedimentation rate, water chemistry, aquatic and terrestrial food resources,

and vegetation characteristics in playa wetlands, (5) investigate the effect of landuse on

body size, demographics, and dynamics of SHP amphibian larvae populations, (6)

137
examine pathogenic incidence between landuses in SHP amphibians, (7) determine the

effect of UV-B radiation on hatching and survival of SHP amphibian larvae, and (8)

determine the effect of landuse on the relative importance of regulafing mechanisms (e.g..

competition and predation) in aquatic and terrestrial amphibian populations. These

studies could be accomplished through multiple concurrent lab and field experiments. A

manipulative study in which populations were monitored in a cropland landscape before

and after grassland restoration, or alternatively in a grassland landscape before and after

disturbance, would be novel and provide strong inferential insight. Results from the

aforementioned research could be used to predict population dynamics using difference

equations or discrete Markov chains. Moreover, landscape changes could be simulated

using the geographic information system to predict the effect of restoration techniques

(e.g., entire grassland restoration, corridor construction) or disturbance on amphibian

populations. Lastly, a long-term amphibian population monitoring program needs to be

established on the SHP, and demographic information of playa amphibians tied to global

declines.

The Southem High Plains is an incredible region to examine the effect of

anthropogenic disturbance on amphibian populations. Thousands of playa wetlands in

cultivated and grassland landscapes allow comparative examination of human impacts on

wildlife populations with substantial statistical replication. Playas also are spatially

stmctured; thus, they are ideal systems to test various hypotheses of metapopulation

theory. Therefore, an experimenter can investigate progressive concepts in conservation

biology and landscape ecology simultaneously in nature—This is a rare scientific

138
opportunity. I encourage continued amphibian research on the Southem High Plains.

Indeed, Ann Anderson's (1997) research and this study only have scratched the surface of

our understanding of basic life history for SHP amphibians and the effect of human

disturbance on their population processes. I am confident the stellar researchers at Texas

Tech University will guarantee the progression of science and our knowledge. Buena

suerte y que la ciencia provea la luz\

139
LITERATURE CITED

Abarbanel, H. D. I. 1996. Analysis of observed chaotic data. Springer-Verlag, New


York, New York, USA.

Agresti, A. 1990. Categorical data analysis. John Wiley and Sons, New York, New
York, USA.

Alford, R. A. 1999. Ecology: resource use, competition, and predation. Pages 240-278
in R. W. McDiarmid and R. Altig, editors. Tadpoles: the biology of anuran
larvae. University of Chicago, Chicago, Illinois, USA.

Alford, R. A., and R. N. Harris. 1988. Effects of larval growth history on anuran
metamorphosis. American Naturalist 131:91-106.

Alford, R. A., and S. J. Richards. 1999. Global amphibian declines: a problem in applied
ecology. Annual Review of Ecology and Systematics 30:133-165.

Allen, B. L., B. L. Harris, K. R. Davis, and G. B. Miller. 1972. The mineralogy and
chemistry of High Plains playa lake soils and sediments. Water Resources Center
Publication WRC-72-4, Texas Tech University, Lubbock, Texas, USA.

Allen, J. C , W. M. Schaffer, and D. Rosko. 1993. Chaos reduces species extincfion by


amplifying local population noise. Nature 364:229-232.

Anderson, A. M. 1997. Habitat use and diet of amphibians breeding in playa wetlands
on the Southem High Plains of Texas. Thesis, Texas Tech University, Lubbock,
USA.

Anderson, A. M., D. A. Haukos, and J. T. Anderson. 1999a. Habitat use by anurans


emerging breeding in playa wetlands. Wildlife Society Bulletin 27:759-769.

Anderson, A. M., D. A. Haukos, and J. T. Anderson. 19996. Diet composition of three


anurans from the playa wetlands of northwest Texas. Copeia 1999:515-520.

Andrewartha, H. G., and L. C. Birch. 1954. The distribution and abundance of animals.
University of Chicago, Chicago, Illinois, USA.

Baker, J. M. R., and V. Waights. 1993. The effect of sodium nitrate on the growth and
survival of toad larvae {Bufo bufo) in the laboratory. HerpetologicalJournal
3:147-148.

140
Baker, J. M. R., and V. Waights. 1994. The effects of nitrate on tadpoles of the tree frog
{Litoria caerulea). Herpetological Journal 4:106-108.

Bardsley, L. S. 1998. Use and selecfion of terrestrial resources by adult common toads
{Bufo bufo) in agricultural landscapes. Dissertation. De Montfort Univerisit>.
Leicester, United Kingdom.

Bascompte, J., and R. V. Sole. 1996. Habitat fragmentation and extinction thresholds in
spatially explicit models. Journal of Animal Ecology 65:465-473.

Beck, C. W., and J. D. Congdon. 2000. Effects of age and size at metamorphosis on
performance and metabolic rates of Southern Toad, Bufo terrestris. metamorphs.
Functional Ecology 14:32-38.

Beebee, T. J. C. 1983. Habitat selection by amphibians across an agricultural


land-heathland transect in Britain. Biological Conservation 27:111-124.

Beebee, T. J. C. 1996. Ecology and conservation of amphibians. Chapman and Hall,


London, UK.

Berger, L. 1989. Disappearance of amphibian larvae in the agricultural landscape.


Ecology International Bulletin 17:65-73.

Berryman, A. A. 1991. Chaos in ecology and resource management: what causes it and
how to avoid it. Pages 23-38 in J. A. Logan and F. P. Hain, editors. Chaos and
insect ecology. Virginia Polytechnic Institute and State University, Virginia
Experiment Station Information Series 91-3, Blacksburg, Virginia, USA.

Berryman, A. A., and J. A. Millstein. 1989. Are ecological systems chaotic-And if not,
why not? Trends in Ecology and Evolution 4:26-28.

Berven, K. A. 1981. Mate choice in the wood frog, Rana sylvatica. Evolution
35:707-722.

Berven, K. A. 1982. The genetic basis of altitudinal variation in the wood frog Rana
sylvatica. I. An experimental analysis of life history traits. Evolution
36:962-983.

Berven, K. A. 1990. Factors affecting population fluctuations in larval and adult stages
of the wood frog {Rana sylvatica). Ecology 71:1599-1608.

Berven, K. A. 1995. Populafion regulation in the wood frog, Rana sylvatica, from three
diverse geographic localities. Australian Journal of Ecology 20:385-392.

141
Berven, K. A., and D. E. Gill. 1983. Interpreting geographic variation in life-history
traits. American Zoologist 23:85-97.

Blair, W. F. 1961. Calling and spawning in a mixed population of anurans. Ecology


42:99-110.

Blaustein, A. R., D. B. Wake, and W. P. Sousa. 1994. Amphibian declines: judging


stability, persistence, and susceptibility of populations to local and global
extinctions. Conservation Biology 8:60-71.

Bolen, E. G., L. M. Smith, and H. L. Schramm, Jr. 1989. Playa lakes: prairie wetlands of
the Southern High Plains. BioScience 39:615-623.

Bonin, J., J.-L. DesGranges, J. Rodrigue, and M. Ouellet. 1997a. Anuran species
richness in agricultural landscapes of Quebec: foreseeing long-term results of road
call surveys. Pages 141-149 m D. M. Green, editor. Amphibians in decline:
Canadian studies of a global problem. Society for the Study of Amphibians and
Reptiles, St. Louis, Missouri, USA.

Bonin, J., M. Ouellet, J. Rodrigue, and J.-L. DesGranges. 19976. Measuring the health
of frogs in agricultural habitats subjected to pesticides. Pages 246-257 in D. M.
Green, editor. Amphibians in decline: Canadian studies of a global problem.
Society for the Study of Amphibians and Reptiles, St. Louis, Missouri, USA.

Boone, M. D., and R. D. Semlitsch. 2001. Interactions of an insecticide with larval


density and predation in experimental amphibian communities. Conservation
Biology \5:22%-23%.

Boutilier, R. G., D. F. Stiffler, and D. P. Toews. 1992. Exchange of respiratory gases,


ions, and water in amphibious and aquatic amphibians. Pages 81-124 m M. E.
Feder and W. W. Burggren, editors. Environmental physiology of the
amphibians. University of Chicago, Chicago, Illinois, USA.

Bowker, R. G., and M. H. Bowker. 1979. Abundance and distribution of anurans in a


Kenyan pond. Copeia \979.27%-2Z5.

Bradford, D. F., C. Swanson, and M. S. Gordon. 1992. Effects of low pH and aluminum
on two declining species of amphibians in the Sierra Nevada, California. Journal
ofHerpetology 26:369-377.

Brady, L. D., and R. A. Griffiths. 2000. Developmental responses to pond desiccation in


tadpoles of British anuran amphibians {Bufo bufo, B. calamita, and Rana
temporaria). Journal of Zoology {London) 252:61-69.

142
Bragg, A. N. 1965. Gnomes of the night: the spadefoot toads. University of
Pennsylvania, Philadelphia, Pennsylvania, USA.

Bridges, C. M. 1997. Tadpole swimming performance and activity by acute exposure to


sublethal levels of carbaryl. Environmental Toxicology and Chemistry
16:1935-1939.

Bridges, C. M., and R. D. Semlitsch. 2000. Variafion in pesficide tolerance of tadpoles


among and within species of Ranidae and pattems of amphibian decline.
Conservation Biology 14:1490-1499.

Brown, J. H., and A. Kodric-Brown. 1977. Tumover rates in insular biogeography:


effects of immigration on extinction. Ecology 58:445^49.

Bruce, R. C , and N. G. Hairston, Sr. 1990. Life-history correlates of body size


differences between two populations of the salamander, Desmognathus monticola.
Journal ofHerpetology 24:124-134.

Burel, F. 1989. Landscape structure effects on carabid beefies spatial pattems in westem
France. Landscape Ecology 2:215-226.

Camp, C. D., J. L. Marshall, and R. M. Ausfin, Jr. 2000. The evolution of adult body-
size in black-bellied salamanders {Desmognathus quadramaculatus). Canadian
Journal of Zoology 78:1712-1722.

Carey, C , N. Cohen, and L. Rollins-Smith. 1999. Amphibian declines: an


immunological perspective. Developmental and Comparative Immunology
23:459-472.

Case, T. J., and M. E. Gilpin. 1974. Interference competition and niche theory.
Proceedings of the National Academy of Sciences, U.S.A. 71:3073-3077.

Chesson, P., and N. Huntly. 1997. The roles of harsh and fluctuating conditions in the
dynamics of ecological communities. American Naturalist 150:519-553.

Clarke, R. D. 1974. Postmetamorphic growth rates in a natural population of Fowler's


toad, Bufo woodhousei fowleri. Canadian Journal of Zoology 52:1489-1498.

Cody, M. L. 1968. On the methods of resource division in grassland bird communities.


American Naturalist 102:107-147.

Collins, J. P., and J. R. Holomuzki. 1984. Intraspecific variation in diet within and
between trophic morphs in larval tiger salamanders {Ambystoma tigrinum
nebulosum). Canadian Journal of Zoology 62:168-174.

143
Connell, J. H. 1980. Diversity and coevolution of competitors, or the ghost of
competition past. Oikos 35:131-138.

Conover, W. J. 1980. Practical nonparametric statistics. Second edition. John Wiley


and Sons, New York, New York, USA.

Com, P. S., and R. B. Bury. 1989. Logging in westem Oregon: responses of headwater
habitats and stream amphibians. Forest Ecology and Management 29:39-57.

Creusere, F. M., and W. G. Whitford. 1976. Ecological relationships in a desert anuran


community. Herpetologica 32:7-18.

Crist, T. O., D. S. Guertin, J. A. Wiens, and B. T. Milne. 1992. Animal movement in


heterogeneous landscapes: an experiment with Eleodes beetles in shortgrass
prairie. Functional Ecology 6:536-544.

Crowson, R. A. 1981. The biology of the Coleoptera. Academic, London, UK.

Daoust, J.-L. 1991. Coping with dehydration of trapped terrestrial anurans.


Herpetological Review 22:95.

Dayton. G. H., and L. A. Fitzgerald. 2001. Competition, predation, and the distributions
of four desert anurans. Oecologia 129:430-435.

DeAngelis, D. L., and J. C. Waterhouse. 1987. Equilibrium and nonequilibrium


concepts in ecological models. Ecological Monographs 57:1-21.

Degenhardt, W. G., C. W. Painter, and A. H. Price. 1996. Amphibians and reptiles of


New Mexico. University of New Mexico, Albuquerque, New Mexico, USA.

deMaynadier, P. G., and M. L. Hunter, Jr. 1995. The relationship between forest
management and amphibian ecology: a review of the North American literature.
Environmental Reviews 3:230-261.

deMaynadier, P. G., and M. L. Hunter, Jr. 1998. Effects of silvicultural edges on the
distribution and abundance of amphibians in Maine. Conservation Biology
12:340-352.

Denver, R. J. 1997. Proximate mechanisms of phenotypic plasticity in amphibian


metamorphosis. American Zoologist 37:172-184.

Denver, R. J., N. Mirhadi, and M. Phillips. 1998. Adaptive plasficity in amphibian


metamorphosis: response of Scaphiopus hammondii tadpoles to habitat
desiccafion. Ecology 79:1859-1872.

144
Dimmitt, M. A. 1975. Terrestrial ecology of spadefoot toads {Scaphiopus): emergence
cues, nutrition, and burrowing habits. Dissertation, University of Califomia,
Riverside, Califomia, USA.

Dimmitt, M. A., and R. Ruibal. 1980a. Exploitafion of food resources by spadefoot


\02ids {Scaphiopus). Copeia 1980:854-862.

Dimmitt, M. A., and R. Ruibal. 19806. Environmental correlates of emergence in


spadefoot toads {Scaphiopus). Journal ofHerpetology 14:21-29.

Dodd, C. K., Jr. 1996. Use of terrestrial habitats by amphibians in the sandhill uplands
of north-central Florida. Alytes 14:42-52.

Dodd, C. K., Jr. 1997. Imperiled amphibians: a historical perspective. Pages 163-200 in
G. W. Benz and D. E. Collins, editors. Aquatic fauna in peril: the southeastern
perspective. Special Publication 1, Southeast Aquatic Research Institute, Lenz
Design and Communications, Decatur, Georgia, USA.

Dodd, C. K., Jr., and D. E. Scott. 1994. Drift fences encircling breeding sites. Pages
125-130 in W. R. Heyer. M. A. Donnelly. R. W. McDiarmid. L.-A. C. Hayek.
and M. S. Foster, editors. Measuring and monitoring biological diversity:
standard methods for amphibians. Smithsonian Institution. Washington, D. C.
USA.

Driscoll, D. A. 1997. Mobility and metapopulation stmcture of Geocrinia alba and


Geocrinia vitellina, two endangered frog species from southwestern Australia.
Australian Journal of Ecology 22:185-195.

Duellman, W. E., and L. Tmeb. 1994. Biology of amphibians. Johns Hopkins


University, Baltimore, Maryland, USA.

Durelli. P., M. Studer. I. Marchand, and S. Jacob. 1990. Population movements of


arthropods between natural and cultivated areas. Biological Conservation
54:193-207.

Edelstein-Keshet, L. 1988. Mathematical models in biology. McGraw-Hill, New York.


New York, USA.

Elkinton, J. S. 2000. Detecting stability and causes of change in populafion density.


Pages 191-212 m L. Boitani and T. K. Fuller, editors. Research techniques in
animal ecology. Columbia University, New York. New York, USA.

145
Ellner, S. P. 2000. Defining chaos for real, noisy data: local Lyapunov exponents and
sensitive response to perturbations. Pages 1-32 in J. N. Perry. R. H. Smith, I. P.
Woiwod, and D. R. Morse, editors. Chaos in real data: the analysis of non-linear
dynamics from short ecological time series. Kluwer Academic Publishers.
Dordrecht, Netherlands.

Ellner, S. P., and P. Turchin. 1995. Chaos in a noisy world: new methods and evidence
from time-series analysis. American Naturalist 145:343-374.

Fahrig. L., and G. Merriam. 1985. Habitat patch connectivity and population survival.
Ecolog}-66:\762-\76^.

Fahrig. L.. and G. Merriam. 1994. Conservation of fragmented populations.


Conservation Biology 8:50-59.

Fahrig. L., and J. Paloheimo. 1988. Determinants of local population size in patchy
habitats. Theoretical Population Biology 34:194-212.

Findlay, C. S., and J. Houlahan. 1997. Anthropogenic correlates of species richness in


southeastem Ontario wetlands. Conservation Biology 11:1000-1009.

Fioramonti, E., R. D. Semlitsch, H.-U. Reyer, and K. Fent. 1997. Effects of triphenyltin
and pH on the growth and development of Rana lessonae and Rana esculenta
tadpoles. Environmental Toxicology and Chemistry 16:1940-1947.

Flowers, M. A., and B. M. Graves. 1995. Prey selecti^ity and size-specific diet changes
in Bufo cognatus and B. woodhousii during earh- postmetamorphic ontogen\.
Journal ofHerpetology 29:608-612.

Fontenot, C. L., Jr. 1999. Reproductive biology of the aquatic salamander yimp/imma
tridactylum in Louisiana. Journal ofHerpetology 33:100-105.

Fomianowicz, D. R.. Jr., and E. D. Brodie, Jr. 1982. Relati\e palatability of members of
the larval amphibian community. Copeia 1982:91-97.

Eraser, D. F. 1976a. Coexistence of salamanders in the genus Plethodon: a \ariation of


the Santa Rosalia theme. Ecology 57:238-251.

Eraser. D. F. 19766. Empirical evidence of the hypothesis of food competition in


salamanders of the genus, Plethodon. Ecology 57:459—471.

146
Freemark, K., and C. Boutin. 1995. Impacts of agricultural herbicide use on terrestrial
wildlife in temperate landscapes: a re\ iew with special reference to North
America. Agriculture, Ecosystems and Environment 52:67-91.

Freund, R. J, and R. C. Littell. 2000. SAS® system for regression. Third edition. SAS
Institute, Gary. North Carolina, USA.

Gabor, C. R. 1995. Correlational test of Mathis' hypothesis that bigger salamanders


have better territories. Copeia 1995:729-735.

Gause, G. F. 1934. The struggle for existence. Williams and Wilkins, Baltimore,
Maryland, USA.

Gehlbach, F. R. 1967. Evolution of the tiger salamander {Ambystoma tigrinum) on the


Grand Canyon rims. Arizona. Yearbook of the American Philosophical Society
1967:266-269.

Gehlbach, F. R., J. R. Kimmel, and W. A. Weems. 1969. Aggregations and body water
relations in tiger salamanders {Ambystoma tigrinum) from the Grand Canyon
rims, Arizona. Physiological Zoology 42:173-182.

Gibbs, J. P. 1998a. Distribution of woodland amphibians along a forest fragmentation


gradient. Landscape Ecology 13:263-268.

Gibbs. J. P. 19986. Amphibian movements in response to forest edges, roads, and


streambeds in southem New England. Journal of Wildlife Management
62:584-589.

Goater, C. P. 1994. Growth and survival of postmetamorphic toads: interactions among


larval history, density, and parasitism. Ecology 75:2264-2274.

Goater, C. P., R. D. Semlitsch, and M. V. Bemasconi. 1993. Effects of body size and
parasite infection on the locomotion performance of juvenile toads, Bufo bufo.
Oikos 66:129-136.

Gomulkiewicz, R., and M. Kirkpatrick. 1992. Quantitative genetics and the evolution of
reaction norms. Evolution 46:390-411.

Gro\'er, M. C. 1998. Infiuence of cover and moisture on abundances of the terrestrial


salamanders Plethodon cinereus and Plethodon glutinosus. Journal of
Herpetology 32:489-497.

147
Gustavson, T. C , V. T. Holliday, and S. D. Hovorka. 1994. Development of playa
basins, Southem High Plains, Texas and New Mexico. Pages 5-14 in L. V.
Urban and A. W. Wyatt, editors. Proceedings of the playa basin symposium.
Texas Tech University, Lubbock, Texas, USA.

Guthery, F. S., and F. C. Bryant. 1982. Status of playas in the Southem Great Plains.
Wildlife Society Bulletin 10:309-317.

Hair, J. D. 1980. Measurement of ecological diversity. Pages 269-275 in S. D.


Schemnitz, editor. Wildlife management techniques manual. Third edition. The
Wildlife Society, Bethesda, Maryland, USA.

Hall, D. L. 1997. Species diversity of aquatic macroinvertebrates in playa lakes: island


biogeographic and landscape influences. Thesis, Texas Tech University,
Lubbock, USA.

Hall, R. J., and E. Kolbe. 1980. Bioconcentration of organophosphoms pesticides to


hazardous levels by amphibians. Journal of Toxicology and Environmental
Health 6:853-860.

Hanski, I. A. 1997. Metapopulation dynamics: from concepts and observations to


predictive models. Pages 69-91 in I. A. Hanski and M. E. Gilpin, editors.
Metapopulation biology: ecology, genetics, and evolution. Academic, San Diego,
California, USA.

Hanson, G. C , P. M. Groffman, and A. J. Gold. 1994. Symptoms of nitrogen saturation


in a riparian wetland. Ecological Applications 4:750-756.

Hardin, G. 1960. Competitive exclusion principle. Science \31:1292-1297.

Hastings, A. 1996. Population biology: concepts and models. Springer-Verlag, New


York, York, USA.

Hastings, A., and S. Harrison. 1994. Metapopulation dynamics and genetics. Annual
Review of Ecology and Systematics 25:167-188.

Hastings, A., C. L. Hom, S. Ellner, P. Turchin, and H. C. J. Godfray. 1993. Chaos in


ecology. Annual Review of Ecology and Systematics 24:1-33.

Haukos, D. A., and L. M. Smith. 1994. The importance of playa wetlands to biodiversity
of the Southem High Plains. Landscape and Urban Planning 28:83-98.

148
Hecnar, S. J. 1995. Acute and chronic toxicity of ammonium nitrate fertilizer to
amphibians from southem Ontario. Environmental Toxicology and Chemistry
14:2131-2137.

Hecnar, S. J., and R. T. M'Closkey. 1996. Regional dynamics and the status of
amphibians. Ecology 77:2091-2097.

Hecnar, S. J., and R. T. M'Closkey. 1997. Pattems of nestedness and species association
in a pond-dwelling amphibian fauna. Oikos 80:371-381.

Hecnar, S. J., and R. T. M'Closkey. 1998. Species richness pattems of amphibians in


southwestem Ontario ponds. Journal of Biogeography 25:763-772.

Henrickson, B.-I. 1990. Predation on amphibian eggs and tadpoles by common


predators in acidified lakes. Holarctic Ecology 13:201-206.

Henttonen, H., J. Tast, J. Viitala, and A. Kaikusalo. 1984. Ecology of cyclic rodents in
northem Finland. Memoranda Societatis pro Fauna et Flora Fennica 60:84-92.

Herbeck, L. A., and D. R. Larsen. 1999. Plethodontid salamander response to


silvicultural practices in Missouri Ozark forests. Conservation Biology
13:623-632.

Hess, G. 1996. Linking extinction to connectivity and habitat destmction in


metapopulation models. American Naturalist 148:226-236.

Heyer, R. W., R. W. McDiarmid, and D. L. Weigmann. 1975. Tadpoles, predation and


pond habitats in the tropics. Biotropica 7:100- 111.

Heyer. W. R., M. A. Donnelly, R. W. McDiarmid, L.-A. C. Hayek, and M. S. Foster.


editors. 1994. Measuring and monitoring biological diversity: standard methods
for amphibians. Smithsonian Institution, Washington, D.C., USA.

Hocking, R. R. 1973. A discussion of the two-way mixed model. The American


Statistician 27:148-152.

Hoff, vS., A. R. Blaustein, R. W. McDiarmid, and R. Altig, 1999. Behavior: interactions


and their consequences. Pages 215-239 in R. W. McDiarmid and R. Altig,
editors. Tadpoles: the biology of anuran larvae. University of Chicago, Chicago,
Illinois, USA.

Holt, R. D. 1977. Predation, apparent competition, and the stmcture of prey


communities. Theoretical Population Biology 12:337-406.

149
Houlahan, J. E., C. S. Findlay, B. R. Schmidt, A. H. Meyers, and S. L. Kuzmin. 2000.
Quantitative evidence for global amphibian population declines. Nature
404:752-755.

Howard, R. D. 1978. The influence of male-defended oviposition sites on early embryo


mortality in bullfrogs. Ecology 59:789-798.

Howard, R. D., and J. R. Young. 1998. Individual variation in male vocal traits and
female mating preference in Bufo americanus. Animal Behaviour 55:1165-1179.

Huey, R. B., and E. R. Pianka. 1983. Temporal separafion of acfivity and interspecific
dietary overlap. Pages 281-290 in R. B. Huey, E. R. Pianka, and T. W. Schoener.
editors. Lizard ecology: studies of a model organism. Harvard University,
Cambridge, Massachusetts, USA.

Hurlbert, S. H. 1969. The breeding migrations and interhabitat wandering of vermilion-


spotted newt Notophthalmus viridescens (Rafinesque). Ecological Monographs
39:465-488.

Hurlbert, S. H. 1978. The measurement of niche overlap and some relatives. Ecology
59:67-77.

Hurlbert, S. H. 1984. Pseudoreplication and the design of ecological field experiments.


Ecological Monographs 54:187-211.

Huston, M. 1979. A general hypothesis of species diversity. American Naturalist


113:81-101.

Hutchinson, G. E. 1957. Concluding remarks. Cold Spring Harbor Symposia on


Quantitative Biology 22:415^27.

Hutchison, V. H., W. G. Whitford, and M. Kohl. 1968. Relation of body size and
surface area to gas exchange in anurans. Physiological Zoology 41:65-85.

Jaeger, R. G. 1980. Fluctuations in prey availability and food limitation for a terrestrial
salamander. Oecologia 44:335-34\.

Jansen, K. P., A. P. Summers, and P. R. Delis. 2001. Spadefoot toads {Scaphiopus


holbrookii holbrookii) in an urban landscape: effects of nonnatural substrates on
burrowing in adults and juveniles. Journal ofHerpetology 35:141-145.

John-Alder, H. B., and P. J. Morin. 1990. Effects of larval density on jumping ability
and stamina in newly metamorphosed Bufo woodhousii fowleri. Copeia
1990:856-860.

150
Johnson, R. A., and D. W. Wichern. 1998. Applied multivariate statistical analysis.
Fourth edition. Prentice Hall, Upper Saddle River. New Jersey, USA.

Jung, R. E. 1993. Blanchard's cricket frogs {Acris crepitans blanchardi) in southwest


Wisconsin. Transactions of the Wisconsin Academy of Sciences, Arts and Letters
81:79-87.

Knutson, M. G., J. R. Sauer, D. A. Olsen, M. J. Mossman, L. M. Hemesath, and M. J.


Lannoo. 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 13:1437-1446.

Kolozsvary, M. B., and R. K. Swihart. 1999. Habitat fragmentation and the distribution
of amphibians: patch and landscape correlates in farmland. Canadian Journal of
Zoology 77:\2SS-1299.

Krupa, J. J. 1986. Multiple egg clutch production in the Great Plains toad. Prairie
Naturalist \S:\5\-\52.

Krupa, J. J. 1994. Breeding biology of the Great Plains toad in Oklahoma. Journal of
Herpetology 28:217-224.

Kmse, K. C , and M. G. Francis. 1977. A predation deterrent in larvae of the bullfrog,


Rana catesbeiana. Transactions of the American Fisheries Society 106:248-252.

Laan, R., and B. Verboom. 1990. Effects of pool size and isolation on amphibian
communities. Biological Conservation 54:251-262.

Laimoo, M. J. 1998. Amphibian conservation and wetland management in the upper


Midwest: a catch-22 for the cricket frog? Pages 330-339 in M. J. Lannoo, editor.
Status and conservation of Midwestern amphibians. University of Iowa, Iowa
City, Iowa, USA.

Lefkovitch, L. P., and L. Fahrig. 1985. Spafial characteristics of habitat patches and
population survival. Ecological Modelling 30:297-308.

Lehtinen, R. M., S. M. Galatowitsch, and J. R. Tester. 1999. Consequences of habitat


loss and fragmentafion for wetland amphibian assemblages. Wetlands 19:1-12.
Leibold, M. A., and H. M. Wilbur. 1992. Interactions between food-web stmcture and
nutrients on pond organisms. Nature 360:341-343.

Littell, R. C , R. J. Freund, and R. C. Spector. 1991. SAS® system for linear models.
Third edifion. SAS Institute, Gary, North Carolina, USA.

151
Littell, R. C , G. A. Milliken, W. W. Stroup, and R. D. Wolfinger. 1996. SAS® system
for mixed models. SAS Institute, Gary, North Carolina, USA.

Loman, J. 1988. Breeding by Rana temporaria; the importance of pond size and
isolation. Memoranda Societatis pro Fauna et Flora Fennica 64:113-115.

Loreau, M. 2000. Biodiversity and ecosystem functioning: recent theoretical advances.


Oikos 9\:3-\7.

Loreau, M., and A. Hector. 2001. Partitioning selection and complementarity in


biodiversity experiments. Nature 412:72-76.

Lotka, A. J. 1925. Elements of physical biology. Williams and Wilkins, Balfimore,


Maryland, USA.

Luo, H.-R., L. M. Smith, B. L. Allen, and D. A. Haukos. 1997. Effects of sedimentafion


on playa wetland volume. Ecological Applications 7:247-252.

MacArthur, R. H., and R. Levins. 1967. The limiting similarity, convergence, and
divergence of coexisting species. American Naturalist 101:337-385.

Marco, A., J. M. Kiesecker, D. P. Chivers, and A. R. Blaustein. 1998. Sex recognition


and mate choice by male westem toads, Bufo boreas. Animal Behaviour
55:1631-1635.

Marsh, D. M., and P. B. Pearman. 1997. Effects of habitat fragmentafion on the


abundance of two species of Leptodactylid frogs in an Andean montane forest.
Conservation Biology 11:1323-1328.

Marsh, D. M., and P. C. Trenham. 2001. Metapopulation dynamics and amphibian


conservation. Conservation Biology 15:40-49.

Martin, D. B., and W. A. Hartman. 1987. The effect of cultivation on sediment


composition and deposition in prairie potholes. Water, Air, and Soil Pollution
34:45-53.

Mathis, A. 1990. Territoriality in a terrestrial salamander: the influence of resource


quality and body size. Behaviour 112:162-175.

May, R. M. 1974. Biological populafions with nonoverlapping generafions: stable


points, stable cycles, and chaos. Science 186:645-647.

May, R. M. 1975a. Deterministic models with chaotic dynamics. Nature 256:\65-\66.

152
May, R. M. 19756. Biological populafions obeying difference equations: stable points,
stable cycles, and chaos. Journal of Theoretical Biology 51:511-524.

May. R. M. 1976a. Simple mathematical models with very complicated dynamics.


Nature 261:459-466.

May, R. M. 19766. Theoretical ecology: principles and applications. Saunders,


Philadelphia, Pennsylvania, USA.

May, R. M., and R. H. MacArthur. 1972. Niche overlap as a ftjnction of environmental


variability. Proceedings of the National Academy of Sciences, U.S.A. 69:1109-
1113.

Maynard Smith, J. 1974. Models in ecology. Cambridge University. Cambridge,


England.

McClanahan, L. L., R. Ruibal, and V. H. Shoemaker. 1994. Frogs and toads in deserts.
Scientific American 270:64-70.

McGarigal, K., and B. J. Marks. 1995. FRAGSTATS: spatial pattem analysis program
for quantifying landscape structure. U.S. Forest Service General Technical
Report PNW-GTR-351.

Mclntyre, N. E. 2000. Community structure of Eleodes beetles (Coleoptera:


Tenebrionidae) in the shortgrass steppe: scale-dependent uses of heterogeneity.
Western North American Naturalist 60:1-15.

Mecham, J. S., M. J. Littlejohn, R. S. Oldham, L. E. Brown, arid J. R. Brown. 1973. A


new species of leopard frog {Rana pipiens complex) from the plains of the central
United States. Texas Tech University Museum Occasional Paper 18:1-11.

Meyer, A. H., B. R. Schmidt, and K. Gossenbacher. 1998. Analysis of three amphibian


populations with quarter-century long time-series. Proceedings of the Royal
Society of London, Series B Biological Sciences 265:523-528.

Miaud, C , D. Sanuy, and J.-N. Avrillier. 2000. Terrestrial movements of the natterjack
toad Bufo calamita (Amphibia, Anura) in a semi-arid, agricultural landscape.
Amphibia-Reptilia 21:357-369.

Milliken, G. A., and D. E. Johnson. 1992. Analysis of messy data. Chapman and Hall,
New York, New York, USA.

Milton, J. S., and J. C. Amold. 1995. Introduction to probability and statistics. Third
edition. McGraw-Hill, New York, New York, USA.

153
Montgomery, D. C. 2001. Design and analysis of experiments. Fifth edifion. John
Wiley and Sons, New York, New York, USA.

Morey, S., and D. Reznick. 2000. A comparative analysis of plasticity in larval


development in three species of spadefoot toads. Ecology 81:1736-1749.

Morey, S., and D. Reznick. 2001. Effects of larval density on postmetamorphic


spadefoot toads {Spea hammondii). Ecology 82:510-522.

Morin, P. J. 1981. Predatory salamanders reverse the outcome of competition among


three species of anuran tadpoles. Science 212:1284-1286.

Morin, P. J. 1983. Predation, competition, and the composition of larval anuran guilds.
Ecological Monographs 53:119-138.

Morin. P. J., S. P. Lawler, and E. A. Johnson. 1988. Compefition between aquatic


insects and vertebrates: interaction strength and higher order interactions.
Ecology 69:\40\-\409.

Murphy, T. D. 1963. Amphibian populations and movements at a small semi-permanent


pond in Orange County, North Carolina. Dissertation, Duke University, Durham.
North Carolina, USA.

Myers, R. H. 1990. Classical and modern regression with applications. Second edition.
PWS-KENT, Boston, Massachusetts, USA.

Naughton, G. P., C. B. Henderson, K. R. Foresman, and R. L. McGraw, II. 2000. Long-


toed salamanders in harvested and intact Douglas-fir forests of westem Montana.
Ecological Applications 10:1681-1689.

Newman, R. A. 1988. Adaptive plasticity in development of Scaphiopus couchii


tadpoles in desert ponds. Evolution 42:774-783.

Newman, R. A. 1989. Developmental plasticity of Scaphiopus couchii tadpoles in an


unpredictable environment. Ecology 70:1775-1787.

Newman, R. A. 1992. Adaptive plasficity in amphibian metamorphosis. BioScience


42:671-678.

Newman, R. A. 1994a. Genefic variafion for phenotypic plasficity in the larval life
history of spadefoot toads {Scaphiopus couchii). Evolution 48:1773-1785.

154
Newman, R. A. 19946. Effects of changing density and food level on metamorphosis of
a desert amphibian, Scaphiopus couchii. Ecology 75:1085-1096.

Newman, R. A. 1999. Body size and diet of recently metamorphosed spadefoot toads
{Scaphiopus couchii). Herpetologica 55:507-515.

Newman, R. A., and A. E. Dunham. 1994. Size at metamorphosis and water loss in a
desert anuran {Scaphiopus couchii). Copeia 1994:372-381.

O'Neill, R. v., B. T. Milne, M. G. Turner, and R. H. Gardner. 1988. Resource ufilizafion


scales and landscape pattern. Landscape Ecology 2:63-69.

Oldham, R. S. 1985. Toad dispersal in agricultural habitats. Bulletin of the British


Ecological Society 16:211-215.

Oldham, R. S., and M. J. S. Swan. 1991. Conservation of amphibians populations in


Britain. Pages 141-157 in A. Seitz and V. Loeschcke, editors. Species
conservation: a population-biological approach. Birkhauser Verlag, Basel,
Switzerland.

Olsen, L. F., and H. Degn. 1985. Chaos in biological systems. Quarterly Review of
Biophysics 18:165-225.

Olson, C. L. 1974. Comparative robustness of six tests in multivariate analysis of


variance. Journal of the American Statistical Association 69:894-908.

Olson, C. L. 1976. On choosing a test statistic in multivariate analysis of variance.


Psychological Bulletin 83:579-586.

Olson, C. L. 1979. Practical considerations in choosing a MANOVA test statistic: a


rejoinder to Stevens. Psychological Bulletin 86:1350-1352.

Osterkamp, W. R., and W. W. Wood. 1987. Playa-lake basins on the Southem High
Plains of Texas and New Mexico. Parti. Hydrologic, geomorphic, and geologic
evidence for their development. Geological Society of America Bulletin
99:215-223.

Paine, R. T. 1969. A note on trophic complexity and community stability. American


Naturalist 103:91-93.

Paton, P. W., and W. B. Crouch III. 2002. Using the phenology of pond-breeding
amphibians to develop conservation strategies. Conservation Biology
16:194-204.

155
Pechmann, J. H. K. 1994. Populafion regulation in complex life cycles: aquatic and
terrestrial density-dependence in pond-breeding amphibians. Dissertation, Duke
University, Durham, North Carolina, USA.

Pechmann, J. H. K., D. E. Scott, J. W. Gibbons, and R. D. Semlitsch. 1989. Influence of


wetland hydroperiod on diversity and abundance of metamorphosing juvenile
amphibians. Wetlands Ecology and Management 1:3-11.

Peters, R. H. 1983. The ecological implications of body size. Cambridge University.


Cambridge, UK.

Peterson, J. A., and A. R. Blaustein. 1991. Unpalatability in anuran larvae as a defense


against natural salamander predators. Ethology Ecology and Evolution 3:63-72.

Petranka, J. W. 1998. Salamanders of the United States and Canada. Smithsonian


Institufion, Washington, D.C., USA.

Petranka. J. W., and C. A. Kennedy. 1999. Pond tadpoles with generalized morphology:
is it time to reconsider their functional roles in aquatic communities? Oecologia
120:621-631.

Petranka, J. W., M. E. Eldridge, and K. E. Haley. 1993. Effects of timber harvesting on


southem Appalachian salamanders. Conservation Biology 7: 363-370.

Pfennig, D. W. 1992. Proximate and functional causes of polyphenism in an anuran


tadpole. Functional Ecology 6:167-174.

Pfennig, D. W., and P. J. Murphy. 2000. Character displacement in polyphenic tadpoles.


Evolution 54:\73S-\749.

Pfennig, D. W., A. Mabry, and D. Orange. 1991. Environmental causes of correlations


between age and size at metamorphosis in Scaphiopus multiplicatus. Ecology
72:2240-2248.

Pianka, E. R. 1966. Latitudinal gradients in species diversity: A review of concepts.


American Naturalist 100:33-46.

Pianka, E. R. 1974. Niche overlap and diffuse competition. Proceedings of the National
Academy of Sciences, U.S.A. 71:2141-2145.

Pianka, E. R. 1976. Competition and niche theory. Pages 114-141 in R. M. May, editor.
Theoretical ecology: principles and applications. W. B. Saunders, Philadelphia,
Pennsylvania, USA.

156
Pianka, E.R. 1994. Evolutionary ecology. Fifth edition. HarperCollins, New York,
New York, USA.

Polls, G. A., and R. D. Holt. 1992. Intraguild predafion: the dynamics of complex
trophic interactions. Trends in Ecology and Evolution 7:151-154.

Pope, S. E., L. Fahrig, and G. Merriam. 2000. Landscape complementation and


metapopulation effects on leopard frog populations. Ecology 80:2326-2337.

Pulliam, H. R. 1988. Sources, sinks, and population regulation. American Naturalist


132:652-661.

Pulliam, H. R. 1996. Sources and sinks: empirical evidence and population


consequences. Pages 45-69 in O. E. Rhodes, Jr., R. K. Chesser, and M. H. Smith,
editors. Population dynamics in ecological space and time. University of
Chicago, Chicago, Illinois, USA.

Reading, C. J. 1990. A comparison of size and body weights of common toads {Bufo
bufo) from two sites in Southem England. Amphibia-Reptilia 11:155-163.

Reading, C. J., and R. T. Clarke. 1995. The effects of density, rainfall and
environmental temperature on body condition and fecundity in the common toad,
Bufo bufo. Oecologia 102:453-459.

Reading, C. J., and R. T. Clarke. 1999. Impacts of climate and density on the duration of
the tadpole stage of the common toad Bufo bufo. Oecologia 121:310-315.

Reh, W., and A. Seitz. 1990. The influence of land use on the genetic stmcture of
populations of the common frog Rana temporaria. Biological Conservation
54:239-249.

Relyea, R. A., and N. Mills. 2001. Predator-induced stress makes the pesticide carbaryl
more deadly to gray treefrog tadpoles {Hyla versicolor). Proceedings of the
National Academy of Sciences, U.S.A. 98:2491-2496.

Richmond, N. D. 1947. Life history of Scaphiopus holbrookii holbrookii {HsLilan). Part


I: larval development and behavior. Ecology 28:53-67.

Ritchie, M. E. 1997. Populations in a landscape context: sources, sinks, and


metapopulations. Pages 160-184 in J. A. Bissonette, editor. Wildlife and
landscape ecology: effects of pattern and scale. Springer, New York, New York,
USA.

157
Rohde, K., and P. P. Rohde. 2001. Fuzzy chaos: reduced chaos in combined dynamics
of several independently chaotic populations. American Naturalist 158:553-556.

Rose, F. L., and D. Armentrout. 1974. Populafion esfimates of Ambystoma tigrinum


inhabiting two playa lakes. Journal of Animal Ecology 43:671-679.

Rose, F. L., and D. Armentrout. 1976. Adaptive strategies of Ambystoma tigrinum


Green inhabiting the Llano Estacado of West Texas. Journal of Animal Ecology
45:713-729.

Rosenzweig, M. L. 1995. Species diversity in space and time. Cambridge University,


Cambridge, UK.

Roughgarden, J. 1998. Primer of ecological theory. Prentice Hall, Upper Saddle River,
New Jersey, USA.

Sabin, T. J., and V. T. Holliday. 1995. Playas and lunettes on the Southem High Plains:
morphometric and spatial relationships. Annals of the Association of American
Geographers 85:286-305.

Savage, R. M. 1952. Ecological, physiological, and anatomical observations of some


species of anuran tadpoles. Proceedings of the Zoological Society of London
122:467-514.

Schaffer, W. M., and M. Kot. 1986. Chaos in ecological systems: the coals that
Newcastle forgot. Trends in Ecology and Evolution 1:5 8-63.

Schneider, D. W., and T. M. Frost. 1996. Habitat duration and community stmcture in
temporary ponds. Journal of the North American Benthological Society
15:64-86.

Schoener, T. W. 1974. Resource partitioning in ecological communities. Science


185:27-39.

Schoener, T. W. 1983. Population and community ecology. Pages 233-239 in R. B.


Huey, E. R. Pianka, and T. W- Schoener, editors. Lizard ecology: studies of a
model organism. Harvard University, Cambridge, Massachusetts, USA.

Scribner, K. T., J. W. Amtzen, N. Cruddace, R. S. Oldham, and T. Burke. 2001.


Environmental correlates of toad abundance and population genetic diversity.
Biological Conservation 98:201-210.

Scott, D. E. 1994. The effects of larval density on adult demographic traits in


Ambystoma opacum. Ecology 75:1383-1396.

158
Scale. D. B. 1980. Influence of amphibian larvae on primar\ production, nutrient flux,
and competition in a pond ecosystem. £co/6>g>' 61:1531-1550.

Semlitsch, R. D. 1985a. Reproducti\e strategy of a facultatively paedomorphic


salamander Ambystoma talpoideum. Oecologia 65:305-313.

Semlitsch. R. D. 19856. Analysis of climatic factors influencing migrations of the


salamander Ambystoma talpoideum. Copeia 1985:477-489.

Semlitsch, R. D. 1987a. Density-dependent growth and fecundity in the paedomorphic


salamander Ambystoma talpoideum. Ecology 68:1003-1008.

Semlitsch, R. D. 19876. Relationship of pond drying to the reproducti\e success of the


salamander Ambystoma talpoideum. Copeia 1987:61-69.

Semlitsch, R. D. 2000. Principles for management of aquatic-breeding amphibians.


Journal of Wildlife Management 64:615-631.

Semlitsch, R. D., D. E. Scott, and J. H. K. Pechmann. 1988. Time and size at


metamorphosis related to adult fitness in Ambystoma talpoideum. Ecology
69:184-192.

Semlitsch, R. D., M. Foglia, A. Mueller, I. Steiner, E. Fioramonfi, and K. Fent. 1995.


Short-term exposure to triphenyltin affects the swimming and feeding behavior of
tadpoles. Environmental Toxicology and Chemistry 14:1419-1423.

Semlitsch, R. D., D. E. Scott, J. H. K Pechmann, and J. W. Gibbons. 1996. Stmcture and


dynamics of an amphibian community: evidence from a 16-year study of natural
ponds. Pages 217-248 in M. L. Cody and J. A. Smallwood, editors. Long-term
studies of vertebrate communities. Academic, San Diego, Califomia, USA.

Shoemaker, V. H., S. S. Hillman, S. D. Hillyard, D. C. Jackson, L. L. McClanahan, P. C.


Withers, and M. L. Wygoda. 1992. Exchange of water, ions, and respirator)
gases in terrestrial amphibians. Pages 125-150 in M. E. Feder and W. W.
Burggren, editors. Environmental physiology of the amphibians. Uni\'ersity of
Chicago, Chicago, Illinois, USA.

Shulenburger, L., Y-C. Lai, T. Yalcinkaya, and R. D. Holt. 1999. Controlling transient
chaos to prevent species extinction. Physics Letters A 260:156-161.

Sinsch, U. 1988. Seasonal changes in the migratory behaviour of the toad Bufo bufo:
direction and magnitude of movements. Oecologia 76:390-398.

159
Sinsch, U. 1990. Migrafion and orientafion in anuran amphibians. Ethology Ecology
and Evolution 2:65-79.

Sinsch, U. 1997. Postmetamorphic dispersal and recmitment of first breeders in a Bufo


ca/am/Ya metapopulafion. Oecologia 112:42-47.

Sinsch, U., and D. Seidel. 1995. Dynamics of local and temporal breeding assemblages
in a Bufo calamita metapopulation. Australian Journal of Ecology 20:351-361.

Sjogren, P. 1991. Extinction and isolation gradients in metapopulations: the case of the
pool frog {Rana lessonae). Biological Journal of the Linnean Society
42:135-147-

Sjogren-Gulve, P. 1994. Distribution and extinction pattems within a northem


metapopulation of the pool frog, Rana lessonae. Ecology 75:1357-1367.

Skelly, D. K. 1996. Pond drying, predators, and the distribution of Pseudacris tadpoles.
Copeia 1996:599-605.

Skelly, D. K. 1997. Tadpole communities. American Scientist S5:36-45.

Skelly, D. K., and E. E. Wemer. 1990. Behavioral and life-historical responses of larval
American toads to an odonate predator. Ecology 71:2313-2322.

Smith, D. C. 1987. Adult recmitment in choms frogs: effects of size and date at
metamorphosis. Ecology 6^:344-350.

Smith, D. C , and J. Van Burskirk. 1995. Phenotypic design, plasticity, and ecological
performance in two tadpole species. American Naturalist 145:211-233.

Smith, L. M., and D. A. Haukos. 2002. Floral diversity in relation to playa wetland area
and watershed disturbance. Conservation Biology 16:in press.

Smith-Gill, S. J., and K. A. Berven. 1979. Predicting amphibian metamorphosis.


American Naturalist 113:563-585.

Snodgrass, J. W., M. J. Komoroski, A. L. Bryan, Jr., and J. Burger. 2000. Relationships


among isolated wetland size, hydroperiod, and amphibian species richness:
implications for wetland regulations. Conservation Biology 14:414-419.

160
Stacey, P. B., V. A. Johnson, and M. L. Taper. 1997. Migration within metapopulations:
the impact upon local populafion dynamics. Pages 267-291 in I. A. Hanski and
M. E. Gilpin, editors. Metapopulation biology: ecology, genetics, and evolution.
Academic Press, San Diego, Califomia, USA.

Stamps, J. A., M. Buechner, and V. V. Krishnan. 1987. The effects of edge permeability
and habitat geometry on emigration from patches of habitat. American Naturalist
129:533-552.

Steinwascher, K. 1979. Competitive interactions among tadpoles: responses to resource


level. Ecology 60:ll72-l\S3.

Stokes, M. E., C. S. Davis, and G. G. Koch. 2000. Second edifion. Categorical data
analysis using the SAS® system. SAS Institute, Gary, North Carolina, USA.

Strong, D. R., Jr. 1983. Natural variability and the manifold mechanisms of
ecological communities. American Naturalist 122:636-660.

Sullivan, B. K., and P. J. Femandez. 1999. Breeding activity, estimated age-structure,


and growth in Sonoran Desert anurans. Herpetologica 55:334-343.

Szacki, J. 1999. Spatially stmctured populations: how much do they match the classic
metapopulation concept? Landscape Ecology 14:369-379.

Taigen, T. L., and F. H. Pough. 1981. Activity metabolism of the toad {Bufo
americanus): ecological consequences of ontogenetic change. Journal of
Comparative Physiology B 144:247-252.

Taylor, P. D., L. Fahrig, K. Henein, and G. Merriam. 1993. Connectivity is a vital


element of landscape stmcture. Oikos 68:571-573.

ter Braak, C. J. F. 1986. Canonical correspondence analysis: a new eigenvector


technique for multivariate direct gradient analysis. Ecology 67:1167-1179.
ter Braak, C. J. F. 1994. Canonical community ordination. Part I: basic theory and
linear methods. Ecoscience 1:127-140.

ter Braak, C. J. F. 1995. Ordinafion. Pages 91-173 in R. H. G. Jongman, C. J. F. ter


Braak, and O. F. R. Van Tongeren, editors. Data analysis in community and
landscape ecology. Cambridge University, Cambridge, UK.

ter Braak, C. J. F., and P. Smilauer. 1998. CANOCO reference manual and user's guide
to Canocofor windows: software for canonical community ordination (version 4).
Microcomputer Power, Ithaca, New York, USA.

161
Tinsley, R. C , and K. Tocque. 1995. The populafion dynamics of a desert anuran.
Scaphiopus couchii. Australian Journal of Ecology 20:376-384.

Toft, C. A. 1985. Resource partitioning in amphibians and reptiles. Copeia 1985:1-21.

Toft, C. A., and W. E. Duellman. 1979. Anurans of the lower Rio LluUapichis,
Amazonian Peru: a preliminary analysis of community stmcture. Herpetologica
35:71-77.

Travis, J. 1984. Anuran size at metamorphosis: experimental test of a model based on


intraspecific competition. Ecology 65:1155-1160.

Turchin, P. 1990. Rarity of density dependence or population regulation with lags?


Nature 344:660-663.

Turchin, P. 1991. Nonlinear modeling of time-series data: limit cycles and chaos in
forest insects, voles, and epidemics. Pages 39-61 in J. A. Logan and F. P. Hain.
editors. Chaos and insect ecology. Virginia Polytechnic Institute and State
University, Virginia Experiment Station Information Series 91-3, Blacksburg,
Virginia, USA.

Turchin, P. 1993. Chaos and stability in rodent population dynamics: evidence from
non-linear time series analysis. Oikos 68:167-172.

Turchin, P. 1996. Nonlinear time-series modelling of vole population fluctuations.


Researches on Population Ecology 38:842-874.

Turchin, P. 1998. Quantitative analysis of movement: measuring and modeling


population redistribution in animals and plants. Sinauer, Sunderland,
Massachusetts, USA.

Turchin, P., and S. P. Ellner. 2000. Modelling time-series data. Pages 33-48 in J. N.
Perry, R. H. Smith, I. P. Woiwod, and D. R. Morse, editors. Chaos in real data:
the analysis of non-linear dynamics from short ecological time series. Kluwer
Academic Publishers, Dordrecht, Netherlands.

Turchin, P., and A. D. Taylor. 1992. Complex dynamics in ecological time series.
Ecology 73:2^9-305.

Turner, F. B. 1960. Postmetamorphic growth in anurans. American Midland Naturalist


64:327-338.

Tumer, M. G., R. H. Gardner, V. H. Dale, and R. V. O'Neill. 1989. Predicfing the


spread of disturbance across heterogeneous landscapes. Oikos 55:121 -129.

162
Ultsch, G. R. 1973. A theoretical and experimental investigation of the relationship
between metabolic rate, body size, and oxygen exchange capacity. Respiration
Physiology n:l43-l60.

Ultsch, G. R. 1974. Gas exchange and metabolism in the Sirenidae (Amphibia:


Caudata)—I. Oxygen consumption of submerged Sirenids as a function of body
size and respiratory surface area. Comparative Biochemistry and Physiology
47A:485-498.

Verrell, P. A. 1982. Male newts prefer large females as mates. Animal Behaviour
30:1254-1255.

Verrell, P. A. 1985. Male mate choice for large, fecund females in the red-spotted newt,
Notophthalmus viridescens: how is the size assessed? Herpetologica 41:3S2-3S6.

Via, S., and R. Lande. 1985. Genotype-environment interaction and the evolution of
phenotypic plasticity. Evolution 39:505-522.

Vos, C. C , and J. P. Chardon. 1997. Landscape resistance and dispersal in fragmented


populations: a case study of the tree frog {Hyla arborea) in an agricultural
landscape. Pages 19-26 in A. Cooper and J. Power, editors. Species dispersal
and land use processes. Proceedings of the sixth annual conference of the
Intemational Association for Landscape Ecology, UK region.

Vos, C. C , and A. H. P. Stumpel. 1995. Comparison of habitat-isolation parameters in


relation to fragmented distribution pattems in the tree frog {Hyla arborea).
Landscape Ecology 11:203-214.

Walls, S. C. 1998. Density dependence in a larval salamander: the effects of interference


and food limitafion. Copeia 1998:926-935.

Walls, S. C , and M. G. Williams. 2001. The effect of community composition on


persistence of prey with their predators in an assemblage of pond-breeding
amphibians. Oecologia 128:134-141.

Walters, B. 1975. Studies of interspecific predation within an amphibian community.


Journal ofHerpetology 9:267-279.

Webb, R. G., and W. L. Roueche. 1971. Life history aspects of the tiger salamander
{Ambystoma tigrinum mavortium) in the Chihuahuan Desert. Great Basin
Naturalist 31:193-212.

163
Welsh, H. W., Jr. 1990. Relictual amphibians and old-growth forests. Conservation
Biology 4:309-319.

Werner, E. E. 1986. Amphibian metamorphosis: growth rate, predation risk, and the
optimal size at transformation. American Naturalist 128:319-341.

Wemer, E. E. 1991. Nonlethal effects of a predator on competitive interactions between


two anuran larvae. Ecology 72:1709-1720.

Wemer, E. E., and B. R. Anholt. 1993. Ecological consequences of the trade-off


between growth and mortality rates mediated by foraging activity. American
Naturalist 142:242-272.

Wemer, E. E., and M. A. McPeek. 1994. Direct and indirect effects of predators on two
anuran species along an environmental gradient. Ecology 75:1368-1382.

Wemer, J. K.. and M. B. McCune. 1979. Seasonal changes in anuran populations in a


northem Michigan pond. Journal ofHerpetology 13:101-104.

Westfall, P. H., and S. S. Young. 1993. Resampling-based multiple testing: examples


and methods for p-value adjustment. John Wiley and Sons, New York, New
York, USA.

Westfall, P. H., R. D. Tobias, D. Rom, R. D. Wolfinger, and Y. Hochberg. 1999.


Multiple comparisons and multiple tests using the SAS® system. SAS Institute,
Gary, North Carolina, USA.

Wiens, J. A. 1977. On competition and variable environments. American Scientist


65:590-597.

Wiens, J. A. 1997- Metapopulation dynamics and landscape ecology. Pages 43-68 in 1.


A. Hanski and M. E. Gilpin, editors. Metapopulation biology: ecology, genetics,
and evolution. Academic, New York, New York, USA.

Wiens, J. A., N. C. Stenseth, B. Van Home, and R. A. Ims. 1993. Ecological


mechanisms and landscape ecology. Oikos 66:369-380.

Wiens, J. A., R. L. Schooley, and R. D. Weeks, Jr. 1997. Patchy landscapes and animal
movements: do beetles percolate? Oikos 78:257-264.

Wiest, J. A., Jr. 1982. Anuran succession at temporary ponds in a post oak-savanna
region of Texas. U. S. Fish and Wildlife Service Research Report 13:39-47.

164
Whitaker, J. O., Jr., D. Rubin, and J. R. Munsee. 1977. Observafions on food habits of
four species of spadefoot toads, genus Scaphiopus. Herpetologica 33:468^75.

Wilbur, H. M. 1977. Interactions of food level and population density in Rana sylvatica.
Ecology 5S:206-209.

Wilbur, H. M. 1980. Complex life cycles. Annual Review of Ecology and Systematics
11:67-93.

Wilbur, H. M. 1984. Complex life cycles and community organization in amphibians.


Pages 195-224 in P. W. Price, C. N. Slobodchikoff, and W. S. Gaud, editors. A
new ecology: novel approaches to interactive systems. John Wiley and Sons.
New York, New York, USA.

Wilbur, H. M. 1987. Regulation of stmcture in complex systems: experimental


temporary pond communities. Ecology 68:1437-1452.

Wilbur, H. M. 1997. Experimental ecology of food webs: complex systems in temporary


ponds. Ecology 7S:2279-2302.

Wilbur, H. M., and J. P. Collins. 1973. Ecological aspects of amphibian metamorphosis.


Science 182:1305-1314.

With, K. A. 1994. Ontogenetic shifts in how grasshoppers interact with landscape


structure: an analysis of movement pattems. Functional Ecology 8:477-485.

With, K. A., and T. O. Crist. 1995. Critical thresholds in species' responses to landscape
structure. Ecology 76:2446-2459.

Woodward, B. D. 1982. Tadpole competition in a desert anuran community. Oecologia


54:96-100.

Woodward, B. D. 1983. Predator-prey interactions and breeding-pond use of temporary-


pond species in a desert anuran community. Ecology 64:1549-1555.

Wright, D. H., B. D. Patterson, G. M. Mikkelson, A. Cufier, and W. Atmar. 1998. A


comparative analysis of nested subset pattems of species composition. Oecologia
113:1-20.

Yanes, M., J. M. Velasco, and F. Suarez. 1995. Permeability of roads and railways to
vertebrates: the importance of culverts. Biological Conservation 71:217-222.

165
Yeargers, E. K., R. W. Shonkwiler, and J. V. Herod. 1996. An introduction to the
mathematics of biology. Birkhauser, Boston. Massachusetts, USA.

Zar, J. H. 1984. Biostatistical analysis. Second edition. Prentice-Hall, Englewood


Cliffs, New Jersey, USA.

166
APPENDIX A

PERCENT COVERAGE OF COVER TYPES IN LANDSCAPES

SURROUNDING PLAYA WETLANDS

167
Table A.l. Percent coverage of cover types in landscapes^ surrounding 16 playa wetlands
between landuses on the Southem High Plains, Texas, 1999 and 2000.
Landuse"
Cover t> pe*^ Cropland Grassland
X SE X oh
Alfalfa Medicago sativa 0.12 0.08 0.02 0.01
Barley Hordeum spp. 0 0 0.04 0.04
Com Zea mays 2.11 1.31 1.07 0.47
Cotton Gossypium hirsutum 60.3 5.37 2.75 1.27
Ha> Phalaris arundinacea 0.43 0.19 1.17 0.67
Kleingrass Panicum coloratum 0.45 0.19 0.34 0.17
Millet Pennisetum glaucum 0.15 0.15 0 0
Oats Arena sativa 0.21 0.11 0.09 0.08
Rye Secale cereale 0.17 0.17 0 0
Sorghum Sorghum vulgare 4.78 1.39 3.21 0.64
Soybean Glycine max 1.14 0.81 0 0
Sunflower Helianthus annuus 0.79 0.24 0 0
Vegetables (various spp.) 0.04 0.02 0 0
Wheat Triticum aestivum 6.95 2.28 10.9 2.96
Fallow 2.55 0.66 1.65 0.52
Buildings 0.96 0.24 0.38 0.12
Maintained county roads 1.67 0.09 1.05 0.09
Interstate 0.32 0.32 0 0
Other anthropogenic structures 0.02 0.02 0.01 0.01
Canyon 0.96 0.96 0 0
Native grass 3.19 0.67 28.8 4.34
Replanted grass 8.47 1.29 46.1 6.78
Playa wetland 3.86 0.32 2.35 0.23
Water 0.29 0.14 0.02 0.02
^Landscapes were 2830-ha circular plots (i.e., 3-km radius) with their origins positioned
at the center of each study playa.
''Percent cover of cotton, county roads, and native and replanted grass differed (P<0.006)
between landuses by Wilcoxon 2-sample nonparametric tests (Conover 1980:216); no test
was performed between landuses when x and SE=0; « = 8 playas/landuse.
'^Other anthropogenic stmctures included cemeteries, quarries, and airstrips; replanted
grass was cultivated previously then revegetated by the Conservation Reserve Program.

168
APPENDIX B

ANOVA TABLES FOR CHAPTER III

169
Table B.l. Statistics for ANOVAs testing differences in relative daily abundance and
diversity of amphibians between landuse t>pes (cropland \s. grassland) and years and
among species at 16 playa wetlands on the Southem High Plains, Texas. May-October
1999 and April-August 2000.
Variable Tests and Effects^'' df^ F P
Abundance 3-wa> ANOVA
Landuse 1 11.97 <0.001
Year 1 10.54 0.002
Species 6 55.67 <0.001
Landuse x Year 1 0.80 0.375
Landuse x Species 6 8.36 <0.001
Year >: Species 6 2.52 0.051
Landuse x Year x Species 6 1.40 0.223
2-wa> ANOVA
NSF
Landuse 20.33 <0.001
Year 2.63 0.131
Landuse x Year 2.34 0.152
BTS
Landuse 1.26 0.283
Year 8.08 0.015
Landuse x Year 0.02 0.895
GPT
Landuse 0.01 0.913
Year 2.35 0.152
Landuse x Year 0.11 0.747
PSF
Landuse 7.94 0.016
Year 0.51 0489
Landuse x Year 3.30 0.094
SCF
Landuse 0.23 0.639
Year 1.99 0.184
Landuse x Year 0.11 0.751
PLF
Landuse 0.156 0.156
Year 0.289 0.289
Landuse x Year 0.151 0.151
WHT
Landuse 0.33 0.576
Year 1.00 0.338
Landuse x Year 0.97 0.345

170
Table B.l. Continued.
Variable Tests and Effects*^^ df F
Abundance 1-way ANOVA
Cropland
Species 6 34.20 <0.001
Grassland
Species 6 18.74 <0.001
1999
Species 6 20.06 <0.001
2000
Species 6 15.56 <0.001
Diversity' 2-way ANOVA
Landuse 1 0.91 0.359
Year 1 0.29 0.602
Landuse x Year 1 1.04 0.329
Overall 3-way ANOVA on abundance was separated b\ species to test for differences
between landuses and years (i.e., 2-way analyses) then by landuse and year (i.e.. l-wa>
analyses) to test the species effect, because species interacted with landuse and year in
the 3-wa>' analysis.
''NSF = New Mexico spadefoot {Spea multiplicata), BTS = barred tiger salamander
{Ambystoma tigrinum mavortium). GPT = Great Plains toad {Bufo cognatus), PSF =
plains spadefoot {S. bombifrons), SCF = spotted chorus frog {Pseudacris clarkii). PLF =
plains leopard frog {Rana blairi), and WHT = Woodhouse's toad {B. woodhousii).
'^Error (denominator) df for 3-way ANOVA, 2-way ANOVAs. and 1-way ANOVAs was
84 (2 landuses x 2 years x 7 species x 3 playas), 12 (2 landuses x 2 >ears x 3 playas). and
49 (7 species x 7 playas), respectively.
'^Diversity was calculated using the Shannon-Weaver algorithm (Hair 1980:273); no
species effect was included in analyses because all species were used in estimation of the
index.

171
APPENDIX C

AGE CLASS .\NOVA TABLES FOR CHAPTER III

172
Table C.l Stafistics for ANOVAs for species (by age class) that differed significantly in
relati\ e dail>' abundance between landuse types (cropland vs. grassland) or years (see
Table B.l) at 16 playa wetlands on the Southern High Plains, Texas, May-October 1999
and April-August 2000.
Species Tests and Effects df
NSF 2-way ANOVA
Metamorph
Landuse 12 6.79 0.023
Year 12 13.99 0.003
Landuse x Year 10.19 0.008
Subadult
Landuse 12 17.18 0.001
Year 12 50.06 <0.001
Landuse x Year 12 21.26 <0.001
Adult
Landuse 12 14.02 0.002
Year 12 0.10 0.758
Landuse x Year 12 048 0.499
PSF Metamorph
Landuse 12 1.93 0.189
Year 12 12.1 0.005
Landuse x Year 12 0.001 0.966
Subadult
Landuse 12 6.74 0.023
Year 12 11.69 0.005
Landuse x Year 12 8.66 0.012
Adult
Landuse 12 3.71 0.078
Year 12 7.88 0.016
Landuse x Year 12 2.79 0.121
BTS Metamorph
Landuse 12 2.07 0.176
Year 12 10.50 0.007
Landuse x Year 12 0.01 0.909
Adult
Landuse 12 0.22 0.648
Year 12 0.76 0.399
Landuse x Year 12 0.01 0.907
NSF 1-way ANOVA
Metamorph
Landuse in 1999 1,6 17.46 0.006
Landuse in 2000 1,6 0.17 0.698
Year in Cropland 1.6 34.84 0.001
Year in Grassland 1,6 0.11 0.747

173
T a b l e d . Continued.
Species Tests and Effects df F P
NSF 1-way ANOVA
Subadult
Landuse in 1999 1,6 0.36 0.570
Landuse in 2000 1,6 22.55 0.003
Year in Cropland 1,6 43.66 <0.001
Year in Grassland 1,6 6.96 0.039
PSF 1-way ANOVA
Subadult
Landuse in 1999 1,6 1.96 0.211
Landuse in 2000 1,6 7.79 0.032
Year in Cropland 1,6 10.59 0.017
Year in Grassland 1,6 1.27 0.302
"^NSF = New Mexico spadefoot {Spea multiplicata), PSF = plains spadefoot {S.
bombifrons), and BTS = barred tiger salamander {Ambystoma tigrinum mavortium).
Overall 2-way ANOVA on abundance of significant species was separated by landuse
and year for metamorph NSF and subadult NSF and PSF (i.e., 1-way analyses), because
landuse and year main effects interacted.

174
APPENDIX D

CROPLAND AND GRASSLAND LANDSCAPES

175
176
Figure D.l. Anthropogenic and natural cover types in landscapes (i.e., 2,830-ha [3-km
radius] circular plots) surrounding 16 playa wetlands on the Southem High Plains, 1999
and 2000. Each study playa is positioned at the landscape plot origin. Crop types
included alfalfa {Medicago sativa), barley {Hordeum spp.), com {Zea mays), cotton
{Gossypium hirsutum), domestic grass {Panicum coloratum), hay {Phalaris
arundinacea), millet {Pennisetum glaucum), oats {Avena sativa), sorghum {Sorghum
vulgare), soybean {Glycine max), sunflower {Helianthus annuus), various vegetable
species, and wheat {Triticum aestivum). Fallow was arable but unplanted land. Non-
playa water was permanent water resulting from anthropogenic modifications of the
landscape (e.g., tailwater pit, effluent treatment pond). Conservation Reserve Program
(CRP) grasslands were cultivated previously and replanted with native and exotic grasses
for erosion control and wildlife habitat.
Cropland Landscapes, 1999

0 1,000 2,000 4,000 6,000 8,000


I Meters

_ _ _ Alfalfa Sorghum Airstrip

I I Com Soybean Buildings

1111111 Cotton Sunflower Maintained Roads

I I Domestic Grass Vegetables Playa Wetland

Hay Wheat Non-playa Water

i l i i Milet Fallow Native Grassland


f^^m
I Oats %>mi 1-27 CRP Grassland

Figure D. 1. Continued.

177
Grassland Landscapes, 1999
N

0 1,000 2,000 44K)a 6.000 8.000


(Meters
Com Hwy385

Cotton Buildings
^?^
Domestic Grass Maintained Roads

I Hay Playa Wetland

Sorghum Native Grassland

Wheat CRP Grassland

Fallow

Figure D.l. Continued.

178
Cropland Landscapes, 2000

0 1,000 2,000 4J0OO 6,000 8,000


I Meters
^ ^ 1 Vegetables Raya Wetland

n u l l Cotton Norvplaya Water

Kss Domestic Grass p• -• "•• j FaHow Native Grassland

r n n a y I j Cemetery CRP Grassland

J H J H Sorghum ^ H H Buildings V\^rte River Canyon

Sunflower ^ ^ H Maintained Roads

Figure D.l. Continued.

179
Grassland Landscapes, 2000

0 1,000 2,000 4,{KK) 6,000 8,000


[Meters
Alfalfa Sofghum Playa W ^ b n d

Barley \/\*ieat Non-playa Water

Com Fallow Native Grassland

Cotton Buildings CRP Grassland


Domestic Grass Gravel Pit Bumt Grassland

J Hay Maintained Roads

t Oats

Figure D.l. Continued.

180

You might also like